Adsorption Studies of Some Anionic Dyes Adsorbed by Chitosan and

21 hours ago - Department of Chemical Engineering, Jadavpur University , Kolkata 700 032 , India. J. Chem. Eng. Data , Article ASAP. DOI: 10.1021/acs...
1 downloads 0 Views 2MB Size
Article pubs.acs.org/jced

Cite This: J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Adsorption Studies of Some Anionic Dyes Adsorbed by Chitosan and New Four-Parameter Adsorption Isotherm Model R. Gopinathan,† Avijit Bhowal,‡ and Chandrasekhar Garlapati*,† †

Department of Chemical Engineering, Pondicherry Engineering College, Pondicherry 605 014, India Department of Chemical Engineering, Jadavpur University, Kolkata 700 032, India

Downloaded via QUEEN MARY UNIV OF LONDON on April 3, 2019 at 17:45:08 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



ABSTRACT: The correlation of adsorption isotherm is crucial for the successful implementation of adsorption technology. In this work, the adsorption behavior of anionic dyes, eosin Y, and indigo carmine from dilute aqueous solution on chitosan was investigated at T = 303, 313, and 323 K and P = 0.1 MPa through adsorption isotherms and thermodynamic parameters such as enthalpy change, entropy change, and Gibbs energy. The experimental data obtained in this study were used to develop a new four-parameter model for correlating liquid-phase adsorption isotherm based on the solid−liquid phase equilibrium. The new model correlates the isotherm as a function of Gibbs energy change and temperature. The overall average absolute relative deviation between the experimental and correlated results was less than 9.0%. Adsorption kinetic studies and effect of pH are reported.

1. INTRODUCTION Dyes are the major source of water pollution from industries such as paper, printing, leather, textile, dye synthesis, cosmetics, and electroplating. Dyes are highly visible even in very low concentration in water and are more difficult to degrade because of their xenobiotic properties and complex aromatic structure.1 The discharge of the toxic dye effluent in water bodies is harmful to flora and fauna and also hinders light penetration and subsequently disturbs the biological system. Practical difficulties prevail in the treatment and disposal of dyes. Physical and chemical treatments such as reverse osmosis, ultrafiltration, ion exchange, adsorption, coagulation, flocculation, and photodecomposition are commonly adopted methods. Due to the complex aromatic structure of dyes, the biological treatments are ineffective.2 The low cost of adsorbent and easy operation make adsorption attractive among all of the methods.3 Abundantly available marine-based biopolymer, chitin, may be used as a source of chitosan, which is a low-cost adsorbent. For many classes of dyes, chitosan shows good affinity. Due to its unique polycationic structure, chitosan finds extensive application in removal of anionic dyes such as acid, as well as reactive and direct dyes from aqueous solution.4 Worldwide, several million tons of seafood are handled, nearly about half of which is generated as shell wastes. Therefore, it is essential to utilize such biomaterials generated from the seafood in a sustainable way. The application of chitin/chitosan in the area of removal of dyes gained significant development.5 For effective implementation of adsorption technology, it is essential to obtain accurate equilibrium adsorption data.6−10 The objectives of this study are in 2-fold. The first objective is to determine the adsorption isotherm of eosin Y and indigo carmine from dilute aqueous solutions onto chitosan at T = 303, 313, and 323 K, P = 0.1 MPa. The second objective is to develop a new adsorption isotherm model to correlate © XXXX American Chemical Society

adsorption results based on the principle of chemical equilibrium and Gibbs energy. The new model consists of four adjustable parameters and correlates the adsorption isotherm as a function of temperature and Gibbs energy. The adjustable parameters are the dilute liquid-phase activity coefficient (γL) and solid-phase activity coefficients (D, E, and F). The significance of the proposed new model is that the dilute liquid-phase activity coefficient can be directly obtained from the experimental adsorption isotherm data.

