Development and Reuse of Amine-Grafted Chitosan Hybrid Beads in

Jan 2, 2018 - A variety of adsorbent materials have been tried day by day to minimize the effects of blue baby syndrome and eutrophication but most of...
0 downloads 9 Views 8MB Size
Article Cite This: J. Chem. Eng. Data XXXX, XXX, XXX−XXX

pubs.acs.org/jced

Development and Reuse of Amine-Grafted Chitosan Hybrid Beads in the Retention of Nitrate and Phosphate Ilango Aswin Kumar and Natrayasamy Viswanathan* Department of Chemistry, Anna University, University College of Engineering - Dindigul, Reddiyarchatram, Dindigul - 624 622, Tamilnadu, India S Supporting Information *

ABSTRACT: A variety of adsorbent materials have been tried day by day to minimize the effects of blue baby syndrome and eutrophication but most of them constitute academic interest because of their properties such as powder form, poor selectivity, not easily separable, and difficult regeneration; hence, they could not be used at field conditions. To overcome such technological bottle-necks, the present study has been carried out with aminefunctionalized magnetic chitosan (AFMCS) composite beads which was prepared and applied for the remediation of nitrate and phosphate in batch scale. The fabricated adsorbent materials were characterized by Fourier transform infrared spectroscopy, Brunauer−Emmett−Teller analysis, transmission electron microscopy, X-ray diffraction, scanning electron microscopy, and energy dispersive X-ray analysis with mapping analysis. The numerous adsorption-influenced parameters such as contact time, adsorbent dosage, initial ions concentration, co-anions, solution pH, and temperature were optimized. The various sorption isotherm models were used to fit the equilibrium data. The feasible mechanism toward nitrate and phosphate removal was interpreted. The efficient regeneration and repeated use of AFMCS composite beads were achieved up to six cycles. The applicability of easily separable AFMCS composite beads at field conditions was checked by collecting nitrate and phosphate contaminated water sample from the nearby area of Dindigul district.

1. INTRODUCTION Nitrogen and phosphorus are essential nutrients which have a significant contribution to biomass growth, energy transport, and agricultural fields. However, the chemical reduction of nitrate into nitrite in potable water could lead to crucial health hazards when it reacts with hemoglobin. This reaction causes methemoglobinemia or blue baby syndrome referring to the blue colored skin of young children.1 In addition, the elevated nitrate content in drinking water is also a main reason for the causes of gastric cancer by reduced nitrate in the form of nitrosamines.2 The phenomonen of algae blooms of the aquatic region is often referred to as eutrophication, which leads to esthetic troubles such as the worsening of reservoirs, coastal region, rivers, and lakes by depleted dissolved oxygen content in water.3 Apart from decomposed organic biomaterials and weathering rocks, nitrate and phosphate vigorously enters into the precious water sources by means of the manufacture of polishers, animal excretion, fertilizers, industrial effluents, and detergents.4 Even though both nitrate and phosphate play a vital role in many biological systems, agricultural field, plant growth and their development, they seem to be toxic when their allowable limits in water is exceeded. The tolerance limits of nitrate and phosphate in drinking water are 40 and 11), respectively.49 Among them, H2PO4− was predominantly observed by the reaction of both HPO42− and hydrogen ion (H+) as given by eq 2.

functionalized adsorbent has more activity toward the removal of toxic pollutants, and also AFMCS composite beads reach equilibrium time at 30 and 40 min for nitrate and phosphate sorption. Overall, the SCs of nitrate and phosphate were rapid in the first 20−40 min, and also not all the adsorbent materials possess high nitrate and phosphate SC except AFMCS composite beads which hold the nitrate and phosphate SCs of 38.40 and 42.95 mg·g−1 respectively. As a result, further studies were limited to AFMCS composite beads with fixed contact time of 30 and 40 min toward nitrate and phosphate removal. 3.3. Effect of Adsorbent Dosage. The various adsorbent dosage (0.025−0.15 g) levels were taken to optimize the amount of adsorbent dosage required for the efficient removal of nitrate and phosphate. The relation between dosage and SCs of nitrate and phosphate are shown in Figure 5C. It was clearly observed that increasing the adsorbent dosage reasonably increases both nitrate and phosphate SC. This is mainly due to the increase in the adsorbent dosage that increases the active sites of adsorbent thereby increasing the SC. At the same time, the addition of adsorbent at a dosage greater than 0.1 g leads to only a mild increase in the SC of nitrate and phosphate. Taking this into consideration, the appropriate adsorbent dosage was fixed as 0.1 g throughout the studies. 3.4. Effect of Solution pH. The solution pH is a significant parameter in the adsorption process because it often affects the active sites of the adsorbent surface. The pH effect toward nitrate and phosphate removal was determined by altering the solution pH from 3 to 11 using HCl/NaOH solution.

