Next Article in Journal
Magnetorheological Elastomer-Based Self-Powered Triboelectric Nanosensor for Monitoring Magnetic Field
Next Article in Special Issue
Photocatalytic Activity of Titanium Dioxide Nanotubes Following Long-Term Aging
Previous Article in Journal
Impedance Spectroscopy Analysis of PbSe Nanostructures Deposited by Aerosol Assisted Chemical Vapor Deposition Approach
Previous Article in Special Issue
Efficient Hydrogen Generation and Total Nitrogen Removal for Urine Treatment in a Neutral Solution Based on a Self-Driving Nano Photoelectrocatalytic System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Efficient Iron Oxide Nanoparticles Immobilized on Cellulose Nanofibril Aerogels for Arsenic Removal from Water

1
Laboratory of Renewable Nanomaterials, School of Forest Resources, University of Maine, 5755 Nutting Hall, Orono, ME 04469, USA
2
Department of Civil, Environmental and Sustainable Engineering, Santa Clara University, 500 El Camino Real, Santa Clara, CA 95053, USA
*
Author to whom correspondence should be addressed.
Nanomaterials 2021, 11(11), 2818; https://doi.org/10.3390/nano11112818
Submission received: 6 October 2021 / Revised: 18 October 2021 / Accepted: 19 October 2021 / Published: 23 October 2021
(This article belongs to the Special Issue Application of Nanomaterials and Nanotechnology in Water Treatment)

Abstract

:
The application and optimal operation of nanoparticle adsorbents in fixed-bed columns or industrial-scale water treatment applications are limited. This limitation is generally due to the tendency of nanoparticles to aggregate, the use of non-sustainable and inefficient polymeric resins as supporting materials in fixed-bed columns, or low adsorption capacity. In this study, magnesium-doped amorphous iron oxide nanoparticles (IONPs) were synthesized and immobilized on the surface of cellulose nanofibrils (CNFs) within a lightweight porous aerogel for arsenic removal from water. The IONPs had a specific surface area of 165 m2 g−1. The IONP-containing CNF aerogels were stable in water and under constant agitation due to the induced crosslinking using an epichlorohydrin crosslinker. The adsorption kinetics showed that both As(III) and As(V) adsorption followed a pseudo second-order kinetic model, and the equilibrium adsorption isotherm was best fitted using the Langmuir model. The maximum adsorption capacities of CNF-IONP aerogel for As(III) and As(V) were 48 and 91 mg As g-IONP−1, respectively. The optimum IONP concentration in the aerogel was 12.5 wt.%, which resulted in a maximum arsenic removal, minimal mass loss, and negligible leaching of iron into water.

Graphical Abstract

1. Introduction

Arsenic is one of the most toxic elements that can be found in groundwater at high concentrations in many locations across the globe [1]. Long-term exposure to arsenic from drinking water is associated with health issues such as cancers, skin lesions, and cardiovascular diseases [2]. Arsenic is a metalloid that has two major oxidation states in an ambient environment: arsenite, As(III), and arsenate, As(V). According to the United States Environmental Protection Agency (US-EPA), the maximum contaminant level (MCL) for arsenic in drinking water is 10 μg L−1.
Arsenic can be removed from water by techniques such as reverse osmosis, oxidation, ion exchange, and coagulation–flocculation [3]. These technologies are efficient but are often costly and are based on non-sustainable materials. Adsorption-based techniques have the potential to effectively treat drinking water if an appropriate and efficient adsorbent is developed [4]. A number of studies have focused on the development of cost-effective and sustainable adsorbents for arsenic removal [5]. Iron oxide-based nanoparticles typically have a high affinity towards arsenic. Crystalline iron oxide nanoparticles such as hematite (α-Fe2O3), magnetite (Fe3O4), or maghemite (γ-Fe2O3) have relatively low specific surface areas (<100 m2 g−1) [4,6]. Amorphous iron oxide nanoparticles (IONPs) tend to have a higher surface area than crystalline nanoparticles; however, their transition to other crystalline phases over time and temperature limits their application [7,8,9]. Doping IONPs with another metal ion could inhibit their phase transition and retain their amorphous nature, hence maintaining the increased surface area [9,10,11]. Another common challenge with using IONPs is their tendency to aggregate, which decreases their specific surface area [12]. This means that the application of nanoparticle adsorbents in fixed-bed columns or at an industrial scale is limited. Researchers have attempted to utilize nanoparticulate aggregates in fixed-bed columns [13,14,15]. In these attempts, the adsorption capacity was often sacrificed because of the increased extent of aggregation.
Cellulose nanomaterials have emerged as promising materials in environmental applications. Cellulose is a linear polysaccharide composed of glucose units connected through β 1–4 glycosidic bonds [16]. Cellulose nanofibrils (CNFs) are generally produced by mechanical refining or other mechanical size reduction methods such as microfluidization, homogenization, or grinding [17]. The diameter of the fibrils is typically 5–50 nm, and the fibrils can be several μm in length [16].
A large number of scientific reports have been published to prove the concept of nanocellulose-based materials for a variety of water and wastewater treatment processes, including the removal of toxic heavy metals and dyes [18]. The ongoing research on nanocellulose-based materials for sustainable water treatment has largely focused on the development of chemically functionalized materials such as TEMPO (2,2,6,6-tetramethylpiperidine-1-oxyl)-oxidized, carboxylated, sulfonated, phosphorylated, and amino-functionalized nanocellulose to enhance the adsorption capacity [19]. The functionalization route remains a challenging task due to the complexity of the process and the low adsorption capacity.
Immobilizing the IONPs on the surface of cellulose fibrils could be a promising solution to tackle the two challenges mentioned above: aggregation of nanoparticles and their limited application in fixed-bed columns. CNFs have the potential to decrease the extent of IONP aggregation owing to their stable aqueous suspension and nanoscale dimensions. In such a case, drying a stable suspension of CNFs and IONPs into a foam or aerogel structure could result in uniformly distributed IONPs within the CNF network with minimal aggregation. A few attempts have been made to incorporate iron oxide nanoparticles into porous CNF structures [20,21]. Although the incorporation of the nano-entities seemed feasible, several issues have been identified. First, the studies lacked adequate characterization of adsorbent stability under wet conditions, which is essential given their intended application in water treatment. Second, information relative to the stability of nanoparticles within the cellulose structures remains to be elucidated. Third, a favorable adsorption capacity was obtained by incorporating a high load of iron oxide nanoparticles within the aerogel (36%) without investigating the structural stability of the adsorbents in water [20].
In the present work, a novel protocol to produce porous CNF structures loaded with uniformly distributed IONPs with minimal aggregation and considerably high specific surface area was studied. The IONPs were immobilized on CNFs, and the adsorbents were tested for their adsorption performance towards As(III) and As(V). In addition, extensive physical and chemical characterizations of the CNF-IONP aerogels were conducted, as well as an assessment of their stability under various IONP loadings.

