Next Article in Journal
Facile Synthesis of Anatase–Rutile Diphase N-doped TiO2 Nanoparticles with Excellent Visible Light Photocatalytic Activity
Next Article in Special Issue
Co-Catalytic Action of Faceted Non-Noble Metal Deposits on Titania Photocatalyst for Multielectron Oxygen Reduction
Previous Article in Journal
Cu-Promoted Iron Catalysts Supported on Nanorod-Structured Mn-Ce Mixed Oxides for Higher Alcohol Synthesis from Syngas
Previous Article in Special Issue
Thickness Effect on Photocatalytic Activity of TiO2 Thin Films Fabricated by Ultrasonic Spray Pyrolysis
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water

1
Central Department of Chemistry, Tribhuvan University, Kathmandu 44600, Nepal
2
Department of Chemistry, Tri–Chandra Multiple Campus, Tribhuvan University, Kathmandu 44600, Nepal
3
Korea Center for Artificial Photosynthesis, Department of Chemistry, Sogang University, Seoul 04107, Korea
4
Carbon Composite Energy Nanomaterials Research Centre, Woosuk University, Wanju, Chonbuk 55338, Korea
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(10), 1125; https://doi.org/10.3390/catal10101125
Submission received: 15 September 2020 / Accepted: 26 September 2020 / Published: 1 October 2020
(This article belongs to the Special Issue TiO2 for Photocatalytic Applications)

Abstract

:
A novel type of adsorbent, TiO2 impregnated pomegranate peels (PP@TiO2) was successfully synthesized and its efficacy was investigated based on the removal of As(III) from water. The adsorbent was characterized using Scanning Electron Microscopy (SEM), Energy Dispersive X-ray Spectrometer (EDS), X-ray Diffraction (XRD) analysis, and Fourier Transform Infrared (FTIR) Spectroscopy, to evaluate its morphology, elemental analysis, crystallinity, and functional groups, respectively. Batch experiments were conducted on PP@TiO2 for As(III) adsorption to assess the adsorption isotherm, effect of pH, and adsorption kinetics. Characterization data suggested that TiO2 was successfully impregnated on the biomass substrate. The equilibrium data better fitted to the Langmuir isotherm model having a maximum adsorption capacity of 76.92 mg/g and better distribution coefficients (KD) in the order of ~103 mL/g. The highest percentage of adsorption was found at neutral pH. The adsorption kinetics followed the pseudo-2nd-order model. X-ray Photoelectron Spectroscopy (XPS) of the adsorption product exhibited that arsenic was present as As(III) and partially oxidized to As(V). PP@TiO2 can work effectively in the presence of coexisting anions and could be regenerated and reused. Overall, these findings suggested that the as-prepared PP@TiO2 could provide a better and efficient alternative for the synergistic removal of As(III) from water.

1. Introduction

Safe water is the basic need for the existence of life. Due to the rapid industrialization, the world’s supply of safe water is decreasing and the pollutant levels are rising steadily [1,2,3]. Arsenic (As) contaminated water is one of the most challenging environmental concerns because of its chronic toxicity and carcinogenicity (black foot disease; skin, liver, bladder, kidney, and lung cancer; stillbirths; and heart attacks) [4]. Naturally occurring As-contaminated groundwater has caused hazardous effects to hundreds of millions of population globally, especially in Bangladesh, India, Pakistan, China, Argentina, Cambodia, Chile, Mexico, New Zealand, Taiwan, Vietnam, the USA, and Nepal as well [5,6]. Arsenic is introduced into groundwater through natural sources such as volcanic emissions, geochemical activity, rock, and soil weathering as well as the anthropogenic sources such as mining activities, combustion of fossils fuel, agricultural chemicals, wood preservatives, acid mine drainage wastewater, and other industrial discharges of metal processing, semiconductor, copper smelting, electroplating, pigments, dyestuff, and paints [7,8,9]. Inorganic arsenic has been classified as a human carcinogenic substance of Group 1 by the International Agency for Research on Cancer (IARC) [5,10]. Based on its exposure concerns and acute lethal effects on human health, WHO and USEPA promulgated the maximum contamination limit (MCL) for arsenic in drinking water not to exceed 10 ppb (0.01 mg/L) [11,12]. However, some South Asian countries have followed MCL of 50 ppb (0.05 mg/L) as their drinking water standard [13,14,15].
Inorganic arsenic species are the most abundant in natural waters. The predominant species of toxic inorganic arsenic in waters are either arsenite (As(III)) or arsenate (As(V)). Under aerobic conditions, e.g., in oxygen-rich surface waters, As(V) is dominant. Under anaerobic conditions, e.g., in groundwater, the reduced As(III) dominates, which is more hazardous and more challenging to sequester from water by most conventionally applied techniques than As(V) [16,17,18,19,20,21]. Therefore, As(III) is generally sequestered by oxidizing it to As(V) followed by adsorption, ion exchange, coagulation-precipitation, or membrane separation processes [22,23,24,25,26]. In developing countries, adsorption is the best method compared to others because it is simple, cheap, and competent with minimal sludge generation and can be used in water having a trace level of contaminants [5,27]. The oxidation pretreatment can be achieved using chemical reagents (e.g., ozone, hydrogen peroxide, potassium permanganate, ferrous salt, or manganese oxide, etc.) [28], but this may result in a significant increase in treatment costs and cause secondary pollutants of residual reagents and by-products [29]. Recently, TiO2 (the most common semiconductor photocatalyst) has been reported as an environment–friendly and cheap material in oxidizing As(III) to As(V) under visible and ultraviolet light irradiation [30,31,32,33,34,35,36]. During the photocatalytic oxidation of As(III), it is oxidized first into As(IV) with superoxide radicals produced at the TiO2 surface and then further oxidized to As(V) by reaction with dioxygen, hydroxyl radicals, or a photogenerated hole [32,33,34]. Moreover, past studies have proposed TiO2 as an adsorbent for the remediation arsenic from aqueous solution [36,37,38,39,40,41]. However, the low adsorption capacities as well as the difficulty of separation and recovery of the small particles from water usually limit the direct application of suspended TiO2 particles in arsenic removal [32].
Like TiO2, waste biomaterials can be utilized as attractive water treatment materials because of being cheap, widely available, non-toxic, and biodegradable. In the last decades, biomass-based adsorbents have been used for sequestration of arsenic from water. Some of these biomasses includes orange waste [20,21], watermelon rind [42], sugarcane bagasse [43,44], pomegranate waste [18], Staphylococcus xylosus [10] etc. Biomass usually is composed of cellulose, pectin, lignin as major constituents and provides multiple functional groups (e.g., –OH, –COOH, and C=O), which can efficiently bind arsenic through complexation and ion exchange. Though biomass has been used as adsorbents for removing pollutants, they have a less adsorptive capacity for As(III) in their raw form. Thus, effective sorbents for arsenic can be synthesized by combining TiO2 and biomass. TiO2 impregnated biomass offers the synergistic benefit of biomass and TiO2, which can not only adsorb As(III) but also catalyze the oxidation of As(III) to As(V). Biomass can also support the dispersion of TiO2 particles increasing the stability and recovery of tiny particles. Several interactions (e.g., H-bonding, electrostatic interaction, and C–Ti–O or C–O–Ti bonds) may be formed between biomass and TiO2 [45]. A few works have been reported on using chitosan/TiO2 composite for remediation of some toxic elements, such as Cd2+, Ni2+, Cu2+, Ag+, and Pb2+ from water [46,47,48]. Past studies have reported that carbon adsorbents derived from agricultural waste are more effective for the removal of heavy metals and antibiotics from water and wastewater [49,50,51]. Cruz et al. [52] conducted an adsorption experiment using agrowaste-derived biochars impregnated with ZnO for removing As(V) and Pb(II) from water. Recently, Pincus et al. [53] reported chitosan-TiO2-Cu composite photocatalyst for the removal of Arsenic.
There are voluminous research works available in the literature regarding the use of agro-waste derived biomass materials [3,10,18,20,21,28,31,42,44] and bare nano-TiO2 [37,38,39,40] for sequestration of heavy metals including arsenic from aqueous medium. However, impregnation of TiO2 into agro-waste derived biomass is still lacking though TiO2 composite with other precursors like activated carbon [32], metal oxides [36], chitosan [45,46,47,48], graphene [54], and carbon nanotube [55] have been studied. To the best of our knowledge, no reports are available in the literature on the synthesis of TiO2 impregnated pomegranate peels (PP@TiO2) for the removal and oxidation of As(III) simultaneously. Because of such circumstances, PP@ TiO2 is supposed to be a novel materials herein studied and its application as an adsorbent for removal of As(III) was proposed.
The objective of the present work was to fabricate TiO2 impregnated pomegranate peels (PP@TiO2) by using a modified sol-gel method and to investigate its adsorption behaviors of As(III) from water under different conditions through batch experiments. The adsorption characteristics of the PP@TiO2 were analyzed by SEM, EDS, XRD, FTIR, and XPS. The adsorption isotherms, kinetics parameters, effect of pH, and effect of common coexisting anions on As(III) adsorption were also investigated and discussed. From the results, the plausible adsorption mechanisms had been proposed. A desorption study was conducted to investigate the regeneration and reusability of the adsorbent. With further development, this research may offer a cost-effective, environmentally friendly, and easy-handle treatment alternative for remediation of arsenic from aqueous medium.

