Skip to content
Publicly Available Published by De Gruyter July 25, 2017

Bio-adsorbent derived from papaya peel waste and magnetic nanoparticles fabricated for lead determination

  • Sahar Abbaszadeh , Hamid Rashidi Nodeh and Sharifah Rafidah Wan Alwi EMAIL logo

Abstract

Determination of toxic lead ions at trace level using solid-based adsorbents has become of interest in recent years. In this work, a novel bio-adsorbent originating from papaya peel waste (PPw) and magnetic nanoparticles (Fe3O4) was developed (Fe3O4/PPw). The new adsorbent was prepared using a one-pot green method and characterized by Fourier transform infrared, X-ray diffractometer, energy-dispersive X-ray spectroscopy and field emission scanning electron microscopy. The synthesized Fe3O4/PPw was used as a magnetic solid-phase extraction (MSPE) sorbent for extraction of lead ions from waste water prior to assessing by flame atomic absorption spectroscopy. The parameters influencing extraction recovery, including desorption solvent, solvent volume, sample volume, extraction time, desorption time, adsorbent dosage, salt effect and pH were optimized. A linear response for the MSPE method was achieved at concentrations from 10 to 100 ng mL−1 with a good coefficient of determination (R2=0.9987). Detection limits and quantitation limit of the MSPE method were observed around 2 ng mL−1 and 6.6 ng mL−1, respectively. The intraday and interday precision (%RSD) was in the range 1.6%–4.5% and 2.3%–7.4%, respectively. The recovery amounts obtained were 91% for tap water, 85.9% for river water and 86% for waste water. The synthesized adsorbent showed a minimum reusability of eight cycles without significant change in the lead determination. The results proved that the new bio-adsorbent (Fe3O4/PPw) is potentially capable to extract the Pb(II) from aqueous media under optimum conditions with a high extraction efficiency.

Introduction

Recently, the great problems of environmental pollution sourced from heavy metal contamination has drawn authorities’ attention worldwide [1], [2]. Among different types of heavy metals, lead (Pb) is characterized as a highly toxic element in an aqueous environment [3], [4]. Lead enters the environment through industrial effluents such as batteries manufacturing, cosmetics and printing industries, usage in military services [5] or from natural sources, for example, lead and coal mining as well as from volcanic activities [6]. Lead can be found as Pb0 or Pb2+ [Pb(II)] in water samples. Its toxicity to the environment and human health has been known for more than two thousand years [7], [8]. Lead poses severe health risks to humans, such as cancer, reproductive system disorders, high blood pressure, heart disease and skin disease [9]. The lead threshold limit in drinking water is considered 10 ng mL−1, as appears in the European Union (EU) standard [10]. Excessive Pb(II) with a concentration far higher than the EU recommended level (129 mg L−1) has been frequently reported in water samples collected from industrial areas and rivers previously [11]. Therefore, monitoring and removal of lead from waste water before discharging to the environment is essential to protect both the environment and human health [12], [13].

Because the amount of Pb(II) existing in water samples are at trace level, so, sample preparation techniques are essential for Pb(II) analysis (i.e. extraction and preconcentration) in water samples. Solid-phase extraction (SPE) is an established method that is commonly applied for preconcentration of Pb(II) ions from aqueous solutions [14], [15], [16]. The SPE methodology offers many advantages, like simple and inexpensive operation, flexibility, high recovery and environmentally friendly procedure [17], [18]. Despite all the benefits of SPE as a standard method, time-consuming analysis and low breakthrough volume make it user unfavourable. However, utilization of magnetic nanoparticles (MNPs) has been reported as a good technique to overcome those SPEs’ drawbacks by prevention of column blocking and channeling as well as fast separation using an external magnet [19], [20]. The extraction technique based on magnetic adsorbents is called magnetic solid-phase extraction (MSPE) [21], [22]. The application of MSPE was successfully used and addressed for extraction of Pb(II) from aqueous media and food samples [23], [24], [25].

However, an adsorbent is the key parameter in the adsorption process for removal and extraction purposes. Recently, employment of adsorbents originating from biomaterial and agricultural waste has become an interesting approach in adsorption methodology due to their natural capability and inherently environmentally friendly characteristics [26], [27]. These bio-adsorbents have also been used successfully for heavy metal analysis in water decontamination processes [4], [28], [29].