2. Materials and Methods

2.1. Materials

CNFs were provided by the Process Development Center (PDC) of the University of Maine (Orono, ME, USA). The 3 wt.% CNFs were produced by mechanically refining bleached softwood Kraft pulp. This is the standard grade of CNFs produced by the PDC and contains 90% fines. In this context, the fines content indicates the percentage of fibers with lengths < 200 µm. PolycupTM (polyamide-epichlorohydrin) 5150 crosslinker (26 wt.%) was supplied by Solenis (Wilmington, DE, USA). Anhydrous ferric chloride (FeCl3; 98%) and magnesium chloride (MgCl2; 99%) were purchased from Alfa Aeser (Haverhill, MA, USA). Anhydrous ethyl alcohol (C2H5OH; 99.5%) was obtained from Acros Organics (Jair Lawn, NJ, USA). HEPES (4-(2-hydroxyethyl)-1-piperazineethanesulfonic acid) buffering agent (≥99.5%), sodium chloride (NaCl; ≥99%), sodium hydroxide (NaOH; ≥97%), hydrochloric acid (HCl; 37%), sodium (meta) arsenite (NaAsO2), and sodium arsenate dibasic heptahydrate (Na2HAsO4.7H2O; ≥98%) were purchased from Sigma-Aldrich (St. Louis, MO, USA). All chemicals and solvents were used without any further purification.

2.2. Synthesis of Mg-Doped Amorphous IONPs

To synthesize the Mg-doped amorphous IONPs, 0.09 M FeCl3 and 0.01 M MgCl2 solutions were prepared by dissolving 1.02 g FeCl3 and 0.066 g MgCl2 salts together in 70 mL of ethanol [22]. A 20 mL 2.3 M ethanolic NaOH solution was prepared by dissolving 1.84 g solid NaOH. Ten mL of 2.3 M ethanolic NaOH solution was added to the salt mixture, which resulted in a red yellowish precipitate of the amorphous Mg-doped ferric hydroxide, (Fe, Mg)x(OH)y. The solution was continuously ultrasonicated by a Branson 450 sonicator (Branson Ultrasonics Corporation; Radnor, PA, USA) for 1 h to break up the aggregates. The reaction mixture was then transferred to a 100 mL Teflon-lined autoclave container and placed in a preheated oven at 150 °C for 2 h. During the heating process, simultaneous nucleation and homogeneous heating further decreased the particle size [9]. Subsequently, the autoclave was cooled to room temperature. The IONP suspension was washed with distilled water 2–3 times using a centrifuge to remove ethanol and other dissolved components until pH 7.0 was reached. Finally, the volume of the IONP suspension was adjusted to 50 mL by adding distilled water, and the suspension was stored at 2–5 °C.

2.3. Preparation of CNF-IONP Adsorbent

Initially, a batch of 1 wt.% CNF suspension was mixed with PolycupTM crosslinker (5 wt.% of the total dry mass of CNFs) and stirred with a magnetic stirrer at 300 rpm for 5 min to ensure a homogeneous distribution of the crosslinker. CNF aerogels containing IONP content of 12.5 wt.% were prepared by mixing IONP and CNF suspensions according to the amounts listed in Table S1. The suspension mixture was mechanically stirred with a magnetic stirrer at 300 rpm followed by continuous ultrasonication for 5 min. Finally, the CNF-IONP suspension mixture was poured into a cylindrical plastic mold (height: 2.5 cm and diameter: 1.3 cm) and freeze-dried using a Harvest Right freeze-dryer (North Salt Lake, UT, USA). The temperature cycles of the freeze dryer were −34.4, −6.7, 4.4, 15.6, and 32.2 °C for 8, 10, 8, 3, and 3 h, respectively. A schematic illustration that describes the synthesis process is presented in Figure S1. After freeze-drying, the CNF-IONP aerogel adsorbents were heated in a vacuum oven (25 mm Hg = 86 kPa) at 105 ± 2 °C to induce crosslinking. The proposed reaction mechanism is presented in Figure S2.

2.4. Characterization

A Panalytical X’Pert PRO X-ray diffractometer (Royston, UK) was used to assess the nature of the nanoparticles and the crystalline structure of the CNF aerogels. The XRD anode material was Cu with Kα at a wavelength of 1.54 nm. The generator voltage and current were 40 kV and 40 mA, respectively. The scan step size and 2θ range were 0.05° s−1 and 10–80°, respectively. The baseline correction, smoothing, and background subtraction for the XRD data were performed using Origin Pro 2021 software.
The surface morphology of CNF aerogels before and after IONP immobilization was assessed using a Zeiss Nvision 40 scanning electron microscope (SEM; Oberkochen, Germany). The aerogel samples of a thickness of 2–4 mm were prepared by slicing through the middle section using a sharp blade. Prior to imaging, all the samples were sputter-coated by a thin layer of gold–palladium and scanned at an accelerating voltage of 3 kV.
Energy-dispersive X-ray spectroscopy (EDS; iXRF model 550i AMRay 1820, Bedford, MA, USA) was used for the elemental mapping of Fe, Mg, C, and O atoms at the CNF-IONP surface. An accelerating voltage of 20 kV was maintained for the EDS analysis.
The specific surface areas (SSA) of freeze-dried CNF aerogel and IONP samples were measured by the Brunauer, Emmett, and Teller (BET) nitrogen adsorption method using an ASAP 2020 instrument (Micromeritics; Norcross, GA, USA). The CNF aerogel and IONP samples were degassed in a vacuum for 5 h at 75 and 130 °C, respectively.
The ATR-FTIR analysis was performed using a PerkinElmer Spectrum Two™ FTIR spectrophotometer (Shelton, CT, USA) to evaluate the nature of the interaction between the CNFs and IONPs in the aerogel. The data obtained for CNF and CNF-IONP aerogels were normalized with respect to the wavenumber 1055 cm−1, which represents the stretching vibration of the cellulose backbone (not altered by the crosslinking reaction).
A Nano ZS90 Zetasizer (Malvern, UK) was used to measure the zeta potentials of CNFs and IONPs in water. The isoelectric points (IEPs) were determined by measuring the zeta potentials at pH 3–11. Relatively low concentrations (~0.1 wt.%) of CNF and IONP suspensions were used with a constant ionic strength of 0.01 M NaCl for all the zeta potential measurements.
The density (ρ, g cm−3) and porosity (%) of the aerogels were calculated by measuring the void volume (v1) with an Accupyc II gas pycnometry system (Norcross, GA, USA), the total volume (v2) using a digital caliper, and corresponding mass (m) according to Equations (1) and (2), respectively.
Density   of   the   aerogel ,   ρ = m v 2
Porosity   ( % ) = ( 1 v 1 v 2 ) × 100
The shape recovery tests were conducted by compressing the aerogels at a pressure of 1.2 kPa using a Dake® manual hydraulic pump. The aerogel was compressed into a 5 mm thin disk and submerged in 80 mL distilled water for 10 s. The aerogel heights before compression (hi) and after submerging in water (hf) were recorded. The shape recovery was calculated using Equation (3).
Shape   recovery   ( % ) = h f h i × 100
The water absorption capacity of the aerogels was determined by calculating the mass differences of aerogels before (Wi) and after (Wf) soaking in 80 mL distilled water for 12 h according to Equation (4).
Water   absorption   capacity   ( g water   g dry   mass 1 ) = ( W f W i ) W i
The mass losses of the CNF and CNF-IONP aerogels were determined gravimetrically after submerging the aerogels into 80 mL distilled water for 12 h with constant agitation and drying in an oven at 75 °C for 5 h. The % of mass loss was calculated by comparing the initial (m1) and after drying (m2) dry mass according to Equation (5).
Mass   loss   ( % ) = ( m 1 m 2 m 1 ) × 100