2. Results and Discussion

2.1. Adsorbent Characterization

The surface morphology of adsorbent was studied through SEM measurements. The SEM images (Figure 1a,b) illustrate the porous nature and irregular morphology of cross-linked pomegranate peels powder (PP) and PP@TiO2. The decrease in porosity in PP@TiO2 (Figure 1b) as compared to PP (Figure 1a) is probably due to the occlusion of pores by TiO2 particles. As shown, some particles of TiO2 are attached to the surface of biomass. Therefore, it is justified that the impregnation of PP by TiO2 has been done effectively. SEM image of As(III) adsorbed PP@TiO2 (Figure 1c) shows that the surfaces of adsorbent are covered because of the adsorption of arsenic and forms layer on the surfaces. EDS spectra of the PP, PP@TiO2, and As(III) adsorbed PP@TiO2, are shown in Figure 1d–f, respectively. As seen in Figure 1d, the main peaks corresponding to C, O, and Si were observed for the PP sample. After impregnation with TiO2 (PP@TiO2), new peaks related to Ti were observed (Figure 1e), suggesting that the TiO2 had been successfully impregnated onto biomass. As a result of arsenic adsorption, another peak belonging to arsenic was observed in the EDX spectra of As(III) adsorbed PP@TiO2 (Figure 1f). This verified that PP@TiO2 have successfully adsorbed arsenic from aqueous solution. The adsorption product was termed as As(III) adsorbed PP@TiO2 and abbreviated as As//PP@TiO2 hereafter.
Figure 2a shows the result of the XRD pattern of PP and PP@TiO2. There were no sharp peaks of PP indicating the amorphous nature. In the case of PP@TiO2, the diffraction peaks attributed to crystalline TiO2 appear at 2θ = 25.2°, 38.1°, 48°, 55°,63°,70°, and 75°. These peaks correspond to the (1 0 1), (0 0 4), (2 0 0), (1 0 5), (2 0 4), (2 2 0), and (2 1 5) planes, respectively. This confirms the successful impregnation of TiO2 in the crystalline anatase form in biomass. The peaks show good agreements with JCPDS, card No. 21-1272 of anatase phase of TiO2 [56,57,58]. The confirmation of the functional groups on adsorbents was done using FTIR spectra as shown in Figure 2b. It was examined in the ATR mode of FTIR. In the case of PP, absorption bands observed at around 3346, 2915, 1735, and 1621 cm−1 are attributed to –OH, –CH2, –COO, and C=O stretching, respectively. The bands with the highest intensity located at 1023 cm−1 correspond to the C–O stretching of the hydroxyl functional group connected to the carbon atom. In the case of PP@TiO2, the absorption peaks between 420 and 700 cm−1 ascribed to the Ti–O vibration representing the interaction of TiO2 with biomass [54,56]. This provides clear evidence about the impregnation of TiO2 and the formation of the Ti–O–C bond. The intensities of bands of the aforementioned oxygen-containing functional groups of PP@TiO2 were significantly decreased as compared to PP, indicating that loading of TiO2 onto PP. After arsenic adsorption, at the FTIR spectra of As//PP@TiO2, a new band observed at 825 cm−1, corresponding to As–O stretching vibrations. This confirms As(III) adsorption onto PP@TiO2. A similar phenomenon has been reported for arsenic adsorption onto nano scaled activated carbon modified by iron and manganese oxide [59].

2.2. Adsorption Isotherm

To investigate the quantitative As(III) adsorption capacity of PP@TiO2, adsorption isotherm experiments were carried out using batch methods within a broad range of As(III) concentrations (V/m =1000 mL/g, pH~7). The initial and equilibrium concentrations of As(III) in solution are measured by inductively coupled plasma-mass (ICP-MS) spectroscopy. The obtained concentrations were used to calculate the % adsorption (A%, from Equation (1)) and the adsorption capacity of PP@TiO2 (qe, from Equation (2)).
%   A   =   C o     C e C o   ×   100
q e   =   ( C o     C e )   V m
where, V, m, Co, and Ce indicate the volume of As(III) solution (mL), amount of PP@TiO2 taken (g), initial concentration (mg/L), and equilibrium concentration (mg/L) respectively. The distribution coefficient (KD) is used to express the chemical binding capacity of an adsorbent and is stated as the ratio of the concentration of arsenic adsorbed in per gram of the adsorbent and its concentration per mL in the aqueous phase. It was expressed according to Equation (3) [60].
K D   =   ( V m ) ( C o C e ) C e
The equilibrium data are presented in Figure 3a. Past studies have reported that adsorbents having a KD value of 103 mL/g are considered good adsorbent [61]. KD values for the adsorption of As(III) obtained are in the order of ~103 mL/g(see Figure 4b), which indicates that As(III) binds efficiently to the adsorbent.
The equilibrium data were fitted using two binding models, Langmuir and Freundlich. The Langmuir model assumes the saturated monolayer adsorption on the homogenous and energetically identical adsorption sites, and no significant interaction between adsorbed species, which is expressed according to Equation (4) in the linear form [62] as:
C e q e   =   1 q m   b   +   C e q m
where, qe (mg/g) is the amount of As(III) adsorbed at equilibrium concentration Ce (mg/L), qm (mg/g) is the maximum adsorption capacity of adsorbent, and b (L/mg) is the Langmuir equilibrium constant related to adsorption energy. The qm and b were determined from the slope and intercept of the Langmuir plot of Ce versus Ce/qe presented in Figure 3b.
Freundlich model considers a heterogeneous sorption surface and sorption sites with a different energy, which can be expressed by Equation (5) in its linear form [63] as:
log   q e   =   log   K F   +   ( 1 n )   log   C e
where KF(mg/g) and n (L/mg) are Freundlich constants that indicate the adsorption capacity and the adsorption intensity, respectively. The two constants KF and n were calculated from the slope (1/n) and intercept (logKF) of the Freundlich isotherm plot of logCe versus logqe presented in Figure 3c. The evaluated values of the isotherm parameters of both models are listed in Table 1. It is obvious from the result that the PP@TiO2 is a suitable adsorbent for the adsorption of As(III) from water. For good adsorbent, the value of Freundlich constant (n) ranged between 1 and 10 (i.e., 1/n < 1). It is evident from Table 1 that the value of ‘n’ was found to be 2.72, indicating a strong adsorbent-adsorbate interaction. This tells us that the result obtained from Freundlich isotherm certainly cannot be neglected. Thus, both Langmuir and Freundlich isotherms showed a good fit to the experimental equilibrium data. The value of correlation coefficients (R2) shows that the Langmuir model is better fitted than the Freundlich isotherm model. A similar observation was reported by Chandra et al. [64] for arsenic adsorption on magnetite reduced graphene oxide and Shehzad et al. [9] for As(III) adsorption on ZrO2 nanosheets. These results indicated that the adsorption of As(III) onto PP@TiO2 was not governed by a single mechanism.
The observed maximum As(III) adsorption capacity, qm, 76.92 mg/g compares well with previously reported materials for Arsenic adsorption, which are listed in Table 2. This comparison indicates that the adsorption capacity of PP@TiO2 is higher than that of previously studied adsorbents [9,16,17,18,21,37,42,44,65,66,67,68,69,70]. Hence, the prepared PP@TiO2 may be a promising material towards the removal of As(III) from water.
The feature of Langmuir isotherm was further justified by the analysis of the dimensionless separation factor, RL, [71], which is given by Equation (6).
R L   =   1 1 + ( b C o )
where b (L/mg) is Langmuir equilibrium constant and Co (mg/L) is the initial concentration of As(III). The value of RL maybe 0, 1, or >1 indicating irreversible, linear, and unfavorable adsorption respectively. From the experimental data for all the concentrations tested, the evaluated RL values were ranges from 0.06 to 0.76 (see Figure 3d). All of these values lie between 0 and 1 (0 < RL< 1), which indicates more favorable adsorption of As(III) onto PP@TiO2.

