Next Article in Journal
Intensification of endo-1,4-Xylanase Extraction by Coupling Microextractors and Aqueous Two-Phase System
Previous Article in Journal
Determination of the Bonding Strength of Finger Joints Using a New Test Specimen Geometry
 
 
Correction published on 15 May 2024, see Processes 2024, 12(5), 1003.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Removal of the Pigment Congo Red from Synthetic Wastewater with a Novel and Inexpensive Adsorbent Generated from Powdered Foeniculum Vulgare Seeds

Department of Chemistry, Faculty of Science, University of Tabuk, Tabuk 71474, Saudi Arabia
Submission received: 23 December 2022 / Revised: 27 January 2023 / Accepted: 30 January 2023 / Published: 2 February 2023 / Corrected: 15 May 2024

Abstract

:
In this research, powdered Foeniculum vulgare seed (FVSP) was treated separately with H2C2O4, ZnCl2, and a mixture of ZnCl2-CuS. The characteristics of the treated and untreated FVSP samples, as well as their abilities to eliminate Congo red (CR) from solutions, were investigated. The influences of the empirical circumstances on CR adsorption by the ideal adsorbent were studied. The thermodynamic, isothermal, and dynamic constants of this adsorption were also inspected. The ideal adsorbent was found to be the FVSP sample treated with a ZnCl2-CuS mixture, which eliminated 96.80% of the CR dye. The empirical outcomes proved that this adsorption was significantly affected by the empirical circumstances, and the second-order dynamic model as well as the Langmuir isotherm model fit the empirical data better than the first-order model and the Freundlich model. The values of Ea (15.3 kJ/mol) and ∆Ho (32.767 kJ/mol ≤ ∆Ho ≤ 35.495 kJ/mol) evidence that CR anions were endothermally adsorbed on Zn/Cu-FVSP via the ionic exchange mechanism. The superior Qmax values (434.78, 625.00, 833.33 mg/g), along with the cheapness and stability of the adsorbent used in this work, are evidence to confirm that this adsorbent will receive special interest in the field of contaminated water purification.

1. Introduction

The industrial products of plastic, rubber, paper, pulp, textiles, pharmaceuticals, cosmetics, and leather are usually colored by dyes [1,2]. As a result, wastewater generated from these industries and the manufacturing of dyes is polluted by various types of dyes [3]. According to a survey, there are more than 105 different dyes commercially accessible; the annual production of these dyes is higher than 700,000 tons; and around 5 to 10% of these dyes are lost in industrial discharges [4,5,6,7]. The direct disposal of the liquid waste of these industries without suitable treatment into rivers and streams causes serious problems in the ecological system due to their toxicity, ability to increase organic loading, and pollution by color, which reduces photosynthesis in aquatic life [8,9]. Additionally, dyes are carcinogenic, mutagenic, and toxic [10,11,12]. Congo red (CR), a mutagen, carcinogen, and poisonous compound, has allergy and platelet aggregation effects, lowers protein serum concentrations, induces thrombocytopenia, and is a hazardous as well as a destructive dye [13,14]. CR, similarly to any synthetic dye, has high solubility in water and high optical and biological stability due to its complicated aromatic forms [15,16]. Thus, it is essential to eliminate this dye from the industries’ effluents before their disposal into the medium of water [17]. For this reason, the fungus of white rot [18], foam of polyurethane, and foam of polypropylene [19] have been used for the biodegradation of CR dissolved in some industries’ sewage. Nanoparticles of some metal oxides, such as ZnO [20], NiO [21], and CoFeO [22], were also applied in the photocatalytic degradation of CR. ZnO and Ni-doped ZnO were also used for methylene blue degradation [23]. It was stated that the effectiveness of both the photocatalytic and biodegradability of CR is unsatisfactory due to the different and complicated aromatic forms of this dye [24]. Consequently, coagulation [25], filtering [26], precipitation [27], and some other conventional methods were also employed for the elution of CR from its solutions. The high expense and low efficiency of these conventional methods confirm that their application is infeasible.
Contrarily, it was stated that the most highly effective and easiest method to design and operate is CR adsorption by clay [28], activated carbons of periwinkle shells [29], coffee waste [30], and aloe vera leaf shells [31], and nanoparticles of NiO [32], Co3O4, CuO, and Mn2O3 [33], MgO [34], etc. As a result, in this investigation, the selected method for eliminating CR from artificial aqueous solutions was adsorption.
Despite the excellent adsorptive ability of CR on activated carbon and nanoparticles of metallic oxides, the usage of these effective adsorbents has become undesirable in recent years due to the elevated costs of their manufacturing [35]. Thus, widely available, efficient, and inexpensive bio-adsorbents, such as Teucrium polium [35], waste from tea [36], shrimp shells [37], shiitake mushrooms [38], pine bark [39], and others, were recently used for eliminating CR from its aqueous solutions. Surfactant-modified waste ash was also used to purify wastewater from nitro and chlorophenol [39]. The findings of these investigations have shown that there are significant variations among these inexpensive adsorbents towards CR adsorption. Therefore, additional studies are needed on CR adsorption by other cheap and effective biosorbents.
The plant of Foeniculum vulgare is referred to as fennel in various countries and shamr in Saudi Arabia and some other Arabian countries. The seeds of this plant are essentially used as a flavoring for tea, drinking water, and food. They are also used as antioxidants [40], antimicrobials [41], and antitumor compounds [42]. According to the findings of clinical trials and animal experiments, the use of the plant Foeniculum vulgare on a regular basis is not toxic or harmful [43].
Gold nanoparticles were extracted from the seeds of fennel and used as catalysts for the photo-degradation of methylene blue and rhodamine B dyes [44]. The stem powder of fennel was also used for the extraction of V2O5–Fe2O3 nanocomposites used for 4-nitrophenol reduction [45]. It was also reported by Bani-Atta [46] that the modified Foeniculum vulgare seed powder (FVSP), in addition to its availability in huge quantities in different countries and its cheapness, has effective adsorptive activity for the removal of potassium permanganate from aqueous solutions.
Therefore, the adsorption effectiveness of CR by the modified FVSP will be investigated for the first time in this research. Oxalic acid (H2C2O4), zinc chloride (ZnCl2), and copper sulfide (CuS) were used for the modification of FVSP. The characterization and uptake performance of the prepared adsorbents were investigated to determine the best sample for CR adsorption.
Then, for the best adsorbent prepared and evaluated in this study, the impact of the experimental circumstances, as well as the constants of thermodynamics, isotherms, and kinetics, were studied for CR adsorption.