In the current study, a new bio-adsorbent based on MNPs and papaya peel waste (Fe3O4/PPw) was developed and applied to determine a trace amount of Pb(II) from preconcentrated Pb(II) aqueous solution. Magnetic Fe3O4 nanoparticles were slowly grown on PPw to produce Fe3O4/PPw using a one-pot green synthesis method. Magnetic properties permit easy separation of the adsorbent using an external magnet. A good adsorption performance of Fe3O4/PPw towards Pb(II) is expected due to the electrostatic interaction. The developed Fe3O4/PPw capacity was evaluated through the batch-mode experiments and tested for the extraction of Pb(II) from waste water, tap water and river water.

Experimental procedure

Material

All the common reagents used in this study were analytical grade. Ammonia solution (25%) and lead(II) nitrate [Pb(NO3)2] were obtained from Merck (Schuchardt, Germany). Iron(III) chloride hexahydrate (FeCl3·6H2O) and iron(II) chloride (FeCl2·4H2O) were obtained from Sigma-Aldrich (St. Louis, MO, USA). Pb(II) standards (1000 μg mL−1) were prepared in double deionized water. The Pb(II) stock solution was obtained by dissolving 0.159 g of Pb(NO3)2 in 100 mL deionized water (1000 μg L−1). The diluted solutions were used for analytical purposes.

Instruments

A flame atomic absorption spectrometer (FAAS) (model 400, Perkin-Elmer, MA, USA) equipped with a lead hollow cathode lamp and pre-set wavelength of 283.31 nm, was utilized to measure the Pb(II) uptake amount. The FAAS instrument was calibrated using direct injection of different concentrations of standard solutions in the range 0.5–20 μg/mL. A Fourier transform infrared instrument (FT-IR) (NICOLET 5700, Thermo Electron Corporation, Japan) was used for FT-IR analysis in transmission mode and all the spectra were recorded in the range of 400–4000 cm−1. The crystalline structure of the prepared adsorbent was studied using a D8-Advance X-ray diffractometer (XRD) (Bruker GmbH, Karlsruhe, Germany) with Cu Kα radiation (λ=1.54060 Å) at 40 kV. Morphology and elemental analysis were performed using a scanning electron microscope (SEM) equipped with energy-dispersive X-ray spectroscopy (EDX) analysis (JSM-6701F, JEOL, Ltd Tokyo, Japan). In addition, the samples were agitated using an orbital shaker (Systec Laboratory equipment) in predefined time intervals.

Preparation of PPw

The ripened fruit of papaya (Carica papaya) was collected from the local market, washed exhaustively with distilled water to remove dirt and then peeled. The peeled samples were dried in the oven for 24 h at 105°C until they became crisp [30]. After that, the oven-dried samples were ground using an electric grinder and sieved using a 45 mesh (355 micron) [31]. The final powder was kept in an airtight container for further synthesis procedure.

Synthesis of PPw-MNPs

A simple one-pot process was used for the preparation of magnetic bio-adsorbent. PPw powder (1 g), FeCl2·4H2O (0.1 g) and FeCl3·6H2O (0.2 g) were mixed in 50 mL de-ionized water (Fig. 1). Then, the mixture was sonicated for 30 min. The sample was heated to 50°C with vigorous stirring and then the heater was turned off. Thereafter, 5 mL of ammonia (25%) was added dropwise following continuous stirring for 5 h at room temperature. The magnetic (Fe3O4/PPw) adsorbent was then washed with distilled water and methanol, and oven dried at 80°C for 24 h. The new bio-adsorbent as the final product was kept in an airtight container for further experiments.

Fig. 1: 
						Schematic procedure for preconcentration of Pb(II) using Fe3O4/PPw adsorbent.
Fig. 1:

Schematic procedure for preconcentration of Pb(II) using Fe3O4/PPw adsorbent.

Sampling method

The Fe3O4/PPw adsorbent was tested for adsorbing Pb(II) ions from waste water, tap water and river water samples. The waste water sample was obtained from an industrial area in Johor Bahru, Malaysia. The river water sample was obtained from Sungai Melana, Johor Bahru and tap water was obtained from the university laboratory (Universiti Teknologi Malaysia main campus). The collected samples were filtered using Whatman filter paper (125 mm) to remove the barriers. In addition, waste water and river water samples were also filtered using Smith filter paper (125 mm) before the adsorption process. The pH was set at 6.0 prior to analysis based on the optimum condition.