2.3. Effect of pH and As(III) Adsorption Mechanism

The % adsorption of As(III) and distribution coefficient, KD onto PP@TiO2evaluated in the pH range of 2–12 are presented in Figure 4a,b, respectively, which show that the effective pH range for As(III) adsorption is from 6 to 9. A maximum KD value of 3.21 × 103 mL/g with the highest As(III) adsorption capacity of 81.4% is observed at pH 7. On increasing pH of the solution from 2 to 7, As(III) adsorption increased and then decreased beyond the pH 7. At lower pH values, As (III) exists only as neutral H3AsO3 species, and the surfaces of PP@TiO2 are protonated. Thus, binding of As (III) on positively charged adsorbent is not favorable [20]. The adsorption of As(III) species took place via both electrostatic attraction and ligand exchange of the hydroxyl ions or water molecules contained in the coordination spheres of impregnated TiO2 from the biomass surface. At higher than 10 pH values, the predominant As(III) species H2AsO3, HAsO32– and AsO33– are repelled by the negatively charged adsorbent surface and therefore lower adsorption of As(III) has been obtained [21]. The decrease in As(III) adsorption can be attributed to the combined effect of electrostatic repulsion and the competition between the hydroxyl ions, and anionic arsenic species for adsorption sites. This is in agreement with the observation reported for As(III) adsorption from aqueous solution using iron oxide/nano-porous carbon magnetic composite [16].
Mechanism of As(III) adsorption using PP@TiO2 can be described as follows. Impregnation with TiO2 results in the generation of additional adsorption sites on the exterior surface of PP@TiO2. It is inferred that during the impregnation of tetravalent Ti, neutralization of all the four positive charges of Ti by the functional groups (–OH, –COOH) present in biomass is difficult because of steric hindrance. Therefore, one or two positive charges of tetravalent Ti bind with a carboxylic acid or other functional groups in pomegranate peel biomass, and the rest are neutralized with hydroxyl ion. Thus, TiO2 in aqueous solution undergoes a hydration reaction to give hydrated titanium oxide, Polymer ≡ Ti(OH)n(H2O)6–n (Equation (7)). At near neutral as well as weakly alkaline conditions i.e., at pH 6 to 9, As(III) exists as anionic species H2AsO3 and HAsO32– as well as neutral H3AsO3 molecule [20]. The adsorption of As(III) on PP@TiO2 may be considered as a ligand exchange mechanism by the substitution of hydroxyl anions or neutral H2O molecules. Thus, the undermentioned mechanism may be responsible for the sorption of anionic species (Equation (8)) and neutral species (Equation (9)) of As(III).
Polymer ≡TiO2 + (6–n)H2O + nH+ → Polymer ≡Ti(OH)n(H2O)6–n,
Polymer ≡Ti(OH)n(H2O)6–n + H2AsO3 → Polymer ≡Ti(OH)n–1(H2AsO3)(H2O)6–n + OH,
Polymer ≡Ti(OH)n(H2O)6–n + H3AsO3 → Polymer ≡Ti(OH)n(H3AsO3)(H2O)5–n + H2O,
Besides, As(III) could be partially oxidized to less toxic and more easily adsorbed As(V) by the simple interaction with TiO2 on the surface of the adsorbent, in batch experiments performed under room conditions [37]. The oxidized As(V) species, H2AsO4 would be adsorbed according to Equation (10).
Polymer ≡Ti(OH)n(H2O)6–n + H2AsO4 → Polymer ≡Ti(OH)n–1(H2AsO4)(H2O)6–n + OH,
A similar mechanism has also been reported by Biswas et al., [20] for the removal of arsenic using Zr(IV) loaded orange waste gel. Joshi et al., [16] also reported a ligand exchange mechanism for As(III) onto Fe3O4/sugarcane bagasse activated carbon composite.

2.4. Adsorption Kinetics

To understand the dynamics of adsorption, the time variation in the amount of adsorption of As(III) on PP@TiO2 was measured. Figure 5a shows the equilibrium concentration and adsorption percentage of As(III) as a function of contact time. Figure 5b exhibits the adsorption capacity (qt) of As(III) with contact time on PP@TiO2. The adsorption for As(III) reaches equilibrium within ~6 h. Therefore, 24 h was chosen as an optimum agitation period to achieve the complete adsorption of As(III) on PP@TiO2 in the subsequent adsorption experiments. The kinetic data were fitted with pseudo-1st-order and pseudo-2nd-order models, which are expressed respectively [72] as Equations (11) and (12) below.
log ( q e     q t )   =   log q e     k 1 2.303   t
t q t   =   1 k 2 q e 2   +   1 q e   t
where qe and qt (mg/g) are the amount of As(III) ions adsorbed per unit mass of PP@TiO2 at equilibrium and time t, respectively, and k1,(min–1) and k2 (g mg−1 min−1) are the rate constants of pseudo-1st-order and pseudo-2nd-order adsorption kinetics model. The pseudo-1st-order kinetic plot of log(qe − qt) versus t (Figure 5c) gives a linear relationship from which k1 and qe can be measured from the slope and intercept, respectively. Similarly, the pseudo-2nd-order kinetic plot t/qt versus t (Figure 5d) gives a linear relationship from which k2 and qe can be measured from the intercept and slope, respectively. The kinetic parameters of As(III) adsorption evaluated based on both equations and R2 are listed in Table 3. The pseudo-2nd-order kinetics is suitable for As(III) adsorption due to the higher value of R2 (0.999), indicating the adsorption is chemisorption.

2.5. X-ray Photoelectron Spectroscopy (XPS) Studies

The surface composition and chemical states of the PP@TiO2 after As(III) adsorption experiments were further analyzed by the XPS. The wide scan spectrum indicated signals from Ti, O, C, and As in the sorption product (Figure 6a). The high-resolution XPS scan of Ti 2p (Figure 6b) showed two broad peaks centered at ~458.6 and 467.3 eV representing the Ti 2p3/2 and Ti 2p1/2 peaks of Ti4+, respectively [56]. Additional peaks at about 457.25 eV and 463.27 are attributed to the Ti 2p3/2 and Ti 2p1/2 peaks of Ti3+, respectively [73]. The Ti 2p spectrum mainly appears in Ti4+ state with a small amount of Ti3+. The occurrence of Ti3+ is due to oxygen deficiency in the lattice of TiO2 [73]. This indicated a reduction of Ti4+ to Ti3+ during the partial oxidation of adsorbed As(III). The high-resolution XPS scan of As 3d (Figure 6c) was composed of two peaks indicating the presence of both As3+ (B.E. = 44.32 eV) and As5+ (B.E. = 46.24 eV) with a qualitative ratio of 75.7% and 24.3%, respectively. Hence, the partial oxidation of As(III) by impregnated TiO2 was verified by XPS using As 3d spectrum [74]. The high-resolution XPS scan of O 1s (Figure 6d) was composed of peaks at ~528.4, 529.9, and 531.8 eV are associated with Ti–O, O-As, and O–H bonds, respectively.
Thus, XPS results revealed that As(III) is adsorbed onto PP@TiO2 and is partially oxidized to As(V). Similar behavior was reported by Tamayo et al. [19] for As(III) removal from aqueous solution on calcium titanate nanoparticles. The partial oxidation of As(III) to As(V) could be due to the presence of physio-sorbed oxygen and surface hydroxyl groups. All results revealed that PP@TiO2, on one hand, can remediate As(III) from water, and on the other hand, it can oxidize some of the As(III) to As(V) at the surface of the adsorbent simultaneously. Previous studies have reported that the degree of photocatalytic oxidation of As(III) to As(V) under the influence of TiO2 photocatalyst highly dependent on various operation factors such as pH, dissolved oxygen, initial As(III) concentration, a wavelength of light, the intensity of light, TiO2 loading, calcination temperature, properties of TiO2 (bandgap, particle size), presence of coexisting ions and batch reactor configuration [32,75]. Bang and Patel [37] reported that around 10% of As(III) is oxidized to less toxic and more easily adsorbed As(V) by the simple interaction with TiO2, in batch experiments performed under room conditions. Yao et al. [32] reported that 0.80 mg/L of As(III) could be completely oxidized to As(V) within 30 min in the presence of 3.0 g/L TiO2/ACF (composite of TiO2 and activated carbon fiber) in the irradiation of UV-light. They observed that the presence of silicate and phosphate ions significantly decreased the oxidation of As(III). Past studies have reported the complete oxidation of As(III) to As(V) by TiO2 in the presence of UV light and oxygen [30,34]. The As(III) adsorption and photo-oxidation capability of biomaterial/TiO2 composite increases as n-TiO2 loading [53]. Nevertheless, additional research is required to investigate detailed As(III) oxidation characteristics of PP@TiO2.

2.6. Effect of Common Coexisting Anions

The effect of common coexisting anions on As(III) adsorption from aqueous solution was examined considering various concentrations of chloride (Cl), sulphate (SO42−), and phosphate (PO43−). As shown in Figure 7, compared to the blank (no coexisting ions), an increase in the concentration of coexisting ions leads to a decrease in As(III) adsorption efficiency for all the tested ions. Minimal interference was observed in the presence of Cl, while the stronger interference was observed for PO43− that reaches % adsorption loss up to 27.9% at its highest concentration (100 mg/L). Hence, phosphate is the strongest competitor with arsenic ions to binding sites on PP@TiO2. This can be attributed to the similar oxyanion structure of both phosphate and arsenic anions [76].