2. Methodology

2.1. Modification of FVSP

The Foeniculum vulgare seeds (FVS) were bought from a local herb shop in Tabuk, KSA, rinsed three times with distilled water, dried for 7 h in an oven at 125 °C, and crushed into powder by a grinder. The powder that resulted was sieved to produce fine particles ranging in size from 150 to 212 µm and was identified as FVSP. In a 2000 mL round-bottom flask, 150 g of FVSP was combined with 1500 mL of 25% w/w ZnCl2 (purity 97%) and refluxed at boiling point for 3 h before being cooled to room temperature. The solid portion of the cooled blend was separated by filtration and boiled with 300 mL of 3 M HCl for an hour to eliminate the excess quantity of ZnCl2. After filtering the new blend (FNSP + HCl), the solid filtrate was repetitively rinsed with distilled water and dehydrated in a 125 °C oven. Eventually, the dried solid was ground, sieved to obtain uniform particles, and then named Zn-FVSP.
The same procedures and circumstances were used for the modification of 150 g of FVSP with 75 g of CuS (purity 99.99%) and 1500 mL of 25% w/w HO2CCO2H (purity 99.99%) to obtain the other two adsorbents, which were named Zn/CuS-FVSP and Ox-FVSP, respectively.
FVSP was modified by ZnCl2, HO2CCO2H, and CuS to improve the porosity and surface area, surface functional groups, and doping of the adsorbent surface with CuS nanoparticles, all of which have a positive impact on the adsorbent’s effectiveness.

2.2. Adsorbent Characterizations

The FVSP, Zn-FVSP, Zn/Cu-FVSP, and Ox-FVSP adsorbents were scanned by both FT-IR (Nicolet iS5 of Thermo Scientific, Waltham, MA, USA) and SEM (at 10 kV), respectively, to recognize the functional groups and morphology of the surfaces of these adsorbents. By using the BET technique at 77.350 K, the porosity and surface area of each of these adsorbents were also estimated.
Furthermore, a pH meter and 1 M, 0.1 M, and 0.01 M HCl and NaOH solutions were used to prepare five 0.04 M Na2CO3 solutions with varying initial pH values (pHi) (2–10). Then, 0.6 g of Zn/Cu-FVSP, as the best adsorbent (ideal adsorbent), was blended with 50 mL of each Na2CO3 solution in 100 mL plastic bottles, and then these bottles were sealed and agitated at 155 rpm and 27 °C for 20 h in a shaker incubator. After this, the Zn/Cu-FVSP was removed from each of the five solutions by filtration, and the pH meter was applied again to measure the pHf (final pH) of each solution. Ultimately, the pHi–pHf values were calculated and graphed as pHi to determine the pHZPC of Zn/Cu-FVSP. This method was previously applied by Bani-Atta [46] for the determination of the pHZPC of modified FVSP.

2.3. Experiments of Adsorption

2.3.1. Identifying the Ideal Adsorbent

To recognize the idealistic adsorbent amongst the four samples that were prepared in this study for CR adsorption, 30 mg of each prepared sample was added individually into four amber glass bottles (50 mL) containing 25 mL of 150 mg/L CR solution. These bottles were then closed and stirred in a shaker incubator for 25 h at a temperature of 27 °C. The mixture in each bottle was then filtered to separate the solid fraction, and a Jenway 6800 UV–Vis spectrophotometer was used to measure the amount of CR residual in the liquid portion at 500 nm. Equation (1) was used to calculate the percentage of CR that had been successfully removed from solutions (%R) by each adsorbent.
% R = ( C i C f ) C i × 100
where Ci and Cf stand for the CR solution’s starting and final concentrations, respectively.

2.3.2. Kinetic and Equilibrium Experiments

All the batch adsorption tests using Zn/Cu-FVSP as the ideal adsorbent and 15 mL of CR solutions were carried out in amber glass bottles (30 mL) and a shaker incubator (155 rpm) under the various experimental conditions described in Table 1. After the predetermined time had passed for each of these experiments, the Zn/Cu-FVSP was removed by filtration, and then the residual concentrations of CR in the liquid phases were measured as described in Section 2.3.1. Equations (1)–(3) were utilized to calculate the values of % R, Qt (mg/L), and Qe (mg/L), respectively.
Q t = ( C i C t ) V m
Q e = V m ( C i C e )
Qt and Qe: uptake amount at time (t) and equilibrium, correspondingly. V: solution volume of CR (L); m: mass of Zn/Cu-FVSP (g); Ci: CR initial concentration; Ce: CR final concentration, and Ct: CR concentration at any time t.
The impact of pH variation on this adsorption was considered only in the range of 5.5 to 11.5, as the CR solution becomes blue at pH levels below 5. The effects of the Zn/Cu-FVSP dose, adsorption period, CR solution concentration, and temperature were investigated at various ranges, such as (0.005–0.03 g), (0–7 h), (20–1000 mg/L), and (27–57 °C), correspondingly.
The practical outcomes for the adsorption of 150 mg/L of CR solution by Zn/Cu-FVSP at 27, 42, and 57 °C and different periods (0–7 h) were analyzed by the linear dynamic Equations (4)–(6) of the 1st and 2nd orders and intra-particle diffusion, respectively.
l o g ( Q e Q t ) = l o g ( Q e ) K 1 t 2.303
t Q t = 1 K 2 ( Q e ) 2 + t Q e
Q t = K d i f t   + B
t: adsorption time (min); K1 and K2: rate constant of the 1st and 2nd order, correspondingly. Kdif and B are the constants of the intra-particle model.
The practical outcomes for the adsorption of CR (20–1000 mg/L) on Zn/Cu-FVSP (0.015 g) at different temperatures (27–57 °C) were examined using the linear versions of the Langmuir (7) and Freundlich (8) models. Equation (9) was also applied to determine the values of the Langmuir separation factor.
l n ( Q e ) = l n ( K F ) + 1 n l n ( C e )
C e Q e = 1 Q m a x K L + C e Q m a x
S f = 1 1 + K L C i
Qmax: maximum uptake capability; KL and KF: Langmuir and Freundlich constants; n: uptake intensity constant; Sf: factor of separation; Ci: the highest initial concentration of CR.
Equations (10) and (11) were used to test the accuracy of the mathematical equations used to evaluate the dynamics and isotherm characteristics of this adsorption.
CFEF = n = 1 n ( ( Q e . exp + Q e . cal ) 2 Q exp )
X 2 = n = 1 n ( ( Q e . exp + Q e . cal ) 2 Q e . cal )
CFEF: fractional error function; X2: chi-square statistic; n: number of experimentations; Qe.cal and Qe.exp: calculated and experimental uptake capacity values, respectively.
By using Equations (12) and (13), it was possible to calculate the coefficients of the thermodynamical (∆Go, ∆So, ∆Ho) for CR adsorption from its solutions (300, 400, 500 mg/L) on the surface of Zn/Cu-FVSP (0.015 g).
ln ( Q e C e ) = H o R T + S o R
G o = H o T   S o
∆Go, ∆So, and ∆Ho: the free energy, entropy, and enthalpy changes, respectively. T (°C): temperature; R: the universal gas constant (0.008314 KJ/K mol).
The linear equation of Arrhenius (Equation (14)) was also applied to estimate the activation energy for the uptake of 150 mg/L of CR solution by Zn/Cu-FVSP at 27, 42, and 57 °C and different periods (0–7 h).
ln ( K ) = l n A E a R T
A: factor of Arrhenius; Ea: activation energy (KJ/ mol).

