Preparation and Characterization of Mn0.4Zn0.6Fe2O4 Nanoparticles

Feb 20, 2018 - The CCD allowed the optimization and the scale-up of the process, which presented a good correlation between large and small scales. Ad...
9 downloads 11 Views 13MB Size
Research Article Cite This: ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

pubs.acs.org/journal/ascecg

Preparation and Characterization of Mn0.4Zn0.6Fe2O4 Nanoparticles Supported on Dead Cells of Yarrowia lipolytica as a Novel and Efficient Adsorbent/Biosorbent Composite for the Removal of Azo Food Dyes: Central Composite Design Optimization Study Arash Asfaram,†,‡ Mehrorang Ghaedi,*,† Kheibar Dashtian,† and Gholam Reza Ghezelbash§ †

Chemistry Department, Yasouj University, Yasouj, 75918-74831, Iran Medicinal Plants Research Center, Yasuj University of Medical Sciences, Yasuj, Iran § Biology Department, Faculty of Science, Shahid Chamran University, Ahvaz, 61357-831351, Iran ‡

S Supporting Information *

ABSTRACT: The removal of hazardous dyes is of great importance to making healthy and drinkable water. Here, a new ferromagnetic composite based on Mn0.4Zn0.6Fe2O4 nanoparticles (NPs) supported on dead Yarrowia lipolytica ISF7 (D-YL-ISF7) was prepared. Nanoparticle aggregation was inhibited using D-YL-ISF7, which causes the availability of more active sites. The dead D-YL-ISF7-supported Mn0.4Zn0.6Fe2O4 nanoparticles (NPs) were characterized by Fourier transform infrared (FT-IR), X-ray diffraction (XRD), field-emission scanning electron microscopy (FESEM), high-resolution transmission electron microscopy (HRTEM), energy dispersive X-ray (EDX), Brunuaer−Emmett−Teller (BET), and vibrating sample magnetometer (VSM) analysis and used as robust adsorbents/biosorbents to simultaneously remove tartrazine (TA) and ponceau 4R (P4R) azo food dyes in their binary solution. Firstorder derivative spectrophotometry was implemented for the simultaneous analysis of dyes in binary mixtures. Central composite design (CCD) was used to evaluate the influence of pH, sonication time, Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 mass, and initial TA and P4R concentrations on the efficiency for the removal of the studied dyes. At optimum conditions (pH 2.0, sonication time 5 min, Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 mass 0.015 g, TA concentration 12 mg L−1 and P4R concentration 16 mg L−1), high removal efficiencies (>99.0%) were obtained for TA and P4R dyes, reasonably well predicted by the model. The CCD allowed the optimization and the scale-up of the process, which presented a good correlation between large and small scales. Adsorption isotherm data fitted well to the Langmuir model. Under ultrasound, the Langmuir adsorption capacity of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 was obtained to be 90.827 mg g−1 for TA and 101.461 mg g−1 for P4R. A pseudo-second-order reaction model was chosen for kinetic study. KEYWORDS: Azo food dyes, Mn0.4Zn0.6Fe2O4-NPs, Dead Yarrowia lipolytica ISF7, Isotherm and kinetic studies, Dye removal



INTRODUCTION Old and modern industries such as textiles, dyestuff, paper, leather, and plastics generate a lot of dye-containing wastewater which enters into the environment and leads to serious problems such as severe damage to human health (liver, digestive, central nervous systems and types of mutations, cancers, allergies, irritation, and itching).1−3 Wastewaters originating from the above-mentioned industries are highly colored and strongly affect underground water quality leading to serious harm to both aquatic and human life.4−6 Therefore, it is necessary and important to treat the wastewaters before they enter into the environment using available techniques such as chemical oxidation, biosorption, membrane filtration, ozonation, photocatalytic degradation, coagulation−flocculation, and adsorption.7,8 Adsorption is the most widely used © 2018 American Chemical Society

due to simplicity in handling, less time consumption, and cost effectiveness. Therefore, different adsorbents such as activated carbon, graphene oxide, multiwalled carbon nanotubes, inorganic materials/minerals, mesoporous silica materials, and biomaterials have been traditionally used to remove dyes.9−11 However, they suffer from drawbacks such as difficulties in separation from the process solution resulting in adsorbent loss, thus makes the regeneration and process expensive.12,13 Therefore, easily recoverable and renewable adsorbents are in high demand. In recent decades, nanomaterials are of high interest because of having large surface area and high number of active sorptive sites. Received: September 11, 2017 Revised: February 3, 2018 Published: February 20, 2018 4549

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

building of second-order models, and additionally, it is possible to estimate higher-order model coefficients if nonsignificant coefficients are removed. Then, the adsorption kinetics and isotherms of TA and P4R adsorption onto magnetic Mn0.4Zn0.6Fe2O4-NPsD-YL-ISF7 nanocomposites as well as its usability for processing TA- and P4R-containing wastewater were investigated. Finally, the adsorption performance of magnetic Mn0.4Zn0.6Fe2O4-NPs-D-YLISF7 bio-nano-composites was compared with other reported adsorbents.

Spinel-type ferrite nanoparticles (NPs) of type MFe2O4 (M: Zn, Mn, Co, Fe, Cu, Ni) have been proven to fulfill easy recoverability because of their magnetic properties suggesting a smart and cost-effective approach for practical application.14 In addition, they are less toxic, highly available in terms of their parent materials with medium chemical stability.15 Even though MFe2O4NPs exhibited high adsorption performance, the medium surface area and the unique magnetic properties of the nanomaterials have led to their aggregation, resulting in lower adsorption efficiency.16−18 In addition, controllability of the dispersion and durability of these materials is essential for the performance and cost-effectiveness. To this end, interactions between the support and active MFe2O4-NPs are required to be adjusted and optimized.19,20 In these cases, high electrically conductive carbonbased materials have been frequently applied as supports for MFe2O4-based materials. This support synergistically combines the properties of both MFe2O4-based materials and carbon materials as well as provides a larger surface area for the diffusion of reactants onto the active sites and enhances their adsorption performance.21 Despite many efforts, difficulties are there in the controllability of the synthesis of such adsorbents. This is due to their complexity, which limits the information on the mechanism of interactions between carbon-based supports and MFe2O4 nanoparticles.22,23 Heretofore, there is no report on the exploration of magnetic MFe2O4 and Yarrowia lipolytica (an unconventional yeast with significant biological relevance) for the adsorption of organic pollutants. Yarrowia lipolytica can utilize various renewable carbon materials and the biomass of yeast.24 Compared to other materials (e.g., carbon nanotubes, graphene, or activated carbon), Yarrowia lipolytica in dead form has attracted tremendous attention for supporting metal oxides attributed to its medium electrical conductivities, medium surface areas, and unique mechanical strength.25 Here, MnZnFe2O4-NPs were loaded on dead Yarrowia lipolytica ISF7 (D-YL-ISF7) and used as a highly flexible, electrically resistive, and super-para-magnetic adsorbent with high magnetic response against radio frequencies.26 In MnZnFe2O4 as soft ferrites, MnFe2O4 has an inverse spinel structure, whereas ZnFe2O4 is usually assumed to be a completely normal spinel.27 The size, charge, and flexibility of these materials as well as their surface energy and morphology, physicochemically affect the amount of dyes adsorbed, rate of adsorption, and the stability of the layer adsorbed.28,29 MnFe2O4 spinel ferrites have a narrow band gap energy (∼1.4 eV), thus it was a soft adsorbent, while ZnFe2O4 band gap energy (∼1.7 eV) also creates a soft adsorbent. Therefore, selection of this material for adsorption of the soft dyes can be assisted for dyes adsorbed to the surface of a soft adsorbent mostly via electrostatic, H bond, soft−soft interactions, and other hydrophobic and or hydrophilic forces.30 Finally, the structure may cause the relaxation and spreading of the soft dye molecules, enhancing the interaction with the soft adsorbent surface.31,32 This new super-ferro-magnetic nanocomposite was applied for the simultaneous adsorption of tartrazine (TA) and ponceau 4R (P4R) as azo food-grade dyes in their binary solution. First-order derivative spectrophotometry was applied for the binary analysis of dyes. Powerful central composite design (CCD) allows the simultaneous study of different experimental factors, their investigative optimization, and interactions, to quantify their influences on one or more properties of interest to multivariate models.33 However, this type of design requires a number of experiments beyond those necessary to calculate these coefficients. CCD has demonstrated that it is possible to obtain the same conclusions from a reduced number of experiments34 as well as allows the