HPO4 2 − + H+ → H 2PO4 −

(pH ∼ 2−7)

(2)

In this condition the protonated sorbent surface might electrostatically bind with H2PO4−.3 This attraction was held up until the pH reached neutral which caused an increase in phosphate sorption as shown in Figure 5D. Even though both HPO42− and PO43− are the species that exist at a basic pH level, a decrease in phosphate sorption was observed due to the competition of available OH− ions in basic pH medium. The F

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 6. Effect of (A) initial adsorbate concentration, (B) temperature, and (C) aggressive co-ions on the nitrate and phosphate SCs of AFMCS composite beads.

that only a slight increase in the removal was achieved. Taking this into an account, the remaining studies were examined for nitrate and phosphate having initial ion concentration of 100 mg·L−1. 3.6. Effect of Temperature. The reaction temperature plays a vital role in the adsorption process. To analyze the effect of temperature on nitrate and phosphate adsorption by AFMCS composite beads, about 0.1 g of AFMCS composite beads was added with 50 mL of 100 mg·L−1 individual nitrate and phosphate solution by varied reaction temperatures which has been fixed as 303, 313, and 323 K. It was clearly observed from Figure 6B, that the nitrate adsorption onto AFMCS composite beads slightly decreased from 38.40 to 35.96 mg·g−1 with an increase in the reaction temperature, suggesting that the nitrate adsorption process is exothermic in nature, whereas in phosphate adsorption a slight increase from 42.95 to 45.73 mg·g−1 suggests that the phosphate adsorption process is endothermic in nature with increasing reaction temperature in the range of 303 to 323 K, respectively.50,51 3.7. Effect of Aggressive Co-anions. The active sites of the adsorbent surface were often afflicted with various common anions such as chloride (Cl−), sulfate (SO42−), and bicarbonate (HCO3−) which occur in natural/wastewater, and these were expected to compete with nitrate and phosphate sorption. To analyze the competing effect of these anions, about 0.1 g of AFMCS composite beads was added into 50 mL of a mixture of 200 mg·L−1 individual co-anions solution with 100 mg·L−1 nitrate and phosphate solution. The effects of these co-anions toward nitrate and phosphate sorption are depicted in Figure 6C. It was observed that all the common anions compete with

pH at zero point charge (pHzpc) of AFMCS composite beads measured as 5.46 clearly denotes that when pH < pHzpc the adsorbent surface was protonated which shows the electrostatic attraction of protonated adsorbent surface with both nitrate and phosphate. However, pH > pHzpc leads to a decrease in nitrate and phosphate sorption because there is a competition for occupying the active sites of the adsorbent by OH− ions with nitrate and phosphate. After the nitrate and phosphate sorption, the equilibrium pH tended toward neutral which shows the suitability of the adsorbent at varied pH environment. 3.5. Influence of Initial Adsorbate Concentration. To verify the influence of initial adsorbate concentration on removal of nitrate and phosphate about 0.1 g of AFMCS composite beads were added into 50 mL of both nitrate and phosphate solutions with different initial ion concentrations in the range of 20−140 mg·L−1, and a fixed contact time of 30 and 40 min was carried out at room temperature. It was noticed from Figure 6A that a higher adsorbate concentration improved nitrate and phosphate sorption. This is mainly due to an elevated adsorbate concentration implying an available energetic, dynamic, driving force to overcome the mass transmit barrier in the aqueous phase at the solid/liquid boundary. This leads to fast transport of ions and therefore the adsorbent reaches saturation more quickly with decreased time. The nitrate SCs were increased from 9.10 to 44.34 mg·g−1, whereas the phosphate SCs were increased from 11.56 to 49.82 mg·g−1. As a result, there was a gradual increase in the removal of nitrate and phosphate observed in the initial ion concentration range from 20 to 100 mg·L−1. After the initial adsorbate concentration is raised to 100 mg·L−1, it is observed G

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 2. Isotherms of AFMCS Composite Beads for Nitrate and Phosphate Adsorption nitrate parameters

303 K

phosphate

313 K

1/n n kF (mg·g−1) (L·mg−1)1/n r sd χ2

0.752 4.463 23.16 0.980 0.098 1.304

0.741 4.496 22.93 0.983 0.102 1.309

Qo (mg.g−1) b (L·g−1) RL r sd χ2

37.45 0.637 2.081 0.989 0.067 0.125

36.21 0.798 2.149 0990 0.072 0.129

kDR (mol2·J2−1) Xm (mg·g−1) E (kJ mol−1) r sd χ2

3.21 × 10−04 12.37 5.971 0.667 1.007 5.009

3.18 × 11.41 6.008 0.669 1.011 5.014

323 K

303 K

Freundlich Isotherm 0.744 5.073 21.45 0.982 0.106 1.311 Langmuir Isotherm 35.01 0.961 2.152 0.990 0.075 0.133 D-R Isotherm 10−04 3.19 × 10−04 10.28 6.657 0.670 1.014 5.013