2. EXPERIMENTAL SECTION 2.1. Materials. The dyes eosin Y (CAS No. 17372-87-1, 88%) and indigo carmine (CAS No. 860-22-0, 98%) were purchased from Nice (India) and Loba Chemie (India), respectively. The chitin was prepared from the waste of crab and shrimp shells. The waste materials, crab and shrimp shells, were collected from the local fish market of Cuddalore, Tamilnadu, which are the marine source of Bay of Bengal, India. The chitosan was obtained by deacetylation of the chitin with 50% sodium hydroxide solution. More details regarding the preparation of chitosan are well described in the literature.11 The degree of deacetylation (DD) is determined by Fourier-transform infrared (FTIR) spectroscopy based on the following equation.12 DD = 97.67 − [26.486(A1655 /A3450)]

(1)

It is based on the relationship attributed by the amide group at the absorbance value of 1655 cm−1 and the corresponding value of hydroxyl band at 3450 cm−1.10 Figure 1 shows the FTIR spectrum of the chitosan sample (Thermo Nicolet, Avatar 370). The band appearing at 3410 cm−1 is due to −NH Received: November 20, 2018 Accepted: March 26, 2019

A

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 1. FTIR spectrum of chitosan sample.

and −OH groups stretching vibration. The band at 2886 cm−1 is attributed to aliphatic C−H stretching. The band at 1650 cm−1 corresponds to the CO stretching. The peak at 1380 cm−1 corresponds to C−O stretching of primary alcoholic group. The peak at 1080 cm−1 is due to C−O−C vibration. The band appearing at 650 is due to N−H stretching vibration. The degree of deacetylation of the chitosan sample was found to be 71.6%. Figure 2 shows the FTIR spectra of chitosan

Figure 3. (a) Chitosan hysteresis isotherm (BET). (b) Pore size distribution of chitosan (BET).

total pore area vs pore diameter is shown in Figure 4a. The particle size distribution of the chitosan sample (Horiba LA90) is shown in Figure 4b. The adsorbent mean particle size is 48.1844 μm. 2.3. Thermogravimetric (TGA) Analysis. A thermogravimetric analyzer (PerkinElmer STA 6000) was employed to measure the weight loss of the chitosan in the temperature range of 40−860 °C with a heating rate of 10 °C min−1 under nitrogen stream. The weight losses of the chitosan sample at various stages are recorded and presented in Figure 5. It is readily seen from the figure that chitosan weight loss occured at 55−101 °C, due to moisture vaporization. Thermal degradation of the chitosan is observed at 250−320 °C. 2.4. Scanning Electron Microscopy (SEM) Examination. SEM (JEOL model JSM-6390LV) images of chitosan and dye-adsorbed chitosan are shown in Figure 6a−c. The SEM images clearly indicate the surface morphology and texture before and after adsorption of dyes. The presence of the dye on the adsorbent surface is observed as layer. This indicates that the dyes are adsorbed on the surface of the chitosan. 2.5. X-Ray Diffraction (XRD) Analysis. XRD (Bruker AXS D8 Advance) image of the chitosan is shown in Figure 7. XRD analysis is employed to examine the crystalline structure. The XRD pattern of the chitosan is presented in the figure.

Figure 2. FTIR spectra of chitosan before and after adsorption of dyes.

before and after adsorption of dyes. FTIR studies before and after adsorption indicate that there is a significant change in the FTIR spectrum due to adsorption of dye molecules over the adsorbent. This confirms that the adsorption of dye molecules happened on the adsorbent surface. 2.2. Brunauer−Emmett−Teller (BET) Analysis. The Brunauer−Emmett−Teller (BET) surface area of the chitosan is determined by surface area analyzer (ASAP 2020 V4.03 (V4.03 H)). The surface area of the chitosan sample is obtained by nitrogen adsorption/desorption at −195.814 °C, and the plot of adsorbed volume versus relative pressure is shown in Figure 3a. The calculated BET surface area is 0.5305 m2·g−1. The pore size distribution is shown in Figure 3b. The B

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 4. (a) Chitosan pore area vs pore diameter. (b) Particle size distribution of chitosan (Horiba LA90).