3. Results and Discussion

3.1. Characterization of the FVSP Adsorbents

Figure 1 displays the images produced by scanning FVSP, Ox-FVSP, Zn-FVSP, and Zn/Cu-FVSP under an electron microscope. These images show that the structures of the FVSP have been broken, crumpled, and melted after modification by these three chemical activation agents. Figure 1 also shows that the gaps are almost absent on the surfaces of FVSP, Ox-FVSP, and Zn-FVSP, but there are various cavities on the surfaces of Zn/Cu-FVSP. Instead of pores, these cavities can be thought of as channels on the Zn/Cu-FVSP surface, which suggests that adsorption has occurred both on the surface and inside the cavities.
The FT-IR results of these adsorbents (FVSP, Ox-FVSP, Zn-FVSP, and Zn/Cu-FVSP) are summarized in Table 2, and the spectra are displayed in Figure 2.
Five identical absorption peaks with a slight shift can be observed for each one of these four adsorbents (Figure 2) (Table 2). These peaks were associated with the functional groups O-H (stretching of hydrogen bond), stretching of C-H, C-H (alkyl), C=O (stretching of aliphatic aldehyde), and stretching of C=C. The absorption peaks of C-H scissoring and C-O-C (ether) can be observed in the cases of Ox-FVSP and Zn-FVSP only, while the peak associated with NO2 stretching appeared in the spectra of FVSP and Zn-FVSP. Moreover, the spectra of Ox-FVSP, FVSP, and Zn-FVSP, respectively, show C-O stretching, C-H out-of-plane bending, as well as in-plane bending. Thus, the effectiveness of the adsorbent towards the adsorbate is positively affected by some functional groups.
Both camphora of Cinnamomum [47] and carbon prepared from the powder of waste toner [48], which were used to adsorb some selected heavy metals and organic contaminants, included most of the functional groups found in these adsorbents.
It was found from the findings of the surface analysis of BET that the surface areas of FVSP, Ox-FVSP, Zn-FVSP, and Zn/Cu-FVSP are 11.7 m2/g, 25.0 m2/g, 68.1 m2/g, and 398.0 m2/g, correspondingly. This means that the surface area is Zn/Cu-FVSP ˃ Zn-FVSP ˃ Ox-FVSP ˃ FVSP. This demonstrates that the hydrothermal treatment of FVSP with CuS and ZnCl2 has a significant impact in increasing the FVSP’s surface area.
Figure 3 displays the correlation between the (pHi–pHf) and pHi values. This figure indicates that the value of pHi–pHf equals zero when the pH solution is 7.3. This means that the surface of the ideal adsorbent (Zn/Cu-FVSP) will be uncharged at a pH solution value of 7.3. Therefore, 7.3 is the pHZPC for this adsorbent. Bani-Atta [46] reported that the pHZPC for Foeniculum vulgare seed powder is 7.2.

3.2. Outcomes of Adsorption

3.2.1. Effectiveness of Adsorbents

The percentages of CR eliminated from solutions by the adsorbents FSVP, Zn-FSVP, Zn/Cu-FSVP, and Ox-FSVP were found to be 34.48%, 65.56%, 96.80%, and 49.13%, respectively (Figure 4). This reveals that the efficacy of these adsorbents decreases as follows:
Zn/Cu-FVSP ˃ Zn-FVSP ˃ OX-FVSP ˃ FSVP
This further demonstrates the superiority (96.80%) of Zn/Cu-FVSP, which also has the largest surface area and porosity (Section 3.1). Moreover, the absence of the functional group of NO2 from the sample modified by ZnCl2 and CuS (Zn/Cu-FVSP) may be the real reason for the superiority of this sample. As a result, Zn/Cu-FVSP was determined to be the ideal adsorbent for CR adsorption.

3.2.2. Impact of the Empirical Circumstances

To determine the ideal mass of Zn/Cu-FVSP required for this adsorption and to analyze the dosage affect, the percentages of CR dye eliminated (% R) from liquids were graphed vs. Zn/Cu-FVSP dosages (Figure 5a). This graph displays that the % R values were significantly increased when the Zn/Cu-FVSP dose rose from 0.005 g to 0.015 g and nearly stabilized when this adsorbent’s mass reached over 0.015 g. When the mass of Zn/Cu-FVSP was augmented between 0.005 and 0.015 g, the effective sites on the Zn/Cu-FVSP surface were enhanced, which increased the % R values [49]. When the Zn/Cu-FVSP mass was increased above 0.015 g, the cluster assembly of Zn/Cu-FVSP particles was responsible for the stabilization of % R [50]. The optimal dose in this study was therefore decided to be 0.015 g of Zn/Cu-FVSP.
It is crucial to examine the effect of the pH solution on this adsorption, because both the surface charge of Zn/Cu-FVSP and the ionization degree of the CR molecules are significantly controlled by changing the pH solution [51], where the adsorbent surface is positively, neutrally, or negatively charged at solution pH values less than, equal to, or greater than the adsorbent’s pHZPC [51]. Moreover, the adsorbate will be in the form of molecules and ions at a pH less than or higher than the adsorbate pKa value, correspondingly. In this work, the pHZPC of Zn/Cu-FVSP was found to be 7.3, and 4 is the pKa value of CR. Then, the Qe (mg/g) values were determined at various pH levels (5.5, 6.5, 7.5, 8.5, 9.5, 10.5, and 11.5) and graphed against the pH (Figure 5b). As displayed in Figure 5b, Qe was lightly and significantly decreased when the pH rose in the ranges of (5.5–7.3) and (7.3–11.5), in that order. In these two ranges of pH (5.5–7.3) and (7.3–11.5), CR will be in the form of ions as its pKa = 4. Thus, the slight drop in Qe can be attributed to the diminishing positive surface charges of Zn/Cu-FVSP (pHZPC = 7.3), which attract the anions of CR (pKa = 4) [51]. Meanwhile, the large increase in the negative surface charges of Zn/Cu-FVSP, which repel the anions of CR, is responsible for the abrupt drop in Qe when the pH is over 7.5 [51].
Comparable outcomes were noted for the adsorption of CR by soybean curd [52].
The effects of temperature and time on Zn/Cu-CR FVSP’s adsorption are shown in Figure 5c. This figure indicates that the CR uptake was progressively enhanced when enhancing the time of agitation in the period of 0 to 90 min, and there was no change in the uptake values after 90 min at each of these three temperatures. This is because all the effective uptake sites were originally empty before progressively accumulating CR anions during the adsorption period, which lasted from 0 to 90 min. Following this, none of the sites were available to adsorb more anions of CR. Figure 5c proves that this adsorption’s equilibrium was reached in 1.5 h, but the other tests were conducted at 9 h to ensure that all uptake constants had essentially been investigated at equilibrium.
Comparable results were stated for the adsorption of CR by Teucrium polium [35].
The plots in Figure 5c also show that this adsorption is positively impacted by temperature. This is a result of the lowering of the CR solution’s viscosity and the increasing transfer rate of CR anions in its solution as the temperature rises [53,54].
Figure 5d proves how the primary CR concentration affects the adsorption of Zn/Cu-FVSP. As displayed in this figure, Qe was considerably and slightly augmented when the CR concentration rose from 20 to 400 mg/L and from 400 to 1000 mg/L, respectively. The first increase can be attributed to the availability of many adsorptive sites, while the second increase was due to the partial saturation of these effective sites.
Additionally, the solid particles of Zn/Cu-FVSP cannot be seen in the CR solutions by the human eye after filtration in each experiment. This proves the high stability of Zn/Cu-FVSP in the solution of CR.