EXPERIMENTAL SECTION

Chemicals and Reagents. Analytical grade tartrazine (TA) and ponceau 4R (P4R) dyes, sodium hydroxide (NaOH), hydrochloric acid (HCl), Zn(NO3)2·6(H2O), Fe(NO)3·9H2O, and Mn(NO3)2·6H2O were obtained from Merck (Darmstadt, Germany). The dye concentration was analyzed by first-derivative spectrophotometry by measuring the absorbance at λmax of 378 nm for TA and 560 nm for P4R (See Figure S1) after the calibration curves were established. Stock solutions (100 mg L−1) of TA and P4R were prepared. Instrumentation. The determination of initial and residual TA and P4R concentrations was performed by a UV−visible spectrophotometer (Lambda 950, PerkinElmer, USA) at 378 and 560 nm (first-derivative spectrophotometry), respectively. A Metrohm 692 pH meter (Metrohm, Herisau, Switzerland) was applied for pH measurements. The adsorption was enhanced by ultrasonication for 5 min in a Tecno-GAZ SPA ultrasonic bath working at 40 kHz and 130 W from Parma, Italy. The samples’ surface morphologies were investigated using field-emission scanning electron microscopy (FESEM, Zeiss SUPRATM 50 VP, Germany) according the following method: powder samples were mounted on a stub of carbon with adhesive, coated with 40−60 nm followed by transferring in stage of setup and observing by the microscope. Transmission electron microscopy (TEM) of the prepared samples was studied according the following method: ethanol was used to disperse samples in an ultrasound bath for 30 min, and a few drops of this were dripped onto carbon grids. After casting it over the grid as a support, the image of particles was acquired. Energy dispersive X-ray spectrometer (Oxford INCA II energy solid-state detector) as a part of the FESEM above was used for the elemental analysis of samples. X-ray powder diffraction (XRD, Philips PW 1800) with a monochromatic Cu X-ray source Kα (λ = 0.154 nm) was used to analyze the crystalline nature of the samples. We characteristically run the XRD over 2θ angles within the range 5−70°. Infrared data were obtained using the KBr pellet technique by mixing 1.0 mg of prepared powder with 0.1 g of KBr. Fourier transform infrared (FTIR), PerkinElmer Spectrum RX-IFTIR, was applied to determine the functional groups of the samples over the wavenumber range of 400−4000 cm−1. The specific surface area (m2 g−1) of each sample was determined by BET instrument (BET, PHS-1020, PHS CHINA) during nitrogen adsorption−desorption. The magnetic characteristics of the samples were obtained by a vibrating sample magnetometer (VSM, LDJ 9600-1, USA) as follows: the sample chamber was warmed up to 300 K, the magnetic field was set to zero (0) Oe, and the sample chamber was ventilated. After applying magnetic field, the magnetic properties were recorded followed by the removal of any sample puck or option that is currently installed in the chamber. Statistica (Ver. 10.0, Stat Soft Inc., Tulsa, USA) was used to apply CCD and to handle the experimental data. Preparation of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7. Yarrowia lipolytica ISF7 powder was prepared according to our previous report.35 Briefly, Yarrowia lipolytica ISF7 and yeast−peptone−glucose (YPG: 2% peptone, 1% yeast extract, 1% glucose) were incubated during a night at 30 °C followed by inoculation of a single colony into a 100 mL Erlenmeyer flask that contained 25 mL of YPG broth at pH 7.0 and incubation in a shaker at 160 rpm at 30 °C for 24 h. Finally, the products were collected and washed with deionized water followed by drying at 80 °C for 12 h. The dead Yarrowia lipolytica ISF7 (D-YL-ISF7) as a product was collected and used as a supporting agent for the deposition of magnetic materials. Mn0.4Zn0.6Fe2O4-NPs-loaded D-YL-ISF7 powder was prepared as follows: After dispersing 1.0 g of D-YL-ISF7 powder in 125 mL of 4550

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

Table 1. Central Composite Design Parameters Obtained for Various Runs, Experimental Values of Five Variables, and Responses for the Adsorption of TA and P4R by Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 independent variables factors

coded

pH sonication time adsorbent mass TA concentration P4R concentration

X1 X2 X3 X4 X5

range and levels (coded) units

−α (−2.0)

low (−1)

central (0)

2.000 0.500 0.005 4.000 4.000

3.000 2.000 0.010 8.000 8.000

4.000 3.500 0.015 12.000 12.000

min g mg L−1 mg L−1 independent variables

high (+1)

+α (+2.0)

5.000 5.000 0.020 16.000 16.000 response (Y, %)

6.000 6.500 0.025 20.000 20.000

X1

X2 (min)

X3 (g)

X4 (mg L−1)

X5 (mg L−1)

run order

pH

sonication time

adsorbent mass

TA concentration

P4R concentration

R % TA

R % P4R

R % TA

R % P4R

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25

5.000 5.000 3.000 5.000 5.000 5.000 3.000 3.000 5.000 3.000 3.000 2.000 6.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000 4.000

5.000 2.000 5.000 5.000 5.000 2.000 2.000 5.000 2.000 5.000 2.000 3.500 3.500 0.500 6.500 3.500 3.500 3.500 3.500 3.500 3.500 3.500 3.500 3.500 3.500

0.010 0.020 0.020 0.020 0.010 0.010 0.020 0.010 0.020 0.020 0.010 0.015 0.015 0.015 0.015 0.005 0.025 0.015 0.015 0.015 0.015 0.015 0.015 0.015 0.015

16.000 16.000 8.000 8.000 8.000 16.000 16.000 16.000 8.000 16.000 8.000 12.000 12.000 12.000 12.000 12.000 12.000 4.000 20.000 12.000 12.000 12.000 12.000 12.000 12.000

8.000 8.000 16.000 8.000 16.000 16.000 16.000 16.000 16.000 8.000 8.000 12.000 12.000 12.000 12.000 12.000 12.000 12.000 12.000 4.000 20.000 12.000 12.000 12.000 12.000

81.560 61.430 96.780 94.750 70.980 41.760 67.870 84.870 48.870 92.320 58.230 97.450 65.342 21.870 99.880 25.880 96.876 99.980 50.870 93.760 78.543 87.560 86.760 84.650 85.320

95.760 88.650 92.080 97.870 75.760 48.970 70.123 89.311 55.450 98.760 48.450 99.800 65.786 35.780 99.876 36.760 99.876 98.760 63.670 98.750 81.870 92.123 91.876 90.423 91.890