Figure 6. SEM images showing the surface morphologies at 3500× magnification. (a) Chitosan, (b) chitosan treated with eosin Y, and (c) chitosan treated with indigo carmine.

mL stoppered conical flask and mixed with a dye solution of known weight and concentration.13 The initial concentrations of eosin Y and indigo carmine bo were 1.45 × 10−4 and 2.15 × 10−4 mol·kg−1, respectively. The samples were shaken for 3 days in a temperature-controlled water bath (at 303, 313, and 323 K). The water bath was maintained at the desired temperature within ±0.1 K. The equilibrium concentration be of the dye in the solution was accurately measured by absorbance of ultraviolet light, using a spectrophotometer (Jasco UV model V-630) at 517 and 611 nm for eosin Y and indigo carmine, respectively. For the spectrophotometer calibration, dye solutions were prepared with known weights of dyes in distilled water. Figures 8 and 9 show the UV response curves and calibration graphs for eosin Y and indigo

Figure 5. Weight loss of the chitosan sample at various stages (TGA curve).

Two characteristic peaks are observed at 2θ of 10 and 20°, which is generally seen for all typical chitosan materials. 2.6. Adsorption Isotherm Measurement. Chitosan ranging from 1 × 10−4 to 6 × 10−4 kg was taken in a 250 C

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Figure 7. XRD analysis of chitosan.

Figure 9. (a) Absorbance versus wavelength for indigo carmine. (b) Absorbance versus concentration for indigo carmine.

quantity of dye adsorbed onto the chitosan, which is expressed by the following equation (qe /mol ·kg −1)(wchitosan /kg) = (wwater /kg)(bo /mol ·kg −1 − be /mol ·kg −1)

(2)

where wwater = kg of water, wchitosan = kg of chitosan; in eq 2, any consistent set of units for the quantities in this mass balance is acceptable. 2.7. Estimation of Thermodynamic Parameters. The thermodynamic parameters such as Gibbs energy (ΔG°), change in enthalpy (ΔH°), and change in entropy (ΔS°) are calculated using the following relations14−16 ΔG°/J·mol−1 = −(R /J·mol−1K−1)(T /K) ln(lm)

(3)

ij (T2/K) ·(T1/K) yz z ΔH °/J·mol−1 = −(R /J·mol−1K−1)jjj j (T /K) − (T /K) zzz 1 k 2 { ij lm2 yz lnjjj zzz jl z (4) k m1 {

Figure 8. (a) Absorbance versus wavelength for eosin Y. (b) Absorbance versus concentration for eosin Y.

carmine, respectively. The moles of solute adsorbed onto chitosan per unit mass of chitosan is denoted by qe. The quantity of dye removed from the solution must equal the D

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data ΔS°/J·mol−1K−1 =

Article

(ΔH °/J·mol−1) − (ΔG°/J·mol−1) T /K

x1γ1L = z1γ1S exp(ΔG1S − L /RT )

where is Gibbs energy change (i.e., − when the solute transferred from liquid to solid solution. The solid-phase activity coefficient can be obtained from truncated Redlich−Kister expansion.19

(5)

where lm, lm1, and lm2 are the Langmuir model constants at different temperatures T, T1, and T2, respectively. For the estimation of Langmuir constant, the Langmuir isotherm model can be used. The Langmuir isotherm model is given by the following expression17 qe =

where qm is the amount of adsorbate adsorbed at complete monolayer coverage, lm is the Langmuir adsorption isotherm constant for adsorbate−adsobent system.

ij exp((1 − z )2 (D + E z + F z 2)) yz 1 1 1 zz x1 = z1jjjj zz L j z γ1 k { ij ij ΔG S − L yzyz 1 jjexpjj zzzzz jj jj zzzz j RT {{ k k

3. NEW THEORETICAL MODEL In this model, we have assumed that solute is uniformly distributed into the adsorbent. The solid phase is treated as a solid solution.18,19 At equilibrium, the chemical potentials of sorbate in liquid and solids can be equated as

a1̂ L

(8)