313 K

323 K

0.903 6.783 33.05 0.998 0.074 0.009

0.895 6.947 32.23 0.997 0.073 0.018

0.899 7.252 34.05 0.999 0.076 0.021

41.49 1.982 3.082 0.978 0.357 0.538

42.28 2.003 3.904 0977 0.358 0.536

44.10 2.041 3.996 0.979 0.356 0.537

5.031 × 10−04 17.84 7.081 0.732 1.974 2.603

4.98 × 10−04 19.34 7.352 0.731 1.976 2.605

5.99 × 10−04 20.89 7.995 0.734 1.979 2.605

of ln qe vs ε2 and also the values for parameters such as E, Xm, KDR with r, χ2 and sd are given in Table 2. It is clear that the Langmuir isotherm was the appropriate model for nitrate adsorption onto AFMCS composite beads, while the Freundlich isotherm was suited for phosphate adsorption onto AFMCS composite beads due to higher r value and lower values of both χ2 and sd. Moreover, the mean free energy of adsorption (E) lying between 5 and 8 kJ·mol−1 from D-R isotherm might suggest that a physical adsorption links the synthesized AFMCS composite beads with that of both nitrate and phosphate. The experimentally measured raw isotherm (c vs q) data were given in the Supporting Information file as Figures S1 and S2, where the experimental data was compared with various isotherms. 3.9. Adsorption Thermodynamics. The significant nature of the nitrate and phosphate adsorption process at varied reaction temperatures of 303, 313, and 323 K was governed by Khan and Singh method.58 The values of various important thermodynamic parameters such as the standard Gibb’s free energy (ΔG°), standard entropy change (ΔS°), and standard enthalpy change (ΔH°) were illustrated in Table 3. It was noticed that the gradual decrease (more negative) in the ΔG° value with increase in temperature for adsorption of both nitrate and phosphate indicates the spontaneous and more favored nature of the adsorption process from 303 to 323 K.59,60 The positive value of ΔS° for both nitrate and phosphate

nitrate sorption, except bicarbonate which has a negligible competition.52 Among the co-anions, sulfate possesses more anionic charge clouds which could easily compete with nitrate, thus the multivalent anion with higher charge density has a tendency to absorb more quickly than a univalent anion such as nitrate. In the case of phosphate, the competing effect was governed by both chloride and sulfate which occupies onto the sorbent surface instead of phosphate.53 Though chloride has maximum electron affinity, sulfate only possesses the significant competing effect toward phosphate sorption because of the comparable ionic size of sulfate with that of phosphate.3 3.8. Adsorption Isotherms. The adsorption isotherms data and their related parameters were summarized by Langmuir,54 Freundlich,55 and Dubinin−Radushkevich (D-R) models.56 Mainly, these three isotherm models are employed to investigate the binding nature of the adsorbent surface with that of the adsorbate in the solid/liquid interphase. The experimental results are shown in Table 2, where the equilibrium SCs are improved with increasing initial ion concentration (80, 100, 120, and 140 mg·L−1) of both nitrate and phosphate with varied reaction temperature ranges of 303, 313, and 323 K. Langmuir isotherm model covers the homogeneous type of adsorbent with uniform adsorption energy during the single layer adsorption process. Langmuir coefficients such as Q° (quantity of adsorbate adsorbed at single layer coverage in mg· g−1) and b (Langmuir constant in L·mol−1) are obtained by the plot of Ce vs Ce/qe.57 The reversible and nonideal Freundlich isotherm of heterogeneous system could be expressed by the linear plot of log qe vs log Ce where the equilibrium concentration (Ce) which provides the values of n (heterogeneity factor or Freundlich intensity constant) and kF (Freundlich capacity constant) are attained from the slope and intercept. The conditions for a favorable adsorption process were signified from n values around 1 to 10 and 1/n values around 0 to 1 in Freundlich isotherm.36 In addition, the appropriate linearity of D-R isotherm was attained by the plot

Table 3. Thermodynamic Parameters of AFMCS Composite Beads for Nitrate and Phosphate Removal thermodynamic parameters ΔGo (kJ mol−1)

ΔHo (kJ mol−1) ΔSo (J mol−1 K−1) H

303 K 313 K 323 K

nitrate

phosphate

−3.53 −3.68 −4.01 −4.52 33.41

−7.19 −7.53 −7.85 9.09 45.19

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 7. Possible mechanism of nitrate and phosphate removal by AFMCS composite beads.

adsorption shows an increase in randomness and disorder at the solid/liquid interface of the adsorbed species.61 Moreover, the negative value of ΔH° for nitrate adsorption (−4.52 kJ mol−1) denotes that the interaction of the nitrate and AFMCS composite beads is exothermic in nature, whereas the positive value of ΔH° for phosphate adsorption (9.09 kJ mol−1) implies the interaction between phosphate and AFMCS composite beads is endothermic in nature.50 3.10. Adsorption Mechanism. The nitrate and phosphate sorption mechanism may depend on the reaction of metal ions, surface functionality, and surface area of AFMCS composite beads.62 The possible mechanism of amine-grafted magnetic chitosan composite beads onto nitrate and phosphate sorption are shown in Figure 7. The sorptive toward nitrate and phosphate removal was mainly governed by both electrostatic attraction and surface complexation. In most cases, the metal oxide has the tendency to form hydrated metal oxides (M− OH) in aqueous medium.63 Similarly, the magnetic iron oxide could also possess the hydrated form as given by eq 3. FeO + H 2O/H+ → Fe(OH)3 → Fe3 + + 3OH−

Table 4. Comparison of Nitrate and Phosphate Adsorption Capacities of the Synthesized Adsorbents with Other Reported Adsorbent Materials adsorption capacity (mg·g−1) serial no.

nitrate

phosphate

1

AFMCS composite beads

38.40

42.95

2

MCS composite beads

12.80

17.36

3

poly(dimethyl diallyl ammonium chloride)/polyacrylamide hydrogel ammonium functionalized mesoporous MCM-48 silica zerovalent iron supported activated carbon chitosan saturated with copper(II) modified commercial activated carbon copper and magnesium impregnated alumina Zr@AlgBent composite beads granular chitosan Fe3+ complex lanthanum incorporated porous zeolite waste solids containing boron impurity stratified wheat straw resin Zr@CSBent biocomposite carbon-silica nano composite

present study present study 2

34.80

39.70

65

4.60

1.75

66

28.86 21.51

67 68

8.00

69

4 5 6 7 8

(3)