2.7. Desorption Study and Reusability of PP@TiO2

Desorption of As(III) from As//PP@TiO2 is important to assess the recovery of toxic arsenic and regeneration of spent adsorbents for repeated use. Desorption tests were carried out by batch method using the adsorbent loaded with As(III) immediately after the adsorption processes. As shown in Figure 4, As(III) is efficiently adsorbed on PP@TiO2 at pH range 6–9 but it is poorly adsorbed at higher pH (>9), indicating that desorption of adsorbed As(III) can be easily achieved by employing the weakly alkaline solution. Consequently, the desorption test was carried out with different concentrations of NaOH ranging from 0.01–0.5 M as presented in Figure 8a. The % desorption (% D) of As(III) was calculated using Equation (13) as follows [77].
%   D   =   D amount A amount   ×   100
where, Aamount and Damount are the adsorbed and desorbed amounts (mg) of As(III) ions, respectively. The desorption of arsenic increases from 29.1% to 97.8% by increasing the concentration of NaOH from 0.01 to 0.1 M and remains constant with a further increase in concentration. Hence, a 0.1 M NaOH solution was found to be a suitable desorbing agent for the regeneration of spent adsorbent. Hydroxyl ions released from the NaOH solution might have replaced the adsorbed arsenic anions onto the PP@TiO2 surface. The desorption of arsenic species is believed to take place by the ligand substitution mechanism by the replacement of adsorbed arsenic species by hydroxyl ions in alkali solution as shown below (Equation (14)).
Polymer ≡Ti(OH)n–1 (H2AsO3)(H2O)6–n + OH → Polymer ≡Ti(OH)n (H2O)6–n + H2AsO3,
To evaluate the reusability of the PP@TiO2, it was subjected to four series of adsorption/desorption cycles. Figure 8b shows that the initial adsorption capacity of As(III) was decreased by 14.5% after four cycles. This result indicated that the PP@TiO2 exhibited excellent stability and reusability. In alkaline solution, the surface of the adsorbent becomes negatively charged and As(III) anions are displaced by OH, leaving the sorption sites before the succeeding adsorption cycle without affecting the chemical stability of the adsorbent. The high desorption efficiency of arsenic from the PP@TiO2 created the low-cost removal technique since both adsorbent and arsenic were regenerated and recycled efficaciously.

3. Materials and Methods

3.1. Materials

Analytical grade chemicals were used without additional purification. The waste biomass of pomegranate peels employed in this study was collected from a local juice vendor, Kathmandu, Nepal. All the solutions in the study were prepared using deionized water. Sodium arsenite (NaAsO2) (Sigma–Aldrich, New Delhi, India), titanium(IV) n-butoxide (Samchun pure chemical, Seoul, Korea) were used. A standard stock solution of As(III) (1000 mg/L) was prepared by dissolving an exact amount of NaAsO2 in deionized water and stored in the refrigerator for further use.

3.2. Synthesis of TiO2 Impregnated Pomegranate Peels (PP@TiO2)

The pomegranate peel waste was kindly donated by a local juice vendor, Kathmandu, Nepal. Pomegranate peels were first washed thoroughly with distilled water. It was then oven-dried at 70–80 °C for 48 h. The dried mass was grounded and sieved through (150 μm) mesh copper sieve. To avoid the dissolution of the adsorbent in aqueous solutions, polyphenol and polysaccharide hydroxyl groups in the biomass were cross-linked through condensation reaction using concentrated sulfuric acid as a dehydrating agent. The cross-linked pomegranate peels were prepared as described in Paudyal et al., [78]. 15 g of this biomass powder was mixed with 30 mL of concentrated H2SO4 (96%) in a round bottom flask and the mixture was agitated for 24 h at 100 °C followed by cooling at room temperature. The black product was then neutralized with sodium bicarbonate and reconditioned by stirring again in 1 M HCl solution. It was washed several times with distilled water till neutrality, after which it was dried overnight in a convection oven at 70 °C. The pomegranate peels powder obtained in this way is abbreviated as PP.
The PP@TiO2 was prepared using a previously reported modified sol-gel method [73,79,80,81]. About 10 mL of titanium(IV) n-butoxide was mixed with 15 mL of ethanol. After stirring for 30 min at room temperature, 2 g of powdered PP was added to it. Then, 15 mL of a solution of 1:1:1 ethanol, deionized water, and acetic acid were added dropwise from the burette to the magnetically stirred mixture. The whole content was further stirred for 4h until sol was obtained. The sol was dried overnight at 80 °C after aging at 40 °C for 2 h. Finally, the dried gel was ground into a powder and calcined at 400 °C in a muffle furnace for 2 h. The resulting PP@TiO2 was stored in a plastic container before being used in the characterization and adsorption experiment.

3.3. Characterizations

To investigate the morphology and elemental composition of adsorbents, SEM images and EDS were carried out on a JEOL JSM–6700F instrument (Jeol Ltd., Tokyo, Japan) equipped with X-ray Energy Dispersive Spectrometer. The crystallinity of PP@TiO2 samples was investigated by XRD, Rigaku Ultima IV X-ray diffractometer (Rigaku Co., Tokyo, Japan)with Cu Kα (λ = 1.54056 Å) radiation. The surface functional groups of samples were analyzed using FTIR spectroscopy on IR AFFINITY-1 Shimadzu (Shimadzu, Kyoto, Japan) spectrometer in the wavenumber range of 400–4000 cm−1.

3.4. Batch Adsorption Studies

The adsorption studies for As(III) were carried out to examine the adsorption behavior on the PP@TiO2 using the batch method (V: m~1000 mL/g). 25 mg (dry weight) of PP@TiO2 was taken in different 50 mL conical flasks containing 25 mL of As (III) solution in the concentration range 10–600 mg/L and shaken for 24 h at 25 °C to attain equilibrium. The pH of the As(III) solution was maintained with 0.1M HCl and 0.1 M NaOH solutions. After adsorption, the PP@TiO2 was separated from the mixture by filtration and dried at 70 °C. The initial and equilibrium concentration of arsenic was analyzed. From the equilibrium concentration, the sorption isotherms and maximum adsorption capacity were evaluated. Two isotherms model, Langmuir, and Freundlich were used to describe the As(III) sorption behavior.
Effect of pH: The influence of pH of the solution on the sorption of arsenic was investigated over the pH range 2–13. 25 mg of PP@TiO2 was shaken for 24 h in 25 mL solution of 20–25 mg/L As(III) with varying pH solutions, prepared by using 0.1M HCl and 0.1M NaOH solutions. The resulting solution was analyzed for the equilibrium concentration.
Adsorption Kinetics: Kinetic studies of As(III) adsorption on to PP@TiO2 were carried out until it reached equilibrium. For each experiment, 25 mg of the PP@TiO2 was taken in a 50 mL conical flask containing 25 mL of ~30 mg/L As(III) aqueous solution (V/m = 1000 mL/g) at pH 7, and the content was shaken well at room temperature. Adsorption experiments of various contact times were performed. After a predetermined time interval, the sample of each flask was immediately filtered and the equilibrium concentration of As(III) was analyzed.
Effects of Common Co-existing Ions: Natural water contains several coexisting ions, which may compete with As(III) for available active sorption sites and affect the As(III) adsorption efficiency of PP@TiO2. The effect of common coexisting anions in an aqueous system such as Cl, SO42−, and PO43− on the adsorption of As(III) was examined by adding various concentrations (10, 50, and 100 mg/L) of each coexisting ions into 25 mg/L As(III) solution as a binary mixture. To prepare various concentrations of coexisting ion solutions, different potassium salts (KCl, K2SO4, and KH2PO4, 99% pure, Merk, Kenilworth, NJ, USA) were used. Additionally, the blank sample was made similarly without adding coexisting anion. The pH of all the solutions was maintained to 7.0 by using 0.1M HNO3 and 0.1M NaOH. A defined amount (1 g/L) of PP@TiO2 was added and the mixtures were shaken for 24 h at 25 °C to attain equilibrium. After filtration, the remaining concentration of As(III) was analyzed.

3.5. X-ray Photoelectron Spectroscopy (XPS) Studies

To elucidate chemical composition and arsenic speciation, the XPS investigation of sorption product was carried out on Kratos Axis Ultra DLD spectrometer (Kratos Analyticals, Manchester, UK) with a monochromatic Al Kα X-ray source (150 W). Samples were mounted using double-sided adhesive tape, and binding energies were referenced to the C 1s binding energy of adventitious carbon taken to be 286.4 eV. For the survey of elements, the wide scans were taken at pass energy of 80 eV, and scans of photoelectron peaks were taken at pass energy of 40 eV.