μ1S = G1oS + RT ln a1̂ S

(9)

=

G1oL

+ RT ln

z1 =

a1̂ L =

f1̂

f1oL

=

fL γ1Lx1 1oL f1

S

a1̂ S

=

f1̂

f1oS

= γ1Sz1

x1 = (10)

qe /mol ·kg −1 1000/12 + (qe /mol ·kg −1)

(18)

be /mol ·kg −1 1000/18 + (be /mol ·kg −1)

(19)

Gibbs energy change required in eq 17 is obtained from eq 3.

f1S f1oS

(17)

where qe is the moles of solute adsorbed per kg of adsorbent. Similarly, x1 is the mole fraction of the solute in the liquid phase, which can be calculated from the experimental data on solute concentration (be). If be is in the units of mol·kg−1, then x1 is calculated by the following expression

where âL1 and âS1 are activity of the solute in liquid and solid phases. For liquid phase, the equation is modified by introducing the activity coefficient as L

(16)

Usually, the adsorption data available in terms of qe and be from such data z1 can be calculated from the following equation

(7)

In eq 7, subscript 1 is the adsorbate and superscripts L and S denote liquid and solid phases, respectively. Equation 7 can be rewritten in terms of activity as μ1L

GoL 1 ))

where (1 − z1) is the mole fraction of the adsorbent; D, E, and F are temperature-dependent parameters related to constants in the relation for excess Gibbs energy GE, and further solidphase activity coefficients are similar to the four-suffix Margules equation for liquid systems.21 Equations 15 and 16 are combined and rearranged to give

(6)

μ1L = μ1S

(GoS 1

ln(γ1S) = (1 − z1)2 (D + E z1 + Fz12)

qmlmCe 1 + lmCe

(15)

ΔGS−L 1

4. RESULTS AND DISCUSSION 4.1. Effect of pH on Anionic Dye Removal. Figure 10 shows the effect of pH for chitosan−eosin Y and indigo

(11)

where x1 is the mole fraction of the solute in the liquid phase, z1 is the mole fraction of the solute in solid solution (i.e., on adsorbent), γL1 is the activity coefficient of the solute in the liquid phase, γS1 is the activity coefficient of the solute in solid phase, and fL1 and foL 1 represent fugacity of pure liquid 1 at temperature T and pressures p and po, respectively. Similarly, both fS1 and foS 1 represent fugacity of pure solid 1 (i.e., solid solution) at temperature T and pressures p and po, respectively. Except in the critical region, pressure has little effect on the properties of liquids as well as on solid solutions; therefore, the 20 S oS ratios fL1 /foL Equations 10 1 and f1/f1 are often taken as unity. and 11 are reduced as a1̂ L = γ1x1

(12)

a1̂ S = γ1z1

(13)

Combing eqs 7−11 gives G1oL + RT ln x1γ1L = G1oS + RT ln z1γ1S

Figure 10. Effect of pH on the removal of eosin Y and indigo carmine dye by chitosan at a dosage of 0.5 g and volume of 5 × 10−4 m3 at T = 303 K.

(14)

Rearranging eq 14 gives E

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 1. Comparison of Degree of Deacetylation, pH, Kinetic Behavior, and Adoption Capacity of the Present Work with the Literature adsorbent−adsorbate system

T/K

chitosan hydrobeads−eosin Y

303 313 323 298 308 303 303

chitosan−indigo carmine chitosan−indigo carmine chitosan−eosin Y chitosan−indigo carmine