Magnetic solid-phase extraction procedure

The preconcentration procedure for Pb(II) ions from water samples was studied following a batch-wise method (Fig. 1). Two hundred milliliter of water samples including the desired concentration (10 ng mL−1) of Pb(II) ions was transferred into Erlenmeyer flasks. pH was adjusted at 6.0 using HCl (0.1 mol L−1) and NaOH (0.1 mol L−1). Then, 50 mg of bio-adsorbent (PPw-MNPs) was added to the Pb(II) solution followed by 20 min shaking (250 rpm) using an orbital shaker. Thereafter, the magnetic bio-adsorbent was separated from the solution by an external magnetic field (Fig. 2). Successively, 3 mL of desorption solvent (0.1 mol L−1 HCl) was used to desorb the Pb(II) ions from the surface of the bio-adsorbent by sonication in 2 min. Finally, desorption solvent including desorbed Pb(II) ions was magnetically separated and analysed using FAAS.

Fig. 2: 
						FT-IR spectra of PPw (a) and Fe3O4/PPw (b).
Fig. 2:

FT-IR spectra of PPw (a) and Fe3O4/PPw (b).

The extracted Pb(II) from water samples was evaluated using percent recovery (R%) and calculated by eq. (1) [32]:

(1) R % = C f V f C i V i × 100

where Ci is the initial concentration (μg mL−1) and Cf is the final concentration of lead obtained by FAAS (μg mL−1), Vi is the initial volume of water sample (mL) and Vf is final volume or desorption solvent volume (mL).

Results and discussion

Characterization of adsorbent

Surface functionality analysis (FT-IR spectroscopy)

Figure 2 shows the FT-IR spectra of PPw and Fe3O4-PPw. Figure 2a, the FT-IR spectra of PPw, shows peaks at 3413 cm−1, 2992 cm−1, 1650 cm−1, 1384 cm−1, 1125 cm−1, 1073 cm−1 and 911 cm−1, which indicate O–H stretching vibration, C–H stretching, C=O stretching for carboxylic groups, C–C/C–N stretching, C–O–C stretching vibrations, C–O stretching vibration alkoxy and C–O, respectively. Figure 2b shows an extra sharp peak at 582 cm−1, which is evidence of the existence of the Fe–O–Fe bond of Fe3O4. The appearance of a new peak (582 cm−1) and disappearance of some peaks (1125 cm−1 and 911 cm−1) in the FT-IR spectra, clearly demonstrates the successful preparation of Fe3O4-PPw. The peaks at 1059 cm−1 and 3413 cm−1 can be linked to the C–N and N–H bonds, respectively [33].

Surface crystallinity analysis (XRD analysis)

The XRD patterns of Fe3O4-PPw are shown in Fig. 3. The XRD diffraction peaks in a curve can be indexed as cubic for Fe3O4 nanoparticles (01-071-6336). Figure 3 shows that the magnetic Fe3O4 nanoparticles XRD signals appeared at angle 2θ (Miller indices) 18.32° (202), 30.14° (220), 35.51° (311), 43.15° (400), 53.44° (422), 57.09° (334), 62.68° (440), 74.16° (533) and 89.86° (400).

Fig. 3: 
							XRD patterns of the magnetic nanocomposite.
Fig. 3:

XRD patterns of the magnetic nanocomposite.

The sharp signals indicate the high crystallinity of Fe3O4 nanoparticles, this trend is reported in previous studies [34]. When Fe3O4 fabricates with the PPw, the XRD diffraction peaks for Fe3O4-PPw (Fig. 3) show a broad signal as characteristic of amorphous PPw [broad peak at 20° (002)] and several signals appeared for magnetic Fe3O4 nanoparticles. These findings indicates the crystalline structure of the nanocomposite.

Surface morphology analysis [Field emission scanning electron microscopy (FESEM)]

The morphology analysis of the developed magnetic bio-adsorbent (Fe3O4/PPw) was performed using FESEM (Fig. 4a and b). Figure 4a illustrate the SEM micrograph of the raw PPw prior to magnetization. After modification (Fig. 4b) demonstrates the appearance of magnetic Fe3O4 nanoparticles dispersed on the PPw surface. In addition, the chemical composition of Fe3O4/PPw adsorbent was analysed using EDX coupled with FESEM equipment (Fig. 5). The EDX spectra clearly shows the expected elements including carbon, oxygen and iron. The presence of iron is corresponding to the magnetic Fe3O4 nanoparticles. These findings confirm the successful preparation of the green Fe3O4/PPw adsorbent.

Fig. 4: 
							FESEM images of raw PPw (a) and micrograph of the Fe3O4/PPw (b).
Fig. 4:

FESEM images of raw PPw (a) and micrograph of the Fe3O4/PPw (b).

Fig. 5: 
							EDX result for elements analyses of Fe3O4/PPw.
Fig. 5:

EDX result for elements analyses of Fe3O4/PPw.