9 10 11

Further, the charge separation of the ferric hydroxide (Fe(OH)3) leads to the formation of the Lewis acidic ferric (Fe3+) metal ion which has the ability to form an electrostatic attraction with nitrate and phosphate, it also coordinates a bond with H2PO4− by surface complexation.64 The outermost stable electronic configuration (2s22p3) of nitrogen is three which means that it could have a maximum of three chemical bonds with other molecules or ions. In addition, it has one lone pair of electrons which leads to making it a Lewis base, but in the case of AFMCS composite beads, the surface nitrogen possesses a positive charge as −NH2+Cl and so it may chemically bind with both nitrate and phosphate by means of electrostatic attraction.39 3.11. Comparative Analysis. The comparative study on adsorption capacities of AFMCS composite beads toward nitrate and phosphate with other reported adsorbent materials available on the market are shown in Table 4. It is observed that the adsorption capacity of synthesized AFMCS composite beads toward nitrate and phosphate sorption was found to possess appreciable adsorption capacity compared with reported adsorbents. 3.12. Field Trial. The applicability of synthesized AFMCS composite beads at field conditions was verified for nitrate and phosphate contaminated water samples which are collected from the nearby area of Dindigul district. Both nitrate and phosphate were responsible for the formation of algae in

12 13 14 15

name of the adsorbent

1.40

30.25 17.20

3 70 71

52.50

72

40.86

73 18 74

2.04

63.20

reference

35.30 11.34

natural/wastewater, since the removal of nitrate and phosphate was carried out by adding 0.1 g of AFMCS composite beads into 50 mL of field water sample, and the results are shown in Table 5. The initial nitrate and phosphate concentration of collected field water was observed as 14.90 and 19.80 mg·L−1, Table 5. Field Applicability of AFMCS Composite Beads AFMCS composite beads

I

water quality parameters

before

after

initial nitrate concentration (mg·L−1) initial phosphate concentration (mg·L−1) pH Cl− (mg·L−1) total hardness (mg·L−1) total dissolved solids (mg·L−1)

14.90 19.80 5.48 368 618 426

nil nil 6.23 184 512 209

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

4. CONCLUSIONS The present study focused on the preparation and easy separation of AFMCS composite beads for enhanced removal of nitrate and phosphate from water. The results showed that the protonated sorbent surface binds with nitrate and phosphate by means of electrostatic attraction. The prepared AFMCS composite beads hold enhanced nitrate and phosphate SCs of 38.40 and 42.95 mg·g−1, respectively. Among the coanions present, the predominant competing effect was governed by sulfate as it occupies the sorbent surface instead of both nitrate and phosphate. The functional analysis, surface morphology, specific surface area, elemental investigation, and crystalline nature of the prepared adsorbents were characterized using FTIR, BET, XRD, TEM, SEM, and EDAX with mapping analysis. The good regularity of experimental data was attained for Langmuir and Freundlich isotherm models toward nitrate and phosphate adsorption. An adsorption thermodynamic study clearly explained the exothermic and endothermic nature of the nitrate and phosphate adsorption process. The adsorption system follows electrostatic attraction and surface complexation mechanism. In addition, the regeneration performance of AFMCS composite beads was observed up to six cycles without significant loss of removal efficiency. The successful field analysis of prepared AFMCS composite beads demonstrated its applicability in nitrate and phosphate contaminated field water sample collected from the nearby area of Dindigul district.

respectively. The removal performance was achieved by reducing the toxic anionic concentration to less than tolerance limit within 5 min. This is because grafted polymeric composite beads are highly reactive toward anionic pollutants. Considerably, the observed removal capacity of nitrate and phosphate is almost nil in the eutrophicated water sample which is shown in Table 5. The functional reactive sites of AFMCS composite beads reduce the concentration of chloride, dissolved solids, and total hardness which are generally present in natural water. The use of AFMCS composite beads possesses the added advantages of high stability, quick removal, and easy separation during filtration which signifies it aptness at field conditions. 3.13. AdsorptionDesorption Cycle. The reuse of prepared AFMCS composite beads were investigated by batch adsorption−desorption cycles. Several reports show that aqueous NaOH was a proficient regenerant for desorbing both nitrate and phosphate from iron oxides sprayed on adsorbent surfaces.75 In addition, Wang et al. have reported the desorption studies of amine-functionalized corn straw by the use of NaOH.36 Taking this, the desorption study and removal efficiencies of AFMCS composite beads were evaluated using NaOH. To examine the reusability of the adsorbent, 0.1 g of nitrate-sorbed and 0.1g of phosphate-sorbed AFMCS composite beads were added into 50 mL of 0.25 M NaOH individual solutions. The mixtures were stirred for 60 min at room temperature, filtered, and then the concentration of the filtrates was determined by UV-vis spectrophotometry. The results are depicted in Figure 8.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.jced.7b00751. Experimentally measured raw isotherm (c vs q) data of prepared AFMCS composite beads where the experimental data was compared with various isotherms toward nitrate and phosphate removal (PDF)



AUTHOR INFORMATION

Corresponding Author

*Tel.: +91-451-2554066. Fax: +91-451-2554066. E-mail: [email protected]. ORCID

Natrayasamy Viswanathan: 0000-0003-4939-3442 Funding

Figure 8. Regeneration performance of prepared AFMCS composite beads.