81.039 60.909 96.259 94.229 70.459 41.239 67.349 84.349 48.349 91.799 57.188 98.232 66.124 22.652 100.66 26.662 97.658 100.76 51.652 94.542 79.325 85.682 85.682 85.682 85.682

95.425 88.315 91.745 97.535 75.425 48.635 69.788 88.976 55.115 98.425 47.780 100.30 66.288 36.282 100.37 37.262 100.37 99.262 64.172 99.252 82.372 91.327 91.327 91.327 91.327

deionized water for 30 min, Fe(NO)3·9H2O, Zn(NO3)2·6(H2O), and Mn(NO3)2·6H2O were added into the obtained suspension followed by the ultrasonication for 30 min. Subsequently, NaOH solution was added drop-by-drop into the above mixture and stirred until the pH value of 9.0 was achieved. The brown aqueous solution obtained was transferred into a Teflon lined autoclave (175 mL) to hydrothermally react at 200 °C for 12 h. Finally, the powders obtained (Mn0.4Zn0.6Fe2O4-NPsD-YL-ISF7) after the hydrothermal reaction were washed with deionized water followed by drying at 80 °C for 12 h. Ultrasound-Assisted Adsorption Procedures. Batch adsorption experiments under ultrasound irradiation were run to investigate the effect of pH, sonication time, Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 mass, and initial dye concentrations. Experiments were run in an Erlenmeyer flask containing 50 mL of binary dye solution in an ultrasonic bath with contact times over the range 0.5−6.5 min. The solution samples were withdrawn in known intervals after the end of the adsorption process, and super-ferro-magnetic Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 adsorbent was separated with external magnetite filings (1.3 T). The residual dye concentrations in withdrawn samples were analyzed for calculating the final concentrations of TA and P4R dyes, applying the first derivative spectrophotometry at 378 and 560 nm, respectively. The following equation was applied to calculate the removal percentage of each dye: removal % =

experimental

predicted

Adsorption isotherm study was conducted over the range of 5−30 mg L−1 of initial concentration of TA and P4R dyes, while other variables were kept constant at optimized condition and at 25 °C until the system reached equilibrium state. Finally, the qe was calculated at equilibrium conditions according to our previous report.36 A fitting procedure was applied to fit the equilibrium experimental data to Langmuir, Dubinin− Radushkevich (DR), Freundlich, and Temkin isotherms to investigate the isotherms of TA and P4R adsorption. Kinetic study was conducted at different sonication times (0.5−6.5 min), meanwhile keeping other factors constant at their optimal values and at 25 °C until the system reached an equilibrium state. The kinetics of adsorption was investigated applying pseudo-first- and second-order as well as intraparticle diffusion and Elovich kinetic models. Comparison of ultrasound waves as mass a transfer device, applied here, with vortex and magnetic stirring as other mass transfer devices (applied according to similar procedures) was performed. Experimental Design Procedure. After the careful consideration of parameters, which could potentially affect the simultaneous adsorption of azo food dyes, five parameters (pH (X1), sonication time (X2), Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 mass (X3), initial TA concentration (X4), and initial P4R (X5) concentration) were selected according to the literature and initial experiments. A CCD approach was applied for overall optimization of the variables for obtaining the best responses. In this approach, each factor was considered at five levels including −α, −1, 0, +1, and +α, as the lowest, low, center, high, and highest levels, respectively (see Table 1). Twelve replicates were

initial conc of dye − final conc of dye (TA or P4R) × 100% initial conc of dye (TA or P4R)

(1) 4551

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

a broad band within 3100−3600 cm−1 and a band within the range of 2900−2960 cm−1, which correspond to the hydroxyl group of alkyl groups in proteins and asymmetric stretching vibration of C−H bonds of carbohydrates present in the cell wall of D-YL-ISF7, respectively. A strong band at 1656 cm−1 corresponds to CC bond. The bands at 1450, 1220, and 1050 cm−1 correspond to C−O stretching of alcohol, ethers, sulfoxide, carbohydrates, or polysaccharide-like materials, respectively. The band observed at 1400 cm−1 corresponds to the starching COOH in the carbocylic acid group of D-YL-ISF7. The final trough observed at 532 cm−1 corresponds to O−C−O and C−O vibrations in scissoring and bending modes, respectively.38 The stability of D-YL-ISF7 as a good supporting agent was investigated by FTIR analysis after exposing ultrasound waves (5 min), where the obtained results were in good compatibility with primary FTIR, which confirms its good stability in the presence of ultrasound waves (see Figure 1b). Figure 1c shows the FTIR spectrum of the pure Mn0.4Zn0.6Fe2O4-NPs, demonstrating broad bands at 3430 and 1620 cm−1 from −OH groups. The bands appearing at 810, 735, and 548 cm−1 are ascribed to the absorptions of Mn−O, Fe−O, and ZnO, respectively.39 After loading the Mn0.4Zn0.6Fe2O4-NPs on D-YL-ISF7, the FTIR spectrum (Figure 1d) showed all the mentioned bands for pure compounds. The stability of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 as an efficient adsorbent was investigated by FTIR analysis after ultrasound wave exposure (5 min), which resulted in good compatibility with primary FTIR and confirmed its good stability in the presence of ultrasound waves (see Figure 1e). Finally, after TA and P4R adsorption onto Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7, the FTIR spectrum (Figure 1f) suggests that the mechanism of

considered for center point. In total, 25 experimental runs were designed in random to reduce systematic errors (see Table 1). The response was modeled with low standard error by fitting experimental data to a quadratic polynomial against coded values of the significant variables (eq 2). It was termed according five individual effects (X1−X5), ten mutual interactions, and five quadratic effects (X12−X52) as follows:35 k

Y = β0 +

k

k

k

∑ βi xi + ∑ ∑ βijxixj + ∑ βiixi 2 + ε i=1

i=1 j=1

i=1

(2)

where regression coefficients of β0, βi, βii, and βij are constant, linear, quadratic, and interaction term coefficients, respectively. Y is the predicted value of response. k is the number of factors, i (1, ..., 5) and j (1, ..., 5) are the index numbers for factors, and ε is the residual error. Analysis of variance (ANOVA) was applied to detect the significant and insignificant terms. The model robustness was investigated by standard deviation (SD), coefficient of variance (CV %), R2 coefficient, Adj-R2, and adequate precision (AP). A confidence level of 95% was considered to predict the p-value of each term.37 The extent of the variation of response against each pair of factors was predicted by corresponding contour plot.



RESULTS AND DISCUSSION Characterization of Mn0.4-Zn0.6-Fe2O4-YL-ISF7. The FTIR spectrum of the D-YL-ISF7 sample (Figure 1a) exhibited

Figure 1. FTIR spectra of D-YL-ISF7 before (a) and after (b) ultrasound exposure (for confirming the support stability), Mn0.4Zn0.6Fe2O4-NPs (c), Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 before (d) and after (e) ultrasound exposure (for confirming its stability), and Mn0.4Zn0.6Fe2O4-NPs-D-YLISF7 after use in adsorption processes in the presence of ultrasound irradiation (f).