Optimization of the extraction method

Different process parameters influencing the Pb(II) extraction performance of the developed adsorbent, such as solution pH, salt effect, adsorbent dosage, sample volume, extraction time, desorption time, type of desorption solvent and volume of solvent were studied. The extraction efficiency (%R) was used to evaluate the MSPE performance under optimized conditions.

Effect of pH

The effect of solution pH (3.0–7.0) on the MSPE performance was studied in 50 mL water solution (100 ng mL−1 of Pb). Solution pH above 7.0 was not studied due to precipitation of Pb(II) in the samples as lead hydroxide. The results presented in Fig. 6 show that high Pb(II) recovery was obtained in the pH range 5.0–6.0. The least recovery of Pb(II) is observed at pH 3–4, is probably due to proton (H+) competition. In addition, repulsion possibly occurs between Pb(II) and positively charged receptors on the surface of the bio-adsorbent. Thus, pH 6.0 is considered as the best operating pH for further studies.

Fig. 6: 
							Effect of pH on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 6:

Effect of pH on percentage extraction of Pb(II) using Fe3O4/PPw.

Effect of salt

The effect of the aqueous solution’s ionic strength on Pb(II) extraction performance of the new adsorbent was studied by adding different concentrations of NaCl; 0.01 M, 0.05 M, 0.1 M and 0.2 M into the samples. From the results (Fig. 7), it is observed that Pb(II) ion extraction efficiency decreases with increasing NaCl concentration. This is probably due to occupation of adsorbent active sites with Na+. According to Hao et al. [35], a decline in adsorption efficiency depends on (1) ionic interaction between adsorbent and metal cations and (2) the activity coefficients of the metal ions are influenced by ionic strength. Therefore, to optimize the Pb(II) adsorption process by using the new adsorbent, the concentration of salts in the aqueous solution as an operational condition should be controlled. Hence, the Pb(II) extraction process using the newly developed adsorbent was followed without the addition of NaCl.

Fig. 7: 
							Effect of salt on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 7:

Effect of salt on percentage extraction of Pb(II) using Fe3O4/PPw.

Effect of adsorbent dosage

One of the key parameters in MSPE method is the amount of adsorbent dosage. The effect of F3O4/PPw dosage on the extraction of the Pb(II) ions was studied under optimized conditions. Figure 8 shows that the extraction efficiency rapidly increased with the increasing dosage from 10 mg to 50 mg, due to increase of adsorption sites. After that, it reached the adsorption equilibrium at 50–200 mg of adsorbent, and no further uptake was observed. This may be due to the lack of Pb(II) ions in the samples and most of 100 ng mL−1 of Pb(II) ions was extracted by 50 mg of F3O4/PPw. Thus, 50 mg of adsorbent dosage is considered as the best dosage for further analysis.

Fig. 8: 
							Effect of adsorbent dosage on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 8:

Effect of adsorbent dosage on percentage extraction of Pb(II) using Fe3O4/PPw.

Effect of extraction and desorption time

Effect of time on the extraction efficiency of the F3O4/PPw-MSPE method was studied considering a range from 1 to 90 min (Fig. 9). The maximum Pb(II) extraction was obtained after 20 min of shaking. After that, the extraction amount remained almost constant. It could be due to the extraction of the entire Pb(II) analytes in the solution by the adsorbent because the amount of analytes in each sample is constant.

Fig. 9: 
							Effect of extraction time on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 9:

Effect of extraction time on percentage extraction of Pb(II) using Fe3O4/PPw.

Effect of desorption time on extraction efficiency of the developed adsorbent was studied considering the time from 1 to 10 min. From the results, 2 min of shaking time revealed the best results and therefore, is considered as the optimum time for desorption of Pb(II) from the adsorbent.

Desorbing solvent selection

Types of desorption solvent (acids), solvent concentration (molarity) and solvent volume (mL) can influence the MSPE performance. Previously reported, the releasing of metal ions from adsorbent is important in the extraction process [36]. The effect of these parameters on the MSPE process using Fe3O4/PPw was evaluated through batch-mode experiments. For metal ions desorption like Pb(II), acidic solutions are more efficient because heavy metal cations are displaced by protons from the binding sites [37]. Therefore, the effect of different diluted acids (0.1 mol L−1), namely, HNO3, H2SO4, HCl and HNO3/HCl (1:1) on Pb(II) extraction efficiency was evaluated (Fig. 10). Among all, HCl (0.1 mol L−1) demonstrates the highest extraction recovery. In addition, the concentration of HCl was also studied in the range from 0.01 mol L−1 to 0.5 mol L−1. 0.1 mol L−1 of HCl that gives the best recovery was chosen as the best solvent for further studies. Furthermore, the effect of solvent volume on extraction efficiency was investigated in the range of 3–10 mL. Because of the dilution factor, the percentage recovery decreased continuously when the solvent volume increased from 3 to 10 mL. Therefore, it was found that 3 mL of 0.1 mol L−1 of HCl is able to complete the desorption process of Pb(II) ions from the adsorbent. Thus, 3 mL of 0.1 M HCl solution was considered as the best desorption solvent volume to desorb Pb(II) ions from the Fe3O4/PPw.