The authors were gratefully acknowledging University Grants Commission (F. No. 43-179/2014(SR)), New Delhi, India, for the provision of financial support to carry out this research work.

The first cycle results show the percentage removal efficiency of nitrate and phosphate as 96.88 and 98.16%, respectively. Further, it was decreased from 95.46 to 40.69% for nitrate and 95.27 to 59.47% for phosphate. Moreover, the SCs of both nitrate and phosphate were also decreased from 37.20 to 15.62 mg·g−1 and 42.16 to 25.54 mg·g−1 at the completion of eight cycles which are shown in Figure 8. This occurs because of the competing effect of OH− ions that occupy AFMCS composite beads surface instead of both nitrate and phosphate.20 As a result, AFMCS composite beads have been utilized up to six cycles with no significant loss in their removal efficiency, which shows the reutilization ability of AFMCS composite beads toward the removal of nitrate and phosphate, minimizing the operating cost of the prepared adsorbent.

Notes

The authors declare no competing financial interest.



REFERENCES

(1) Chatterjee, S.; Woo, S. H. The Removal of Nitrate from Aqueous Solutions by Chitosan Hydrogel Beads. J. Hazard. Mater. 2009, 164, 1012−1018. (2) Zheng, Y.; Wang, A. Nitrate Adsorption using Poly(Dimethyl Diallyl Ammonium Chloride)/ Polyacrylamide Hydrogel. J. Chem. Eng. Data 2010, 55, 3494−3500. (3) Aswin Kumar, I.; Viswanathan, N. Fabrication of Metal Ions Cross-Linked Alginate Assisted Biocomposite Beads for Selective Phosphate Removal. J. Environ. Chem. Eng. 2017, 5, 1438−1446.

J

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Leaf Extract in the Presence of CTAB Surfactant. Chemosphere 2016, 159, 23−31. (26) Feng, Y.; Ji-Lai, G.; Guang-Ming, Z.; Qiu-Ya, N.; Hui-Ying, Z.; Ching-Gang, N.; Jiu-Hua, D.; Yan, M. Adsorption of Cd(II) and Zn(II) from Aqueous Solutions using Magnetic Hydroxyapatite Nanoparticles as Adsorbents. Chem. Eng. J. 2010, 162, 487−494. (27) Padilla-Rodriguez, A.; Hernandez-Viezcas, J. A.; Peralta-Videa, J. R.; Gardea-Torresdey, J. L.; Perales-Perez, O.; Roman-Velazquez, F. R. Synthesis of Protonated Chitosan Flakes for the Removal of Vanadium(III, IV and V) Oxyanions from Aqueous Solutions. Microchem. J. 2015, 118, 1−11. (28) Balan, V.; Verestiuc, L. Strategies to Improve Chitosan Hemocompatibility: A Review. Eur. Polym. J. 2014, 53, 171−188. (29) Anirudhan, T. S.; Rijith, S.; Tharun, A. R. Adsorptive Removal of Thorium (IV) from Aqueous Solutions using Poly(Methacrylic Acid) Grafted Chitosan/Bentonite Composite Matrix: Process Design and Equilibrium Studies. Colloids Surf., A 2010, 368, 13−22. (30) Xie, J.; Li, C.; Chi, L.; Wu, D. Chitosan Modified Zeolite as a Versatile Adsorbent for the Removal of Different Pollutants from Water. Fuel 2013, 103, 480−485. (31) Pandi, K.; Viswanathan, N. A Novel Metal Coordination Enabled in Carboxylated Alginic Acid for Effective Fluoride Removal. Carbohydr. Polym. 2015, 118, 242−249. (32) Jiang, Y.-U.; Yu, X.-Y.; Luo, T.; Jia, Y.; Liu, J.-H.; Huang, X.-J. γFe2O3 Nanoparticles Encapsulated Millimeter-Sized Magnetic Chitosan Beads for Removal of Cr(VI) from Water: Thermodynamics, Kinetics, Regeneration and Uptake Mechanisms. J. Chem. Eng. Data 2013, 58, 3142−3149. (33) Periyasamy, S.; Gopalakannan, V.; Viswanathan, N. Fabrication of Magnetic Particles Imprinted Cellulose Based Biocomposites for Chromium(VI) Removal. Carbohydr. Polym. 2017, 174, 352−359. (34) Ahmad, M.; Manzoor, K.; Chaudhuri, R. R.; Ikram, S. Thiocarbohydrazide Cross-Linked Oxidized Chitosan and Poly(Vinyl Alcohol): A Green Framework as Efficient Cu(II), Pb(II) and Hg(II) Adsorbent. J. Chem. Eng. Data 2017, 62, 2044−2055. (35) Hong-Tao, F.; Fan, X.; Li, J.; Guo, M.; Zhang, D.; Yan, F.; Sun, T. Selective Removal of Arsenic(V) from Aqueous Solution using a Surface-ion-Imprinted Amine-Functionalized Silica Gel Sorbent. Ind. Eng. Chem. Res. 2012, 51, 5216−5223. (36) Wang, H.; Xu, X.; Ren, Z.; Gao, B. Removal of Phosphate and Chromium(VI) from Liquids by Amine-Crosslinked Nano-Fe3O4 Biosorbent Derived from Corn Straw. RSC Adv. 2016, 6, 47237− 47248. (37) Lopez-Ramon, M. V.; Stoeckli, F.; Moreno-Castilla, C.; Carrasco-Marin, F. On the Characterization of the Acidic and Basic Surface Sites on Carbons by Various Techniques. Carbon 1999, 37, 1215−1221. (38) APHA. Standard Methods for the Examination of Water and Wastewater; American Public Health Association: Washington, DC, USA, 2005. (39) Yang, D.; Qiu, L.; Yang, Y. Efficient Adsorption of Methyl Orange using a Modified Chitosan Magnetic Composite Adsorbent. J. Chem. Eng. Data 2016, 61, 3933−3940. (40) Muthu Prabhu, S.; Viswanathan, N.; Meenakshi, S. Defluoridation of Water using Chitosan Assisted Ethylenediamine Functionalized Synthetic Polymeric Blends. Int. J. Biol. Macromol. 2014, 70, 621−627. (41) Pandi, K.; Viswanathan, N. Remediation of Fluoride from Drinking Water using Magnetic Iron Oxide Coated Hydrotalcite/ Chitosan Composite. Int. J. Biol. Macromol. 2017, 104, 1569−1577. (42) Lu, W.; Shen, Y.; Xie, A.; Zhang, W. Green Synthesis and Characterization of Super Paramagnetic Fe3O4 Nanoparticles. J. Magn. Magn. Mater. 2010, 322, 1828−1833. (43) Xu, X.; Gao, B.; Tan, X.; Zhang, X.; Yue, D. T.; Yue, Q. Uptake of Perchlorate from Aqueous Solutions by Amine-Crosslinked Cotton Stalk. Carbohydr. Polym. 2013, 98, 132−138. (44) Sundaraganesan, N.; Kalaichelvan, S.; Meganathan, C.; Joshua, B. D.; Cornard, J. FT-IR, FT-Raman Spectra and Ab Initio HF and