degree of deacetylation

kinetic behavior

maximum dye removal pH

qm/mol·kg−1

second order

4.0−8.0

116.8 109.6 105.3 0.154 0.0128 0.0297 0.0501

82.48 50.0 71.6

second order second order

carmine dye system. The percentage of dye removed varied from 97 to 92% for the pH range 4−12. The maximum uptake of dye was observed at pH 6. For indigo carmine dye, the removal varies from 97 to 72% when the solution pH varied from 4.0 to 12.0 with maximum dye removal at pH 6.0. The dye removal efficiency at pH 4.0 was slightly lower compared to pH 6.0 because of the solubility of chitosan.11 A comparison of the influence of pH on dye adsorption with chitosan and modified chitosan reported in the literature is presented in Table 1. 4.2. Kinetic Studies. Kinetic studies are important as they provide information about the mechanism of adsorption. Experiment is performed in a beaker fitted with a stirrer containing 500 mL of dye solution at a temperature of 303 K. Samples are withdrawn at various time intervals and the dye concentration is measured. The kinetic data are modeled by pseudo-first-order, second-order, and intraparticle diffusion kinetic model. Pseudo-second order fits better for the systems studied and is reported in Table 2. The relevant mathematical

indigo carmine

qe(exp) (mg g−1) k2 (g·mg−1 min−1) qe(cal) (mg g−1) R2

22.82 0.218 22.84 0.999

14.23 0.351 14.39 0.999

12 32 this study this study

be attributed to the difference in degree of deacetylation and modification of the adsorbent. In Table 5, the thermodynamic parameters are reported. The negative value of enthalpy change indicates that the adsorption of dyes onto chitosan is exothermic.4 The negative value of Gibbs energy change indicates that the adsorption process is spontaneous.23−25 The positive value for entropy change indicates increase in randomic condition at solid-solution interface with the adsorption of dyes on chitosan.26,27 The model parameters γL, D, E, and F were estimated by minimizing the absolute value of average relative deviation between experimental and predicted x1. These parameters are obtained by correlating the experimental adsorption data with the model eq 17 and are shown in Table 6 along with AARD%, which is defined as

name of the compound eosin Y

31

Figure 11. Pseudo-second-order adsorption kinetics of eosin Y and indigo carmine dye on chitosan at a dose of 2 g and solution volume of 5 × 10−4 m3 at T = 303 K. Dye concentration: 100 mg L−1.

Table 2. Second-Order Kinetic Rate Constants Obtained for Chitosan−Dye System at 303 K kinetic model

6.0 6.0

ref

equation can be seen elsewhere. 22 The parameter q e determined from the experiment and the model are comparable. Figure 11 shows kinetic plots for the eosin Y and indigo carmine dyes. A comparison of kinetic studies of dye adsorption with chitosan and modified chitosan reported in the literature is presented in Table 1. 4.3. Temperature Effect of Anionic Dyes Adsorption and Model Results. The equilibrium adsorption isotherms data of eosin Y and indigo carmine from dilute aqueous solutions on chitosan determined at T = 303, 313, and 323 K and P = 0.1 MPa are shown in Table 3. The adsorption capacity of the chitosan toward the adsorption of dyes is found to increase with increasing temperature for both eosin y and indigo carmine. The isotherms are correlated with the new model (i.e., eq 17). The new model requires Gibbs energy change; therefore, it is estimated with the help of Langmuir isotherm model, as discussed in Section 3.15 Table 4 presents the parameter values of Langmuir isotherm. The Langmuir adsorption capacities of chitosan of the present work and the literature reported values of chitosan and modified chitosan are reported in Table 1. The variation in adsorption capacity may

AARD% =

cal exp 100 N |x − x | ∑i =i 1 1i x exp1i , Ni 1i

where Ni is the number of

adsorption equilibrium data points, x1 represents mole fraction of adsorbate in liquid phase, and superscripts cal and exp denote the calculated and experimental values, respectively.28 The AARD values obtained for the new model range from 1.01 to 9.0 in Figure 12, and Figure 13 shows the experimental and new model predictions for eosin Y and indigo carmine, respectively. The model proposed in this study is employed to correlate adsorption equilibrium data of phenol-activated carbon obtained from the literature,29 and the correlation coefficients are found to be γL = 21.64, D = 4.3395, E = 62.346, and F = 1658.6 with an AARD of 3.36%. Figure 14 shows the correlation ability of the new model and indicates the applicability of the proposed model to data available in the literature. The validation of the proposed model with phenol as F