Figure 2. XRD pattern of D-YL-ISF7 (a), Mn0.4Zn0.6Fe2O4-NPs (b), and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 (c). 4552

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering the dye adsorption involves H bonding and electrostatic, soft− soft, and π−π interactions. The peaks corresponding to CN, NN, C−N, SO, and aromatic CC and CH of TA and P4R were observed at 1610, 1680, 1420, 1150, 700−900, and 1410 cm−1, respectively, demonstrating the successful adsorption of TA and P4R dyes onto Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7.40 The pattern of X-ray diffraction taken from D-YL-ISF7 (Figure 2a) showed distinct crystalline peaks around 2θ values of 10° and 19° due to a plenty of −OH and −COOH groups present in the YL-ISF7 structure, which could form stronger inter- and intramolecular hydrogen bonds and indicate that the D-YL-ISF7 structure has certain regularity. The XRD pattern of Mn0.4Zn0.6Fe2O4-NPs sample (Figure 2b blue line) shows that peaks that appeared at 18.15, 29.86, 35.17, 36.79, 42.74, 46.79, 53.02, 56.51, 62.05, 65.23, and 66.27° correspond to the (111), (022), (113), (224), (004), (133), (224), (115), (044), (135), and (244) planes, respectively, and reveals the typical information from the cubic structure of Mn0.4Zn0.6Fe2O4-NPs (reference code: 96-200-9104). In addition, it indicates that the powder specimens obtained are single phase and the hydrothermal method is feasible to synthesize Mn0.4Zn0.6Fe2O4-NPs.41 Moreover, the absence of any extra peak shows the single-phase cubic structure of all samples with the space group of Fd3m. As shown, the XRD peaks from pure Mn0.4Zn0.6Fe2O4-NPs are very narrow having a low bandwidth, which are due to their aggregation and

Figure 4. EDS spectra of D-YL-ISF7 (a) and Mn0.4Zn0.6Fe2O4-NPs-DYL-ISF7 (b).

lower dispersion power. As mentioned above, the controllability of the dispersion and durability of these materials is essential for

Figure 3. FESEM images of D-YL-ISF7 (a), Mn0.4Zn0.6Fe2O4-NPs (b), Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 (c), and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 after use in adsorption processes in the presence of ultrasound (d). 4553

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

cubic shapes with different size distributions; the unique magnetic properties of the NPs have led to such a morphology (Figure 3b). After loading the Mn0.4Zn0.6Fe2O4 nanoparticles onto dead Yarrowia lipolytica ISF7, the aggregation was suppressed and more active sites were available (Figure 3c). In addition, the FESEM image of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 after recycling was performed, and the results is shown in Figure 3d. It is seen that after three cycles the Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 morphology has changed little, which indicates the optimal stability of this absorbent. EDS peaks corresponding to carbon, potassium, oxygen, and copper taken from D-YL-ISF7 samples (Figure 4a) and the peaks as an indication of Mn, Zn, Fe, C, O, K, and Cu taken from the Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 sample successfully confirm the synthesis of the desired product (Figure 4b). In addition, to adequately confirm the presence of the Mn0.4Zn0.6Fe2O4-NP component in this material, it was dissolved in HNO3:HCl followed by the determination of Zn(II), Mn(II), and Fe(III) by atomic absorption spectroscopy (AAS). The stoichiometric values of 0.4, 0.6, and 2.0 respectively obtained for Mn, Zn, and Fe in addition to the lattice constant value obtained confirm the successful incorporation of these materials in accordance with the stoichiometry used. In other word, the atomic ratio of Mn, Zn, and Fe elements is approximately 0.4:0.6:2 from AAS analysis.

their performance and cost effectiveness. Here, interactions between the support and active material are required to be adjusted and optimized. Therefore, after loading the Mn0.4Zn0.6Fe2O4-NPs onto D-YL-ISF7, the XRD pattern of the product (see Figure 2c) becomes very wide with a high bandwidth. In addition, overlap can be seen between the narrow band peaks in pure Mn0.4Zn0.6Fe2O4NPs, and the new peaks in the XRD pattern of the final product in the presence of D-YL-ISF7 are formed. These results showed that D-YL-ISF7 is a good supporting and dispersion agent. The reference code for cubically structured Mn0.4Zn0.6Fe2O4NPs in the presence of D-YL-ISF7 is 96-200-9104. As seen in Figure 2c, the pattern corresponds to YL-ISF7 at 10° and 19° (with a little shift). The surface morphologies of the pure D-YL-ISF7, Mn0.4Zn0.6Fe2O4 nanoparticles, and Mn0.4Zn0.6Fe2O4-NPs-DYL-ISF7 were characterized by FESEM. Figure 3a shows an FESEM image acquired from pure D-YL-ISF7, which revealed a superficial area with lightly ordered and porous structure. The FESEM image shown in Figure 3b indicates the severe agglomeration of Mn0.4Zn0.6Fe2O4 nanoparticles. It could be explained as the aggregation of smaller primary nanoparticles into secondary ones possibly because of their very low dimensionality, high surface energies, and magnetic properties. In addition, the Mn0.4Zn0.6Fe2O4 nanoparticles exhibit a mixture of spherical and

Figure 5. HRTEM images of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 with different magnifications. (inset in b and c) TEM image of a single Mn0.4Zn0.6Fe2O4-NP attached on the D-YL-ISF7 nanosheet. 4554

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

The synthesized D-YL-ISF7 and Mn0.4Zn0.6Fe2O4-NP samples showed isotherms of type-III which indicates high adsorption at high relative pressure (P/P0) and shows that the materials are porous (see Figure 6a and b). The BET study resulted in specific surface areas of 86.44 and 13.41 m2 g−1 for D-YL-ISF7 and Mn0.4Zn0.6Fe2O4-NPs, respectively. The low surface area for Mn0.4Zn0.6Fe2O4-NPs is due to the particle aggregation and agglomeration. In this case, D-YL-ISF7 with high electrical conductivity and medium surface area has been used as a matrix for Mn0.4Zn0.6Fe2O4-NPs, which synergistically combines the properties of both Mn0.4Zn0.6Fe2O4-NPs and D-YL-ISF7 as well as provides a considerably larger surface area of 493.13 m2 g−1 (see Figure 6c). In addition, the synthesized Mn0.4Zn0.6Fe2O4-NPs-DYL-ISF7 samples showed typical type-I behavior, indicating the microporosity of such a material (see Figure 6c). Figure 7a shows the magnetic hysteresis taken from Mn0.4Zn0.6Fe2O4-NPs and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7

The surface morphology of the Mn0.4Zn0.6Fe2O4-NPs-D-YLISF7 nanocomposite was investigated by HRTEM. Figure 5 shows the TEM images of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 nanocomposite at different magnifications. It is clear to see that the Mn0.4Zn0.6Fe2O4-NPs are uniformly dispersed on the D-YL-ISF7 nanosheets with much reduced aggregation. For the Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7, the quasi-two-dimensional D-YL-ISF7 nanosheets act as supporting substrates for homogeneous anchoring of Mn0.4 Zn 0.6Fe 2O 4 -NPs, building a Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 hetero architecture. It is speculated that the introduction of D-YL-ISF7 into Mn0.4Zn0.6Fe2O4NPs could improve the uniformity of the Mn0.4Zn0.6Fe2O4 nanoparticle distribution on the D-YL-ISF7 nanosheets and suppress the aggregation of Mn0.4Zn0.6Fe2O4-NPs nanoparticles. It can be seen that the typical particle size of Mn0.4Zn0.6Fe2O4-NPs in the nanocomposite is less than 15 nm, which causes the high surface area of Mn0.4Zn0.6Fe2O4-NPs onto D-YL-ISF7 compared with the pure Mn0.4Zn0.6Fe2O4-NPs. As shown in the insets of Figure 5b and c, the HRTEM of a single Mn0.4Zn0.6Fe2O4 nanoparticle exhibits well-resolved lattice fringes with an interplane distance of around 0.275 nm, which can be attributed to the (113) plane of the Mn0.4Zn0.6Fe2O4-NP crystal, indicating the crystalline feature of the nanoparticles.