3.2.3. Kinetics of Adsorption

Figure 6a–c provide the plots of log (Qe-Qt) vs. t (pseudo-first-order), plots of t/Qt vs. t (pseudo-first-order), and plots of Qt vs t1/2 (intra-particle diffusion), respectively. The parameters of these three dynamic models, such as Qe, K1, K2, Kdif, and B, were calculated by utilizing the intercepts and slopes of these linear plots. To evaluate the accuracy of these two mathematical kinetic models (first- and second-order modes), the values of the statistic parameters CFEF and X2 were also computed using Equations (10) and (11). The calculated values of the kinetic and statistical parameters, along with the values of the R2 (coefficient of correlation) and the experimental values of Qe, are included in Table 3. As shown in Table 3, the R2 values in the first order (0.760 ˂ R2 ˂ 0.802) are significantly lower than those in the second order (R2 ≥ 0.998), and the experimental Qe (Qe.exp) values are close to those predicted by the dynamic model of the second order (Qe.cal) but far from those predicted by the dynamic model of the first order. Additionally, the second-order model’s statistical parameter values (CFEF and X2) are much lower than those of the first-order model. These findings strongly support the success of the second-order model in analyzing the kinetic experimental data of this adsorption, as well as the failure of the first-order model in analyzing them.
These data reveal that the rate of CR adsorption on Zn/Cu-FVSP may either be controlled by chemical interactions such as electronic sharing or by an ionic exchange process between CR and Zn/Cu-FVSP [54]. The final evidence for the type of this adsorption mechanism in terms of whether it is a chemical interaction such as electronic sharing or an ionic exchange process will be obtained from the thermodynamic data.
Comparable outcomes for CR adsorption by Teucrium polium [35], the adsorption of methylene blue by both Azadirachta indica [55] and Nitraria retusa [56], and the adsorption of methylene blue by Ocimum basilicum [57] have also been reported.
Figure 6c shows that each of these plots has two linear sections with significant R2 values but is not linear for the entire time span, with none passing from the origin point. These findings show that intra-particle diffusion and some other mechanisms control this adsorption. Additionally, the low values of the B constant (Table 3), which are related to the thickness of the adsorbent layer, demonstrate that the external diffusion of CR molecules has a minor role in controlling this adsorption rate.

3.2.4. Adsorption Isotherms

The isotherm curves resulting from the plotting of Ce/Qe vs. Ce (Langmuir equation) and ln (Qe) vs. ln (Ce) (Freundlich equation) for the adsorption of CR onto Zn/Cu-FVSP are shown in Figure 7a,b, respectively. The slopes and intercepts of the curves in Figure 7a,b, respectively, were used to calculate the constants of the Langmuir (Qmax and KL) and Freundlich (KF and n) equations. Along with the isotherm constants, the values of Sf, X2, and CFEF were also computed and are summarized in Table 4. As demonstrated by Table 4, R2 in the case of the Langmuir model has values that are slightly higher than those of the Freundlich model. Moreover, the Freundlich model’s statistical parameter (X2 and CFEF) values are significantly greater than the Langmuir model’s. These results strongly support the success of the Langmuir model in describing the isothermal properties of this adsorption. This confirms that the anions of CR were adsorbed onto the homogeneous surface of Zn/Cu-FVSP as a monolayer, with no interaction effects between them [32]. According to the values of 1/n and Sf that are less than one and higher than zero (Table 4), the empirical circumstances applied in this work were appropriate for the adsorption of CR onto Zn/Cu-FVSP [58].
Table 4 also shows that the computed Qmax values at 27, 42, and 57 °C are 434.78, 625.00, and 833.33 mg/g, in that order. These higher adsorption capacities were due to the higher surface area and the presence of effective surface functional groups on the surface of Zn/Cu-FVSP.
These high values of Qmax indicate the superior effectiveness of Zn/Cu-FVSP towards the adsorption of CR. The low cost and superior effectiveness of this adsorbent further demonstrate that Zn/Cu-FVSP will be of special interest in the field of contaminated water purification.

3.2.5. Thermodynamics of CR Adsorption

The linear plots in Figure 8 were used to calculate ∆So and ∆Ho for this uptake process, which were then used to calculate the ∆Go values using Equation (13). The values of these thermodynamical constants (∆So, ∆Go, and ∆Ho) are summarized in Table 5. At each of these three CR concentrations, the values of ∆Go are negative and increase with increasing temperature, as shown in Table 5. This is sufficient evidence that the anions of CR can feasibly and spontaneously adsorb on Zn/Cu-FVSP, and that this adsorption would reduce the free energy of Gibbs [37]. Moreover, the ∆Ho (32.767 kJ/mol ≤ ∆Ho ≤ 35.495 kJ/mol) values (Table 5) are in the 24–38 kJ/mol range, confirming that CR anions were adsorbed on Zn/Cu-FVSP via the ionic exchange mechanism [59,60]. The values of ∆So are also positive, which proves that the driving force behind this uptake is significantly more inflected by ∆So than ∆Ho [61].
Figure 9 displays the values of lnK plotted vs. 1/T. The slope of the linear curve shown in this figure (Figure 9) was used to estimate the energy of activation (Ea, KJ/mol) for this adsorption. The obtained Ea value (15.3 kJ/mol) from this work is in the 0.2–38 kJ/mol range. This is another example that confirms that this adsorption is caused by an ionic exchange mechanism [62].

3.3. The Performance of Cheap Adsorbents toward CR

Table 6 contains the Qmax values for Zn/Cu-FVSP and the other inexpensive adsorbents utilized in the prior literature to adsorb CR. This table shows that the Zn/Cu-FVSP developed and used in this study had the highest Qmax. The high stability, cheapness, and exceptional efficacy of the Zn/Cu-FVSP utilized in this study confirm the fact that this adsorbent will receive special interest in the field of contaminated water purification.

4. Conclusions

In this investigation, H2C2O4, ZnCl2, and CuS were used as chemical agents to modify powdered Foeniculum vulgare seed (FVSP). To select the best adsorbent, unmodified and modified samples were examined for their ability to remove CR from solutions. It was noted that Zn/Cu-FSVP gave the highest percent CR removal (96.80%). Along with the thermodynamics, kinetics, and isotherms, the influences of the empirical conditions on the adsorption of CR by Zn/Cu-FVSP were also examined. In this adsorption, it was revealed that the ideal Zn/Cu-FVSP dose, ideal solution pH, ideal temperature, ideal contact time, and ideal concentration of CR are 0.015 g, 7.3, 57 °C, 90 min, and 1000 mg/L, respectively. The obtained outcomes indicate that the second-order dynamic model and Langmuir isotherm model fit the empirical data better than the first-order model and Freundlich model, respectively. The thermodynamic results revealed that anions of CR spontaneously adsorb on Zn/Cu-FVSP. Furthermore, the values of Ea and ∆Ho prove that the anions of CR were endothermally adsorbed on Zn/Cu-FVSP via the ionic exchange mechanism. The superior computed values of Qmax (434.78, 625.00, and 833.33 mg/g), along with the cheapness and stability of the adsorbent used in this work, are sufficient evidence to confirm that Zn/Cu-FVSP will receive special interest in the field of contaminated water purification.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

The devices required to perform this study were provided by the University of Tabuk’s Nanotechnology Research Unit, for which the author is grateful.