2.5. Arsenic Adsorption Experiments

All batch arsenic adsorption experiments were carried out in 100 mL Falcon tubes filled up to 80 mL at room temperature (~25 °C) under constant agitation (VWR Scientific Products rocking platform model 100, Radnor, PA, USA). For the kinetic experiments, 3 mL of solution was removed at specific time intervals. Samples were acidified immediately after collection by 1% v/v of concentrated HNO3, and the arsenic and iron concentrations were measured by a Thermo Scientific™ Element 2™ ICP-MS (Waltham, MA, USA) calibrated using SLRS-6, a river certified reference material from National Research Council of Canada. The initial arsenic concentration for adsorption kinetic studies was ~300 µg L−1 for both As(III) and As(V). For the kinetic experiments involving both As species, the initial IONP concentrations were 16, 31, and 63 mg L−1. Samples were collected for analysis at different time intervals up to 12 h. For the equilibrium adsorption experiments, the initial arsenic concentration ranged from 0.055 to 15.89 mg L−1 for As(III), and 0.073 to 21.65 mg L−1 for As(V). All arsenic solutions were prepared in a 0.01 M HEPES buffer with 0.05 M NaCl. The pH remained constant at 7.0–7.2 throughout all experiments. A set of equilibrium adsorption experiments were conducted at different pH values and at an initial As(V) and As(III) concentration of 1 mg L−1 with a dosage of 63 mg of IONP L−1. The role of IONP concentration in the CNF-IONP aerogel was investigated by varying the wt.% of IONPs between 1% and 30% and adjusting the CNF concentration accordingly. A concentration of 7.2 mg L−1 of As(V) solution was used to perform the analysis with 12 h of contact time under constant agitation.

3. Results and Discussion

3.1. Arsenic Adsorption Experiments

Figure 1 shows the adsorption kinetics of As(III) and As(V) for three CNF-IONP sorbent concentrations. Given that CNFs alone did not remove significant As(III) and As(V) concentrations, the reported sorbent dosages are those of IONPs. As(III) and As(V) concentrations decreased rapidly in the first 1–2 h depending on the sorbent dosage due to the increased number of available IONP surface sites. For both arsenic species, near-equilibrium adsorption was reached within ~3 h. IONP dosages of 63 mg L−1 resulted in the removal of As(III) and As(V) to <10 µg L−1 after 12 h of contact time.
The adsorption kinetic data were best represented by a pseudo second-order rate equation,
t q t = t q e + 1 K ad   q e 2  
where qe and qt are the adsorbed arsenic (mg-As g-IONP−1) at equilibrium and at time t, respectively, and Kad is the pseudo second-order rate constant (g-IONP mg-As.min−1). The Kad values for As(III) and As(V) uptake increased with increasing adsorbent dosage (Table 1), indicating a faster uptake rate with a higher adsorbent concentration.
The equilibrium adsorption of As(III) and As(V) onto CNF-IONP aerogels at pH 7 was investigated (Figure 2). Experimental data for both arsenic species were well represented by the Langmuir isotherm model (Equation (7)), which is applicable to a monolayer adsorption behavior assuming a uniform surface adsorption energy [23].
q e = q max K L C e 1 + K L C e
where qe is the mass of As(III) or As(V) adsorbed at equilibrium per mass of IONPs (mg-As g-IONP−1), qmax is the maximum adsorption capacity (mg-As g-IONP−1), Ce is the equilibrium As(III) or As(V) concentration (mg L−1) in post-adsorption water samples, and KL is the Langmuir constant for adsorption. The maximum adsorption capacities of the CNF-IONP aerogel adsorbent were 47.75 and 90.90 mg-As g-IONP−1 for As(III) and As(V), respectively.
The constant KL in Equation (7) indicates the sorption affinity of a solute toward a sorbent. Our results show that As(V) has a significantly higher affinity toward the IONPs than As(III) (Table 2). The difference in affinity between the two arsenic species is especially evident at low aqueous arsenic concentrations that are representative of natural waters, indicating that As(III) is more mobile in the environment [24]. This may be attributed to the largely uncharged form of As(III) at pH 7 (H3AsO3 ⇔ H+ + H2AsO3, pK = 9.15) compared to As(V) that is negatively charged at the same pH (H3AsO4 ⇔ H+ + H2AsO4, pK = 2.3; H2AsO4 ⇔ H+ + HAsO42−, pK = 7.16). The negatively charged As(V) species interact more favorably with the positively charged surface of IONPs, resulting in a higher adsorption affinity and capacity (Table 2).
Findings from previous studies shown in Table 3 highlight the importance of functionalizing the micro- and nano-sized cellulosic materials to result in an acceptable affinity towards arsenic. The table also shows the superior performance of our adsorbent to other iron oxide-based nanoparticles used for arsenic removal. However, comparisons with iron oxide nanoparticles incorporated into cellulosic-based structures must be interpreted with caution. Of particular interest is the work of Yu et al. and Dong et al. [20,21], whose studies investigated iron oxide-based nanoparticles in cellulosic structures. The maximum adsorption capacity obtained herein is comparable to that of Yu et al. However, it is worth noting that the Fe2O3 content used in their study was 36%, which is approximately three times that of the IONP content used in this study (12.5%). The maximum adsorption capacity obtained by our adsorbent is considerably higher than that obtained by Dong et al. which could be attributed to the lower iron oxide content in their work (9.4%).
The pH-dependent dissolved arsenic speciation affects its adsorption behavior. Figure 3A shows that a high removal rate (>93%) was achieved for both arsenic species within an acidic to neutral pH (3–7), but drastically dropped after pH 7. Given the circumneutral pH of drinking water, it can be expected that the CNF-IONP adsorbent can be used effectively to remove arsenic. Figure 3B shows the measured zeta potentials of CNFs and IONPs. The CNF surface is negatively charged within the pH range 3–11 due to the presence of carboxylate groups on the surface of CNFs [36]. The IONPs, however, have a low isoelectric point (IEP) of 4.6 compared to the 7–9 range reported for synthetic iron oxides [37]. This is due to the positive charge deficit brought about by the presence of divalent Mg2+ ions in the iron oxide structure [38]. The pH-dependent adsorption of arsenic species onto the IONPs may be explained by the opposite electrostatic charges between arsenic and the IONP surface at low pH values that favor adsorption [39] and electrostatic repulsion between the deprotonated surface groups of IONPs and the anionic arsenic species at higher pH values that result in diminished adsorption [40].
The experiments to determine the effect of the IONP loading in the aerogel on adsorption were conducted using As(V). Figure 4A shows that the percent of arsenic removal was enhanced with the increase in IONPs up to 12.5 wt.%, above which, increasing the IONP concentration resulted in a decrease in arsenic removal. Note that the results presented earlier were obtained with 12.5 wt.% IONPs. A one-way ANOVA statistical test showed differences among the five groups of different IONP concentrations in CNF-IONP aerogel. Post hoc analysis indicated that the 12.5 wt.% sample exhibited maximum arsenic removal. The decrease in arsenic removal with an increase in IONP loading suggests that higher IONP loadings may result in densely packed IONPs in the aerogel, resulting in a decreased surface area for adsorption. High-resolution SEM images of CNF-IONP aerogels at 12.5 and 25 wt.% IONP loadings (Figure 4B,C) show that an increasing IONP concentration can lead to particle aggregation at the surface. A similar observation was reported in a study where increasing the magnetite concentration beyond a threshold value decreased Cr(VI) adsorption, and the observation was attributed to the aggregation of nanoparticles and the reduction in the availability of active adsorption sites [41].
The stability of IONPs in the aerogel was investigated through the measurement of leached iron upon soaking the aerogels in water under constant agitation. For the 12.5 wt.% IONPs in the CNF-IONP aerogel, an insignificant amount (<10 µg L−1) of iron was leached at pH 7, which is below the WHO and US-EPA MCL of 300 µg L−1 for iron concentration in drinking water.