Figure 7. (a) Magnetic hysteresis curves of Mn0.4Zn0.6Fe2O4-NPs and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 and (b) zero point charge plot for Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 adsorbent.

nanocomposite at 300 K. As seen, the magnetization (M)−magnetic field (H) relation is approximately nonlinear, which implies that both Mn0.4Zn0.6Fe2O4-NPs and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 behave as ferrimagnetic materials at 300 K. On the other hand, from the hysteresis the saturation magnetization (Ms) of Mn0.4Zn0.6Fe2O4-NPs (23 emu g−1) was shown to be lower than that of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 (35 emu g−1). This may be attributed to the aggregation and agglomeration of Mn0.4Zn0.6Fe2O4-NPs and the size effect of Mn0.4Zn0.6Fe2O4NPs in pure and composite form.42 As known, coercivity is an

Figure 6. N2 adsorption−desorption isotherms of D-YL-ISF7 (a), Mn0.4Zn0.6Fe2O4-NPs (b), and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 (c). 4555

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering indication of anisotropy, which is in turn considerably dependent on the morphology, surface, size, and interface exchange coupling effects. Therefore, this limited data hardly predicts the effective anisotropies of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7.43 After running detailed experiments and their qualitative analysis, it is suggested that the increase in Hc and Mr/Ms of the Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 may come from the unique nonaggregated Mn0.4Zn0.6Fe2O4-NPs, which create the high anisotropy. In addition, the surface−interface exchange coupling may be affected by the interfaces between the ferrimagnetic Mn0.4Zn0.6Fe2O4-NPs as a good dispersant and D-YL-ISF7 as the supporting agent. This observation may interestingly promote the applicability of such a nanocomposite in magnetic solid phase extraction and adsorption processes. The adsorbent charge distribution may significantly affect the adsorption process, and contacting the solute ions with the surface may change the corresponding forces and mechanism. The adsorbent surface charge is neutral at a so-called pH which is the zero point of charge (pHzpc). At a pH above the pHzpc, the adsorbent surface charge changes to negative, which thus causes the adsorption of positive dyes on the adsorbent surface via electrostatic interactions while a reverse phenomenon happens at pH values below this.44 In this work, the value of the pHzpc of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 was obtained using the plot of ΔpH (final pH − initial pH) against initial pH (Figure 7b). The pHzpc measurement resulted in the value of 4.6 for Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7. Therefore, at pH values below 4.6, the surface of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 is positively charged, which favorably enhances the adsorption of anionic dye (acidic dye). Determination of TA and P4R Dye Concentrations in Binary Solution. Considerable overlap observed between the zero-order spectra of TA and P4R over the wavelength range of 300−800 nm (Figure 8a) does not allow the direct spectrophotometric determination of each dye concentration in respective maximum wavelength. Therefore, the first derivative spectra (Figure 8b) of each dye in single solution with vanishing crossing points was applied for their simultaneous determination. At the desired wavelengths, calibration graphs (Figure 8c) were obtained and the best linear response resulting in an approximately zero intercept, with the least influence from the concentration of other component(s), was chosen for determination of each dye. According to the text above, TA quantification was performed at 378 nm where the first-order derivative (dA/dλ) value for P4R is zero. In a similar way, P4R determination was carried out at 560 nm where the dA/dλ value of TA is zero. RSM and Model Fitting. Table 1 presents the CCD matrix, experimental, and predicted responses values (R% TA (Y1) and R% P4R (Y2)). ANOVA (Table 2) resulted in Fisher values (F-value) of 159.61 and 344.54 for R% TA and P4R, respectively, implying that the model is highly significant. The large F-value signifies that the regression model can describe the variation in responses. Table 2 presents the criteria on the fit quality indicating the models’ adequacy. The determination coefficients (R2 = 0.9987 for R% of TA and R2 = 0.9994 for R% of P4R) were shown by the ANOVA of quadratic regression models, which indicate that the models failure is only for 0.13 and 0.06% of the total variations (Figure 9a). The values of adjusted determination coefficient (Adj-R2 = 0.9925 and 0.9965 for TA and P4R, respectively) also confirmed the high significance of the models.45 In addition, good model fitting was analyzed by plotting the residual against observed values, which indicated that all data points fall within the limits

Figure 8. Absorption zero (a) and first-derivative (b) spectra of tartrazine (TA, 15 mg L−1) and ponceau 4R (P4R, 15 mg L−1) and (c) calibration curve at 378 nm for TA and 560 nm for P4R in the range of 1−30 mg L−1 of both dyes.

with no trend in residual data (Figure 9b). All points of the raw residual were observed to be within the range of −2.0 to +2.0 for the TA and P4R adsorption onto Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7. The responses values (TA and P4R removal percentages) were predicted by the following quadratic models Y1(R %TA) = − 3.29 − 28.26X1 + 56.84X 2 + 32300X3 − 30.15X4 − 4.09X5 − 1769X1X3 + 4.03X1X4 − 2223X 2X3 − 0.49X 2X5 − 547.5X3X5 + 0.90X4X5 − 2.669X 2 2 − 235200X32 − 0.1480X4 2

(3)

Y2(R %P4R) = − 141.8 − 28.1X1 + 63.5X 2 + 32810X3 − 22.7X4 + 10.4X5 − 1070X1X3 + 4.3X1X4 − 2425X 2X3 + 0.9X 2 X4 − 1.0X 2X5 − 785.5X3X5 + 0.4X4X5 − 2.0X12 − 2.6X 2 2 − 225100X32 − 0.2X4 2

(4)

The positive coefficients for X2, X3, X1X4, X2X4, and X4X5 indicate favorable effects on the responses (R% TA (Y1) and R% P4R (Y2)), 4556

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

Table 2. Analysis of Variance of the Regression Coefficients of the Fitted Quadratic Equations for the Tartrazine (R% TA) and Ponceau 4R (R% P4R) as a Function of the Independent Variablesa R% TA source of variation model X1 X2 X3 X4 X5 X1X2 X1X3 X1X4 X1X5 X2X3 X2X4 X2X5 X3X4 X3X5 X4X5 X12 X22 X32 X42 X52 residual lack of fit pure error cor total

a

R% P4R

regression coefficients

Df

SS

MS

F-value

P-value

SS

MS

F-value

P-value

R% TA

R% P4R

20 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 4 1 3

12605.938 515.462 3042.780 2520.216 1205.896 115.779 0.005 265.229 879.521 11.583 942.620 169.805 29.028 0.692 406.514 695.355 19.825 931.948 893.333 144.954 2.528 15.796 10.521 5.275

630.297 515.462 3042.780 2520.216 1205.896 115.779 0.005 265.229 879.521 11.583 942.620 169.805 29.028 0.692 406.514 695.355 19.825 931.948 893.333 144.954 2.528 3.949 10.521 1.758

159.612 130.532 770.531 638.201 305.372 29.319 0.001 67.165 222.723 2.933 238.702 43.000 7.351 0.175 102.943 176.086 5.020 235.999 226.221 36.707 0.640