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

Table 3. Experimental Final Concentration of Solute in the Solution at Equilibrium be; Moles of Solute Adsorbed onto the Chitosan, Per Unit Mass of Chitosan at Equilibrium qe, at T = 303, 313, and 323 K and P = 0.1 MPaa T/K 303 dye compound

313

323

106 × be/mol·kg−1

qe/mol·kg−1

106 × be/mol·kg−1

qe/mol·kg−1

106 × be/mol·kg−1

qe/mol·kg−1

90.5 51.8 26.5 12.7 7.62 6.1 149.4 100.6 68.1 51.7 41.0 34.5

0.027 0.023 0.020 0.017 0.014 0.012 0.033 0.029 0.024 0.020 0.017 0.015

88.9 50.7 25.3 12.1 7.3 5.8 147.0 99.0 64.9 49.8 39.5 33.2

0.028 0.024 0.020 0.017 0.014 0.012 0.034 0.029 0.025 0.021 0.018 0.015

88.5 49.4 24.1 11.6 6.9 5.6 146.1 96.4 62.6 48.0 37.7 31.7

0.029 0.024 0.020 0.017 0.014 0.012 0.035 0.030 0.025 0.021 0.018 0.015

eosin Y

indigo carmine

a The molar densities of water at T = 303, 313, and 323 K are 55 252.33, 55 061.80, and 54 830.02 mol·m−3, respectively.20 Standard uncertainties u are as follows: u(T) = ±0.1 K, u(p) = ±0.0013 MPa, and the relative standard uncertainties ur in ur(be) = 0.03, ur(qe) = 0.03.

Table 4. Langmuir Isotherm Parameters Langmuir constants dye compound

T/K

10 × qm/mol·kg−1

102 × lm/kg·mol−1

R2

eosin Y

303 313 323 303 313 323

2.97 3.07 3.11 5.01 5.29 5.33

9.25 9.13 9.41 1.31 1.28 1.35

0.99 0.99 0.99 0.99 0.99 0.99

indigo carmine

2

Table 5. Thermodynamic Parameters for the Adsorption of Dye Compounds on Chitosan dye compound eosin Y

indigo carmine

T/K 303 313 323 303 313 323

−ΔG°/J·mol−1 ΔS°/J·mol−1 K−1 −ΔH°/J·mol−1 29 750 30 697 31 759 25 823 26 599 27 597

89.69 89.85 90.35 69.75 69.99 70.92

2560.18

Figure 12. Mole fraction of eosin Y on solid phase (z1) against mole fraction of eosin Y in liquid phase (x1). 4679.25

adsorbate is considered in this study due to experimental data. The literature reported value is 20.43.30 The predicted γL value experimental value, which indicates the proposed model.

5. CONCLUSIONS Adsorption of anionic dyes eosin Y and indigo carmine from aqueous solutions on chitosan at T = 303, 313, and 323 K and P = 0.1 MPa was determined. A new four-parameter model was proposed for correlating adsorption isotherm based on the principle of solid−liquid phase equilibrium. The new model successfully correlates the isotherm as a function of Gibbs energy and temperature. The overall average absolute relative

availability of its experimental γL is close to the success of the

Table 6. Correlation Parameters for Dye Compounds Obtained Using New Model eq 17 correlation parameters dye compound