3.2. Characterization of the Adsorbent

Figure 5 shows the XRD patterns for freeze-dried IONP, CNF-IONP, and CNF aerogels. The XRD pattern of freeze-dried IONPs showed no peaks, indicating the amorphous nature of the nanoparticles. The XRD peaks of both CNF and CNF-IONP aerogels were similar. The broad peak between 15° and 16.5° corresponds to Miller indices (1–10) and (110), respectively, and the sharp peak around 22.5° corresponds to the lattice diffraction of the (200) plane [42]. The low-intensity peak around 35° corresponds to the plane (004) of cellulose I [43]. A noticeable decrease was observed in peak intensities for cellulose in CNF-IONP aerogel resulting from the incorporation of the IONPs.
Additional SEM analysis was conducted to evaluate the internal structure and morphology of IONP, CNF, and CNF-IONP aerogels (Figure 6). Figure 6A shows a photograph of the CNF-IONP aerogel. The IONP suspension precipitated in the absence of CNFs due to their tendency to aggregate (Figure 6B). However, in the presence of CNFs, the IONPs did not precipitate. Instead, the IONPs were immobilized on the cellulose fibrils (Figure 6B). Without IONPs, CNFs form a highly porous structure resulting from freeze-drying of CNF suspensions (Figure 6C). The role of CNFs in immobilizing the IONPs on their surfaces was further verified by observing the IONPs in the presence and absence of CNFs using SEM. Figure 6D shows the freeze-dried IONP aggregates. The stable CNF suspension immobilized the IONPs on the fibrils, resulting in uniformly distributed nanoparticles within the aerogel (Figure 6E). This finding was further supported by EDS analysis (Figure 6F) of the aerogel, which verified the uniform distribution of the IONPs.
The porosity values of CNF and CNF-IONP aerogels were 98% and 95%, and their densities were 0.009 ± 0.01 and 0.013 ± 0.03 g cm3, respectively. The water absorption capacities for the CNF and CNF-IONP aerogels were 96.5 ± 4.9 and 74.9 ± 2.6 g g−1, respectively. These results are in agreement with previously reported CNF aerogels of similar densities [44].
The freeze-dried IONPs exhibited a considerably high specific surface area of 165 m2 g−1. The BET surface area of CNF aerogel without additives is 13 m2 g−1, which is typical for this type of material [45]. The minimal aggregation and uniform distribution of the IONPs in the aerogel suggest that the surface area of the IONPs was not compromised. The surface area of the amorphous IONPs in this study is also higher than those of crystalline iron oxides such as hematite and maghemite [4,6] and comparable to other doped iron oxide nanoparticle adsorbents [4].
The structural stability of CNF-IONP aerogels when soaked in water and under constant agitation was assessed through mass loss measurements. The results showed that the <3% mass loss of aerogels was not statistically different (p > 0.05) under various IONP loadings (Figure 7). This result shows that regardless of the loading of IONPs, they maintain their contact with the fibrils, which renders the aerogel stable even with constant agitation for 12 h. We further supported this result by assessing the leaching of iron under the same conditions. The iron concentration in water after soaking the aerogels increased by <10 μg L−1 at pH 7. Although the leaching of iron is minimal, this result revealed the importance of rinsing the aerogels with water prior to their intended application.
In addition, the aerogels exhibited an interesting shape recovery pattern resulting from the addition of the epichlorohydrin crosslinker. The key role of the crosslinker is to impart adequate stability upon soaking the aerogels in water. Typically, CNF-based aerogels without a crosslinker tend to disintegrate in water. Here, when the CNF-IONP dry aerogel is compressed into a thin disk, at least 50% of the shape is recovered upon soaking in water. The shape recovery percentage ranged from 50% to 80% depending on the quantity of IONPs (Figure 7); a higher loading of IONPs decreased the shape recovery. Typically, in crosslinked CNF-only aerogels, this phenomenon is attributed to the high surface tension of water, which creates large capillary forces that expand the fibrils. Additionally, the diffusion of water into the pores coupled with partial swelling of amorphous cellulose augment the extent of recovery [44,45,46]. It is not surprising that the presence of the IONPs decreased the shape recovery percentage, which is presumably due to the partial disruption of the CNF network, which renders the recovery more challenging.
To investigate the interaction between CNFs and IONPs, FTIR spectroscopy was performed (Figure 8). In freeze-dried IONP sample, the peak at ~590 cm−1 is attributed to Fe–O band vibration [47,48]. The broad peaks at 3200–3500 and ~1640 cm−1 were assigned to the stretching and bending vibrations of the OH group, respectively [49]. On the other hand, CNF and CNF-IONP aerogels exhibited almost similar IR spectra. The broad peak around 3380 cm−1 corresponds to OH stretching of CNFs [50]. The N–H stretching, which originated from the crosslinking reaction [51], overlaps with the OH stretching peak. Figure S2 shows the structure of the crosslinker and the proposed crosslinking reaction. The peak at 2905 cm−1 is attributed to the aliphatic C–H stretching [52]. The peaks at 1635 and 1550 cm−1 correspond to amide I and II, respectively [53]. In addition, the broad peak around 1250 cm−1 may be assigned to C–O stretching vibration for ester bonds due to the crosslinking [53]. The intensity of this peak is lower in non-crosslinked CNF samples compared to the crosslinked samples (Figure S3).
Although the CNF-IONP aerogels showed excellent stability in water aided by the crosslinker, no characteristic peak at ~1735 cm−1, which is attributed to C=O stretching vibration for the ester group [53], was observed. However, this peak was visible in the crosslinked CNF aerogels, which can be disrupted by the presence of IONPs. The FTIR spectra of the CNF aerogels before and after crosslinking are presented in Figure S3. Besides this, no noticeable change was observed in the FTIR spectra of CNF-IONP aerogel compared to CNF aerogel, which indicates that there is no evidence of chemical reaction between CNFs and IONPs. It is somewhat surprising that despite the absence of evidence for a chemical reaction, IONPs are retained within the CNF structure remarkably well. This suggests that, in addition to Van der Waals interactions between CNFs and IONPs, other factors may be contributing to the stability of IONPs within the aerogel such as mechanical interlocking [54]. It also suggests that the attraction forces between CNFs and IONPs are stronger than those between the IONPs.