3.6. Desorption Study and Reusability of PP@TiO2

For the desorption study, As(III) adsorbed PP@TiO2 was prepared by shaking 250 mg of PP@TiO2 with 250 mL of As(III) solution (25 mg/L) at pH 7 for 24 h at room temperature. After filtration, the filtrate was examined for residual As(III) concentration. From initial and residual concentration, the adsorbed amount of As(III) was evaluated. The filter cake was dried. The dried As(III) adsorbed PP@TiO2 was used for the elution test. The adsorbed arsenic was effectively eluted by the alkali solution. To identify the optimum concentration of alkali, 25 mg of exhausted PP@TiO2 was shaken with 25 mL of various concentrations of NaOH solution ranging from 0.01 to 0.5 M batch-wise, similar to adsorption studies. The sample of each flask was filtered and the filtrate was analyzed for the desorbed amount of arsenic. The % desorption was calculated. Four cycles of adsorption/desorption of As(III) were carried out to investigate the reusability of PP@TiO2. The PP@TiO2 samples after the adsorption process were stirred with 0.1M NaOH solution followed by filtration, washing with deionized water, and drying. The regenerated PP@TiO2 was again used for the adsorption experiments to evaluate reusability up to four times.

3.7. Analysis of Arsenic Concentration

For all the above experiments, the initial and equilibrium concentration of arsenic in the solution was analyzed by Inductively Coupled Plasma-Mass Spectrometer (ICP-MS) (Agilent Technologies, 7900 ICP-MS, Santa Clara, CA, USA) with argon plasma, and internal standards of germanium and scandium. For the calibration of the instrument, a repeating As(III) standard and blank were analyzed every 10 samples. Before analysis, all samples were acidified with 1% HNO3 and analyzed within 48 h.

4. Conclusions

In summary, we herein synthesized PP@TiO2 using a modified sol-gel method, which was then applied as an efficient adsorbent for removal of As(III) from water. Based on the results of characterizations, it was reasonable to conclude the effective loading of TiO2 into biomass and successful adsorption of As(III) onto PP@TiO2. The maximum adsorption of As(III) on PP@TiO2 was observed at neutral pH with a high distribution coefficient of 4.0 × 103 mL/g. The equilibrium data were better followed by the Langmuir-model, which indicated that the monolayer adsorption of As(III) occurred on a homogenous surface of the PP@TiO2. The maximum adsorption capacity for As(III) was evaluated as 76.92 mg/g, which was found to be higher than many other agro-waste based bio-adsorbents and bare TiO2 reported in the literature. Kinetics data showed that the adsorption followed the pseudo-second-order model, suggesting the process might be chemisorptions with higher initial adsorption rates. The XPS analysis of PP@TiO2 after the As(III) adsorption process revealed that the adsorbed species was present as As(III) and partially oxidized to As(V). Moreover, the PP@TiO2 can work effectively in the presence of various coexisting anions, and the exhausted PP@TiO2 can be easily separated for regeneration and reuse. PP@TiO2 is a safe substance for human health and a feasible low-cost material. Thus, PP@TiO2 investigated in this study can be considered an efficient, cost-effective, environmentally friendly, and reusable adsorbent for removing As(III) from water.