Fig. 10: 
							Effect of solvent types on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 10:

Effect of solvent types on percentage extraction of Pb(II) using Fe3O4/PPw.

In addition, the synthesized adsorbent was regenerated using HCl (0.1 mol L−1) through the adsorption–desorption cycles and after each cycle, the adsorbent was washed using 5 mL HCl (0.1 mol L−1) and 5 mL deionized water. A good reusability potential (>84%) of the Fe3O4/PPw was observed for up to 10 cycles of regeneration.

Effect of sample volume

To carry out the current experiments, each water sample typically contains a trace amount of metal ions, therefore, a high EF is essential to proceed the analysis. To obtain high EF, a large volume of water sample is needed. The maximum applicable volume for water samples (sample volume) in MSPE method was studied in a batch process. Accordingly, different sample volumes from 10 to 200 mL [containing 0.1 ng mL−1 of Pb(II) ions] were studied with the MSPE method under optimum conditions [pH 6, 50 mg adsorbent, 20 min adsorption time and 3 mL HCl (0.1 mol L−1) as desorption solvent]. As presented in Fig. 11, the extraction recovery, regardless of volume, has almost the same efficiency rate when sample volume differs from 10 to 200 mL. The EF was calculated as the ratio of the highest sample volume (200 mL) to the final volume of solvent (3 mL) [38] and it is found to be 66 for the Pb(II) ions.

Fig. 11: 
							Effect of different sample volume on percentage extraction of Pb(II) using Fe3O4/PPw.
Fig. 11:

Effect of different sample volume on percentage extraction of Pb(II) using Fe3O4/PPw.

Effect of co-existing ions

There are various types of metal and non-metal ions in water, the aqueous solutions, i.e. Cl, NO3, SO42−, PO43−, Na+, K+, Cu2+, Zn2+, Ca2+, Fe3+ and Al3+. To conduct the selectivity study of the proposed MSPE method, the effect of co-existing ions was also evaluated in this study according to Wang et al.’s previous work [39]. A 200 mL sample of Pb(II) [0.1 ng mL−1 of Pb(II)] containing other co-existing ions of 2000-fold Cl, NO3, PO43−, Na+, K+, Fe3+ and Al3+; 1000-fold Zn2+ and Ca2+; 500-fold SO42−, was used to run the batch-mode experiments. The results obtained show high recovery (81%) for Pb(II) in the presence of various co-existing ions using the developed adsorbent (Fe3O4/PPw).

Analytical performance

Before MSPE method validation, the FAAS instrument was calibrated using different concentrations of Pb(II) standard solutions. However, in the linear range of 0.5–20 mg L−1, the coefficient of determination (R2) obtained was 0.999 for Pb(II) ions. The standard deviation for the blank sample (SDblank, distilled water) was obtained in the range 0.0006–0.003 ng mL−1 (n=5). The instrument’s limit of detection (LOD) was calculated at 0.2 ng mL−1, (LOD=3×SDblank).

The analytical performance of the MSPE method was validated in terms of linearity, LOD, limit of quantification (LOQ), EF, intraday and interday precision (repeatability and reproducibility) and accuracy (real sample recovery). These parameters were studied under optimum condition [pH 6, 50 mg adsorbent, 20 min adsorption time, 3 mL HCl (0.1 mol L−1) as desorption solvent and 200 mL sample volume]. The MSPE method based on Fe3O4/PPw was calibrated at different concentrations of Pb(II) with the linearity range from 10 to 100 ng mL−1. The coefficient of determination (R2) obtained was 0.9961. LOD and LOQ were calculated to be 2 ng mL−1 (3×SD/m) and 6.6 ng mL−1 (10×SD/m), respectively, where SD is the standard deviation of the blank and m is the slope of the method calibration curve. The LOD obtained (2 ng mL−1) is well below the maximum residue limit (MRL) set by the EU method (10 ng mL−1) for Pb(II). Thus, this method is applicable for the detection of Pb(II) at the set EU method. The EF obtained was 66 (ratio of sample volume to desorption solvent volume). The satisfactory precision (RSD%) of the MSPE technique was achieved by using intraday and interday method in the range 1.7%–4.9% (n=3) and 3.6%–7.2% (n=9), respectively. This exhibits the good repeatability and reproducibility of the Fe3O4/PPw-MSPE.