4. Conclusions

A novel and highly efficient CNF-IONP aerogel was prepared by incorporating IONPs into CNFs by simultaneous freeze-drying. This method addresses the previous issues of nanoparticle aggregation to enable the application of nanoparticle adsorbents in fixed-bed columns at the industrial scale. The amorphous nature of the IONPs was verified by means of XRD analysis. The IONPs had a remarkably high specific surface area (165 m2 g−1). The 12.5 wt.% IONPs in the CNF aerogel resulted in a maximum arsenic removal and uniform distribution on the cellulose fibrils. The maximum adsorption capacities of CNF-IONP aerogel for As(III) and As(V) were 48 and 91 mg As g-IONP1, respectively. The adsorbent was stable in water and under constant agitation with negligible mass loss. In addition, the adsorbent exhibited a partial shape recovery functionality that could be advantageous for transportation purposes. Overall, this study will also provide a near-term alternative to efficient and bio-based commercial adsorbents for arsenic removal from water.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/nano11112818/s1, Figure S1: The summary of preparatory scheme for CNF-IONP aerogel; Table S1: Summary of aerogel formulation before and after the incorporation of IONPs; Figure S2: Proposed reaction mechanism between cellulose and PolycupTM; Figure S3: The FTIR spectrum of crosslinked and non-crosslinked CNF aerogel.

Author Contributions

Conceptualization, I.H., M.T. and A.A.; Methodology, M.M.R., I.H., M.T., and A.A.; Validation. M.M.R.; Formal Analysis, M.M.R.; Investigation, M.M.R.; Resources, I.H., M.T., and A.A.; Writing—Original Draft Preparation, M.M.R.; Writing—Review and Editing, M.M.R., I.H., M.T., and A.A.; Supervision, I.H. and M.T.; Project Administration, I.H.; Funding Acquisition, I.H. All authors have read and agreed to the published version of the manuscript.

Funding

United States Department of Agriculture (USDA) Forest Services. USDA National Institute of Food and Agriculture, McIntire-Stennis. Project number ME042120 through the Maine Agricultural & Forest Experiment Station. Maine Agricultural and Forest Experiment Publication Number 3860.