Conflicts of Interest

The author has no relevant financial or non-financial interest to disclose.

References

  1. Hu, Z.; Chen, H.; Ji, F.; Yuan, S. Removal of Congo Red from aqueous solution by cattail root. J. Hazard. Mater. 2010, 173, 292–297. [Google Scholar] [CrossRef]
  2. Litefti, K.; Freire, M.S.; Stitou, M.; González-Álvarez, J. Adsorption of an anionic dye (Congo red) from aqueous solutions by pine bark. Sci. Rep. 2019, 9, 16530. [Google Scholar] [CrossRef] [PubMed]
  3. Ravi, K.; Deebika, B.; Balu, K. Decolorization of aqueous dye solutions by a novel adsorbent: Application of statistical designs and surface plots for the optimization and regression analysis. J. Hazard. Mater. 2005, B122, 75–83. [Google Scholar]
  4. McMullan, G.; Meehan, C.; Conneely, A.; Kirby, N.; Robinson, T.; Nigam, P.; Banat, I.; Marchant, R.; Smyth, W. Microbial decolorization and degradation of textiles dyes. Appl. Microbiol. Biotechnol. 2001, 56, 81–87. [Google Scholar] [CrossRef] [PubMed]
  5. Lee, J.W.; Choi, S.P.; Thiruvenkatachari, R.; Shim, W.G.; Moon, H. Evaluation of the performance of adsorption and coagulation processes for the maximum removal of reactive dyes. Dye. Pigment. 2006, 69, 196–203. [Google Scholar] [CrossRef]
  6. Yao, Z.; Wang, L.; Qi, J. Biosorption of methylene blue from aqueous solution using a bioenergy forest waste: Xanthoceras sorbifolia seed coat. Clean-Soil Air Water 2009, 37, 642–648. [Google Scholar] [CrossRef]
  7. Sen, T.K.; Afroze, S.; Ang, H.M. Equilibrium, kinetics and mechanism of removal of methylene blue from aqueous solution by adsorption onto pine cone biomass of Pinus radiata. Water Air Soil Pollut. 2011, 218, 499–515. [Google Scholar] [CrossRef]
  8. Tan, I.A.W.; Hameed, B.H.; Ahmad, A.L. Equilibrium and kinetic studies on basic dye adsorption by oil palm fiber activated carbon. Chem. Eng. J. 2007, 127, 111–119. [Google Scholar] [CrossRef]
  9. Ghaly, A.E.; Ananthashankar, R.; Alhattab, M.; Ramakrishnan, V.V. Production, characterization and treatment of textile effluents: A critical review. J. Chem. Eng. Process Technol. 2014, 5, 1000182. [Google Scholar]
  10. Chen, K.C.; Wu, J.Y.; Huang, C.C.; Liang, Y.M.; Hwang, S.C.J. Decolorization of azo dye using PVA-immobilized microorganisms. J. Biotechnol. 2003, 101, 241–252. [Google Scholar] [CrossRef]
  11. Gong, R.; Ding, Y.; Li, M.; Yang, C.; Liu, H.; Sun, Y. Utilization of powdered peanut hull as biosorbent for removal of anionic dyes from aqueous solution. Dyes Pigment. 2005, 64, 187–192. [Google Scholar] [CrossRef]
  12. Chen, H.; Zhao, J. Adsorption study for removal of Congo red anionic dye using organo-atapulgite. Adsorption 2009, 15, 381–389. [Google Scholar] [CrossRef]
  13. Cheng, Z.L.; Zhang, X.; Guo., X.; Jiang, X.; Li, T. Adsorption Behavior of Direct Red 80 and Congo Red onto Activated Carbon/Surfactant: Process Optimization, Kinetics and Equilibrium. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2015, 137, 1126–1143. [Google Scholar] [CrossRef] [PubMed]
  14. Frid, P.; Anisimov., A.V.; Popovic., N. Congo red and protein aggregation in neurodegenerative diseases. Brain Res. Rev. 2007, 53, 135–160. [Google Scholar]
  15. Mittal, A.; Mittal, J.; Malviya, A.; Gupta, V.K. Adsorptive removal of hazardous anionic dye ‘‘Congo red” from wastewater using waste materials and recovery by desorption. J. Colloid Interface Sci. 2009, 340, 16–26. [Google Scholar] [CrossRef] [PubMed]
  16. Dawood, S.; Sen, T.K. Removal of anionic dye Congo red from aqueous solution by raw pine and acid-treated pinecone powder as adsorbent: Equilibrium, thermodynamic, kinetics, mechanism and process design. Water Res. 2012, 46, 1933–1945. [Google Scholar] [CrossRef]
  17. Wasti, A.; Awan, M.A. Adsorption of Textile Dye onto Modified Immobilized Activated Alumina. J. Assoc. Arab. Univ. Basic Appl. Sci 2016, 20, 26–31. [Google Scholar] [CrossRef]
  18. Chakraborty, S.; Basak, B.; Dutta, S.; Bhunia, B.; Dey, A. Decolorization and biodegradation of congo red dye by a novel white rot fungus Alternata CMERI F6. Bioresour. Technol. 2013, 147, 662–666. [Google Scholar] [CrossRef]
  19. Sonwani, R.K.; Swain, G.; Giri, B.S.; Singh, R.S.; Rai, B.N. Biodegradation of Congo red dye in a moving bed biofilm reactor: Performance evaluation and kinetic modeling. Bioresour. Technol. 2020, 302, 122811. [Google Scholar] [CrossRef]
  20. Fowsiya, J.; Madhumitha, G.; Al-Dhabi, N.A.; Arasu, M.V. Photocatalytic degradation of Congo red using Carissa edulis extract capped zinc oxide nanoparticles. J. Photochem. Photobiol. B Biol. 2016, 162, 395–401. [Google Scholar] [CrossRef]
  21. Bhat, S.A.; Zafar, F.; Mondal, A.H.; Abdul, K.; Mirza, A.U.; Khan, S.; Mohammad, A.; Haq, Q.M.; Rizwanul, N.N. Photocatalytic degradation of carcinogenic Congo red dye in aqueous solution, antioxidant activity and bactericidal effect of NiO nanoparticles. J. Iran. Chem. Soc. 2020, 17, 215–227. [Google Scholar] [CrossRef]
  22. Ali, N.; Said, A.; Ali, F.; Raziq, F.; Ali, Z.; Bilal, M.; Reinert, L.; Begum, T.; Iqbal, H.M.N. Photocatalytic Degradation of Congo Red Dye from Aqueous Environment Using Cobalt Ferrite Nanostructures: Development, Characterization, and Photocatalytic Performance. Water Air Soil Pollut. 2020, 231, 50. [Google Scholar] [CrossRef]
  23. Ahmad, I.; Aslam, M.; Jabeen, U.; Zafar, M.N.; Malghani, M.N.K.; Alwadai, N.; Alshammari, F.H.; Almuslem, A.S.; Ullah, Z. ZnO and Ni-doped ZnO photocatalysts: Synthesis, characterization and improved visible light driven photocatalytic degradation of methylene blue. Inorg. Chim. Acta 2022, 543, 121167. [Google Scholar] [CrossRef]