(4) Liang, X.; Wang, Z.; Zhang, Y.; Zhu, C.; Lin, L.; Xu, L. No-Tillage Effects on N and P Exports Across a Rice-Planted Watershed. Environ. Sci. Pollut. Res. 2016, 23, 8598−8609. (5) Guidelines for Drinking Water Quality. 4th ed.; World Health Organization (WHO): Geneva, Switzerland, 2011. (6) Galalgorchev, H. Who guidelines for drinking-water quality. IWSA specialized conference on quality aspects of water supply, 1992; Vol. 11, pp 1−16. (7) Ghafari, S.; Hasan, M.; Aroua, M. K. Bio-Electrochemical Removal of Nitrate from Water and Wastewater - A Review. Bioresour. Technol. 2008, 99, 3965−3974. (8) Rezania, B.; Cicek, N.; Oleszkiewicz, J. A. Kinetics of HydrogenDependent Denitrification under Varying pH and Temperature Conditions. Biotechnol. Bioeng. 2005, 92, 900−906. (9) Yang, G. C. C.; Hsaio-Lan, L. Chemical Reduction of Nitrate by Nanosized Iron: Kinetics and Pathways. Water Res. 2005, 39, 884− 894. (10) Wang, W.; Wei, H.-Z.; Jiang, S.-Y.; Eastoe, C. J.; Guo, Q.; Lin, Y.-B. Adsorption Behavior of Metasilicate on N-methyl D-glucamine Functional Groups and Associated Silicon Isotope Fractionation. Langmuir 2016, 32, 8872−8881. (11) Quan, X.; Ye, C.; Xiong, Y.; Xiang, J.; Wang, F. Simultaneous Removal of Ammonia, P and COD from Anaerobically Digested Piggery Wastewater using an Integrated Process of Chemical Precipitation and Air Stripping. J. Hazard. Mater. 2010, 178, 326−332. (12) Liu, N.; Zhang, Q.; Qu, R.; Zhang, W.; Li, H.; Wei, Y.; Feng, L. Nanocomposite Deposited Membrane for Oil-in-Water Emulsion Separation with In Situ Removal of Anionic Dyes and Surfactants. Langmuir 2017, 33, 7380−7388. (13) Bhatnagar, A.; Sillanpaa, M. A Review of Emerging Adsorbents for Nitrate Removal from Water. Chem. Eng. J. 2011, 168, 493−504. (14) Chen, J. P.; Meng-Loong, C.; Zhang, B. Effects of Competitive Ions, Humic Acid, and pH on Removal of Ammonium and Phosphorous from the Synthetic Industrial Effluent by Ion Exchange Resins. Waste Manage. 2002, 22, 711−719. (15) Mangwandi, C.; Albadarin, A. B.; Glocheux, Y.; Walker, G. M. Removal of Ortho-Phosphate from Aqueous Solution by Adsorption onto Dolomite. J. Environ. Chem. Eng. 2014, 2, 1123−1130. (16) Pandi, K.; Viswanathan, N. In Situ Fabrication of Magnetic Iron Oxide over Nano-Hydroxyapatite Gelatin Eco-Polymeric Composite for Defluoridation Studies. J. Chem. Eng. Data 2016, 61, 571−578. (17) Lopez, D.; Buitrago, R.; Sepulveda-Escribano, A.; RodriguezReinoso, F.; Mondragon, F. Low-Temperature Catalytic Adsorption of NO on Activated Carbon Materials. Langmuir 2007, 23, 12131− 12137. (18) Aswin Kumar, I.; Viswanathan, N. Development of Multivalent Metal Ions Imprinted Chitosan Biocomposites for Phosphate Sorption. Int. J. Biol. Macromol. 2017, 104, 1539−1547. (19) Memon, S.; Bhatti, A. A.; Memon, N. New Calix[4]arene Appended Amberlite XAD-4 Resin with Versatile Perchlorate Removal Efficiency. J. Chem. Eng. Data 2013, 58, 2819−2827. (20) Wang, N.; Feng, J.; Chen, J.; Wang, J.; Yan, W. Adsorption Mechanism of Phosphate by Polyaniline/TiO2 Composite from Wastewater. Chem. Eng. J. 2017, 316, 33−40. (21) Xu, J.; Pu, Y.; Qi, W.-K.; Yang, X. J.; Tang, Y.; Wan, P.; Fisher, A. Chemical Removal of Nitrate from Water by Aluminum-Iron Alloys. Chemosphere 2017, 166, 197−202. (22) Pourzamani, H. R.; Mengelizadeh, N.; Jalil, M.; Moosavian, Z. Nitrate Removal from Aqueous Solutions by Magnetic Nanoparticle. J. Environ. Health. Sustain. Dev. 2017, 2, 187−195. (23) Choi, J.; Chung, J.; Lee, W.; Kim, J.-O. Phosphorous Adsorption on Synthesized Magnetite in Wastewater. J. Ind. Eng. Chem. 2016, 34, 198−203. (24) Salam, M. A.; Fageeh, O.; Al-Thabaiti, S. A.; Obaid, A. Y. Removal of Nitrate Ions from Aqueous Solution using Zero-Valent Iron Nanoparticles Supported on High Surface Area Nanographenes. J. Mol. Liq. 2015, 212, 708−715. (25) Cao, D.; Jin, X.; Gan, L.; Wang, T.; Chen, Z. Removal of Phosphate using Iron Oxide Nanoparticles Synthesized by Eucalyptus K