Acknowledgments

This work is financially supported by United States Department of Agriculture (USDA) Forest Services. This project was also supported by the USDA National Institute of Food and Agriculture, McIntire-Stennis Project number ME042120 through the Maine Agricultural & Forest Experiment Station. Maine Agricultural and Forest Experiment Publication Number 3860. The authors would like to acknowledge Brian Frederick and George Bernhardt for their help with BET and XRD analyses. The authors would also like to thank Kelly Edwards, Michael Handley, and Emma Perry for their help and training on SEM and ICP-MS.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Shankar, S.; Shanker, U. Shikha Arsenic Contamination of Groundwater: A Review of Sources, Prevalence, Health Risks, and Strategies for Mitigation. Sci. World J. 2014, 2014, 304524. [Google Scholar] [CrossRef] [PubMed]
  2. Ratnaike, R.N. Acute and chronic arsenic toxicity. Postgrad. Med. J. 2003, 79, 391–396. [Google Scholar] [CrossRef] [PubMed]
  3. Nicomel, N.R.; Leus, K.; Folens, K.; van der Voort, P.; Du Laing, G. Technologies for Arsenic Removal from Water: Current Status and Future Perspectives. Int. J. Environ. Res. Public Health 2015, 13, 62. [Google Scholar] [CrossRef]
  4. Hao, L.; Liu, M.; Wang, N.; Li, G. A critical review on arsenic removal from water using iron-based adsorbents. RSC Adv. 2018, 8, 39545–39560. [Google Scholar] [CrossRef]
  5. Mohan, D.; Pittman, C.U. Arsenic removal from water/wastewater using adsorbents—A critical review. J. Hazard. Mater. 2007, 142, 1–53. [Google Scholar] [CrossRef]
  6. Macera, L.; Taglieri, G.; Daniele, V.; Passacantando, M.; D’Orazio, F. Nano-Sized Fe(III) Oxide Particles Starting from an Innovative and Eco-Friendly Synthesis Method. Nanomaterials 2020, 10, 323. [Google Scholar] [CrossRef] [Green Version]
  7. Das, S.; Hendry, M.J.; Essilfie-Dughan, J. Transformation of Two-Line Ferrihydrite to Goethite and Hematite as a Function of pH and Temperature. Environ. Sci. Technol. 2011, 45, 268–275. [Google Scholar] [CrossRef]
  8. Liao, X.; Zhu, J.; Zhong, W.; Chen, H.-Y. Synthesis of amorphous Fe2O3 nanoparticles by microwave irradiation. Mater. Lett. 2001, 50, 341–346. [Google Scholar] [CrossRef]
  9. Phu, N.D.; Ngo, D.-T.; Hoang, L.H.; Luong, N.H.; Chau, N.; Hai, N.H. Crystallization process and magnetic properties of amorphous iron oxide nanoparticles. J. Phys. D Appl. Phys. 2011, 44, 345002. [Google Scholar] [CrossRef]
  10. Baltpurvins, K.A.; Burns, R.C.; Lawrance, G.A.; Stuart, A.D. Effect of Ca2+, Mg2+, and Anion Type on the Aging of Iron(III) Hydroxide Precipitates. Environ. Sci. Technol. 1997, 31, 1024–1032. [Google Scholar] [CrossRef]
  11. Iacob, M.; Cazacu, M.; Turta, C.; Doroftei, F.; Botko, M.; Čižmár, E.; Zeleňáková, A.; Feher, A. Amorphous iron–chromium oxide nanoparticles with long-term stability. Mater. Res. Bull. 2015, 65, 163–168. [Google Scholar] [CrossRef]
  12. Bushell, M.; Beauchemin, S.; Kunc, F.; Gardner, D.; Ovens, J.; Toll, F.; Kennedy, D.; Nguyen, K.; Vladisavljevic, D.; Rasmussen, P.E.; et al. Characterization of Commercial Metal Oxide Nanomaterials: Crystalline Phase, Particle Size and Specific Surface Area. Nanomaterials 2020, 10, 1812. [Google Scholar] [CrossRef] [PubMed]
  13. Casentini, B.; Gallo, M.; Baldi, F. Arsenate and arsenite removal from contaminated water by iron oxides nanoparticles formed inside a bacterial exopolysaccharide. J. Environ. Chem. Eng. 2019, 7, 102908. [Google Scholar] [CrossRef] [Green Version]
  14. Ghosh, A.; Chakrabarti, S.; Ghosh, U.C. Fixed-bed column performance of Mn-incorporated iron(III) oxide nanoparticle agglomerates on As(III) removal from the spiked groundwater in lab bench scale. Chem. Eng. J. 2014, 248, 18–26. [Google Scholar] [CrossRef]
  15. Hu, X.; Ding, Z.; Zimmerman, A.; Wang, S.; Gao, B. Batch and column sorption of arsenic onto iron-impregnated biochar synthesized through hydrolysis. Water Res. 2014, 68, 206–216. [Google Scholar] [CrossRef]
  16. Moon, R.J.; Martini, A.; Nairn, J.; Simonsen, J.; Youngblood, J. Cellulose nanomaterials review: Structure, properties and nanocomposites. Chem. Soc. Rev. 2011, 40, 3941–3994. [Google Scholar] [CrossRef]
  17. Yi, T.; Zhao, H.; Mo, Q.; Pan, D.; Liu, Y.; Huang, L.; Xu, H.; Hu, B.; Song, H. From Cellulose to Cellulose Nanofibrils—A Comprehensive Review of the Preparation and Modification of Cellulose Nanofibrils. Materials 2020, 13, 5062. [Google Scholar] [CrossRef] [PubMed]
  18. Voisin, H.; Bergström, L.; Liu, P.; Mathew, A.P. Nanocellulose-Based Materials for Water Purification. Nanomaterials 2017, 7, 57. [Google Scholar] [CrossRef] [PubMed]
  19. Mohammed, N.; Grishkewich, N.; Tam, K.C. Cellulose nanomaterials: Promising sustainable nanomaterials for application in water/wastewater treatment processes. Environ. Sci. Nano 2018, 5, 623–658. [Google Scholar] [CrossRef]
  20. Yu, X.; Tong, S.; Ge, M.; Zuo, J.; Cao, C.; Song, W. One-step synthesis of magnetic composites of cellulose@iron oxide nanoparticles for arsenic removal. J. Mater. Chem. A 2012, 1, 959–965. [Google Scholar] [CrossRef]
  21. Dong, F.; Xu, X.; Shaghaleh, H.; Guo, J.; Guo, L.; Qian, Y.; Liu, H.; Wang, S. Factors influencing the morphology and adsorption performance of cellulose nanocrystal/iron oxide nanorod composites for the removal of arsenic during water treatment. Int. J. Biol. Macromol. 2019, 156, 1418–1424. [Google Scholar] [CrossRef] [PubMed]
  22. Tang, W.; Su, Y.; Li, Q.; Gao, S.; Shang, J.K. Superparamagnetic magnesium ferrite nanoadsorbent for effective arsenic (III, V) removal and easy magnetic separation. Water Res. 2013, 47, 3624–3634. [Google Scholar] [CrossRef] [PubMed]
  23. Stumm, W.; Morgan, J.J. Aquatic Chemistry, 3rd ed.; Wiley-Interscience: New York, NY, USA, 1996. [Google Scholar]
  24. Smedley, P.; Kinniburgh, D. A review of the source, behaviour and distribution of arsenic in natural waters. Appl. Geochem. 2002, 17, 517–568. [Google Scholar] [CrossRef] [Green Version]
  25. Tian, Y.; Wu, M.; Liu, R.; Wang, D.; Lin, X.; Liu, W.; Ma, L.; Li, Y.; Huang, Y. Modified native cellulose fibers—A novel efficient adsorbent for both fluoride and arsenic. J. Hazard. Mater. 2011, 185, 93–100. [Google Scholar] [CrossRef] [PubMed]
  26. Liu, H.; Li, P.; Qiu, F.; Zhang, T.; Xu, J. Controllable preparation of FeOOH/CuO@WBC composite based on water bamboo cellulose applied for enhanced arsenic removal. Food Bioprod. Process. 2020, 123, 177–187. [Google Scholar] [CrossRef]
  27. Singh, K.; Sinha, T.; Srivastava, S. Functionalized nanocrystalline cellulose: Smart biosorbent for decontamination of arsenic. Int. J. Miner. Process. 2015, 139, 51–63. [Google Scholar] [CrossRef]
  28. Najib, N.; Christodoulatos, C. Removal of arsenic using functionalized cellulose nanofibrils from aqueous solutions. J. Hazard. Mater. 2018, 367, 256–266. [Google Scholar] [CrossRef] [PubMed]