Author Contributions

Conceptualization, M.R.P., K.N.G., and H.P.; methodology, B.R.P., R.L.A., and S.B.; validation, S.K.G., H.P., and M.R.P.; formal analysis, H.P. and B.R.P.; data curation, B.P., S.B., and B.R.P.; XPS investigation and XPS data analysis, A.R.K.; writing—original draft preparation, B.R.P., and R.L.A.; writing—review and editing, B.P., M.P., and M.R.P.; Supervision, H.P. and M.R.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Traditional Culture Convergence Research Program through the National Research Foundation of Korea (NRF) funded by the Ministry of Science, CT & Future Planning (2018M3C1B5052283) and the National Research Foundation of Korea (NRF) grant funded by the Korea Government (MSIT) (No. NRF-2019R1A2C1004467).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Shannon, M.A.; Bohn, P.W.; Elimelech, M.; Georgiadis, J.G.; Marinas, B.J.; Mayes, A.M. Science and technology for water purification in the coming decades. Nature 2008, 452, 301–310. [Google Scholar] [CrossRef] [PubMed]
  2. Rathore, E.; Biswas, K. Selective and ppb level removal of Hg(II) from water: Synergistic role of graphene oxide and SnS2. J. Mater. Chem. A 2018, 6, 13142–13152. [Google Scholar] [CrossRef]
  3. Raj, K.R.; Kardam, A.; Srivastava, S. Development of polyethyleneimine modified Zea mays as a high capacity biosorbent for the removal of As (III) and As (V) from aqueous system. Int. J. Miner. Process. 2013, 122, 66–70. [Google Scholar] [CrossRef]
  4. Ratnaike, R.N. Acute and chronic arsenic toxicity. Postgrad. Med. J. 2003, 79, 391–396. [Google Scholar] [CrossRef] [PubMed]
  5. Mohan, D.; Pittman, C.U., Jr. Arsenic removal from water/wastewater using adsorbents—A critical review. J. Hazard. Mater. 2007, 142, 1–53. [Google Scholar] [CrossRef]
  6. Shankar, S.; Shanker, U. Arsenic contamination of groundwater: A review of sources, prevalence, health risks, and strategies for mitigation. Sci. World J. 2014, 2014, 1–18. [Google Scholar] [CrossRef] [PubMed]
  7. Wang, S.; Gao, B.; Zimmerman, A.R.; Li, Y.; Ma, L.; Harris, W.G.; Migliaccio, K.W. Removal of arsenic by magnetic biochar prepared from pinewood and natural hematite. Bioresour. Technol. 2015, 175, 391–395. [Google Scholar] [CrossRef]
  8. Ogata, F.; Nagai, N.; Toda, M.; Otani, M.; Saenjum, C.; Nakamura, T.; Kawasaki, N. Removal of arsenic(III) ion from aqueous media using complex nickel-aluminum and nickel-aluminum-zirconium hydroxides. Water 2020, 12, 1697. [Google Scholar] [CrossRef]
  9. Shehzad, K.; Ahmad, M.; He, J.; Liu, T.; Xu, W.; Liu, J. Synthesis of ultra-large ZrO2 nanosheets as novel adsorbents for fast and efficient removal of As (III) from aqueous solutions. J. Colloid Interface Sci. 2019, 533, 588–597. [Google Scholar] [CrossRef]
  10. Aryal, M.; Ziagova, M.; Liakopoulou-Kyriakides, M. Study on arsenic biosorption using Fe (III)–treated biomass of Staphylococcus xylosus. Chem. Eng. J. 2010, 162, 178–185. [Google Scholar] [CrossRef]
  11. WHO. Arsenic in Drinking-water. Background document for development of WHO Guidelines for Drinking-water Quality. In Guidelines for Drinking-water Quality; (No. WHO/SDE/WSH/03.04/75/Rev/1); World Health Organization: Geneva, Switzerland, 2011; Available online: http://www.who.int/water_sanitation_health/dwq/chemicals/arsenic.pdf (accessed on 16 April 2020).
  12. USEPA. Drinking Water Requirements for States and Public Water System. 2017. Available online: https://www.epa.gov/dwreginfo/chemical–contaminant–rules (accessed on 16 April 2020).
  13. Bureau of Indian Standard (BIS). Indian Standards, Drinking Water Specification, 2nd ed.; Bureau of Indian Standards: New Delhi, India, 2012.
  14. National Drinking Water Quality Standards (NDWQS), Nepal. Nepal Gazette (B.S. 2063–03–12); Government of Nepal/Ministry of Physical Planning: Kathmandu, Nepal, 2006.
  15. Sanjoy, K.M.; Anjali, P.; Trasankar, P. Arsenic removal from real-life groundwater by adsorption on laterite soil. J. Hazard. Mater. 2008, 153, 811–820. [Google Scholar] [CrossRef]
  16. Joshi, S.; Sharma, M.; Kumari, A.; Shrestha, S.; Shrestha, B. Arsenic removal from water by adsorption onto iron oxide/nano–porous carbon magnetic composite. Appl. Sci. 2019, 9, 3732. [Google Scholar] [CrossRef] [Green Version]
  17. Chowdhury, T.; Zhang, L.; Zhang, J.; Aggarwal, S. Removal of arsenic (III) from aqueous solution using metal-organic framework-graphene oxide nanocomposite. Nanomaterials 2018, 8, 1062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Thapa, S.; Pokhrel, M.R. Removal of As(III) from aqueous solution using Fe(III) loaded pomegranate waste. J. Nepal Chem. Soc. 2012, 30, 29–36. [Google Scholar] [CrossRef]
  19. Tamayo, R.; Espinoza-González, R.; Gracia, F.; Rodrigues-Filho, U.P.; Flores, M.; Sacari, E. As(III) removal from aqueous solution by calcium titanate nanoparticles prepared by the sol-gel method. Nanomaterials 2019, 9, 733. [Google Scholar] [CrossRef] [Green Version]
  20. Biswas, B.K.; Inoue, J.I.; Inoue, K.; Ghimire, K.N.; Harada, H.; Ohto, K.; Kawakita, H. Adsorptive removal of As (V) and As (III) from water by a Zr (IV)-loaded orange waste gel. J. Hazard. Mater. 2008, 154, 1066–1074. [Google Scholar] [CrossRef]
  21. Ghimire, K.N.; Inoue, K.; Makino, K.; Miyajima, T. Adsorptive removal of arsenic using orange juice residue. Sep. Sci. Technol. 2002, 37, 2785–2799. [Google Scholar] [CrossRef]
  22. Yu, L.; Peng, X.; Ni, F.; Li, J.; Wang, D.; Luan, Z. Arsenite removal from aqueous solutions by γ-Fe2O3-TiO2 magnetic nanoparticles through simultaneous photocatalytic oxidation and adsorption. J. Hazard. Mater. 2013, 246, 10–17. [Google Scholar] [CrossRef]
  23. Ortega, A.; Oliva, I.; Contreras, K.E.; González, I.; Cruz-Díaz, M.R.; Rivero, E.P. Arsenic removal from water by hybrid electro–regenerated anion exchange resin/electrodialysis process. Sep. Purif. Technol. 2017, 184, 319–326. [Google Scholar] [CrossRef]
  24. Wickramasinghe, S.R.; Han, B.; Zimbron, J.; Shen, Z.; Karim, M.N. Arsenic removal by coagulation and filtration: Comparison of groundwaters from the United States and Bangladesh. Desalination 2004, 169, 231–244. [Google Scholar] [CrossRef]
  25. Litter, M.I.; Morgada, M.E.; Bundschuh, J. Possible treatments for arsenic removal in Latin American waters for human consumption. Environ. Pollut. 2010, 158, 1105–1118. [Google Scholar] [CrossRef] [PubMed]
  26. Sharma, V.K.; Sohn, M. Aquatic arsenic: Toxicity, speciation, transformations, and remediation. Environ. Int. 2009, 35, 743–759. [Google Scholar] [CrossRef] [PubMed]
  27. Pokhrel, M.R.; Poudel, B.R.; Aryal, R.L.; Paudyal, H.; Ghimire, K.N. Removal and recovery of phosphate from water and wastewater using metal–loaded agricultural waste-based adsorbents: A review. JIST 2019, 24, 77–89. [Google Scholar] [CrossRef]
  28. Amin, M.N.; Kaneco, S.; Kitagawa, T.; Begum, A.; Katsumata, H.; Suzuki, T.; Ohta, K. Removal of arsenic in aqueous solutions by adsorption onto waste rice husk. Ind. Eng. Chem. Res. 2006, 45, 8105–8110. [Google Scholar] [CrossRef]
  29. Yao, S.; Liu, Z.; Shi, Z. Arsenic removal from aqueous solutions by adsorption onto iron oxide/activated carbon magnetic composite. J. Environ. Health Sci. Eng. 2014, 12, 58. [Google Scholar] [CrossRef] [Green Version]
  30. Bissen, M.; Vieillard-Baron, M.M.; Schindelin, A.J.; Frimmel, F.H. TiO2-catalyzed photooxidation of arsenite to arsenate in aqueous samples. Chemosphere 2001, 44, 751–757. [Google Scholar] [CrossRef]
  31. Guan, X.H.; Du, J.S.; Meng, X.G.; Sun, Y.K.; Sun, B.; Hu, Q.H. Application of titanium dioxide in arsenic removal from water: A review. J. Hazard. Mater. 2012, 215, 1–16. [Google Scholar] [CrossRef]
  32. Yao, S.; Jia, Y.; Shi, Z.; Zhao, S. Photocatalytic oxidation of arsenite by a composite of titanium dioxide and activated carbon fiber. Photochem. Photobiol. 2010, 86, 1215–1221. [Google Scholar] [CrossRef]
  33. Lee, H.; Choi, W. Photocatalytic oxidation of arsenite in TiO2 suspension: Kinetics and mechanisms. Environ. Sci. Technol. 2002, 36, 3872–3878. [Google Scholar] [CrossRef]
  34. Ferguson, M.A.; Hoffmann, M.R.; Hering, J.G. TiO2-photocatalyzed As (III) oxidation in aqueous suspensions: Reaction kinetics and effects of adsorption. Environ. Sci. Technol. 2005, 39, 1880–1886. [Google Scholar] [CrossRef]
  35. Dutta, P.K.; Pehkonen, S.O.; Sharma, V.K.; Ray, A.K. Photocatalytic oxidation of arsenic (III): Evidence of hydroxyl radicals. Environ. Sci. Technol. 2005, 39, 1827–1834. [Google Scholar] [CrossRef] [PubMed]