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

DFT Calculations of 4-N,N′-Dimethylamino Pyridine. Spectrochim. Acta, Part A 2008, 71, 898−906. (45) Elanchezhiyan, S. SD.; Sivasurian, N.; Meenakshi, S. Enhancement of Oil Recovery using Zirconium-Chitosan Hybrid Composite by Adsorptive Method. Carbohydr. Polym. 2016, 145, 103−113. (46) Cheng, F.; Chia-Hao, S.; Yu-Sheng, Y.; Chen-Sheng, Y.; ChiauYuang, T.; Chao-Liang, W.; Ming-Ting, W.; Dar-Bin, S. Characterization of Aqueous Dispersions of Fe3O4 Nanoparticles and their Biomedical Applications. Biomaterials 2005, 26, 729−738. (47) Zhang, Q.; Zhang, Z.; Teng, J.; Huang, H.; Peng, Q.; Jiao, T.; Hou, L.; Li, B. Highly Efficient Phosphate Sequestration in Aqueous Solutions using Nanomagnesium Hydroxide Modified Polystyrene Materials. Ind. Eng. Chem. Res. 2015, 54, 2940−2949. (48) Yan, L. Y.; Yang, K.; Shan, R. R.; Yan, T.; Wei, J.; Yu, S. J.; Yu, H. Q.; Du, B. Kinetic, Isotherm and Thermodynamic Investigations of Phosphate Adsorption onto Core-Shell Fe3O4@LDHs Composites with Easy Magnetic Separation Assistance. J. Colloid Interface Sci. 2015, 448, 508−516. (49) Shi, J.; Long, C.; Li, A. Selective Reduction of Nitrate into Nitrogen using Fe-Pd Bimetallic Nanoparticle Supported on Chelating Resin at Near-Neutral pH. Chem. Eng. J. 2016, 286, 408−415. (50) Bhatnagar, A.; Ji, M.; Choi, Y. H.; Jung, W.; Lee, S. H.; Kim, S. J.; Lee, G.; Suk, H.; Kim, H. S.; Min, B.; Kim, S. H.; Jeon, B. H.; Kang, J. W. Removal of Nitrate from Water by Adsorption onto Zinc Chloride Treated Activated Carbon. Sep. Sci. Technol. 2008, 43, 886− 907. (51) Bhatnagar, A.; Minocha, A. K. Biosorption Optimization of Nickel Removal from Water using Punica Granatum Peel Waste. Colloids Surf., B 2010, 76, 544−548. (52) Chauhan, K.; Kaur, J.; Singh, P.; Sharma, P.; Sharma, P.; Chauhan, G. S. An Efficient and Regenerable Quaternary Starch for Removal of Nitrate from Aqueous Solutions. Ind. Eng. Chem. Res. 2016, 55, 2507−2519. (53) Qiu, H.; Liang, C.; Yu, J.; Zhang, Q.; Song, M.; Chen, F. Preferable Phosphate Sequestration by Nano-La(III) (Hydr)Oxides Modified Wheat Straw with Excellent Properties in Regeneration. Chem. Eng. J. 2017, 315, 345−354. (54) Langmuir, I. The Constitution and Fundamental Properties of Solids and Liquids. J. Am. Chem. Soc. 1916, 38, 2221−2295. (55) Freundlich, H. M. F. Over the Adsorption in Solution. J. Phys. Chem. 1906, 57, 385−470. (56) Dubinin, M. M.; Zaverina, E. D.; Radushkevich, L. V. Sorption and Structure of Active Carbons. J. Phys. Chem. 1947, 21, 1351−1362. (57) Jiang, H.; Chen, P.; Luo, S.; Tu, X.; Cao, Q.; Shu, M. Synthesis of Novel Nanocomposite Fe3O4/ZrO2/Chitosan and its Application for Removal of Nitrate and Phosphate. Appl. Surf. Sci. 2013, 284, 942− 949. (58) Khan, A. A.; Singh, R. P. Adsorption Thermodynamics of Carbofuran on Sn(IV) Arsenosilicate in H+, Na+ and Ca2+ Forms. Colloids Surf. 1987, 24, 33−42. (59) Tofik, A. S.; Taddesse, A. M.; Tesfahun, K. T.; Girma, G. G. FeAl Binary Oxide Nanosorbent: Synthesis, Characterization and Phosphate Sorption Property. J. Environ. Chem. Eng. 2016, 4, 2458− 2468. (60) Pahlavanzadeh, H.; Katal, R.; Mohammadi, H. Synthesize of Polypyrrole Nanocomposite and its Application for Nitrate Removal from Aqueous Solution. J. Ind. Eng. Chem. 2012, 18, 948−956. (61) Banu, H. T.; Meenakshi, S. One Pot Synthesis of Chitosan Grafted Quaternized Resin for the Removal of Nitrate and Phosphate from Aqueous Solution. Int. J. Biol. Macromol. 2017, 104, 1517−1527. (62) Yao, Y.; Gao, B.; Chen, J.; Yang, L. Engineered Biochar Reclaiming Phosphate from Aqueous Solutions: Mechanisms and Potential Application as a Slow-Release Fertilizer. Environ. Sci. Technol. 2013, 47, 8700−8708. (63) Chiou, C.; Lin, Y.; Chen, H.; Chang, C.; Chang, S. Adsorption of Phosphate in Aqueous Solution by Magnetite Modified with Diethylenetriamine. J. Nanosci. Nanotechnol. 2015, 15, 2850−2857. (64) Fang, L.; Huang, L.; Holm, P. E.; Yang, X.; Hansen, H. C. B.; Wang, D. Facile Upscaled Synthesis of Layered Iron Oxide Nanosheets