  29. Yu, X.; Tong, S.; Ge, M.; Wu, L.; Zuo, J.; Cao, C.; Song, W. Synthesis and characterization of multi-amino-functionalized cellulose for arsenic adsorption. Carbohydr. Polym. 2013, 92, 380–387. [Google Scholar] [CrossRef]
  30. Anirudhan, T.S.; Senan, P.; Suchithra, P.S. Evaluation of Iron(III)-Coordinated Amino-Functionalized Poly(Glycidyl Methacrylate)-Grafted Cellulose for Arsenic(V) Adsorption from Aqueous Solutions. Water Air Soil Pollut. 2011, 220, 101–116. [Google Scholar] [CrossRef]
  31. Anirudhan, T.S.; Divya, L.; Parvathy, J. Arsenic adsorption from contaminated water on Fe(III)-coordinated amino-functionalized poly(glycidylmethacrylate)-grafted TiO2-densified cellulose. J. Chem. Technol. Biotechnol. 2012, 88, 878–886. [Google Scholar] [CrossRef]
  32. Yousif, A.M.; Zaid, O.F.; Ibrahim, I. Fast and selective adsorption of As(V) on prepared modified cellulose containing Cu(II) moieties. Arab. J. Chem. 2016, 9, 607–615. [Google Scholar] [CrossRef]
  33. Zeng, H.; Zhai, L.; Qiao, T.; Yu, Y.; Zhang, J.; Li, D. Efficient removal of As(V) from aqueous media by magnetic nanoparticles prepared with Iron-containing water treatment residuals. Sci. Rep. 2020, 10, 9335. [Google Scholar] [CrossRef]
  34. Joshi, S.; Kumari, A.; Banjara, A.; Sharma, M. Use of iron oxide/activated carbon magnetic composite for adsorptive removal of arsenic from water. Int. J. Adv. Eng. 2019, 1, 9–16. [Google Scholar]
  35. Wen, Z.; Zhang, Y.; Dai, C.; Chen, B.; Guo, S.; Yu, H.; Wu, D. Synthesis of ordered mesoporous iron manganese bimetal oxides for arsenic removal from aqueous solutions. Microporous Mesoporous Mater. 2014, 200, 235–244. [Google Scholar] [CrossRef]
  36. Liu, P.; Sehaqui, H.; Tingaut, P.; Wichser, A.; Oksman, K.; Mathew, A.P. Cellulose and chitin nanomaterials for capturing silver ions (Ag+) from water via surface adsorption. Cellulose 2013, 21, 449–461. [Google Scholar] [CrossRef]
  37. Parks, G.A. The Isoelectric Points of Solid Oxides, Solid Hydroxides, and Aqueous Hydroxo Complex Systems. Chem. Rev. 1965, 65, 177–198. [Google Scholar] [CrossRef]
  38. Mohapatra, M.; Hariprasad, D.; Anand, S.; Mishra, B. Mg-doped nano ferrihydrite—A new adsorbent for fluoride removal from aqueous solutions. Appl. Surf. Sci. 2012, 258, 4228–4236. [Google Scholar] [CrossRef]
  39. Farrell, J.; Chaudhary, B.K. Understanding Arsenate Reaction Kinetics with Ferric Hydroxides. Environ. Sci. Technol. 2013, 47, 8342–8347. [Google Scholar] [CrossRef]
  40. Mondal, P.; Balomajumder, C.; Mohanty, B. A laboratory study for the treatment of arsenic, iron, and manganese bearing ground water using Fe3+ impregnated activated carbon: Effects of shaking time, pH and temperature. J. Hazard. Mater. 2007, 144, 420–426. [Google Scholar] [CrossRef]
  41. Padmavathy, K.; Madhu, G.; Haseena, P. A study on Effects of pH, Adsorbent Dosage, Time, Initial Concentration and Adsorption Isotherm Study for the Removal of Hexavalent Chromium (Cr (VI)) from Wastewater by Magnetite Nanoparticles. Procedia Technol. 2016, 24, 585–594. [Google Scholar] [CrossRef]
  42. French, A.D. Idealized powder diffraction patterns for cellulose polymorphs. Cellulose 2013, 21, 885–896. [Google Scholar] [CrossRef]
  43. Hafez, I.; Amini, E.; Tajvidi, M. The synergy between cellulose nanofibrils and calcium carbonate in a hybrid composite system. Cellulose 2020, 27, 3773–3787. [Google Scholar] [CrossRef]
  44. Zhang, W.; Zhang, Y.; Lu, C.; Deng, Y. Aerogels from crosslinked cellulose nano/micro-fibrils and their fast shape recovery property in water. J. Mater. Chem. 2012, 22, 11642–11650. [Google Scholar] [CrossRef]
  45. González-Ugarte, A.S.; Hafez, I.; Tajvidi, M. Characterization and properties of hybrid foams from nanocellulose and kaolin-microfibrillated cellulose composite. Sci. Rep. 2020, 10, 1–12. [Google Scholar] [CrossRef]
  46. Qing, Y.; Wu, Y.; Cai, Z.; Li, X. Water-Triggered Dimensional Swelling of Cellulose Nanofibril Films: Instant Observation Using Optical Microscope. J. Nanomater. 2013, 2013, 594734. [Google Scholar] [CrossRef] [Green Version]
  47. Battisha, I.; Afify, H.; Ibrahim, M. Synthesis of Fe2O3 concentrations and sintering temperature on FTIR and magnetic susceptibility measured from 4 to 300K of monolith silica gel prepared by sol–gel technique. J. Magn. Magn. Mater. 2006, 306, 211–217. [Google Scholar] [CrossRef]
  48. Lassoued, A.; Lassoued, M.S.; Dkhil, B.; Gadri, A.; Ammar, S. Structural, optical and morphological characterization of Cu-doped α-Fe2O3 nanoparticles synthesized through co-precipitation technique. J. Mol. Struct. 2017, 1148, 276–281. [Google Scholar] [CrossRef]
  49. Darezereshki, E. One-step synthesis of hematite (α-Fe2O3) nano-particles by direct thermal-decomposition of maghemite. Mater. Lett. 2011, 65, 642–645. [Google Scholar] [CrossRef]
  50. Kondo, T. The assignment of IR absorption bands due to free hydroxyl groups in cellulose. Cellulose 1997, 4, 281–292. [Google Scholar] [CrossRef]
  51. Rodriguez, D.M.P. Aminosilane-Functionalized Cellulosic Polymers for Increased Carbon Dioxide Sorption. Master’s Thesis, Georgia Institute of Technology, Atlanta, GA, USA, 2010. [Google Scholar]
  52. Zhang, C.; Su, J.; Zhu, H.; Xiong, J.; Liu, X.; Li, D.; Chen, Y.; Li, Y. The removal of heavy metal ions from aqueous solutions by amine functionalized cellulose pretreated with microwave-H2O2. RSC Adv. 2017, 7, 34182–34191. [Google Scholar] [CrossRef] [Green Version]
  53. Obokata, T.; Isogai, A. The mechanism of wet-strength development of cellulose sheets prepared with polyamideamine-epichlorohydrin (PAE) resin. Colloids Surf. A Physicochem. Eng. Asp. 2007, 302, 525–531. [Google Scholar] [CrossRef]
  54. Zhangab, Y.; Liab, J.; Maab, N.; Mengab, Z.; Suiab, G. Processing cellulose@Fe3O4 into mechanical, magnetic and biodegradable synapse-like material. Compos. Part. B Eng. 2019, 177, 107432. [Google Scholar] [CrossRef]
Figure 1. As(III) and As(V) adsorption kinetics with three different IONP loadings within CNF aerogel. The reported dosages are the IONP concentrations. The fitted aqueous arsenic concentrations were determined from the difference between the initial and adsorbed arsenic concentrations. The inset graphs are the pseudo second-order kinetic fits for the corresponding dosage of adsorbent according to Equation (6).
Figure 1. As(III) and As(V) adsorption kinetics with three different IONP loadings within CNF aerogel. The reported dosages are the IONP concentrations. The fitted aqueous arsenic concentrations were determined from the difference between the initial and adsorbed arsenic concentrations. The inset graphs are the pseudo second-order kinetic fits for the corresponding dosage of adsorbent according to Equation (6).
Nanomaterials 11 02818 g001
Figure 2. Equilibrium adsorption isotherms of As(III) and As(V). The adsorbent dosage was 63 mg of IONPs L−1 of arsenic ion.