X3 > X2X3 > X1X4 > X22 > X3X5 > X32 > X4 > X1 > X42 > X4X5 > X5 > X2X5 > X12 > X1X3 > X2X4 (Figure S2b). As it is given in Table 2, the coefficient of variances (CV %) for the TA and P4R adsorption have been found to be 2.651 and 1.545%, respectively. The model reproducibility was tested by CV %, which is obtained from the ratio of standard error of estimate to the mean observed response value. In general, the ratio less than 10% indicates the reasonable reproducibility of a model.46 An indication of signal-to-noise is adequate precision (AP), which was obtained to be 42.89 and 56.35 for the adsorption of TA and P4R, respectively. The fact that the AP values are larger than 4.0 proves the adequacy of the signal and confirms the well prediction by the model designed using CCD. Influence of the Parameters on the Responses. The contour plot of removal (%) of each dye versus two variables (Figure 10) was obtained, while holding other factors constant at their central level. The combined effects of pH and sonication time on R% of TA are shown in Figure 10a−c. An increasing 4557

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

Statistica to approach 1 as most desirable value (Figure S3) Here, the purpose is the optimization of the adsorption process for two dyes onto Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 adsorbent within reduced time and with low costs. Therefore, the optimization was numerically performed for simultaneous optimization of both responses. For two responses, the optimum conditions were obtained at pH 2.0, Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 mass of 0.015 g, sonication time of 5 min, TA concentration of 12 mg L−1 and P4R concentration of 16 mg L−1 with desirability value of 0.9948. At this condition, the corresponding experimental values (n = 5) for adsorption efficiency (R%) of TA and P4R were found to be 99.98% ± 1.45 and 98.75% ± 2.08, respectively, in good consistency with the predicted values of 101.69 and 99.215%, respectively. Good correlation between these results obtained from desirability functions method, indicates that CCD in complementary with desirability functions effectively optimize the adsorption process for the removal of TA and P4R dyes. Equilibrium Isotherm Study. The adsorption mechanism can be explained by conventional equilibrium isotherms such as Langmuir, Freundlich, Dubinin−Radushkevich, and Temkin. The linearized equations with known definitions and constants presented in the literature47−50 were evaluated (Table 3). The Langmuir isotherm was linearized (see Table 3), and the KL and Qm values were obtained from the intercept and slope of the plot of Ce/qe versus Ce (plots not shown), respectively. High correlation coefficient (>0.992) and lower χ2 ( 1, 0 < RL < 1 indicate the unfavorability, favorability of the adsorption, respectively. In addition, RL = 0 and RL = 1 suggest the irreversibility and linearity of adsorption, respectively. The values of RL for TA and P4R adsorption onto Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 were less than 1.0 and greater than zero indicating favorable adsorption. The Freundlich constant, KF, as an indication of the extent of adsorption, and the value of 1/n as an indication of the effectiveness of adsorption were determined using the intercept and slope of the linear plot of ln qe versus ln Ce, respectively (see Table 3). For a favorable adsorption, n values are within the range 0 < n < 1, while n > 1 and n = 1 indicate unfavorability and linearity of adsorption, respectively.51 Moreover, n = 0 shows irreversibility of adsorption process. Using mathematical calculations, it has been shown (Table 3) that the values of n between 1 and 3 obtained from the Freundlich isotherm indicate the effectiveness of adsorption. The Temkin isotherm was applied to determine the heat of adsorption and the interaction between adsorbent and adsorbate. The related β and KT parameters were obtained using the plot of qe against ln Ce. The correlation coefficient (99.0%) of dyes at 25 °C

Table 5. Comparison of Contact Time, pH, and Adsorption Capacity of Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 with Other Available Adsorbents and Various Methods dye

method

adsorbent

pH

P4R

photo-Fenton photocatalytic degradation magnetic stirrer assisted magnetic stirrer assisted degradation electrochemical degradation magnetic stirrer assisted ultrasound assisted magnetic stirrer assisted vortex assisted photodegradation magnetic stirrer assisted magnetic stirrer assisted magnetic stirrer assisted magnetic stirrer assisted photodegradation ultrasound assisted magnetic stirrer assisted vortex assisted

H2O2 and Fe2+ TiO2 Mg/Al-layered double hydroxide polyamidoamine−cyclodextrin cross-linked copolymer ozonation electrogenerated H2O2 (EO-H2O2) alkali boiled tilapia fish scales Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 nanomolecularly imprinted polymer nanophotocatalyst melamine-formaldehyde-tartaric acid resin activated carbon biosorbents of Lantana camara multiwalled carbon nanotubes (CNTs) CNTs were decorated with silver nanoparticles titanium dioxide Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7

3.0

TA

a

9.0 5.0 4.7 3.0 7.5 2.0 2.0 2.0 2.0 6.7 2.0 3.0 3.0 11.0 2.0 2.0 2.0

Qm (mg g−1)

45.67 254.3

134.40 101.461 57.045 32.480

time (min)

ref

15 70 120 720 120 240 240

53 54 13 2 55 51 1 this worka

5

21.60 90.90 52.24 84.04

300 90 60 60 60 60

90.827 54.705 24.450

5

56 4 9 3 3 5 this worka

Experimental conditions: initial dye concentration 5−30 mg L−1, adsorbent mass 0.015 g, V 50 mL, pH 2.0, contact time 5 min, T 25 °C. 4561