Field application

To describe the field applicability of the developed Fe3O4/PPw-MSPE method, it was applied for the identification of Pb(II) ions in water samples (tap, river and industrial waste water). The water sample was spiked with Pb(II) standard solution (10 ng mL−1) then batch-wise extraction was used at optimum conditions. In addition, water samples were analysed using the proposed MSPE method without addition of a standard solution (unspiked). The found native Pb(II) ions in unspiked samples and relative recoveries (%RR) for spiked samples were calculated and listed in Table 1. In addition, %RSDs (n=3) for unspiked and spiked samples are shown in brackets. Table 1 shows the native Pb(II) ions that are found in the waste water is 40× higher than MRLs level. However, it is found to be much lower than the maximum allowance in tap water and river water. For spiked samples, %RR were obtained in the range 75.3%–98.5%. This demonstrates that the proposed Fe3O4/PPw-MSPE method was suitable for determination of trace Pb(II) ions in different real water samples. However, the low recovery (75.3%) for industrial waste water is probably due to the presence of possible matrices.

Table 1:

Pb(II) ions analysis in real water samples in terms of relative recovery (%RR) and precision (%RSD, n=3).

MSPE adsorbent Real water samples Native Pb(II) found in unspiked samples (ng mL−1) %RR (±RSD%, n=3)
Spiked 10 ng mL−1 of Pb(II)
Fe3O4/PPw Tap water 0.4 88.6 (2.7)
River water 1.3 85.9 (3.2)
Industrial waste water 400 86.9 (9.3)

Comparison with other results

For validation of this study, the result of this work was compared with other reports with a similar adsorbent. Table 2 gives a comparison of the lowest concentration obtained by using the in-house magnetic SPE sorbent (Fe3O4/PPw) compared with the results obtained using other kinds of magnetic graphene oxide-based sorbent. The sensitivity of the proposed method is better than other related works, between 10 and 100 times. The LOD obtained using the proposed adsorbent in this study was higher than the others. Because the ICP-OES instrument (used in the last study, stated in Table 2, of EG-TMSP/Fe3O4) is more sensitive than the FAAS (used in this study), a lower LOD is demonstrated for Fe3O4/PPw.

Table 2:

Comparison of LOD for current study using Fe3O4/PPw-MSPE with other sorbents.

SPE sorbent LODs (ng mL−1) Matrix Detector Reference
Fe3O4/PPw 2 Water FAASf This study
M-PhCPa 2.7 Water FAAS [40]
MWCNTs-NiFe2O4b 0.5 Water FAAS [41]
Fe3O4/surfactant 0.74 Water, soil FAAS [42]
M-MOFc 1.1 Water, fish FAAS [43]
L-IIPd 0.9 Water, food FAAS [44]
MOF-Ag12 0.5 Water FAAS [45]
EG-TMSP/Fe3O4e 0.08 Water, food ICP-OESg [46]
  1. aMagnetic phosphorus-containing polymer, bmultiwalled carbon nanotubes decorated with NiFe2O4 nanoparticles, cmagnetic metal–organic framework, dlead imprinted polymer, eethylene glycol bis-mercaptoacetate/3-(trimethoxysilyl)-1-propanethiol-coated Fe3O4 nanoparticles, fflame atomic absorption spectrometry and ginductively coupled plasma atomic emission spectroscopy.

Conclusion

Bio-adsorbent based on magnetic nanoparticles and papaya peel waste (F3O4/PPw) has been synthesized using a one-pot co-precipitation method. The proposed magnetic adsorbent was proven successful for preconcentration of Pb(II) ions from aqueous media. The experimental results achieved using the F3O4/PPw-MSPE reveals low LOD (5 ng mL−1) and high enrichment factor (66). The proposed method was applied successfully for adsorption of Pb(II) from real water samples (industrial waste water, tap and river water). Pb(II) ions were found to be 400 ng mL−1, 1.3 ng mL−1 and 0.4 ng mL−1 in waste water, tap and river water, respectively. However, the spiked Pb(II) in the three water samples is well identified with high recovery of more than 85.9% obtained. To conclude, this study reveals the feasibility and potential of F3O4/PPw as new bio-adsorbent sorbents for MSPE techniques. In addition, it can also help as a new platform for future development of other bio-adsorbent for water treatment and add value to used waste material for heavy metal pollution control in environmental applications.