T/K

γ

eosin Y

303 313 323 303 313 323

1.2713 1.1984 1.0501 1.3870 1.3413 1.2064

indigo carmine

L

D

E

3.6836 3.3882 3.3285 5.5657 4.9958 4.9432

7.3066 × 103 9.1874 × 103 8.3297 × 103 −4.2567 × 103 −2.8902 × 103 −3.2483 × 103 G

F 6.9949 1.8692 3.3307 1.3013 9.9278 1.0047

AARD (%) × × × × × ×

106 106 106 107 106 107

8.52 9.00 8.47 1.01 2.1 1.95

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(2) Ansari, R.; Mosayebzadeh, Z. Removal of eosin Y, an anionic dye, from aqueous solutions using conducting electroactive polymers. Iran. Polym. J. 2010, 19, 541−551. (3) Liu, T.; Li, Y.; Du, Q.; Sun, J.; Jiao, Y.; Yang, G.; Wang, Z.; Xia, Y.; Zhang, W.; Wang, K.; Zhu, H.; Wu, D. Adsorption of methylene blue from aqueous solution by graphene. Colloids Surf., B 2012, 90, 197−203. (4) Kumar, M. N. V. R. A review of chitin and chitosan applications. React. Funct. Polym. 2000, 46, 1−27. (5) Gerente, C.; Lee, V. K. C.; Cloirec, P. Le; McKay, G. Application of chitosan for the removal of metals from wastewaters by adsorption − Mechanisms and models review. Crit. Rev. Environ. Sci. Technol. 2007, 37, 41−127. (6) Foo, K. Y.; Hameed, B. H. Chem. Eng. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2− 10. (7) Fernandes, D. L. A.; Xavier, M. R. B.; Silva, A. D.; et al. Dynamic and equilibrium adsorption experiments. J. Chem. Educ. 2005, 82, 919−923. (8) Lisec, O.; Hugo, P.; Morgenstern, A. S. Frontal analysis method to determine competitive adsorption isotherms. J. Chromatogr. A 2001, 908, 19−34. (9) Gopinathan, R.; Bhowal, A.; Garlapati, C. Adsorption characteristics of activated carbon for the reclamation of colored effluents containing orange G and new solid − liquid phase equilibrium model. J. Chem. Eng. Data 2017, 62, 558−567. (10) Gopinathan, R.; Bhowal, A.; Garlapati, C. Thermodynamic study of some basic dyes adsorption from aqueous solutions on activated carbon and new correlations. J. Chem. Thermodyn. 2017, 107, 182−188. (11) Subramani, S. E.; Thinakaran, N. Isotherm, kinetic and thermodynamic studies on the adsorption behavior of textile dyes onto chitosan. Process Saf. Environ. Prot. 2017, 106, 1−10. (12) Prado, A. G. S.; Torres, J. D.; Faria, E. A.; Dias, S. C. L. Comparative adsorption studies of indigo carmine dye on chitin and chitosan. J. Colloid Interface Sci. 2004, 277, 43−47. (13) Cooney, D. O. Adsorption Design for Waste Water Treatment; Lewis Publishers: NY, 1999. (14) Mittal, A.; Malviya, A.; Kaur, D.; Mittal, J.; Kurup, L. Studies on the adsorption kinetics and isotherms for the removal and recovery of methyl orange from waste waters using waste materials. J. Hazard. Mater. 2007, 148, 229−240. (15) Watkins, R.; Weiss, D.; Dubbin, W.; Peel, K.; Coles, B.; Arnold, T. Investigations into the kinetics and thermodynamics of Sb (III) adsorption on goethite (α- FeOOH). J. Colloid Interface Sci. 2006, 303, 639−646. (16) Kim, T. K.; Son, Y. A.; Lim, Y. J. Thermodynamic analysis of 1,4 − diaminoanthraquinone adsorption on polyethylene terephthalate in alkane media. Dyes Pigm. 2007, 72, 246−250. (17) Langmuir, I. The adsorption of gases on plane surface of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361−1403. (18) Khan, A. R.; Riazi, M. R.; Al-Roomi, Y. A. A thermodynamic model for liquid adsorption isotherms. Sep. Purif. Technol. 2000, 18, 237−250. (19) Riazi, M. R.; Khan, A. R. A thermodynamic model for gas adsorption isotherms. J. Colloid Interface Sci. 1999, 210, 309−319. (20) Perry, R. H.; Green, D. W. Perry’s Chemical Engineer’s Handbook, McGraw Hill: NY, 1997. (21) Prausnitz, J. M.; Lichtenthaler, R. N.; De Azevedo, E. G. Molecular Thermodynamics of Fluid-Phase Equilibria, Prentice Hall PTR: Upper Saddle River, NJ, 1999. (22) Mall, I. D.; Srivastava, V. C.; Agarwal, N. K. Removal of orange − G and methyl violet dyes by adsorption onto bagasse fly ash − kinetic study and equilibrium isotherm analyses. Dyes Pigm. 2006, 69, 210−223. (23) Srivastava, V. C.; Mall, I. D.; Mishra, I. M. Adsorption thermodynamics and isosteric heat of adsorption of toxic metal ions onto bagasse fly ash (BFA) and rice husk ash (RHA). Chem. Eng. J. 2007, 132, 267−278.