  36. Zhang, F.S.; Itoh, H. Photocatalytic oxidation and removal of arsenite from water using slag–iron oxide-TiO2 adsorbent. Chemosphere 2006, 65, 125–131. [Google Scholar] [CrossRef] [PubMed]
  37. Bang, S.; Patel, M.; Lippincott, L.; Meng, X. Removal of arsenic from groundwater by granular titanium dioxide adsorbent. Chemosphere 2005, 60, 389–397. [Google Scholar] [CrossRef] [PubMed]
  38. López Paraguay, M.Z.; Cortes, J.A.; Pérez-Robles, J.F.; Alarcón-Herrera, M.T. Adsorption of arsenite from groundwater using titanium dioxide. Clean (Weinh.) 2014, 42, 713–721. [Google Scholar] [CrossRef]
  39. Pena, M.E.; Korfiatis, G.P.; Patel, M.; Lippincott, L.; Meng, X.G. Adsorption of As(V) and As(III) by nanocrystalline titanium dioxide. Water Res. 2006, 39, 2327–2337. [Google Scholar] [CrossRef]
  40. Pirila, M.; Martikainen, M.; Ainassaari, K.; Kuokkanen, T.; Keiski, R.L. Removal of aqueous As(III) and As(V) by hydrous titanium dioxide. J. Colloid Interface Sci. 2011, 353, 257–262. [Google Scholar] [CrossRef]
  41. Wei, Z.G.; Liang, K.; Wu, Y.; Zou, Y.D.; Zuo, J.H.; Arriagada, D.C.; Pan, Z.C.; Hu, G.H. The effect of pH on the adsorption of arsenic(III) and arsenic(V) at the TiO2 anatase [1 0 1] surface. J. Colloid Interface Sci. 2016, 562, 252–259. [Google Scholar] [CrossRef]
  42. Shakoor, M.B.; Niazi, N.K.; Bibi, I.; Shahid, M.; Sharif, F.; Bashir, S.; Shaheen, M.S.; Wang, H.; Ok, Y.S.; Rinklebe, J. Arsenic removal by natural and chemically modified watermelon rind in aqueous solutions and groundwater. Sci. Total Environ. 2018, 645, 1444–1455. [Google Scholar] [CrossRef]
  43. Pehlivan, E.; Tran, H.T.; Ouédraogo, W.K.I.; Schmidt, C.; Zachmann, D.; Bahadir, M. Sugarcane bagasse treated with hydrous ferric oxide as a potential adsorbent for the removal of As (V) from aqueous solutions. Food Chem. 2013, 138, 133–138. [Google Scholar] [CrossRef]
  44. Gupta, A.; Vidyarthi, S.R.; Sankararamakrishnan, N. Concurrent removal of As (III) and As (V) using green low cost functionalized biosorbent-Saccharum officinarum bagasse. J. Environ. Chem. Eng. 2015, 3, 113–121. [Google Scholar] [CrossRef]
  45. Wiącek, A.E.; Gozdecka, A.; Jurak, M. Physicochemical characteristics of chitosan-TiO2 biomaterial. 1. Stability and swelling properties. Ind. Eng. Chem. Res. 2018, 57, 1859–1870. [Google Scholar] [CrossRef]
  46. Li, Q.; Su, H.; Tan, T. Synthesis of ion-imprinted chitosan–TiO2 adsorbent and its multi-functional performances. Biochem. Eng. J. 2008, 38, 212–218. [Google Scholar] [CrossRef]
  47. Wu, S.; Kan, J.; Dai, X.; Shen, X.; Zhang, K.; Zhu, M. Ternary carboxymethyl chitosan-hemicellulose-nanosized TiO2 composite as effective adsorbent for removal of heavy metal contaminants from water. Fibers Polym. 2017, 18, 22–32. [Google Scholar] [CrossRef]
  48. Tao, Y.; Ye, L.; Pan, J.; Wang, Y.; Tang, B. Removal of Pb (II) from aqueous solution on chitosan/TiO2 hybrid film. J. Hazard. Mater. 2009, 161, 718–722. [Google Scholar] [CrossRef] [PubMed]
  49. Carvalho, J.; Araújo, J.; Castro, F. Alternative low-cost adsorbent for water and wastewater decontamination derived from eggshell waste: An overview. Waste Biomass Valor. 2011, 2, 157–167. [Google Scholar] [CrossRef] [Green Version]
  50. Özsin, G.; Kılıç, M.; Apaydın-Varol, E.; Pütün, A.E. Chemically activated carbon production from agricultural waste of chickpea and its application for heavy metal adsorption: Equilibrium, kinetic, and thermodynamic studies. Appl. Water Sci. 2019, 9, 56. [Google Scholar] [CrossRef] [Green Version]
  51. Yu, J.; Kang, Y.; Yin, W.; Fan, J.; Guo, Z. Removal of antibiotics from aqueous solutions by a carbon adsorbent derived from protein-waste-doped biomass. ACS Omega 2020. [Google Scholar] [CrossRef]
  52. Cruz, G.J.F.; Mondal, D.; Rimaycuna, J.; Soukup, K.; Gomez, M.M.; Solis, J.L.; Lang, J. Agrowaste derived biochars impregnated with ZnO for removal of arsenic and lead in water. J. Environ. Chem. Eng. 2020, 8, 103800. [Google Scholar] [CrossRef]
  53. Pincus, L.N.; Melnikov, F.; Yamani, J.S.; Zimmerman, J.B. Multifunctional photoactive and selective adsorbent for arsenite and arsenate: Evaluation of nano titanium dioxide–enabled chitosan cross-linked with copper. J. Hazard. Mater. 2018, 358, 145–154. [Google Scholar] [CrossRef]
  54. Fausey, C.L.; Zucker, I.; Shaulsky, E.; Zimmerman, J.B.; Elimelech, M. Removal of arsenic with reduced graphene oxide-TiO2-enabled nanofibrous mats. Chem. Eng. J. 2019, 375, 122040. [Google Scholar] [CrossRef]
  55. Liu, H.; Zuo, K.; Vecitis, C.D. Titanium dioxide-coated carbon nanotube network filter for rapid and effective arsenic sorption. Environ. Sci. Technol. 2014, 48, 13871–13879. [Google Scholar] [CrossRef]
  56. Pant, B.; Pant, H.R.; Park, M. Fe1−xS modified TiO2 NPs embedded carbon nanofiber composite via electrospinning: A potential electrode material for supercapacitors. Molecules 2020, 25, 1075. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  57. Pant, B.; Pant, H.R.; Park, M.; Liu, Y.; Choi, J.W.; Barakat, N.A.M.; Kim, H.Y. Electrospun CdS-TiO2 doped carbon nanofibers for visible-light-induced photocatalytic hydrolysis of ammonia borane. Cata. Commun. 2014, 50, 63–68. [Google Scholar] [CrossRef]
  58. Choi, C.; Hwang, K.J.; Kim, Y.J.; Kim, G.; Park, J.Y.; Jin, S. Rice-straw-derived hybrid TiO2–SiO2 structures with enhanced photocatalytic properties for removal of hazardous dye in aqueous solutions. Nano Energy 2016, 20, 76–83. [Google Scholar] [CrossRef]
  59. Gallios, G.P.; Tolkou, A.K.; Katsoyiannis, I.A.; Stefusova, K.; Vaclavikova, M.; Deliyanni, E.A. Adsorption of arsenate by nanoscaled activated carbon modified by iron and manganese oxides. Sustainability 2017, 9, 1684. [Google Scholar] [CrossRef] [Green Version]
  60. Barkat, M.; Nibou, D.; Chegrouche, S.; Mellah, A. Kinetics and thermodynamics studies of chromium(VI) ions adsorption onto activated carbon from aqueous solutions. Chem. Eng. Process. Process Intensif. 2009, 48, 38–47. [Google Scholar] [CrossRef]
  61. Yantasee, W.; Warner, C.L.; Sangvanich, T.; Addleman, R.S.; Carter, T.G.; Wiacek, R.J.; Fryxell, G.E.; Timchalk, C.; Warner, M.G. Removal of heavy metals from aqueous systems with thiol functionalized superparamagnetic nanoparticles. Environ. Sci. Technol. 2007, 41, 5114–5119. [Google Scholar] [CrossRef]
  62. Langmuir, I. The constitution and fundamental properties of solids and liquids. JACS 1916, 38, 2221–2295. [Google Scholar] [CrossRef] [Green Version]
  63. Freundlich, H. Uber die adsorption in losungen. J. Phys. Chem. 1916, 57, 385–470. [Google Scholar] [CrossRef]
  64. Chandra, V.; Park, J.; Chun, Y.; Lee, J.W.; Hwang, I.C.; Kim, K.S. Water-dispersible magnetite-reduced graphene oxide composites for arsenic removal. ACS Nano 2010, 4, 3979–3986. [Google Scholar] [CrossRef]
  65. Feng, L.; Cao, M.; Ma, X.; Zhu, Y.; Hu, C. Superparamagnetic high–surface–area Fe3O4 nanoparticles as adsorbents for arsenic removal. J. Hazard Mater. 2012, 217–218, 439–446. [Google Scholar] [CrossRef] [PubMed]
  66. Chen, W.; Parette, R.; Zou, J.; Cannon, F.S.; Dempsey, B.A. Arsenic removal by iron modified activated carbon. Water Res. 2007, 41, 1851–1858. [Google Scholar] [CrossRef] [PubMed]
  67. Lenoble, V.; Bouras, O.; Deluchat, V.; Serpaud, B.; Bollinger, J.C. Arsenic adsorption onto pillared clays and iron oxides. J. Colloid Interface Sci. 2002, 255, 52–58. [Google Scholar] [CrossRef] [PubMed]
  68. Liu, X.; Ao, H.; Xiong, X.; Xiao, J.; Liu, J. Arsenic removal from water by iron–modified bamboo charcoal. Water Air Soil Pollut. 2012, 223, 1033–1044. [Google Scholar] [CrossRef]
  69. Setyono, D.; Valiyaveettil, S. Chemically modified sawdust as renewable adsorbent for arsenic removal from water. ACS Sustain. Chem. Eng. 2014, 2, 2722–2729. [Google Scholar] [CrossRef]
  70. Manju, G.N.; Raji, C.; Anirudhan, T.S. Evaluation of coconut husk carbon for the removal of arsenic from water. Water Res. 1998, 32, 3062–3070. [Google Scholar] [CrossRef]
  71. Xia, C.L.; Jing, Y.; Jia, Y.Z.; Yue, D.Y.; Ma, J.; Yin, X.J. Adsorption properties of congo red from aqueous solution on modified hectorite: Kinetic and thermodynamic studies. Desalination 2011, 265, 81–87. [Google Scholar] [CrossRef]
  72. Ghaedi, M.; Hajati, S.; Karimi, F.; Barazesh, B.; Ghezelbash, G. Equilibrium, kinetic and isotherm of some metal ion biosorption. J. Ind. Eng. Chem. 2013, 19, 987–992. [Google Scholar] [CrossRef]
  73. Cai, X.; Li, J.; Liu, Y.; Yan, Z.; Tan, X.; Liu, S.; Zeng, G.; Gu, Y.; Hu, X.; Jiang, L. Titanium dioxide-coated biochar composites as adsorptive and photocatalytic degradation materials for the removal of aqueous organic pollutants. J. Chem. Technol. Biotechnol. 2018, 93, 783–791. [Google Scholar] [CrossRef]