Article note

A collection of invited papers based on presentations at the 6th international IUPAC Conference on Green Chemistry (ICGC-6), Venice (Italy), 4–8 September 2016.


Acknowledgment

The authors would like to thank the UNESCO/PhosAgro/IUPAC financial support through the UNESCO/PhosAgro/IUPAC grant in Green Chemistry by contract number 4500254540, Fundamental Research Grant R.J130000.7809.4F488 from Ministry of Education and Universiti Teknologi Malaysia for facilitations.

References

[1] G. Ungureanu, S. Santos, R. Boaventura, C. Botelho. J. Environ. Manage.151, 326 (2015).Search in Google Scholar

[2] R. J. Gohari, W. J. Lau, T. Matsuura, A. F. Ismail. Sep. Purif. Technol.118, 64 (2013).Search in Google Scholar

[3] P. Geetha, M. S. Latha, S. S. Pillai, M. Koshy. Ecotoxicol. Environ. Saf.122, 17 (2015).Search in Google Scholar

[4] N. Wang, X. Xu, H. Li, Q. Wang, L. Yuan, H. Yu. Bioresour. Technol.233, 58 (2017).Search in Google Scholar

[5] R. R. Ray. Proceedings of the Zoological Society, Springer, p. 161 (2016).Search in Google Scholar

[6] Y. Wang, L. Li, C. Luo, X. Wang, H. Duan. Int. J. Biol. Macromol.86, 505 (2016).Search in Google Scholar

[7] R. J. Gohari, W. J. Lau, T. Matsuura, E. Halakoo, A. F. Ismail. Sep. Purif. Technol.120, 59 (2013).Search in Google Scholar

[8] K. Chen, J. He, Y. Li, X. Cai, K. Zhang, T. Liu, Y. Hu, D. Lin, L. Kong, J. Liu. J. Colloid Interface Sci.494, 307 (2017).Search in Google Scholar

[9] R. Hu, X. Wang, S. Dai, D. Shao, T. Hayat, A. Alsaedi. Chem. Eng. J.260, 469 (2015).Search in Google Scholar

[10] Euoropean Comunitity. Off. J. Eur. CommunitiesL330, 32 (1998).Search in Google Scholar

[11] J. Chawla, R. Kumar, I. Kaur. J. Water Supply Res. Technol.64, 641 (2015).Search in Google Scholar

[12] C. J. Madadrang, H. Y. Kim, G. Gao, N. Wang, J. Zhu, H. Feng, M. Gorring, M. L. Kasner, S. Hou. ACS Appl. Mater. Interfaces4, 1186 (2012).Search in Google Scholar

[13] H. Sereshti, M. Kermani, M. Karimi, S. Samadi. Clean–Soil, Air, Water42, 1089 (2014).Search in Google Scholar

[14] E. Yavuz, Ş. Tokalıoğlu, H. Şahan, Ş. Patat. Food Chem.194, 463 (2016).Search in Google Scholar

[15] E. Yilmaz, I. Ocsoy, N. Ozdemir, M. Soylak. Anal. Chim. Acta906, 110 (2016).Search in Google Scholar

[16] Ş. Tokalıoğlu, E. Yavuz, S. Demir, Ş. Patat. Food Chem.237, 707 (2017).Search in Google Scholar

[17] W. A. Wan Ibrahim, H. Rashidi Nodeh, M. M. Sanagi. Crit. Rev. Anal. Chem.46, 267 (2016).Search in Google Scholar

[18] M. Behbahani, F. Najafi, M. M. Amini, O. Sadeghi, A. Bagheri, P. G. Hassanlou. J. Ind. Eng. Chem.20, 2248 (2014).Search in Google Scholar

[19] H. Bagheri, A. Roostaie, M. Y. Baktash. Anal. Chim. Acta816, 1 (2014).Search in Google Scholar

[20] H. Bagheri, H. Piri-Moghadam, S. Rastegar. RSC Adv.4, 52590 (2014).Search in Google Scholar

[21] W. A. Wan Ibrahim, H. Rashidi Nodeh, H. Y. Aboul-Enein, M. M. Sanagi. Crit. Rev. Anal. Chem.45, 270 (2015).Search in Google Scholar

[22] W. I. Wan Aini, R. N. Hamid. ICOWOBASS-RAFSS, Ibnu Sina Institute for Fundamental Science Studies, Universiti Teknologi Malaysia, Johor Bahru, Malaysia, p. 134 (2013).Search in Google Scholar

[23] M. Khan, E. Yilmaz, B. Sevinc, E. Sahmetlioglu, J. Shah, M. R. Jan, M. Soylak. Talanta146, 130 (2016).Search in Google Scholar