  24. Velkova, Z.Y.; Kirova, G.K.; Stoytcheva, M.S.; Gochev, V.K. Biosorption of Congo red and methylene blue by pretreated waste Streptomyces fradiae biomass—Equilibrium, kinetic and thermodynamic studies. J. Serb. Chem. Soc. 2018, 83, 107–120. [Google Scholar] [CrossRef]
  25. Kristianto, H.; Tanuarto, M.Y.; Prasetyo, S.; Sugih, A.K. Magnetically assisted coagulation using iron oxide nanoparticles-Leucaena leucocephala seeds’ extract to treat synthetic Congo red wastewater. Int. J. Environ. Sci. Technol. 2020, 17, 3561–3570. [Google Scholar] [CrossRef]
  26. Hou, T.; Guo, K.; Wang, Z.; Zhang, X.-F.; Feng, Y.; He, M.; Yao, J. Glutaraldehyde and polyvinyl alcohol crosslinked cellulose membranes for efficient methyl orange and Congo red removal. Cellulose 2019, 26, 5065–5074. [Google Scholar] [CrossRef]
  27. Mao-Xu, Z.; Li, L.; Hai-Hua, W.; Zheng, W. Removal of an anionic dye by adsorption/precipitation processes using alkaline white mud. J. Hazard. Mater. 2007, 149, 735–741. [Google Scholar]
  28. Kausar, A.; Iqbal, M.; Javed, A.; Aftab, K.; Bhatti, H.N.; Nouren, S. Dyes adsorption using clay and modified clay: A review. J. Mol. Liq. 2018, 256, 395–407. [Google Scholar] [CrossRef]
  29. Ikhazuangbe, P.M.O.; Adama, K.K.; Akintoye, G.I. Adsorption of Congo Red Dye onto Activated Carbon from Periwinkle Shell. Niger. J. Eng. Sci. Res. 2020, 3, 63–75. [Google Scholar]
  30. Lafi, R.; Montasser, I.; Hafiane, A. Adsorption of Congo red dye from aqueous solutions by prepared activated carbon with oxygen-containing functional groups and its regeneration. Adsorpt. Sci. Technol. 2019, 37, 160–181. [Google Scholar] [CrossRef]
  31. Khaniabadi, Y.O.; Mohammadi, M.J.; Shegerd, M.; Sadeghi, S.; Saeedi, S.; Basiri, H. Removal of Congo red dye from aqueous solutions by a low-cost adsorbent: Activated carbon prepared from Aloe vera leaves shell. Environ. Health Eng. Manag. 2017, 4, 29–35. [Google Scholar] [CrossRef]
  32. Darwish, A.; Rashad, M.; Al-Aoh, H.A. Methyl orange adsorption comparison on nanoparticles: Isotherm, kinetics, and thermodynamic studies. Dyes Pigment. 2019, 160, 563–571. [Google Scholar] [CrossRef]
  33. Zafar, B.; Shafqat, S.S.; Zafar, M.N.; Haider, S.; Sumrra, S.H.; Zubair, M.; Alwadai, N.; Alshammari, F.H.; Almuslem, A.S.; Akhtar, M.S. NaHCO3 assisted multifunctional Co3O4, CuO and Mn2O3 nanoparticles for tartrazine removal from synthetic wastewater and biological activities. Mater. Today Commun. 2022, 33, 104946. [Google Scholar] [CrossRef]
  34. Priyadarshini, B.; Patra, T.; Sahoo, T.R. An efficient and comparative adsorption of Congo red and Trypan blue dyes on MgO nanoparticles: Kinetics, thermodynamics and isotherm studies. J. Magnes. Alloy. 2020, 16, 53. [Google Scholar] [CrossRef]
  35. Alamrani, N.A.; Al-Aoh, H.A. Elimination of Congo Red Dye from Industrial Wastewater Using Teucrium polium L. as a Low-Cost Local Adsorbent. Adsorpt. Sci. Technol. 2021, 2021, 5728696. [Google Scholar] [CrossRef]
  36. Foroughi-Dahr, M.; Abolghasemi, H.; Esmaili, M.; Shojamoradi, A.; Fatoorehchi, H. Adsorption Characteristics of Congo Red from Aqueous Solution onto Tea Waste. Chem. Eng. Commun 2014, 202, 181–193. [Google Scholar] [CrossRef]
  37. Zhou, Y.; Ge, L.; Fan, N.; Xia, M. Adsorption of Congo red from aqueous solution onto shrimp shell powder. Adsorp. Sci. Technol. 2018, 36, 1310–1330. [Google Scholar] [CrossRef]
  38. Yang, K.; Li, Y.; Zheng, H.; Luan, X.; Li, H.; Wang, Y.; Du, Q.; Sui, K.; Li, H.; Xia, Y. Adsorption of Congo red with hydrothermal treated shiitake mushroom. Mater. Res. Express 2020, 7, 015103. [Google Scholar] [CrossRef]
  39. Mushtaq, S.; Aslam, Z.; Ali, R.; Aslam, U.; Naseem, S.; Ashraf, M.; Bello, M.M. Surfactant modified waste ash for the removal of chloro and nitro group substituted benzene from wastewater. Water. Sci. Technol. 2022, 86, 1969–1980. [Google Scholar] [CrossRef]
  40. Roby, M.H.H.; Sarhan, M.A.; Selim, K.A.H.; Khalel, K.I. Antioxidant and antimicrobial activities of essential oil and extracts of fennel (Foeniculum vulgare L.) and chamomile (Matricaria chamomilla L.). Ind. Crops Prod. 2013, 44, 437–445. [Google Scholar] [CrossRef]
  41. Purkayastha, S.; Narain, R.; Dahiya, P. Evaluation of antimicrobial and phytochemical screening of Fennel, Juniper and Kalonji essential oils against multi drug resistant clinical isolates. Asian Pac. J. Trop. Biomed. 2012, 2, S1625–S1629. [Google Scholar] [CrossRef]
  42. Badgujar, S.B.; Patel, V.V.; Bandivdekar, A.H. Foeniculum vulgare Mill: A review of its botany, phytochemistry, pharmacology, contemporary application, and toxicology. Biomed. Res. Int. 2014, 2014, 842674. [Google Scholar] [CrossRef] [PubMed]
  43. Shah, A.H.; Qureshi, S.; Ageel, A.M. Toxicity studies in mice of ethanol extracts of Foeniculum vulgare fruit and Ruta chalepensis aerial parts. J. Ethnopharmacol. 1991, 34, 167–172. [Google Scholar] [CrossRef] [PubMed]
  44. Choudhary, M.K.; Kataria, J.; Sharma, S.A. Biomimetic synthesis of stable gold nanoparticles derived from aqueous extract of Foeniculum vulgare seeds and evaluation of their catalytic activity. Appl. Nanosci. 2017, 7, 439–447. [Google Scholar] [CrossRef]
  45. Yulizar, Y.; Apriandanu, D.O.B.; Al Jabbar, J.L. Facile one-pot preparation of V2O5-Fe2O3 nanocomposites using Foeniculum vulgare extracts and their catalytic property. Inorg. Chem. Commun. 2021, 123, 108320. [Google Scholar] [CrossRef]
