In the current research, phenol red dyes were adsorbed from the synthesized solutions using the adsorbent produced from Mespilus Germanica leaves in the solution pH of 2 to 11, stirring rate of 0-700 rpm, temperature of 25–50°C, dosage of adsorbent range of 0.25-5 g/L, phenol red initial concentration between 10 and 100 mg/L, and contact time of 5-160 minutes. The samples were tested at the end of batch adsorption tests and the concentration of the residual dye was measured using the spectrophotometer apparatus.
The adsorption percentage (R) and the adsorption capacity of the adsorbed pollutants per the dry adsorbent mass (mg/g) (qe) are calculated by equations 1 and 2, respectively:
\(R=\frac{\left(\text{C}o-{\text{C}}_{\text{e}}\right)}{{C}_{o}}\times 100\) (1) \({\text{q}}_{\text{e}}=\frac{\left({C}_{o}-{\text{C}}_{\text{e}}\right)}{\text{w}}\times \text{V}\) (2)
where, C0 (mg/ml), Ce (mg/ml), V (ml), and W (g) are phenol red dyes initial and final concentrations in solutions, solution volume, and adsorbent mass, respectively.
3.1. Characterization of the Adsorbent
As mentioned before, in the current research the structure and surface of the adsorbent and the effect of adsorption of pollutants on the adsorbent’s structure were analyzed through FTIR, XRD, SEM and EDX/Map, BET, and Raman techniques.
In Fig. 1, FTIR spectra of the produced activated carbon samples, AC/Fe2(MoO4)3 composite prior to and after the adsorption of phenol red dyes are illustrated. As it can be seen in Fig. 1a, there are three peaks in 3834 cm− 1, 3743 cm− 1, 3429 cm− 1 which are attributed to O-H free and hydrogen bonds in activated carbon structure [23–25]. Additionally, there is also an intense adsorption peak in 1428 cm− 1 which is attributed to C = C bonds of aromatic content and alkene of the activated carbon [26]. There is a shift in the adsorptive peaks of AC/Fe2(MoO4)3 composite which were appeared in the activated carbon structure (Fig. 1b); the adsorptive peaks seen in 3834 cm− 1, 3743 cm− 1, and 3429 cm− 1 were moved to 3850 cm− 1, 3742 cm− 1, and 3389 cm− 1. It should be mentioned that these peaks were associated to O-H stretching vibrations [27]. Additionally, there are other adsorptive peaks in 1618 cm− 1 and 1689 cm− 1 which are the result of δ(H2O) and C = O vibrations of oxalate groups present in this composite structure [28]. Moreover, there are some other peaks in 1743 cm− 1 and 1390 cm− 1 which are attributed to tension vibrations of C = O and C-O bonds, respectively [27, 29]. The intense adsorptive peaks seen in 800 cm− 1 and 430 cm− 1 are related to Mo(VI)-O and Fe(III)-O in AC/Fe2(MoO4) structure [27, 30, 31]. The FTIR spectra of the composite after the adsorption of phenol red dyes is also shown in Fig. 1c. As it can be observed, the intensity and location of the AC/Fe2(MoO4) composite peaks prior to adsorption changed as the result of adsorption of phenol red dyes on this composite.
The XRD patterns of the activated carbon and AC/Fe2(MoO4)3 composite in the 2ϴ range of 5–80° are presented in Fig. 2. The peaks which are appeared in the two structures with different intensities approve the crystalline and semi-crystalline phases in their structures; as a result, it can be concluded that both of materials have crystalline structures. The peaks seen in 23.12° and 43.16° in the activated carbon’s XRD pattern are attributed to graphite structure’s (002) and (001) crystalline phases [32–34]. Additionally, there are other peaks in 29.40° (220), 36.04° (311), 39.44° (222), and 64.8° (440) [32, 35]. In the XRD pattern of AC/Fe2(MoO4)3 composite, there are some peaks in 18.68°, 58.76°, 31.28°, 34.32°, 39.36°, and 45.44° which are associated to (112), (202), (222), (116), (511), and (225) crystalline phases of Fe2(MoO4)3 structure, respectively [27, 28]. There are some other peaks in 2ϴ values of 47.08°, 49.32°, 54.04°, 58.04°, and 59.84° which are attributed to (312), (012), (226), (211), and (137) crystalline phases, respectively [36].
In order to study the surface properties of the Mespilus Germanica leaves derived activated carbon and their variations after modification by Fe2(MoO4)3 and also the surface morphology of AC/Fe2(MoO4)3 composite before and after the adsorption of phenol red dyes SEM and EDX/Map analyses were used and the results are presented in Fig. 3. SEM micrographs show the produced activated carbon from Mespilus Germanica leaves has a porous structure which increases its capability in the adsorption of pollutants (Fig. 3a). Besides, the results of EDX/Map of the activated carbon showed that carbon (74.26%) and oxygen (25.74%) are its major constituents (Fig. 3b and c). Comparing the SEM image of AC/Fe2(MoO4)3 composite (Fig. 3d) with the SEM of the activated carbon (Fig. 3a), it can be observed that iron molybdate particles were properly located on the activated carbon surface and the synthesis process was fulfilled properly. Additionally, EDX/Map results approve the presence of iron and molybdenum in the synthesized adsorbent structure (Fig. 3e and f).
The comparison between the SEM micrographs of AC/Fe2(MoO4)3 composite before (Fig. 3e) and after (Fig. 3g) the adsorption of phenol red dyes shows that the accumulation of particles increased and the pores and porosity on the surface of the adsorbent decreased, which can be the result of diffusion of dye molecules into the pores and layers of the composite. In addition, the increase of carbon percentage in AC/Fe2(MoO4)3 composite after the adsorption of phenol red dyes can be seen in results of EDX/Map analysis (Fig. 3h and i). General, the results of SEM and EDX/Map techniques of AC/Fe2(MoO4)3 composite approved its proper synthesis and ability for the adsorption of phenol red dyes from aqueous solution.
The result of BET analysis of the activated carbon and AC/Fe2(MoO4)3 composite is reported in Table 1. As it can be seen, the particle size of the composite adsorbent is approximately equal to the size of the activated carbon particles. However, the surface area of the composite is 45.978 m²/g which is higher than the surface area of the activated carbon (39.706 m²/g) which leads to its higher adsorption efficiency. Moreover, the composite’s pore volume (0.0199 cm³/g) is higher than that of activated carbon (0.0172 cm³/g). These parameters affect the adsorption efficiency of the adsorbents directly and their higher values for the AC/Fe2(MoO4)3 composite approves its proper synthesis and capability in adsorption of pollutants from aqueous media.
Table 1
Parameter | | AC | AC/Fe2(MoO4)3 |
BET surface area | | 39.706 m²/g | 45.978 m²/g |
micropore volume | | 0.0172 cm³/g | 0.0199 cm³/g |
Adsorption average pore diameter | 5.454 nm | 5.529 nm |
Raman spectroscopy which is based on the intensity and measurement of surface bonds [37] is used for the characterization of activated carbon produced from Mespilus Germanica leaves and AC/Fe2(MoO4)3 composite (Fig. 4). There is a characteristic bond in 1600 cm− 1 (G bond) of the Raman spectrum of the activated carbon which is attributed to sp2 vibrations of carbon atoms present in its hexagonal structure and the sp3 which is related to irregularities in the carbon atom structure [38, 39]. The mentioned peak is also observed in the composite’s spectrum with a slight variation which approves the presence of carbon in its structure. Additionally, the lower intensity can be related to location of iron, molybdenum, and oxygen particles on the activated carbon’s surface [40, 41].
3.2. Effect of Solution pH
One of the most critical parameters of adsorption process is solution pH. This parameter plays a key role in promotion of adsorption by changing the status of the adsorbent’s active sites [42–45]. The effect of solution pH on the removal of phenol red dyes from aqueous solutions was investigated in the range of 2–11 and the results are illustrated in Fig. 5. Considering this diagram, by decrease of pH from 11 to 3 the electrostatic adsorption between the positive tales of amine group and hydroxyl and the negative ions of dyes happen as the result of protonation of amine and hydroxyl groups of the adsorbent which leads to higher adsorption efficiencies [46, 47]. By further decrease of pH, the competition between H+ ions of the solution and the protonated amine groups of the adsorbent for the electrostatic adsorption of anionic ions increased which is followed by lower adsorption efficiencies [48]. In high pH values, due to presence of OH− ions, the number of protonated amine groups on the adsorbent surface decreased; as a result, the total charge of the adsorbent surface becomes negative due to the presence of hydroxyl groups which lowers the adsorption power of the adsorbent due to repulsion between the negative charged surface and dye ions [46, 49]. Considering these data, the optimal pH value for the adsorption of phenol red dyes by AC/Fe2(MoO4)3 composite is equal to 3.
3.3 Stirring Rate Effect
Stirring rate effect (0-700 rpm) on the adsorption of phenol red dyes by AC/Fe2(MoO4)3 composite was investigated in the optimal pH = 3, 30°C, dosage of adsorbent of 1 g/L, initial concentration of 10 mg/L, and in 60 min (Fig. 6). A positive relationship between the stirring rate and adsorption efficiency was observed; because the probability of collisions between the dye molecules and the adsorbent surface increases in higher stirring rates from o to 500 rpm. However, there is not any sensible change in the adsorption efficiency between 500 to 700 rpm, while the process cost increases dramatically. A similar behavior was reported by Modirshahla et al. [50]. Considering these data, the optimal stirring rate for the adsorption of phenol red dyes by AC/Fe2(MoO4)3 composite is 500 rpm.
3.4. Effect of Temperature
In adsorption processes, the highest operating temperature is substantially 50°C in order to prevent the degradation and corruption of the adsorbent [51]. Regarding this fact, in this research the maximum temperature in adsorption of phenol red by AC/Fe2(MoO4)3 composite was considered equal to 50°C.
Temperature supplies the required energy in the form of heat and its effect on adsorption efficiency depends of endothermic/exothermic nature of the system [52]. The effect of temperature was investigated in this work in the range of 25–50°C and the results are illustrated in Fig. 7. It can be observed that the current adsorption process is temperature dependent and the adsorption efficiency dropped from 94.31–85.32% in the range of 25–50°C; as a result, the process is exothermic. Such a falling trend of adsorption percentage with temperature can be attributed to the increase of the dye molecules tendency for leaving the adsorbent surface (solid phase) and moving freely in the solution [53, 54] or degradation of the adsorbent active sites in high temperatures [55]. On the other hand, the solubility of the phenol red dyes may be increased by temperature which results in lower adsorption efficiencies [56]. General, 25°C is the optimal temperature for the current adsorption process by AC/Fe2(MoO4)3 composite.
3.5. Effect of Adsorbent Dosage
As a significant parameter of adsorption, the adsorbent dosage is used to determine the adsorbent capacity [57]. In this research, the adsorbent dosage effect was investigated between 0.25 mg/L and 5 mg/L (Fig. 8). It can be observed that by increasing the adsorbent dosage from 0.25 g/L to 1 g/L, the efficiency saw a rise from 28.76–99.23% and became approximately constant for dosages values higher than 1 g/L. Such a trend can be attributed to the higher contact area and higher number of active sites [58]. Moreover, the constant efficiency for the adsorbent dosage values higher than 1 g/L can be the result of low number of he remained dye molecules in the solution. Such a trend was observed in other adsorption processes [56, 59]. Considering the achieved results, the optimal adsorbent dosage was 1 g/L.
3.6. Effect of Dye Initial Concentration and Contact Time
Initial concentration of the pollutant acts as a major role in description of the adsorption mechanism [60]. Figure 9 shows the effect of dye initial concentration (10–100 mg/L). It can be clearly seen that the removal efficiency in 10 mg/L was higher than the other concentrations in all of the contact times. The fall of the efficiency with the increase in the phenol red dye concentration can be the result of attachment of all of the dye molecules on the adsorbent surface in low concentrations, while in higher concentrations the active sites are filled and saturated completely by dye molecules; therefore, the adsorption of the remained molecules is impossible [61].
Contact time affects the adsorption efficiency and determines the kinetic nature of the process[62]. In this research, the effect of contact time on the adsorption of phenol red dyes were investigated between 5 and 160 min and the results are illustrated in Fig. 9. There is a linear initial raising phase (fastest phase of adsorption) in which the pollutant particles diffuse into the surface of the adsorbent. The second phase is the inter-particle diffusion, and the third one is the diffusion of the particles into the small pores of the adsorbent and achievement of the equilibrium time. The rapid adsorption of pollutants in initial phase is the result of the high number of unsaturated active sites on the surface of the adsorbent. Additionally, the concentration gradient of phenol red in the solution in initial moments is high and then decreases gradually with time[63]. Consequently, the concentration gradient and the number of unsaturated active sites on the adsorbent surface decrease with time which results in a constant adsorption efficiency. Overall, 60 min is the optimal contact time for the removal of phenol red dyes from aqueous solutions by AC/Fe2(MoO4)3 composite since the adsorption efficiency is constant behind this point in all concentration values.
3.7. Regeneration Experiments
Regeneration capability of an adsorbent is significant from the economic viewpoint[64]. Reusability studies on the AC/Fe2(MoO4)3 composite adsorbent after the adsorption of phenol red dyes from the synthetic wastewater was carried out for 5 cycles in the optimal experimental conditions determined in this work and the results are depicted in Fig. 10. For this purpose, prior to washing by methanol and drying in the oven at 120°C, the recovery of the used adsorbent was done by vacuum filtration. Regarding Fig. 10, the removal efficiency of the regenerated adsorbent declined gradually in the first three adsorption/desorption cycles, after which the decline was considerable. This trend is the result of deposition of dye molecules on the surface of the adsorbent and the probable blockage of the active sites [65]. Considering these data, it is possible to use AC/Fe2(MoO4)3 adsorbent for 3 adsorption/desorption stages of phenol red dyes from the synthesized wastewater and can be considered as an appropriate one for the removal of such pollutants.
3.8. Comparison with similar works
Table 2 summarizes a comparison between the removal percentage of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent which was achieved in the optimal operating conditions of the present work and the values in the adsorption studies published earlier. Considering these data, the current adsorption efficiency is in an acceptable agreement with others.
Table 2
Comparison of the current adsorption efficiency with similar works
Adsorbent | Dye | Removal (%) | Ref. |
Kola Nut (Cola acuminata) Shell Activated Carbon | Phenol red | 89.95% | [14] |
Iron Nanoparticle | Phenol red | 94.9% | [16] |
titanium oxide (TiO2) + activated carbon (AC) | Congo red and phenol red | 100.24% | [17] |
Carbon-CuO nanocomposite | Phenol red | 99.98% | [18] |
GO-Fe3O4 hybrids | Phenol red | 68% | [66] |
Spinel Co1-xMxFe2O4 Nano Composites (M = Cd, Ag) | Phenol red | 84.2% of Cd and 75.2% of Ag | [19] |
TiN-WN composite particles | Methylene blue | 90% | [67] |
Tea waste/Fe3O4 magnetic composite (TWMC) | Crystal violet | 98.7% | [68] |
Carbon composite lignin-based | Azo dyes congo red and Eriochrome blue black R | 99% | [69] |
Modified nitrate intercalated MgAl LDH adsorbent | Methyl orange | 98.22% | [70] |
Hollow mesoporous manganese dioxide nanoadsorbents with strong negative charge and their ultra-efficient adsorption | Crystal violet and methylene blue | 95% | [71] |
A novel adsorbent (PA-β-CD) | Methylene blue, basic green 4, astrazon pink, and crystal violet | 90% | [72] |
Current adsorbent | Phenol red | 95.1% | This work |
3.9. Kinetic Study
In order to evaluate the adsorption progress with respect to time, the kinetic behavior of the process is studied which determines the diffusion of adsorbate into the adsorbent’s pores [66]. In addition, it focuses on the rate of the chemical process which is characterized by the parameters affect the rate of the reactions [67]. Kinetics of adsorption process includes the precise investigation of the experimental conditions of the chemical reactions which finally helps finding the appropriate equilibrium time.
In order to investigate the kinetic behavior of phenol red dyes adsorption using AC/Fe2(MoO4)3 composite adsorbent, pseudo first order and pseudo second order kinetic models [68] (Table 3) were applied and the results and kinetic parameters are reported in Figs. 11 and Table 4.
Table 3
Pseudo-first-order and pseudo-second-order kinetic models equations
Kinetic Model | Equation |
Pseudo-first-order linear form | \(\text{l}\text{n}\left({\text{q}}_{\text{e}}- {\text{q}}_{\text{t}}\right)= \text{l}\text{n}{\text{q}}_{\text{e}}- {\text{K}}_{1}\text{t}\) | (3) |
Pseudo-second-order linear form | \(\frac{\text{t}}{{\text{q}}_{\text{t}}}=\frac{1}{{\text{K}}_{2}{\text{q}}_{\text{e}}^{2}}+\frac{\text{t}}{{\text{q}}_{\text{e}}}\) | (4) |
Table 4
Pseudo first-order and pseudo second-order kinetic parameters for the adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent
Kinetic model | Parameters | Concentration (mg/L) |
10 | 20 | 30 | 50 | 70 | 100 |
Pseudo-first- Order | qe. Cal(mg/g) | 8.22 | 19.652 | 28.617 | 41.650 | 74.829 | 127.753 |
k1(1/min) | 0.047 | 0.050 | 0.044 | 0.034 | 0.042 | 0.0453 |
R2 | 0.972 | 0.982 | 0.992 | 0.99 | 0.966 | 0.957 |
Pseudo-second-order | qe. Cal(mg/g) | 10.661 | 20.964 | 31.348 | 52.631 | 72.993 | 99.010 |
k2(g/mg.min) | 0.087 | 0.0781 | 0.061 | 0.048 | 0.036 | 0.029 |
R2 | 0.997 | 0.997 | 0.995 | 0.992 | 0.983 | 0.986 |
| qe.exp(mg/g) | 9.781 | 19.074 | 27.852 | 45.465 | 60.228 | 78.96 |
Where qt, K1, and K2 are the adsorbed molecule mass per gram of the adsorbent at given time t (mg/g), pseudo-first-order (1.min− 1), and pseudo-second-order (g.mg− 1 g− 1) constants, respectively. It should be noted that K1 and qe values are determined from the slope and intercept of ln(qe-qt) versus t diagram and K2 can be found from the slope and intercept of t/qt versus t linear diagram.
Considering the reported data, pseudo second order model performed better in description of the kinetic behavior of the process compared to pseudo first order model in all the studied concentrations since the former had higher correlation coefficient than the latter. Additionally, the calculated adsorption capacities achieved by the pseudo-first order kinetic model was lower than the values achieved by the pseudo second order model. In addition, these calculated capacities are in accordance with the data reported in the literature [74].
3.10. Adsorption Equilibrium
Adsorption equilibrium study is carried out using the experimental data and various models. An isotherm model with its unique constants give valuable information about the surface properties and the adsorption capacity of the adsorbents [69]. These isotherms describe the interactions between the adsorbent surface and adsorbates [65]. In the current study, Langmuir [70], Freundlich [71], and Dubinin-Radushkevich [72] isotherm models were used to investigate the equilibrium behavior of the adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent in the concentration range of 10–100 mg/L and the results are presented in Fig. 12 and Table 6. The equations of these models are presented in Table 5. Langmuir isotherm is based on the hypothesis that a monolayer adsorption occurs on the active sites of a homogeneous adsorbent, while Freundlich model, as an empirical one, is assumed on the adsorption on a heterogeneous surface [73]. Additionally, Dubinin-Radushkevich model determines the physical or chemical nature of the adsorption processes [74].
Table 5
Equations of Langmuir, Freundlich, and Dubinin-Radushkevich isotherms
Isotherm Model | Definition | Equation |
Langmuir | Linear Form | \(\frac{1}{{q}_{e}}=\frac{1}{{K}_{L}{q}_{m}}\frac{1}{{C}_{e}}+\frac{1}{{q}_{m}}\) | (5) |
Adsorption Intensity | \({\text{R}}_{\text{L}}=\frac{1}{1+{{\text{K}}_{\text{L}}\text{C}}_{\text{i}}}\) | (6) |
Freundlich | Linear Form | \(\text{ln}{\text{q}}_{\text{e}}=\frac{1}{\text{n}}\text{ln}{\text{C}}_{\text{e}}+\text{ln}{\text{K}}_{\text{f}}\) | (7) |
Dubinin-Radushkevich | Linear Form | \(\text{ln}{q}_{e}=\text{ln}{q}_{m}-\beta {\epsilon }^{2}\) | (8) |
Polanyi potential | \(\epsilon =RT\text{ln}\left(1+\frac{1 }{{C}_{e}}\right)\) | (9) |
Adsorption Mean Free Energy | \(E=\frac{1}{\sqrt{2\beta }}\) | (10) |
Table 6
The parameters of Langmuir, Freundlich, and Dubinin-Radushkevich isotherms in the adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent
Models | Parameters | Value |
Langmuir | qm(mg/g) | 78.740 |
| KL (L/mg) | 0.164 |
R2 | 0.984 |
RL | 0.057–0.378 |
Freundlich | n | 2.046 |
| Kf (mg/g (L/mg)1/n) | 13.705 |
R2 | 0.988 |
Dubinin–Radushkevich (D-R) | E (kJ/moL) | 1.389 |
| qm (mg/g) | 42.679 |
βx10− 6(mol2/J2) | 0.259 |
R2 | 0.763 |
In equations 5 and 6, qm and KL are Langmuir constants and denote the maximum adsorption capacity (mg/g) and adsorption energy (L/mg), respectively. Moreover, RL is the adsorption intensity factor and for 0 < RL<1 and RL>1 the process is favorable (and reversible) and unfavorable, respectively, while it is considered favorable (and linear) and reversible for RL=1 and RL=0, respectively [68]. It should be noted that using the slope and intercept of 1/qe versus 1/Ce linear plot, the values of qm and KL, are calculated. Additionally, n and Kf are Freundlich model constants and denote the rate of adsorption and the process nonlinearity [75]. The slope and intercept of the lnqe vs. lnCe diagram is used to calculate Kf and 1/n. In Dubinin–Radushkevich isotherm model (equations (8–10)) β (mol2/kJ2) and ε are Dubinin–Radushkevich isotherm constant and Polanyi potential, respectively, while T is temperature (K), R is the universal gas constant, and E is the adsorption mean free energy (kJ/mol). While the process is considered chemical for E > 8 kJ/mol, it is physical in other E values [68]. In order to determine β and qm, lnqe versus ε2 diagram is plotted
Considering the values reported in Table 6 and Fig. 12a, Langmuir maximum adsorption capacity, qm, was 78.74 mg/g which might be considered a notable value for the maximum adsorption capacity [68, 75]. Additionally, the correlation coefficient of this isotherm model was equal to 0.984 which reflects this model’s capability in description of the equilibrium of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent. Moreover, the calculated RL values were between zero and one; as a result, the current adsorption process was favorable and reversible.
Figure 12b shows that Freundlich isotherm model parameter n is equal to 2.046 which denotes the physical nature of the current adsorption process. Moreover, the higher R2 (0.988) of this model in comparison with the correlation coefficient of Langmuir model approves the higher capability of Freundlich model in description of the equilibrium behavior of phenol red dyes and these dyes were adsorbed on a heterogeneous surface.
Dubinin–Radushkevich adsorption isotherm model was used to evaluate the adsorption energy and the type of the reaction through the calculation of Polanyi Potential (Fig. 12c). Considering the data reported in Table 6, qm, β and R2, are 42.679 mg/g, 0.259 mol2/J2, and 0.763 respectively. In addition, the adsorption energy, E, was lower than 8 kJ/mol. Therefore, the adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent is physical. General, in the current research Freundlich isotherm model performed better than Langmuir and Dubinin–Radushkevich adsorption isotherms in investigation of the equilibrium behavior of adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent.
3.10. Adsorption Thermodynamics
Thermodynamic nature of the adsorption process can be investigated through the calculation of Gibbs free energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°) is applied to investigate the spontaneous nature of the process. The required equations for the thermodynamic study of the adsorption process are presented in Table 7. As it can be seen in Eq. 13 (van’t Hoff equation), equilibrium constant is a function of temperature, standard enthalpy change, and standard entropy change [65].
Table 7
Thermodynamic parameters equations
Definition | Equation |
Equilibrium constant | \({\text{K}}_{\text{e}}=\frac{{\text{q}}_{\text{e}}}{{\text{C}}_{\text{e}}}\) (11) |
Gibbs free energy | \({\varDelta \text{G}}^{0}=-\text{R}\text{T}\text{l}\text{n}{\text{K}}_{\text{e}}\) (12) |
van’t Hoff equation | \(\text{ln}{\text{K}}_{\text{e}}=-\frac{\varDelta {\text{H}}^{0}}{\text{R}\text{T}}+\frac{{\varDelta \text{S}}^{0}}{\text{R}}\) (13) |
The values of ΔH° and ΔS° is determined from the slope and intercept of lnKe versus 1/T [75]. Figure 13 and Table 8 illustrate the thermodynamic parameters of the current adsorption process. According to the reported values, Gibbs free energy for the adsorption of phenol red dyes in the range of 25–50°C was negative; therefore, this adsorption process is spontaneous and thermodynamically desirable. It should be mentioned that Gibbs free energy increased with temperature which denotes that the spontaneity of the phenol red dyes adsorption decreased with temperature. Moreover, the negative ΔH° value of the process shows that the current adsorption process was exothermic. Finally, the entropy (ΔS°) of the process was negative; consequently, the random collisions on the adsorbent surface and in the solution decreased during the adsorption process [76].
Table 8. Thermodynamic parameters for the adsorption of phenol red dyes using AC/Fe2(MoO4)3 composite adsorbent
T(oC)
|
ΔG°(kJ/moL)
|
ΔH°(kJ/moL)
|
ΔS°(J/moL.K)
|
|
|
25
|
-6.635
|
-3.764
|
-9.963
|
|
30
|
-6.098
|
|
35
|
-5.629
|
|
40
|
-5.573
|
|
45
|
-5.038
|
|
50
|
-4.376
|
|