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering

(12) Zhang, P.; Lo, I.; O’Connor, D.; Pehkonen, S.; Cheng, H.; Hou, D. High efficiency removal of methylene blue using SDS surfacemodified ZnFe2O4 nanoparticles. J. Colloid Interface Sci. 2017, 508, 39− 48. (13) An, Z.-H.; Peng, S.-C.; Tang, H.-J.; Chen, T.-H.; Ma, M.-H. Treatment of simulated wastewater containing ponceau-4R by synthesizing Mg/Al-layered double hydroxide in situ. J. Hefei Univ. Technol. 2007, 10, 2−15. (14) Ojha, S.; Nunes, W. C.; Aimon, N. M.; Ross, C. A. Magnetostatic Interactions in Self-Assembled CoxNi1−xFe2O4/BiFeO3 Multiferroic Nanocomposites. ACS Nano 2016, 10, 7657−7664. (15) Yao, Y.; Cai, Y.; Lu, F.; Wei, F.; Wang, X.; Wang, S. Magnetic recoverable MnFe2O4 and MnFe2O4-graphene hybrid as heterogeneous catalysts of peroxymonosulfate activation for efficient degradation of aqueous organic pollutants. J. Hazard. Mater. 2014, 270, 61−70. (16) Mahmoud, M. E.; Abdou, A. E. H.; Mohamed, S. M. S.; Osman, M. M. Engineered staphylococcus aureus via immobilization on magnetic Fe3O4-phthalate nanoparticles for biosorption of divalent ions from aqueous solutions. J. Environ. Chem. Eng. 2016, 4, 3810−3824. (17) Jagadeesan, D.; Mansoori, U.; Mandal, P.; Sundaresan, A.; Eswaramoorthy, M. Hollow Spheres to Nanocups: Tuning the Morphology and Magnetic Properties of Single-Crystalline α-Fe2O3 Nanostructures. Angew. Chem., Int. Ed. 2008, 47, 7685−7688. (18) Mahmoud, M. E.; Abdelwahab, M. S.; Abdou, A. E. Enhanced removal of lead and cadmium from water by Fe3O4-cross linked-Ophenylenediamine nano-composite. Sep. Sci. Technol. 2016, 51, 237− 247. (19) Zhang, Z.; Ji, Y.; Li, J.; Tan, Q.; Zhong, Z.; Su, F. Yolk bishell Mn x Co1−x Fe2O4 hollow microspheres and their embedded form in carbon for highly reversible lithium storage. ACS Appl. Mater. Interfaces 2015, 7, 6300−6309. (20) Zhang, X.-J.; Wang, G.-S.; Cao, W.-Q.; Wei, Y.-Z.; Liang, J.-F.; Guo, L.; Cao, M.-S. Enhanced microwave absorption property of reduced graphene oxide (RGO)-MnFe2O4 nanocomposites and polyvinylidene fluoride. ACS Appl. Mater. Interfaces 2014, 6, 7471− 7478. (21) Vadiyar, M. M.; Kolekar, S. S.; Chang, J.-Y.; Ye, Z.; Ghule, A. V. Anchoring ultrafine znfe2o4/c nanoparticles on 3D ZnFe2O4 nanoflakes for boosting cycle stability and energy density of flexible asymmetric supercapacitor. ACS Appl. Mater. Interfaces 2017, 9, 26016−26028. (22) Mahmoud, M. E.; Yakout, A. A.; Hussein, K. H.; Osman, M. M. Magnetic accumulation and extraction of Cd (II), Hg (II) and Pb (II) by a novel nano-Fe3O4-coated-dioctylphthalate-immobilized-hydroxylamine. J. Environ. Chem. Eng. 2015, 3, 843−851. (23) Guo, Y.; Zhang, L.; Liu, X.; Li, B.; Tang, D.; Liu, W.; Qin, W. Synthesis of magnetic core−shell carbon dot@ MFe2O4 (M= Mn, Zn and Cu) hybrid materials and their catalytic properties. J. Mater. Chem. A 2016, 4, 4044−4055. (24) Liu, H.-H.; Ji, X.-J.; Huang, H. Biotechnological applications of yarrowia lipolytica: Past, present and future. Biotechnol. Adv. 2015, 33, 1522−1546. (25) Teber, A.; Cil, K.; Yilmaz, T.; Eraslan, B.; Uysal, D.; Surucu, G.; Baykal, A. H.; Bansal, R. Manganese and Zinc Spinel Ferrites Blended with Multi-Walled Carbon Nanotubes as Microwave Absorbing Materials. Aerospace 2017, 4, 2. (26) Irfan, S.; Ajaz-un-Nabi, M.; Jamil, Y.; Amin, N. In Synthesis of Mn1‑xZnxFe2O4 ferrite powder by co-precipitation method. IOP Conf. Ser.: Mater. Sci. Eng. 2014, 60, 012048. (27) Szotek, Z.; Temmerman, W.; Ködderitzsch, D.; Svane, A.; Petit, L.; Winter, H. Electronic structures of normal and inverse spinel ferrites from first principles. Phys. Rev. B: Condens. Matter Mater. Phys. 2006, 74, 174431. (28) Demirbas, A. Agricultural based activated carbons for the removal of dyes from aqueous solutions: a review. J. Hazard. Mater. 2009, 167, 1−9. (29) Li, S.; Li, H.; Liu, J.; Zhang, H.; Yang, Y.; Yang, Z.; Wang, L.; Wang, B. Highly efficient degradation of organic dyes by palladium nanoparticles decorated on 2D magnetic reduced graphene oxide nanosheets. Dalton Trans. 2015, 44, 9193−9199.

in very short time (5 min) under ultrasound irradiation that is a big advantage of the ultrasound-assisted method and Mn0.4Zn0.6Fe2O4-NPs-D-YL-ISF7 composite.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acssuschemeng.7b03205. Related descriptions about chemical structures and general characteristics of dyes, Pareto chart, and desirability profiles (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel./Fax: +98-741-2223048. E-mail: [email protected]; [email protected] (M.G.). ORCID

Arash Asfaram: 0000-0002-2937-5477 Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS The Yasouj University (Yasouj, Iran) is gratefully acknowledged for supporting this work. REFERENCES

(1) Zhu, K.; Gong, X.; He, D.; Li, B.; Ji, D.; Li, P.; Peng, Z.; Luo, Y. Adsorption of Ponceau 4R from aqueous solutions using alkali boiled Tilapia fish scales. RSC Adv. 2013, 3, 25221−25230. (2) Li, N.; Lei, X.-M. Adsorption of ponceau 4R from aqueous solutions by polyamidoamine−cyclodextrin crosslinked copolymer. J. Inclusion Phenom. Mol. Recognit. Chem. 2012, 74, 167−176. (3) Goscianska, J.; Pietrzak, R. Removal of tartrazine from aqueous solution by carbon nanotubes decorated with silver nanoparticles. Catal. Today 2015, 249, 259−264. (4) Baraka, A. Adsorptive removal of tartrazine and methylene blue from wastewater using melamine-formaldehyde-tartaric acid resin (and a discussion about pseudo second order model). Desalin. Water Treat. 2012, 44, 128−141. (5) Gupta, V. K.; Jain, R.; Nayak, A.; Agarwal, S.; Shrivastava, M. Removal of the hazardous dyetartrazine by photodegradation on titanium dioxide surface. Mater. Sci. Eng., C 2011, 31, 1062−1067. (6) Mahmoud, M. E.; Saad, E. A.; Soliman, M. A.; Abdelwahab, M. S. Synthesis and surface protection of nano zerovalent iron (NZVI) with 3aminopropyltrimethoxysilane for water remediation of cobalt and zinc and their radioactive isotopes. RSC Adv. 2016, 6, 66242−66251. (7) Asfaram, A.; Ghaedi, M.; Ghezelbash, G. R.; Pepe, F. Application of experimental design and derivative spectrophotometry methods in optimization and analysis of biosorption of binary mixtures of basic dyes from aqueous solutions. Ecotoxicol. Environ. Saf. 2017, 139, 219−227. (8) Dotto, G. L.; Vieira, M. L.; Pinto, L. A. Kinetics and mechanism of tartrazine adsorption onto chitin and chitosan. Ind. Eng. Chem. Res. 2012, 51, 6862−6868. (9) Gautam, R. K.; Gautam, P. K.; Banerjee, S.; Rawat, V.; Soni, S.; Sharma, S. K.; Chattopadhyaya, M. C. Removal of tartrazine by activated carbon biosorbents of Lantana camara: kinetics, equilibrium modeling and spectroscopic analysis. J. Environ. Chem. Eng. 2015, 3, 79−88. (10) Sharma, Y. C.; Upadhyay, S. N. Removal of a cationic dye from wastewaters by adsorption on activated carbon developed from coconut coir. Energy Fuels 2009, 23, 2983−2988. (11) Mahmoud, M. E.; Ahmed, S. B.; Osman, M. M.; Abdel-Fattah, T. M. A novel composite of nanomagnetite-immobilized-baker’s yeast on the surface of activated carbon for magnetic solid phase extraction of Hg (II). Fuel 2015, 139, 614−621. 4562