  74. Ajith, N.; Bhattacharyya, K.; Ipte, P.R.; Satpati, A.K.; Tripathi, A.K.; Verma, R.; Swain, K.K. Interaction of arsenic (III) and arsenic (V) on manganese dioxide: XPS and electrochemical investigations. J. Environ. Sci. Health A 2019, 54, 277–285. [Google Scholar] [CrossRef]
  75. Chong, M.N.; Jin, B.; Chow, C.W.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  76. Su, H.; Lv, X.; Zhang, Z.; Yu, J.; Wang, T. Arsenic removal from water by photocatalytic functional Fe2O3–TiO2 porous ceramic. J. Porous Mater. 2017, 24, 1227–1235. [Google Scholar] [CrossRef]
  77. Paudyal, H.; Ohto, K.; Kawakita, H.; Inoue, K. Recovery of fluoride from water through adsorption using orange–waste gel, followed by desorption using saturated lime water. J. Matter. Cycles Waste Manag. 2020, 22, 1484–1491. [Google Scholar] [CrossRef]
  78. Paudyal, H.; Pangeni, B.; Inoue, K.; Harada, H.; Kawakita, H.; Ohto, K.; Alam, S. Adsorptive removal of trace concentration of fluoride using orange waste treated using concentrated sulfuric acid. IJMSA 2017, 6, 212. [Google Scholar] [CrossRef]
  79. Zhang, S.; Lu, X. Treatment of wastewater containing reactive brilliant blue KN-R using TiO2/BC composite as heterogeneous photocatalyst and adsorbent. Chemosphere 2018, 206, 777–783. [Google Scholar] [CrossRef]
  80. Zhang, H.; Wang, Z.; Li, R.; Guo, J.; Li, Y.; Zhu, J.; Xie, X. TiO2 supported on reed straw biochar as an adsorptive and photocatalytic composite for the efficient degradation of sulfamethoxazole in aqueous matrices. Chemosphere 2017, 185, 351–360. [Google Scholar] [CrossRef]
  81. Pant, B.; Park, M.; Park, S.J. Recent advances in TiO2 films prepared by sol-gel methods for photocatalytic degradation of organic pollutants and antibacterial activities. Coatings 2019, 9, 613. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (a) SEM image of PP; (b) SEM image of PP@TiO2; (c) SEM image of As(III) adsorbed PP@TiO2; (d) EDS spectra of PP; (e) EDS spectra of PP@TiO2; (f) EDS spectra of As//PP@TiO2.
Figure 1. (a) SEM image of PP; (b) SEM image of PP@TiO2; (c) SEM image of As(III) adsorbed PP@TiO2; (d) EDS spectra of PP; (e) EDS spectra of PP@TiO2; (f) EDS spectra of As//PP@TiO2.
Catalysts 10 01125 g001
Figure 2. (a) XRD spectra of PP and PP@TiO2; (b) FTIR spectra of PP, PP@TiO2, and As// PP@TiO2.
Figure 2. (a) XRD spectra of PP and PP@TiO2; (b) FTIR spectra of PP, PP@TiO2, and As// PP@TiO2.
Catalysts 10 01125 g002
Figure 3. (a) Adsorption isotherm of As(III) using PP@TiO2 from water; (b) modeling of experimental data using Langmuir isotherm; (c) modeling of experimental data using Freundlich isotherm; (d) variation of RL with initial As(III) concentration. (Conditions: V/m = 1000 mL/g, concentration of As(III) = 10–600 mg/L, pH = 7, T = 25 °C, contact time = 24 h).
Figure 3. (a) Adsorption isotherm of As(III) using PP@TiO2 from water; (b) modeling of experimental data using Langmuir isotherm; (c) modeling of experimental data using Freundlich isotherm; (d) variation of RL with initial As(III) concentration. (Conditions: V/m = 1000 mL/g, concentration of As(III) = 10–600 mg/L, pH = 7, T = 25 °C, contact time = 24 h).
Catalysts 10 01125 g003
Figure 4. Effect of pH for the adsorption of As(III) using PP@TiO2 from water: (a) Adsorption % of As(III) by PP@TiO2; (b) Distribution coefficient (KD) of As(III) adsorption at different pH. (Conditions: m = 25 mg, V = 25 mL, V:m = 1000 mL/g, T = 25 °C, contact time = 24 h).
Figure 4. Effect of pH for the adsorption of As(III) using PP@TiO2 from water: (a) Adsorption % of As(III) by PP@TiO2; (b) Distribution coefficient (KD) of As(III) adsorption at different pH. (Conditions: m = 25 mg, V = 25 mL, V:m = 1000 mL/g, T = 25 °C, contact time = 24 h).
Catalysts 10 01125 g004
Figure 5. Adsorption kinetics curves for As(III) using PP@TiO2: (a) equilibrium concentration and % adsorption of As(III) as a function of time; (b)adsorption capacity (qt) with contact time; (c) pseudo-1st-order kinetics plot for As(III) adsorption; (d) pseudo-2nd-order kinetics plot for As(III) adsorption. (Conditions: m = 25 mg, V = 25 mL, V/m = 1000 mL/g, concentration of As(III) = 30.20 mg/L, pH = 7, T= 25 °C).
Figure 5. Adsorption kinetics curves for As(III) using PP@TiO2: (a) equilibrium concentration and % adsorption of As(III) as a function of time; (b)adsorption capacity (qt) with contact time; (c) pseudo-1st-order kinetics plot for As(III) adsorption; (d) pseudo-2nd-order kinetics plot for As(III) adsorption. (Conditions: m = 25 mg, V = 25 mL, V/m = 1000 mL/g, concentration of As(III) = 30.20 mg/L, pH = 7, T= 25 °C).
Catalysts 10 01125 g005
Figure 6. XPS spectra of PP@TiO2 after As(III) adsorption: (a) Wide scan; (b) Ti 2p; (c) As 3d; (d) O 1s.
Figure 6. XPS spectra of PP@TiO2 after As(III) adsorption: (a) Wide scan; (b) Ti 2p; (c) As 3d; (d) O 1s.
Catalysts 10 01125 g006
Figure 7. Effect of common coexisting anions on adsorption of As(III) using PP@TiO2. (Conditions: Coexisting ions concentration = 10, 50, 100 mg/L; initial As(III) concentration = 25 mg/L; pH = 7; V:m = 1000 mL/g, T = 25 °C, contact time = 24 h).
Figure 7. Effect of common coexisting anions on adsorption of As(III) using PP@TiO2. (Conditions: Coexisting ions concentration = 10, 50, 100 mg/L; initial As(III) concentration = 25 mg/L; pH = 7; V:m = 1000 mL/g, T = 25 °C, contact time = 24 h).
Catalysts 10 01125 g007
Figure 8. (a) Desorption of arsenic from arsenic loaded PP@TiO2. (Conditions: weight of loaded PP@TiO2 = 25 mg, volume of eluent = 25 mL, V/m = 1000 mL/g, adsorbed amount of arsenic = 20.34 mg/L, T = 25 °C, contact time = 24 h); (b) variation of the adsorption capacity of the PP@TiO2 in subsequent adsorption-desorption cycles of As(III).
Figure 8. (a) Desorption of arsenic from arsenic loaded PP@TiO2. (Conditions: weight of loaded PP@TiO2 = 25 mg, volume of eluent = 25 mL, V/m = 1000 mL/g, adsorbed amount of arsenic = 20.34 mg/L, T = 25 °C, contact time = 24 h); (b) variation of the adsorption capacity of the PP@TiO2 in subsequent adsorption-desorption cycles of As(III).
Catalysts 10 01125 g008
Table 1. Langmuir and Freundlich isotherm parameters for the adsorption of As(III) onto PP@TiO2.
Table 1. Langmuir and Freundlich isotherm parameters for the adsorption of As(III) onto PP@TiO2.
AdsorbateAdsorbentLangmuir ModelFreundlich Model
qm (mg/g)b (L/mg)R2KF (mg/g)nR2
As(III)PP@TiO276.920.030.9998.722.720.954
Table 2. Comparison of the maximum adsorption capacity (qm) of different adsorbents for As(III).
Table 2. Comparison of the maximum adsorption capacity (qm) of different adsorbents for As(III).
AdsorbentOptimum
pH
qm
(mg/g)
Reference
Orange juice residue10.068.16[21]
Watermelon rind8.23.40[42]
Thiol functionalized sugarcane bagasse728.57[44]
Granular titanium dioxide732.4[37]
Fe3O4 nanoparticles746.06[65]
Iron–modified activated carbon7.6–8.038.8[66]
Amorphous iron hydroxide6–828.0[67]
Fe3O4/sugarcane bagasse activated carbon composite86.69[16]
ZrO2 nanosheets674.9[9]
Iron modified bamboo charcoal4–57.23[68]
Fe(III) loaded pomegranate waste950.0[18]
Al-based MOF graphene–oxide nanocomposite6.165.0[17]
ZrO2–sawdust729.0[69]
Copper–impregnated coconut husk carbon6.520.35[70]
TiO2 impregnated pomegranate peels (PP@TiO2)776.92This study
Table 3. Kinetic parameters for the adsorption of As(III) onto PP@TiO2.
Table 3. Kinetic parameters for the adsorption of As(III) onto PP@TiO2.
OrderAdsorbateR2qe (exp)
(mg/g)
qe (cal)
(mg/g)
k1
(min–1)
k2
(mg/g/min)
Pseudo–2ndAs(III)0.99924.725.511.32 × 10−3
Pseudo–1stAs(III)0.93224.715.459.85 × 10−3

Share and Cite

MDPI and ACS Style

Poudel, B.R.; Aryal, R.L.; Bhattarai, S.; Koirala, A.R.; Gautam, S.K.; Ghimire, K.N.; Pant, B.; Park, M.; Paudyal, H.; Pokhrel, M.R. Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water. Catalysts 2020, 10, 1125. https://doi.org/10.3390/catal10101125

AMA Style

Poudel BR, Aryal RL, Bhattarai S, Koirala AR, Gautam SK, Ghimire KN, Pant B, Park M, Paudyal H, Pokhrel MR. Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water. Catalysts. 2020; 10(10):1125. https://doi.org/10.3390/catal10101125

Chicago/Turabian Style

Poudel, Bhoj Raj, Ram Lochan Aryal, Sitaram Bhattarai, Agni Raj Koirala, Surendra Kumar Gautam, Kedar Nath Ghimire, Bishweshwar Pant, Mira Park, Hari Paudyal, and Megh Raj Pokhrel. 2020. "Agro-Waste Derived Biomass Impregnated with TiO2 as a Potential Adsorbent for Removal of As(III) from Water" Catalysts 10, no. 10: 1125. https://doi.org/10.3390/catal10101125

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