Figure 13. Mole fraction of indigo carmine on solid phase (z1) against mole fraction of indigo carmine in liquid phase (x1).

Figure 14. Mole fraction of phenol on solid phase (z) against the mole fraction of phenol in liquid phase (x).

deviation between the experimental and correlated results was less than 9.0 %.



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. Tel.: +91 413 2655281. Fax: +91 413 2655101. ORCID

Chandrasekhar Garlapati: 0000-0002-6259-476X Notes

The authors declare no competing financial interest.

■ ■

ACKNOWLEDGMENTS The authors acknowledge Sophisticated Test and Instrumentation Center (STIC), Cochin, India, for analytical facilities. REFERENCES

(1) Parida, K. M.; Sahu, S.; Reddy, K. H.; Sahoo, P. C. A kinetic, thermodynamic, and mechanistic approach toward adsorption of methylene blue over water − washed manganese nodule leached residues. Ind. Eng. Chem. Res. 2011, 50, 843−848. H

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX

Journal of Chemical & Engineering Data

Article

(24) Freitas, A. F.; Mendes, M. F.; Coelho, G. L. V. Thermodynamic study of fatty acids adsorption on different adsorbents. J. Chem. Thermodyn. 2007, 39, 1027−1037. (25) Goscianska, J.; Marciniak, M.; Pietrzak, R. The effect of surface modification of mesoporous carbons on auramine − O dye removal from water. Adsorption 2016, 22, 531−540. (26) Ö zcan, A. S.; Ö zcan, A. Adsorption of acid dyes from aqueous solutions onto acid − activated bentonite. J. Colloid Interface Sci. 2004, 276, 39−46. (27) Lataye, D. H.; Mishra, I. M.; Mall, I. D. Adsorption of α picoline onto rice husk ash and granular activated carbon from aqueous solution: Equilibrium and thermodynamic study. Chem. Eng. J. 2009, 147, 139−149. (28) Garlapati, C.; Madras, G. Temperature independent mixing rules to correlate the solubilities of antibiotics and anti − inflammatory drugs in SCCO2. Thermochim. Acta 2009, 496, 54−58. (29) Hamdaoui, O.; Naffrechoux, E. Modeling of adsorption isotherm of phenol and chlorophenols onto granular activated carbon part I. Two parameters models and equations allowing determination of thermodynamic parameters. J. Hazard. Mater. 2007, 147, 381−394. (30) Lyman, W. J.; Reehl, W. F.; Rosenblatt, D. H. Handbook of Chemical Property Estimation Methods, McGraw-Hill: NY, 1982. (31) Chatterjee, S.; Chatterjee, S.; Chatterjee, B. P.; Das, A. R.; Guha, A. K. Adsorption of model anionic dye eosin Y from aqueous solution by chitosan hydrobeads. J. Colloid Interface Sci. 2005, 288, 30−35. (32) Dos Anjos, F. S. C.; Vieira, E. F. S.; Cestari, A. R. Interaction of indigo carmine dye with chitosan evaluated by adsorption and thermochemical data. J. Colloid Interface Sci. 2002, 253, 243−246.

I

DOI: 10.1021/acs.jced.8b01102 J. Chem. Eng. Data XXXX, XXX, XXX−XXX