and Their Application in Phosphate Removal. J. Mater. Chem. A 2015, 3, 7505−7512. (65) Saad, R.; Belkacemi, K.; Hamoudi, S. Adsorption of Phosphate and Nitrate Anions on Ammonium-Functionalized MCM-48: Effects of Experimental Conditions. J. Colloid Interface Sci. 2007, 311, 375− 381. (66) Khalil, A. M. E.; Eljamal, O.; Amen, T. W. M.; Sugihara, Y.; Matsunaga, N. Optimized Nano-Scale Zero-Valent Iron Supported on Treated Activated Carbon for Enhanced Nitrate and Phosphate Removal from Water. Chem. Eng. J. 2017, 309, 349−365. (67) Dai, J.; Yang, H.; Yan, H.; Shangguan, Y.; Zheng, Q.; Cheng, R. Phosphate Adsorption from Aqueous Solutions by Disused Adsorbents: Chitosan Hydrogel Beads After the Removal of Copper(II). Chem. Eng. J. 2011, 166, 970−977. (68) Mazarji, M.; Aminzadeh, B.; Baghdadi, M.; Bhatnagar, A. Removal of Nitrate from Aqueous Solution using Modified Granular Activated Carbon. J. Mol. Liq. 2017, 233, 139−148. (69) Jain, S.; Bansiwal, A.; Biniwale, R. B.; Milmille, S.; Das, S.; Tiwari, S.; Antony, P. S. Enhancing Adsorption of Nitrate using Metal Impregnated Alumina. J. Environ. Chem. Eng. 2015, 3, 2342−2349. (70) Hu, Q.; Chen, N.; Feng, C.; Hu, W. Nitrate Adsorption from Aqueous Solution using Granular Chitosan-Fe3+ Complex. Appl. Surf. Sci. 2015, 347, 1−9. (71) He, Y.; Lin, H.; Dong, Y.; Wang, L. Preferable Adsorption of Phosphate using Lanthanum-Incorporated Porous Zeolite: Characteristics and Mechanism. Appl. Surf. Sci. 2017, 426, 995−1004. (72) Olgun, A.; Atar, N.; Wang, S. Batch and Column Studies of Phosphate and Nitrate Adsorption on Waste Solids Containing Boron Impurity. Chem. Eng. J. 2013, 222, 108−119. (73) Xu, X.; Gao, B.; Tan, X.; Zhang, X.; Yue, Q.; Wang, Y.; Li, Q. Nitrate Adsorption by Stratified Wheat Straw Resin in Lab-Scale Columns. Chem. Eng. J. 2013, 226, 1−6. (74) Muthu, M.; Ramachandran, D.; Hasan, N.; Jeevanandam, M.; Gopal, J.; Chun, S. C. Unprecedented Nitrate Adsorption Efficiency of Carbon-Silicon Nanocomposites Prepared from Bamboo Leaves. Mater. Chem. Phys. 2017, 189, 12−21. (75) Tu, Y.-J.; You, C.-F. Phosphorus Adsorption onto Green Synthesized Nano-Bimetal Ferrites: Equilibrium, Kinetic and Thermodynamic Investigation. Chem. Eng. J. 2014, 251, 285−292.

L

DOI: 10.1021/acs.jced.7b00751 J. Chem. Eng. Data XXXX, XXX, XXX−XXX