Figure 2. Equilibrium adsorption isotherms of As(III) and As(V). The adsorbent dosage was 63 mg of IONPs L−1 of arsenic ion.
Nanomaterials 11 02818 g002
Figure 3. Effect of pH on arsenic removal percentage. Initial arsenic ion concentration: 1 mg L−1, IONP dosage: 63 mg of IONP L−1 (A) and zeta potential measurements of CNF and IONP suspensions at different pH values (B). Error bars represent standard deviation from triplet measurements.
Figure 3. Effect of pH on arsenic removal percentage. Initial arsenic ion concentration: 1 mg L−1, IONP dosage: 63 mg of IONP L−1 (A) and zeta potential measurements of CNF and IONP suspensions at different pH values (B). Error bars represent standard deviation from triplet measurements.
Nanomaterials 11 02818 g003
Figure 4. As(V) adsorption with different wt.% of IONPs in CNF-IONP aerogel. The error bars represent standard deviation from triplet measurements (A); SEM micrographs (scale bar: 1 µm) of 12.5 wt.% (B) and 25 wt.% (C) of IONPs in CNF-IONP aerogel.
Figure 4. As(V) adsorption with different wt.% of IONPs in CNF-IONP aerogel. The error bars represent standard deviation from triplet measurements (A); SEM micrographs (scale bar: 1 µm) of 12.5 wt.% (B) and 25 wt.% (C) of IONPs in CNF-IONP aerogel.
Nanomaterials 11 02818 g004
Figure 5. XRD spectra of the freeze–dried IONP, CNF–IONP, and CNF aerogels.
Figure 5. XRD spectra of the freeze–dried IONP, CNF–IONP, and CNF aerogels.
Nanomaterials 11 02818 g005
Figure 6. Photograph of CNF-IONP aerogel (A); from left to right: CNF suspension of 1 wt.% solids content, aggregated and precipitated IONP suspension, and IONPs immobilized on CNFs (B); SEM image of CNF aerogel macrostructure without IONP (C); SEM image of IONP (D); SEM image of 12.5% CNF-IONP aerogel (E); Corresponding EDS elemental mapping for CNF-IONP (scale bar: same as (C)) (F). Note the different scales.
Figure 6. Photograph of CNF-IONP aerogel (A); from left to right: CNF suspension of 1 wt.% solids content, aggregated and precipitated IONP suspension, and IONPs immobilized on CNFs (B); SEM image of CNF aerogel macrostructure without IONP (C); SEM image of IONP (D); SEM image of 12.5% CNF-IONP aerogel (E); Corresponding EDS elemental mapping for CNF-IONP (scale bar: same as (C)) (F). Note the different scales.
Nanomaterials 11 02818 g006
Figure 7. Shape recovery and mass loss of CNF-IONP aerogels at various IONP loadings. Error bars represent standard deviation from triplet measurements.
Figure 7. Shape recovery and mass loss of CNF-IONP aerogels at various IONP loadings. Error bars represent standard deviation from triplet measurements.
Nanomaterials 11 02818 g007
Figure 8. FTIR spectra for freeze-dried IONP, crosslinked CNF, and CNF-IONP aerogels.
Figure 8. FTIR spectra for freeze-dried IONP, crosslinked CNF, and CNF-IONP aerogels.
Nanomaterials 11 02818 g008
Table 1. Summary of kinetic parameters for As(III) and As(V) adsorption on IONPs in CNF-IONP aerogel.
Table 1. Summary of kinetic parameters for As(III) and As(V) adsorption on IONPs in CNF-IONP aerogel.
Adsorption ParametersAs(III)As(V)
Initial As Concentration (mg-As L−1)0.02960.0302
Adsorbent Loading (mg-IONP L-As−1)163163163163
qe (mg-As g-IONP−1)15.879.574.8520.419.784.78
Kad (g-IONP mg-As·min−1)7.6 × 10−42.1 × 10−31.35 × 10−26.7 × 10−43.2 × 10−31.67 × 10−2
r20.990.990.990.990.990.99
Table 2. Summary of Langmuir isotherm model fits for arsenic adsorption onto the CNF-IONP aerogel.
Table 2. Summary of Langmuir isotherm model fits for arsenic adsorption onto the CNF-IONP aerogel.
As(III)As(V)
qmax (mg-As g-IONP1)47.7590.90
KL2.367.95
r20.990.99
Table 3. Arsenic adsorption capacities of cellulose- and iron oxide-based adsorbents from the literature.
Table 3. Arsenic adsorption capacities of cellulose- and iron oxide-based adsorbents from the literature.
AdsorbentAs Concentration Range, C0 (mg L−1)Maximum as Adsorption Capacity, qmax (mg g−1)BET Surface Area (m2 g−1)pHReference
Cellulose-g-PDMAEMAAs(III) & As(V): 0.05–8.9As(III): 8.96; As(V): 27.93<10[25]
FeOOH/CuO@WBCAs(III): 20–200As(III): 76.13.5[26]
DETA-g-DA-NCCAs(III) & As(V): 0.005–50As(III): 10.56; As(V): 12.067.5[27]
Functionalized CNFsAs(V): 0.025–40As(V): 24.90.164–8[28]
Cellulose-g-GMA-b-TEPAAs(III): ~5–35; As(V): ~20–100As(III): 5.71; As(V): 75.133.68As(III): 7 & As(V): 5[29]
Fe(III)-AM-PGMACellAs(V): 25–400As(V): 78.839.96[30]
AM-Fe-PGDCAs(V): 10–400As(V): 105.4731.66[31]
Cell-N-CuAs(V): 100–700As(V): 98.98.4[32]
CNs/Fe2O3 nanorodAs(III): 13.87, As(V): 15.717 & 3[21]
Cellulose@Fe2O3 composites *As(III): 64.33, As(V): 89.191137[20]
iMNPAs(V): 1–10As(V): 12.74145.56.6[33]
Fe3O4/AC compositeAs(III): 2–120As(III): 7.58[34]
OMIM As(III) & As(V): 1–100As(III): 67.89, As(V): 93.541543[35]
IONP@CNF-IONP aerogel *As(III): 0.055–15.9; As(V): 0.073–21.7As(III): 47.75; As(V): 90.901657This work
Abbreviations: Cellulose-g-PDMAEMA: native cellulose fibers modified with poly(N,N-dimethyl aminoethyl methacrylate), FeOOH/CuO@WBC: iron(III) oxyhydroxide/copper oxide composite based on water bamboo cellulose, DETA-g-DA-NCC: diethylene triamine grafted dialdehyde nanocrystalline cellulose, Functionalized CNFs: cellulose nanofibrils modified with trimethylammonium chloride, Cellulose-g-GMA-b-TEPA: glycidyl methacrylate grafted cellulose modified with tetraethylenepentamine, Fe(III)-AM-PGMACell: iron(III)-coordinated amino-functionalized poly(glycidyl methacrylate)-grafted cellulose, AM-Fe-PGDC: Fe(III)-coordinated amino-functionalized poly(glycidylmethacrylate)-grafted TiO2-densified cellulose, Cell-N-Cu: copper containing modified cellulose, CNs/Fe2O3 nanorod: cellulose nanocrystals/iron oxide nanorod composite, Cellulose@Fe2O3 composites: composite material containing magnetic Fe2O3 and cellulose iMNP: magnetic nanoparticles prepared from iron containing sludge, Fe3O4/AC composite3: iron oxide/commercial activated carbon composite, OMIM: mesoporous iron manganese bimetal oxides. * The maximum adsorption capacities are based on the iron oxide only.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rahman, M.M.; Hafez, I.; Tajvidi, M.; Amirbahman, A. Highly Efficient Iron Oxide Nanoparticles Immobilized on Cellulose Nanofibril Aerogels for Arsenic Removal from Water. Nanomaterials 2021, 11, 2818. https://doi.org/10.3390/nano11112818

AMA Style

Rahman MM, Hafez I, Tajvidi M, Amirbahman A. Highly Efficient Iron Oxide Nanoparticles Immobilized on Cellulose Nanofibril Aerogels for Arsenic Removal from Water. Nanomaterials. 2021; 11(11):2818. https://doi.org/10.3390/nano11112818

Chicago/Turabian Style

Rahman, Md Musfiqur, Islam Hafez, Mehdi Tajvidi, and Aria Amirbahman. 2021. "Highly Efficient Iron Oxide Nanoparticles Immobilized on Cellulose Nanofibril Aerogels for Arsenic Removal from Water" Nanomaterials 11, no. 11: 2818. https://doi.org/10.3390/nano11112818

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