  46. Bani-Atta, S.A. Potassium permanganate dye removal from synthetic wastewater using a novel, low-cost adsorbent, modified from the powder of Foeniculum vulgare seeds. Sci. Rep. 2022, 12, 4547. [Google Scholar] [CrossRef]
  47. Wang, C.; Wang, H. Carboxyl functionalized Cinnamomum camphora for removal of heavy metals from synthetic wastewater-contribution to sustainability in agroforestry. J. Clean. Prod. 2018, 184, 921–928. [Google Scholar] [CrossRef]
  48. Huang, R.; Yang, J.; Cao, Y.; Dionysiou, D.D.; Wang, C. Peroxymonosulfate catalytic degradation of persistent organic pollutants by engineered catalyst of self-doped iron/carbon nanocomposite derived from waste toner powder. Sep. Purif. Technol. 2022, 291, 120963. [Google Scholar] [CrossRef]
  49. Kuchekar, S.R.; Patil, M.P.; Gaikwad, V.B.; Han, S.-H. Synthesis and characterization of silver nanoparticles using Azadirachta indica (Neem) leaf extract. Int. J. Eng. Sci. Invent. 2017, 6, 47–55. [Google Scholar]
  50. Nekouei, F.; Nekouei, S.; Tyagi, I.; Gupta, V.K. Kinetic, thermodynamic and isotherm studies for acid blue 129 removal from liquids using copper oxide nanoparticle-modified activated carbon as a novel adsorbent. J. Mol. Liq. 2015, 201, 124–133. [Google Scholar] [CrossRef]
  51. Chakraborty, S.; Chowdhury, S.; Saha, P.D. Adsorption of crystal violet from aqueous solution onto NaOH-modified rice husk. Carbohyd. Polym. 2011, 86, 1533–1541. [Google Scholar] [CrossRef]
  52. Zhang, Z.; Li, Y.; Du, Q.; Li, Q. Adsorption of Congo red from aqueous solutions by porous soybean curd xerogels. Pol. J. Chem. Technol. 2018, 20, 95–102. [Google Scholar] [CrossRef]
  53. Hameed, B.H.; Ahmad, A.A. Batch adsorption of methylene blue from aqueous solution by garlic peel, an agricultural waste biomass. J. Hazard. Mater. 2009, 164, 870–875. [Google Scholar] [CrossRef] [PubMed]
  54. Wanyonyi, W.C.; Onyari, J.M.; Shiundu, P.M. Adsorption of Congo red dye from aqueous solutions using roots of Eichhornia crassipes: Kinetic and equilibrium studies. Energy Procedia 2014, 50, 862–869. [Google Scholar] [CrossRef]
  55. Mustafa, S.K.; Al-Aoh, H.A.; Bani-Atta, S.A.; Alrawashdeh, L.R.; Aljohani, M.M.; Alsharif, M.A.; Darwish, A.; Al-Shehri, H.; Ahmad, M.A.; Al-Tweher, J.N.; et al. Enhance the adsorption behavior of methylene blue from wastewater by using ZnCl2 modified neem (Azadirachta indica) leaves powder. Desalination Water Treat. 2021, 209, 367–378. [Google Scholar] [CrossRef]
  56. Al-Aoh, H.A.; Aljohani, M.M.; Darwish, A.; Ahmad, M.A.; Bani-Atta, S.A.; Alsharif, M.A.; Mahrous, Y.; Mustafa, S.K.; Al-Shehri, H.; Alrawashdeh, L.R.; et al. A potentially low-cost adsorbent for methylene blue removal from synthetic wastewater. Desalination Water Treat. 2021, 213, 431–440. [Google Scholar] [CrossRef]
  57. Bani-Atta, S.A.; Al-Aoh, H.A.; Aljohani, M.M.H.; Keshk, A.A.; Al-Shehri, H.S.; Mustafa, S.K.; Alamrani, N.A.; Darwish, A.A.A.; Sobhi, M. Methylene Blue sorption by the chemically modified Ocimum basilicum leaves powder. Desalination Water Treat. 2021, 222, 237–245. [Google Scholar] [CrossRef]
  58. Ceglowski, M.; Schroeder, G. Removal of heavy metal ions with the use of chelating polymers obtained by grafting pyridine-pyrazole ligands onto polymethylhydrosiloxane. Chem. Eng. J. 2015, 259, 885–893. [Google Scholar] [CrossRef]
  59. Yaneva, Z.L.; Georgieva, N.V. Insights into Congo rerd adsorption on agro-industrial materials-Spectral, equilibrium, kinetic, thermodynamic, dynamic and desorption studies. Int. Rev. Chem. Eng. 2012, 4, 127–146. [Google Scholar]
  60. Cardoso, N.F.; Lima, E.C.; Royer, B.; Bach, G.L.; Dotto, M.V.; Pinto, L.A.A.; Calvete, T. Comparison of Spirulina platensis microalgae and commercial activated carbon as adsorbents for the removal of reactive Red 120 dye from aqueous effluents. J. Hazard. Mater. 2012, 241, 146–153. [Google Scholar] [CrossRef]
  61. Hou, H.; Zhou, R.; Wu, P.; Wu, L. Removal of Congo red dye from aqueous solution with hydroxyapatite/chitosan composite. Chem. Eng. J. 2012, 211, 336–342. [Google Scholar] [CrossRef]
  62. Rahchamani, J.; Mousavi, H.Z.; Behzad, M. Adsorption of methyl violet from aqueous solution by polyacrylamide as an adsorbent: Isotherm and kinetic studies. Desalination 2011, 267, 256–260. [Google Scholar] [CrossRef]
  63. Salahuddin, N.; Abdelwahab, M.A.; Akelah, A.; Elnagar, M. Adsorption of Congo red and crystal violet dyes onto cellulose extracted from Egyptian water hyacinth. Nat. Hazards 2021, 105, 1375–1394. [Google Scholar] [CrossRef]
  64. Kaur, S.; Rani, S.; Mahajan, R.K. Adsorption kinetics for the removal of hazardous dye Congo red by biowastes materials as adsorbents. J. Chem. 2013, 628582, 12. [Google Scholar] [CrossRef]
Figure 1. SEM micrographs of the adsorbent used in the CR adsorption [46].
Figure 1. SEM micrographs of the adsorbent used in the CR adsorption [46].
Processes 11 00446 g001
Figure 2. FT-IR spectra of the adsorbent used in the Congo red adsorption.
Figure 2. FT-IR spectra of the adsorbent used in the Congo red adsorption.
Processes 11 00446 g002
Figure 3. Zn/Cu-FVSP pHZPC.
Figure 3. Zn/Cu-FVSP pHZPC.
Processes 11 00446 g003
Figure 4. The percent of Congo red removal by the four adsorbents used in this work.
Figure 4. The percent of Congo red removal by the four adsorbents used in this work.
Processes 11 00446 g004
Figure 5. (a) Impact of the adsorbent dose; (b) impact of pH; (c) impact of agitation time; (d) impact of Congo red concentration.
Figure 5. (a) Impact of the adsorbent dose; (b) impact of pH; (c) impact of agitation time; (d) impact of Congo red concentration.
Processes 11 00446 g005