[24] R. Sharma, A. Sarswat, C. U. Pittman, D. Mohan. RSC Adv.7, 8606 (2017).Search in Google Scholar

[25] E. Yılmaz, M. Soylak. Talanta158, 152 (2016).Search in Google Scholar

[26] E.-K. Guechi, O. Hamdaoui. Desalin. Water Treat.51, 3371 (2013).Search in Google Scholar

[27] K. Wantala, N. Pojananukij, P. Sriprom, T. Kumsaen, A. Neramittagapong, S. Neramittagapong. Key Engineering Materials, Trans. Tech. Publ, Ukraine, p. 101 (2013).Search in Google Scholar

[28] T. Jayaramudu, K. Varaprasad, H. C. Kim, A. Kafy, J. W. Kim, J. Kim. Carbohydr. Polym.171, 183 (2017).Search in Google Scholar

[29] M. W. Yap, N. M. Mubarak, J. N. Sahu, E. C. Abdullah. J. Ind. Eng. Chem.45, 287 (2017).Search in Google Scholar

[30] S. Abbaszadeh, S. R. Wan Alwi, N. Ghasemi, H. Rashidi Nodeh, C. Colin Webb, I. I. Muhamad. Chem. Eng. Trans.45, 961 (2015).Search in Google Scholar

[31] S. Abbaszadeh, S. R. W. Alwi, C. Webb, N. Ghasemi, I. I. Muhamad. J. Clean. Prod.118, 210 (2016).Search in Google Scholar

[32] S. Mahpishanian, H. Sereshti. J. Chromatogr. A1443, 43 (2016).Search in Google Scholar

[33] N. Mohamad, N. A. Buang, N. A. Mahat, J. Jamalis, F. Huyop, H. Y. Aboul-Enein, R. A. Wahab. J. Ind. Eng. Chem.32, 99 (2015).Search in Google Scholar

[34] H. Rashidi Nodeh, W. A. Wan Ibrahim, M. A. Kamboh, M. M. Sanagi. RSC Adv.5, 76424 (2015).Search in Google Scholar

[35] L. Hao, H. Song, L. Zhang, X. Wan, Y. Tang, Y. Lv. J. Colloid Interface Sci.369, 381 (2012).Search in Google Scholar

[36] J. P. Chen, L. Yang. Ind. Eng. Chem. Res.44, 9931 (2005).Search in Google Scholar

[37] A. Gundogdu, D. Ozdes, C. Duran, V. N. Bulut, M. Soylak, H. B. Senturk. Chem. Eng. J.153, 62 (2009).Search in Google Scholar

[38] T. Daşbaşı, Ş. Saçmacı, A. Ülgen, Ş. Kartal. Food Chem.174, 591 (2015)Search in Google Scholar

[39] Y. Wang, H. Chen, J. Tang, G. Ye, H. Ge, X. Hu. Food Chem.181, 191 (2015).Search in Google Scholar

[40] E. Yilmaz, R. M. Alosmanov, M. Soylak. RSC Adv.5, 33801 (2015).Search in Google Scholar

[41] A. A. Ensafi, S. Rabiei, B. Rezaei, A. R. Allafchian. Anal. Methods5, 3903 (2013).Search in Google Scholar

[42] N. Jalbani, M. Soylak. Ecotoxicol. Environ. Saf.102, 174 (2014).Search in Google Scholar

[43] M. R. Sohrabi, Z. Matbouie, A. A. Asgharinezhad, A. Dehghani. Microchim. Acta180, 589 (2013).Search in Google Scholar

[44] M. Behbahani, P. G. Hassanlou, M. M. Amini, H. R. Moazami, H. S. Abandansari, A. Bagheri, S. H. Zadeh. Food Anal. Methods8, 558 (2015).Search in Google Scholar

[45] M. Salarian, A. Ghanbarpour, M. Behbahani, S. Bagheri, A. Bagheri. Microchim. Acta181, 999 (2014).Search in Google Scholar

[46] M. H. Mashhadizadeh, M. Amoli-Diva, M. R. Shapouri, H. Afruzi. Food Chem.151, 300–5 (2014).Search in Google Scholar

Published Online: 2017-07-25
Published in Print: 2018-01-26

©2018 IUPAC & De Gruyter. This work is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International License. For more information, please visit: http://creativecommons.org/licenses/by-nc-nd/4.0/

Downloaded on 23.5.2024 from https://www.degruyter.com/document/doi/10.1515/pac-2017-0503/html
Scroll to top button