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563

Research Article

ACS Sustainable Chemistry & Engineering (30) Chu, W. Dye removal from textile dye wastewater using recycled alum sludge. Water Res. 2001, 35, 3147−3152. (31) Benavidez, T. E.; Torrente, D.; Marucho, M.; Garcia, C. D. Adsorption of soft and hard proteins onto OTCEs under the influence of an external electric field. Langmuir 2015, 31, 2455−2462. (32) Boulanouar, O.; Khatyr, A.; Herlem, G.; Palmino, F.; Sanche, L.; Fromm, M. Soft adsorption of densely packed layers of dna-plasmid 3 1, 3-diaminopropane complexes onto highly oriented pyrolitic graphite designed to erode in water. J. Phys. Chem. C 2011, 115, 21291−21298. (33) Dil, E. A.; Ghaedi, M.; Asfaram, A. The performance of nanorods material as adsorbent for removal of azo dyes and heavy metal ions: Application of ultrasound wave, optimization and modeling. Ultrason. Sonochem. 2017, 34, 792−802. (34) Nasiri Azad, F.; Ghaedi, M.; Dashtian, K.; Montazerozohori, M.; Hajati, S.; Alipanahpour, E. Preparation and characterization of MWCNTs functionalized by N-(3-nitrobenzylidene)-N[prime or minute]-trimethoxysilylpropyl-ethane-1,2-diamine for the removal of aluminum(iii) ions via complexation with eriochrome cyanine R: spectrophotometric detection and optimization. RSC Adv. 2015, 5, 61060−61069. (35) Asfaram, A.; Ghaedi, M.; Ghezelbash, G. R. Biosorption of Zn2+, Ni2+ and Co2+ from water samples onto Yarrowia lipolytica ISF7 using a response surface methodology, and analyzed by inductively coupled plasma optical emission spectrometry (ICP-OES). RSC Adv. 2016, 6, 23599−23610. (36) Jamshidi, M.; Ghaedi, M.; Dashtian, K.; Ghaedi, A. M.; Hajati, S.; Goudarzi, A.; Alipanahpour, E. Highly efficient simultaneous ultrasonic assisted adsorption of brilliant green and eosin B onto ZnS nanoparticles loaded activated carbon: Artificial neural network modeling and central composite design optimization. Spectrochim. Acta, Part A 2016, 153, 257−267. (37) Asfaram, A.; Ghaedi, M.; Yousefi, F.; Dastkhoon, M. Experimental design and modeling of ultrasound assisted simultaneous adsorption of cationic dyes onto ZnS: Mn-NPs-AC from binary mixture. Ultrason. Sonochem. 2016, 33, 77−89. (38) Amaral, P. F. F.; Da Silva, J. M.; Lehocky, M.; Barros-Timmons, A. M. V.; Coelho, M. A. Z.; Marrucho, I. M.; Coutinho, J. A. P. Production and characterization of a bioemulsifier from Yarrowia lipolytica. Process Biochem. 2006, 41, 1894−1898. (39) Ibrahim, I.; Ali, I. O.; Salama, T. M.; Bahgat, A.; Mohamed, M. M. Synthesis of magnetically recyclable spinel ferrite (MFe2O4, M= Zn, Co, Mn) nanocrystals engineered by sol gel-hydrothermal technology: High catalytic performances for nitroarenes reduction. Appl. Catal., B 2016, 181, 389−402. (40) Tawa, K.; Kamada, K.; Kiyohara, K.; Ohta, K.; Yasumatsu, D.; Sekkat, Z.; Kawata, S. Photoinduced reorientation of azo dyes bonded to polyurethane studied by polarized FT-IR spectroscopy. Macromolecules 2001, 34, 8232−8238. (41) Koenig, U.; Chol, G. Röntgenbeugungs-und Neutronenbeugungsuntersuchungen an Ferriten der reihe MnxZn1−xFe2O4. J. Appl. Crystallogr. 1968, 1, 124−126. (42) Hu, X.-W.; Liu, S.; Qu, B.-T.; You, X.-Z. Starfish-shaped Co3O4/ ZnFe2O4 hollow nanocomposite: synthesis, supercapacity, and magnetic properties. ACS Appl. Mater. Interfaces 2015, 7, 9972−9981. (43) Lima, E., Jr; Winkler, E. L.; Tobia, D.; Troiani, H. E.; Zysler, R. D.; Agostinelli, E.; Fiorani, D. Bimagnetic CoO Core/CoFe2O4 shell nanoparticles: synthesis and magnetic properties. Chem. Mater. 2012, 24, 512−516. (44) Jamshidi, M.; Ghaedi, M.; Dashtian, K.; Hajati, S.; Bazrafshan, A. A. Sonochemical assisted hydrothermal synthesis of ZnO: Cr nanoparticles loaded activated carbon for simultaneous ultrasound-assisted adsorption of ternary toxic organic dye: derivative spectrophotometric, optimization, kinetic and isotherm study. Ultrason. Sonochem. 2016, 32, 119−131. (45) Arabi, M.; Ghaedi, M.; Ostovan, A. Development of a lower toxic approach based on green synthesis of water-compatible molecularly imprinted nanoparticles for the extraction of hydrochlorothiazide from human urine. ACS Sustainable Chem. Eng. 2017, 5, 3775−3785.

(46) Singh, K. P.; Gupta, S.; Singh, A. K.; Sinha, S. Experimental design and response surface modeling for optimization of Rhodamine B removal from water by magnetic nanocomposite. Chem. Eng. J. 2010, 165, 151−160. (47) Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361−1403. (48) Freundlich, H. Uber die adsorption in lusungen. Z. Phys. Chem. 1907, 57, 385. (49) Dubinin, M.; Serpinsky, V. Isotherm equation for water vapor adsorption by microporous carbonaceous adsorbents. Carbon 1981, 19, 402−403. (50) Temkin, M.; Pyzhev, V. Recent modifications to Langmuir isotherms. Acta Physico-Chim. Sinica 1940, 12, 217−222. (51) Thiam, A.; Brillas, E.; Garrido, J. A.; Rodríguez, R. M.; Sirés, I. Routes for the electrochemical degradation of the artificial food azocolour Ponceau 4R by advanced oxidation processes. Appl. Catal., B 2016, 180, 227−236. (52) Zhou, Z.; Lin, S.; Yue, T.; Lee, T.-C. Adsorption of food dyes from aqueous solution by glutaraldehyde cross-linked magnetic chitosan nanoparticles. J. Food Eng. 2014, 126, 133−141. (53) Benincá, C.; Peralta-Zamora, P.; Camargo, R. C.; Tavares, C. R. G.; Zanoelo, E. F.; Igarashi-Mafra, L. Kinetics of oxidation of ponceau 4R in aqueous solutions by Fenton and photo-Fenton processes. React. Kinet., Mech. Catal. 2012, 105, 293−306. (54) Sadik, W. A.; Nashed, A. W.; El-Demerdash, A.-G. M. Photodecolourization of ponceau 4R by heterogeneous photocatalysis. J. Photochem. Photobiol., A 2007, 189, 135−140. (55) Benincá, C.; Peralta-Zamora, P.; Tavares, C. R. G.; Igarashi-Mafra, L. Degradation of an azo dye (ponceau 4R) and treatment of wastewater from a food industry by ozonation. Ozone: Sci. Eng. 2013, 35, 295−301. (56) Arabzadeh, N.; Khosravi, A.; Mohammadi, A.; Mahmoodi, N. M. Enhanced photodegradation of hazardous tartrazine by composite of nanomolecularly imprinted polymer-nanophotocatalyst with high efficiency. Desalin. Water Treat. 2016, 57, 3142−3151.

4563

DOI: 10.1021/acssuschemeng.7b03205 ACS Sustainable Chem. Eng. 2018, 6, 4549−4563