Figure 6. Kinetic plots of the first order (a), second order (b), and intra-particle diffusion (c) for CR adsorption by Zn/Cu-CR FVSP.
Figure 6. Kinetic plots of the first order (a), second order (b), and intra-particle diffusion (c) for CR adsorption by Zn/Cu-CR FVSP.
Processes 11 00446 g006
Figure 7. Isotherm plots of Langmuir (a) and Freundlich (b) for adsorption of CR by Zn/Cu-FVSP.
Figure 7. Isotherm plots of Langmuir (a) and Freundlich (b) for adsorption of CR by Zn/Cu-FVSP.
Processes 11 00446 g007
Figure 8. Thermodynamic plots for adsorption of CR by Zn/Cu−FVSP.
Figure 8. Thermodynamic plots for adsorption of CR by Zn/Cu−FVSP.
Processes 11 00446 g008
Figure 9. The relationship between lnK and 1/T.
Figure 9. The relationship between lnK and 1/T.
Processes 11 00446 g009
Table 1. Empirical circumstances of adsorption.
Table 1. Empirical circumstances of adsorption.
Type of Empirical Condition Concentration of CR (mg/L)Contact Time (hour)Adsorbent Dose (g)pHTemperature (°C)
Adsorbent dosage60250.005–0.03727
Adsorption time and temp1500–70.015727–57
Solution pH300250.0155.5–11.527
CR conc20–100090.015727–57
Table 2. FT-IR results.
Table 2. FT-IR results.
FT-IR PeakFVSPOx-FVSPZn-FVSPZn/Cu-FVSP
Wavenumber (cm−1)AssignmentWavenumber (cm−1)AssignmentWavenumber (cm−1)AssignmentWavenumber (cm−1)Assignment
13302.49Stretching of O-H, hydrogen bond3363.67Stretching of O-H, hydrogen bond3365.41Stretching of O-H, hydrogen bond3364.91Stretching of O-H, hydrogen bond
22922.38Stretching of C-H (alkyl)2922.41Stretching of C-H (alkyl)2922.09Stretching of C-H (alkyl)2921.60Stretching of C-H (alkyl)
32853.30Stretching of C-H2853.46Stretching of C-H2853.24Stretching of C-H2852.82Stretching of C-H
41742.92Stretching of C=O (aliphatic aldehydes)1743.17Stretching of C=O (aliphatic aldehydes1743.11Stretching of C=O (aliphatic aldehydes1738.88Stretching of C=O (aliphatic aldehydes
51603.69C=C stretching1620.36N-H (1°-amide) II band1629.16N-H (1°-amide) II band1625.67N-H (1°-amide) II band
6---------------1457.37C-H scissoring1457.52C-H scissoring----------------
71375.62NO2 stretching1317.52Stretching of C-O1370.97NO2 stretching----------------
81029.09Bending of C-H in-plane1064.92C-O-C (ether)1064.94C-O-C (ether)------------------
9----------------779.46Bending of C-H out-of-plane--------------------------------
Table 3. Dynamic constant for CR adsorption onto Zn/Cu-FVSP.
Table 3. Dynamic constant for CR adsorption onto Zn/Cu-FVSP.
Temperature (°C)
274257
Qe.exp (mg/g)77.616101.6075124.926
Dynamic model of 1st order
Qe.cal (mg/g)40.76635.21335.221
K1 (1/h)0.01730.01380.0136
R20.7610.7840.801
CFEF17.49543.38564.414
X233.310125.190228.473
Dynamic model of 2nd order
Qe.cal (mg/g)82.645105.263128.205
K2 (g/mg. h)0.000570.000790.00099
R20.9980.9980.999
Rate0.0470.0830.127
CFEF0.3260.1320.086
X20.3060.1270.084
Dynamic model of intra-particle diffusion (first region)
Kdif7.3956.5087.402
B0.34730.21648.423
R20.9910.9720.976
Dynamic model of intra-particle diffusion (second region)
Kdif0.1750.1860.212
B74.22797.894120.680
R20.8770.9760.970
Table 4. Isotherm constants for CR adsorption onto Zn/Cu-FVSP.
Table 4. Isotherm constants for CR adsorption onto Zn/Cu-FVSP.
Temperature (°C)
274257
Qe.exp (mg/g)313.92443.70609.57
Langmuir
Qmax (mg/g)434.78625.00833.33
Qe (mg/g)318.567427.732595.062
KL (L/mg)0.0039960.0038980.006397
Sf0.20020.20420.1352
R20.9980.9940.999
CFEF0.0690.5750.345
X20.0680.5960.354
Freundlich
Qe (mg/g)655.05661.46831.86
KF (mg/g)(L/mg)1/n4.6024.0158.681
n1.3171.2381.308
1/n0.75920.80750.7646
R20.9740.9840.980
CFEF370.714106.87781.058
X2177.65571.69159.398
Table 5. Thermodynamic constants for CR adsorption onto Zn/Cu-FVSP.
Table 5. Thermodynamic constants for CR adsorption onto Zn/Cu-FVSP.
CR Concentration (mg/L)∆Ho
(kJ/mol)
∆So
(kJ/mol)
∆Go (KJ/mol)R2
27 °C42 °C57 °C
30032.7670.1098−0.184−1.831−3.4790.942
40034.9030.1153−0.323−1.406−3.1350.936
50035.4950.1161−0.673−1.068−2.8090.957
Table 6. Qmax values of CR adsorption onto Zn/Cu-FVSP and other low-cost adsorbents used in the prior literature.
Table 6. Qmax values of CR adsorption onto Zn/Cu-FVSP and other low-cost adsorbents used in the prior literature.
AdsorbentsQmax (mg/g)References
Zn/Cu-FVSP434.78
625.00
833.33
27 °C
42 °C
57 °C
This study
Cattail rot38.79 [1]
Pine bark3.9050 °C[2]
Shrimp shell powder232.00
264.60
288.20
30 °C
40 °C
50 °C
[37]
Modified shiitake mushroom 217.90
211.90
209.40
20 °C
35 °C
50 °C
[38]
Soybean curd69.00
69.20
69.90
25 °C
40 °C
55 °C
[52]
Roots of Eichhornia crassipes1.60 [54]
Egyptian hyacinth cellulose 230.00 [63]
Tea waste32.30 [36]
Charcoal of Eichhornia 56.8 [64]
Charcoal of groundnut 117.6 [64]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

AL-Aoh, H.A. Removal of the Pigment Congo Red from Synthetic Wastewater with a Novel and Inexpensive Adsorbent Generated from Powdered Foeniculum Vulgare Seeds. Processes 2023, 11, 446. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11020446

AMA Style

AL-Aoh HA. Removal of the Pigment Congo Red from Synthetic Wastewater with a Novel and Inexpensive Adsorbent Generated from Powdered Foeniculum Vulgare Seeds. Processes. 2023; 11(2):446. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11020446

Chicago/Turabian Style

AL-Aoh, Hatem A. 2023. "Removal of the Pigment Congo Red from Synthetic Wastewater with a Novel and Inexpensive Adsorbent Generated from Powdered Foeniculum Vulgare Seeds" Processes 11, no. 2: 446. https://0-doi-org.brum.beds.ac.uk/10.3390/pr11020446

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop