Characterization and analysis of insolubilized Tafila humic acid samples
XRD characterization
In the X-ray diffraction analysis of INaTAHA, a distinct peak emerged at 2θ = 20.11°, which is considered an unusual characteristic associated with aliphatic or saturated ring structures, a phenomenon previously discussed by Khalili (1993) and El-Eswed (2005). Additionally, X-ray diffraction revealed another prominent peak at 2θ = 24.87°, indicating the probable presence of condensed aromatic carbons, as posited by El-Eswed. Furthermore, INaTAHA exhibited an additional peak at 2θ = 26.50°, a feature typically associated with a graphite-like structural arrangement, in line with Khalili's observations in 1993. The existence of an aromatic structure was further corroborated by the presence of peaks around 34.90°, which also suggested the potential presence of graphite-like layers, consistent with Khalili's findings.
Infrared Spectroscopy
The substantial metal ion binding capability exhibited by Humic Acids (HAs) is commonly ascribed to the abundance of oxygen-containing functional groups within HAs, notably including carboxyl (COOH), phenolic (OH), and carbonyl (C = O) groups (Piccolo and Stevenson 1982). In order to elucidate the role of these functional groups in the metal ion binding process, FT–IR spectroscopy was employed to analyze the INaTAHA, as well as their respective complexes (Supplementary material Appendix, Figure S1).
INaTAHA exhibits a notable abundance of aliphatic constituents, as indicated by the pronounced intensity of absorption bands at 2927 cm− 1 and 2837 cm− 1 in its FT-IR spectrum. The broad absorption band observed at 3435 cm− 1, attributed to the stretching vibration of -OH groups, demonstrated a shift to lower frequencies upon metal complexation. This shift implies the involvement of these OH groups in the binding of metal ions to the HA structure. Additionally, a robust band at 1718 cm− 1, corresponding to C = O stretching, was identified in the FT-IR spectrum. Notably, no significant shift in these absorption bands was detected following complexation with metal ions (Fig. 4 and Supplementary material Appendix, Table S2). Two distinctive bands associated with INaTAHA, specifically a sharp band at 1605 cm− 1 and a strong band at 1398 cm− 1, were attributed to the C = O stretching vibration of -COOH groups. These bands are indicative of carboxylate stretching and suggest the presence of sodium ions incorporated into the INaTAHA structure as a consequence of INaTAHA's treatment with 1M NaNO3 for a duration of 48 hrs (Pandey et al. 1999). Moreover, the weak band at 1221 cm− 1 reflects C-O stretching in protonated carboxylic acids within INaTAHA. This band notably diminished to very weak intensities upon complexation with metal ions. The presence of a peak at 1061 cm− 1, within the region corresponding to C-O stretching in alcohols and Si-O stretching in quartz (Supplementary material Appendix, Table S2) did not exhibit significant shifts upon complexation, suggesting that alcohol groups do not participate in metal ion binding (Alberts et al. 1998). The FTIR spectra of IHA from Ajloun and this one are in agreement (Khalili and Al-Banna 2015) and that from Azraq (Khalili et al., 2016).
Thermal properties
Under an inert nitrogen atmosphere, the thermal stability of the INaTAHA was examined using TGA and DSC instruments. Figure 5 shows a TGA curve for INaTAHA with a two-step weight loss. The mass loss was seen in the first step, which took place at temperatures below 150°C, is probably caused by dehydration processes (Chen et al. 2008; Wu et al. 2014).
This initial mass decrease is also associated with the partial decomposition of aliphatic groups, in accordance with findings by Kolokassidou et al. (2007). The weight loss that followed, which was seen between 150°C and 500°C, was caused by both the extensive breakdown of aliphatic structures and the decarboxylation of the humic material. This stage also includes the decomposition of aromatic structures and the degradation of phenolic, amino, and alcoholic groups (Kolokassidou et al. 2007; Al-Banna and Khalili 2015). At 779.4°C, the TGA curve additionally showed a residual mass percentage of 59.25%. According to Khalili et al. (2016), this residual mass could be explained by the presence of sodium in INaAZHA as well as possible contributions from metal oxides and silicate impurities.
Figure 5. TGA thermogram of INaAZHA
In Fig. 6, a prominent endothermic peak is observed at 95.0°C for INaTAHA. This peak represents the lipid fraction of humic acid melting process. In particular, the aliphatic parts of the lipid fraction have properties similar to polymethylene chains, with a semi-crystalline structure resembling polyethylene. The presence of a broad endothermic peak suggests the existence of a more diverse and heterogeneous mixture of compounds (Chilom et al. 2005). Furthermore, the appearance of peaks at 497.5°C can be attributed to the decomposition of aromatic structures within INaTAHA. This decomposition behavior aligns with findings reported in relation to IHA from the Ajloun region (Al-Banna and Khalili 2015) and that from the Azraq region (Khalili et al. 2016).
Elemental Analysis
Quantifying the elemental composition of INaTAHA was done through elemental analysis. The results detailing the percentage of each element present in INaTAHA are presented in Table 1.
Table 1
Elemental analysis of INaTAHA and another IHA from different sites.
Site
|
% C
|
% N
|
% H
|
Reference
|
INaTAHA
|
25.10
|
2.72
|
1.36
|
(This work)
|
Azraq IHA
|
48.08
|
2.02
|
3.86
|
(Khalili, et al., 2016).
|
Ajloun IHA
|
23.04
|
1.03
|
2.73
|
(Al-Banna and Khalili, 2015)
|
The percentages (%) of carbon, nitrogen, and hydrogen found in INaTAHA are notably lower compared to TAHA. This difference in composition can be attributed to the decarboxylation of humic acid and the extensive decomposition of aliphatic structures that occur when subjected to heating at 330°C for 1.5 hours (Kolokassidou et al. 2007). Additionally, the presence of sodium in INaTAHA may contribute to a reduction in the percentage of carbon (C) observed.
Scanning Electron Microscopy (SEM) Screening
Insolubilized Tafila humic acid
Based on the SEM micrograph of INaTAHA, as illustrated in Fig. 7, it is evident that INaTAHA exhibits an irregular morphology characterized by a random distribution of particles. The micrograph further reveals a partially molten and porous structure with a fluffy appearance. In contrast to solid humic acids, INaTAHA displays larger pores.
Figure 7. SEM micrograph for INaTAHA with magnification 20000x.
EDS of INaTAHA with Pb(II), Zn(II) and Cd(II)
EDS analysis was conducted on INaTAHA samples that were loaded with Pb(II), Zn(II), and Cd(II) ions (Supplementary material Appendix, Figure S2). EDS provides valuable information by identifying the types and weight percentages of each element present at specific points within the SEM micrographs. Upon conducting EDS analysis, distinctive peaks corresponding to Pb, Zn, and Cd were observed within the INaTAHA sample, along with the presence of various other elements. Supplementary material Appendix, Table S3 displays each element's percentage composition following normalization. It could be concluded from the results that the percentage of Pb(II) in the INaTAHA-Pb complex surpasses the respective percentages of Zn and Cd in their respective complexes. This observation aligns harmoniously with the adsorption outcomes, emphasizing the superior adsorption capacity of INaTAHA for Pb(II) ions compared to Zn(II) and Cd(II) ions. Additionally, it is noteworthy that the carbon content (% C) within the Pb complex is comparatively lower than that found in the Zn and Cd complexes. This difference can be attributed to the higher quantity of adsorbed Pb(II) relative to Zn(II) and Cd(II), as elucidated earlier.
Kinetics Studies
INaTAHA's metal ion adsorption kinetics were methodically examined at intervals of 15 minutes, 0.5, 1, 2, 4, 6, 8, 12, 16, 18, 24, 48, and 72 hours, with a focus on different contact times. For the three different metal ions: Pb(II), Cd(II), and Zn(II), these investigations were carried out in controlled environments with pH 6.0 and temperatures of 25.0°C, 35.0°C, and 45.0°C. Each metal ion was started at a concentration of 30 mg/L. The experimental data showed that equilibrium was reached after 24 hours of contact time for all three metal ions (Pb(II), Zn(II), and Cd(II)).
The calculated and experimental qe values were compared in order to evaluate the precision of the adsorption predictions, as shown in Table 2. According to this analysis, the pseudo-second-order reaction kinetic model and the adsorption of Pb(II), Zn(II), and Cd(II) onto INaTAHA are well-aligned. Moreover, it is significant that for all three metal ions, the correlation coefficients connected to the pseudo-second-order reaction kinetic model are greater than those connected to the pseudo-first-order reaction kinetic model.
Table 2
Calculated and experimental qe values for pseudo-first- and pseudo-second- order adsorption kinetics of Pb(II), Zn(II), and Cd(II) onto INaTAHA
|
Metal ion
|
Pb(II)
|
Zn(II)
|
Cd(II)
|
|
25.0°C
|
35.0°C
|
45.0°C
|
25.0°C
|
35.0°C
|
45.0°C
|
25.0°C
|
35.0°C
|
45.0°C
|
pseudo 1st order kinetics
|
qe calculated
|
3.19
|
5.75
|
5.92
|
3.62
|
7.48
|
6.18
|
2.02
|
1.99
|
4.06
|
R2
|
0.9479
|
0.8644
|
0.7484
|
0.9522
|
0.9432
|
0.8741
|
0.2099
|
0.1595
|
0.6941
|
pseudo 2nd order kinetics
|
k2 (g/mg.min)
|
0.088
|
0.125
|
0.221
|
0.045
|
0.111
|
0.193
|
0.087
|
0.096
|
0.239
|
qe calculated
|
17.66
|
20.92
|
21.45
|
11.27
|
17.54
|
15.15
|
2.45
|
3.57
|
6.11
|
R2
|
0.9999
|
0.9997
|
0.9998
|
0.9997
|
0.9987
|
0.9993
|
0.9935
|
0.9968
|
0.9914
|
|
qe experimental
(mg/g)
|
17.66
|
20.88
|
21.46
|
11.26
|
17.38
|
15.22
|
2.78
|
3.88
|
7.03
|
The activation energy (Ea) values provided in Table 3 reveal that the activation energy for Cd(II) is lower than that for Zn(II) and Pd(II), implying that the interaction of Cd(II) is relatively more facile when compared to Zn(II) and Pb(II). These activation parameters play a pivotal role in predicting how the adsorption of Pb(II), Zn(II), and Cd(II) may respond to changes in temperature. Specifically, Ea values recorded were 38.2 kJ/mol for Zn(II), 54.8 kJ/mol for Pb(II), and 34.8 kJ/mol for Cd(II), as detailed in Table 3. These values collectively suggest that the adsorption process is characterized by physisorption, signifying physical interactions.
Table 3
Activation parameters for the adsorption of Pb(II), Zn (II) and Cd(II) onto INaTAHA.
Metal ion
|
Ea
kJ/mol
|
∆H‡ kJ/mol
|
∆S‡
J/mol.K
|
∆G‡ kJ/mol
|
A
|
Pb(II)
|
34.77
|
132.2
|
-197.2
|
190.9
|
1.07E + 05
|
Zn(II)
|
54.80
|
204.1
|
-196.9
|
262.8
|
1.98E + 08
|
Cd(II)
|
38.16
|
139.8
|
-197.2
|
198.6
|
3.80E + 05
|
Moreover, the fact that the activation enthalpy (∆H‡) is positive indicates that the adsorption process is endothermic and that energy is used up in the process. The fact that the free energy of activation (∆G‡) is positive indicates that energy is required for the adsorption reactions to convert reactants into products. In terms of the frequency of collisions between reactant molecules, the parameter Α, also referred to as the pre-exponential factor, is temperature-dependent (Laidler and Meiser 1995). Ultimately, the fact that the activation entropy (∆S‡) is negative suggests that the adsorption produces more order by forming an activated complex. This implies that a related mechanism underlies the adsorption of Pb(II), Cd(II), and Zn(II) on INaTAHA (Al 2012).
Sorption studies of Pb(II), Zn (II) and Cd(II) by INaTAHA
The adsorption isotherms concerning INaTAHA with Pb(II), Zn(II), and Cd(II) ions exhibit characteristics consistent with the L-type isotherm, as described in the figures (not presented here). This behavior indicates that the ratio of metal ion concentrations remaining in the solution to concentrations of metal ions adsorbed onto the solid decreases as the concentration of the solute (metal ions) increases. This finding implies that, even at lower concentrations, the adsorbent exhibits a greater affinity for the adsorbate (Limousin et al. 2007). However, the adsorbate's affinity decreases as concentration increases. According to Stevenson (1982) and Giles et al. (1960), this phenomenon happens because it gets harder for solute molecules to find available vacant sites as more sites within the substrate (INaTAHA) become occupied. Interestingly, this result is consistent with earlier studies that used these metal ions and Ajloun IHA as reported by Al-Banna and Khalili in 2015.
The adsorption capacity (qm) was calculated using the Langmuir isotherm, and the Langmuir isotherms for Pb(II), Zn(II), and Cd(II) showed good linearity (not shown here). This linearity suggests that the metal ions were adsorbed at sites with uniform binding energies, forming a monolayer on the surface of the adsorbent without inter-adsorbed ion interaction, in accordance with the Langmuir model (Adebowale et al. 2005). The adsorption capacities of INaTAHA for the metal ions were found to be enhanced by higher temperatures and pH levels, as indicated by the qm values obtained from the Langmuir model. These findings support the trend in metal uptake that shows Pb(II) > Zn(II) > Cd(II). It is important to note that the qm values from the Langmuir model and the q/m values from the Dubinin-Radushkevich (D-R) model differ significantly. The two models' different definitions of qm account for this discrepancy. The qm denotes the maximum adsorption at monolayer coverage in the Langmuir model and the maximum adsorption within the sorbent's total specific micropore volume in the D-R model (Xu 2008). Specific values for β, q/m, and E for Pb(II), Zn(II), and Cd(II), respectively are providd in Supplementary material Appendix, Figure S4-S6.
The adsorption of metal ions onto INaTAHA conforms well to the Freundlich model, as evidenced by the excellent linearity observed. This indicates the presence of non-uniform and non-specific adsorption sites, suggesting a complex adsorption process involving multiple mechanisms. The data also fit both the Langmuir and Freundlich isotherm models, implying the existence of both monolayer and multilayer adsorption, as well as homogeneous and heterogeneous sites on the surface. This observation is in accordance with the notion of various possible adsorption sites on the surface of INaTAHA (Jiang et al. 2010). The values of (n) in the Freundlich model represent the favorability of adsorption. Values greater than one indicate favorable adsorption (Unuabonah et al. 2008). the n values for Pb(II) range from 2.050 to 2.142, for Zn(II) from 1.538 to 1.576, and for Cd(II) from 1.503 to 3.108, all indicating favorable adsorption. These results are consistent with previous findings by Al-Banna and Khalili in 2015, which established that the adsorption of metal ions by INaTAHA is a favorable physical process. The (KF) values, representing adsorption constants in the Freundlich model, tend to increase with rising temperature and pH for all three metal ions. These trends align with the metal uptake observations, with Pb(II) > Cd(II) > Zn(II).
The relative selectivity of heavy metals and their sorption capacity can be attributed to various metal properties, including ionic radii, atomic weight, electronegativity, hydrolysis constants, distribution coefficients, and the principles of hard and soft acid-base interactions. The highest sorption capacity for Pb(II) can be explained by the hard and soft acid-base principle, wherein soft ligands containing sulfur provide polarizable sites that attract soft metal ions. Conversely, hard ligands with less polarizable sites, such as those containing oxygen, exhibit affinities for hard metal ions like alkali and alkaline earth metal ions. The chelating group with oxygen as a donor atom is considered a hard ligand. In this context, Pb(II) is categorized as a hard metal, while Zn(II) is soft, and Cd(II) is even softer. The obtained sequences align with the first hydrolysis constants and distribution coefficients of the metal cations (Veeresh et al. 2003).
The Dubinin-Radushkevich (D-R) isotherm model, being more versatile than the Langmuir isotherm, provides insights into the overall adsorption mechanism. As seen in Supplementary material Appendix, Figure S4-S6, the values of E from the D–R isotherm fall below 8.00 kJ/mol, suggesting that physical forces significantly influence the adsorption process (Donat et al. 2005).
Thermodynamics Results
To gain a comprehensive understanding of the underlying mechanisms involved in the metal uptake process, it is imperative to estimate the thermodynamic parameters, namely ΔG° (change in Gibbs free energy), ΔH° (change in enthalpy), and ΔS° (change in entropy). These thermodynamic functions provide insights into the spontaneity, endothermic or exothermic nature, and changes in disorder associated with the adsorption process. Two commonly used equations relate these thermodynamic functions. The first equation (Eq. 1) involves the distribution coefficient (Kd), representing the ratio of metal ions adsorbed by the polymer at equilibrium to those remaining in the solution after adsorption:
The equation also takes into account the absolute temperature (T) and the gas constant (R). The second equation (Eq. 2) allows the estimation of ΔH° and ΔS° through a linear plot of ln Kd versus 1/T:
The slope of this plot provides ΔH°/R, while the intercept yields ΔS°/R. These values can then be substituted into (Eq. 2) to calculate ΔG°. The Ln Kd values were determined from the intercept of Ln (qe/Ce) versus qe plots. The calculated Kd values are summarized in Supplementary material Appendix, Table S7. The Gibbs free energy (ΔG°) is a crucial parameter that indicates the spontaneity of the adsorption process. More negative ΔG° values suggest a highly energetically favorable adsorption process (Dakiky et al. 2002). In this study, the negative ΔG° values (Table 5) confirm the feasibility and spontaneity of the adsorption process. Notably, at pH 4, ΔG° values are small and positive, signifying that adsorption is less favorable at low pH. In this case, adsorption of metal ions by INaTAHA requires a minimal amount of energy to convert reactants into products (Scheckel and Sparks 2001).
According to Table 5, the adsorption of metal ions is endothermic, which means that as temperature rises, so does the adsorption capacity. This is indicated by the positive values of enthalpy change (ΔH°). This observation is in line with the findings for the Dubinin-Radushkevich (D-R) model's metal ion uptake, Langmuir parameters (qm), Freundlich constants (KF), and (q/m) values. The fact that the metal ions are well-solvated is one reason why the heats of adsorption could be endothermic. More energy is needed for the ions to dehydrate during adsorption than for the ions to attach to the surface exothermically (Donat et al. 2005).
During the adsorption process, there is more randomness at the solid/solution interface when ΔS° values are positive. The adsorbent's affinity for the metal ions is indicated by the positive entropy change (Table 5). It is explained by the fact that the adsorbate species displaces the adsorbed water molecules, giving the water molecules more translational energy than the adsorbate ions lose. The system can now be more randomly distributed thanks to this phenomenon. Furthermore, according to Güzel et al. (2008), the dehydration of metal ions adds to the general rise in system randomness. Therefore, the positive entropy change indicates that entropy drives the adsorption process, which in turn promotes disorder in the system. These results are consistent with findings from the adsorption of Th(IV) and U(VI) ions by Ajloun IHA, as reported by Al-Banna and Khalili in 2015.
Table 5
Thermodynamic Functions for the adsorption of Pb(II), Zn(II) and Cd(II) by INaTAHA, these values calculated at 25.0°C.
Metal ion
|
pH
|
ΔG° (kJ/mol)
|
ΔH° (kJ/mol)
|
ΔS° (J/mol.K)
|
Pb(II)
|
4
|
-0.176
|
31.028
|
105.029
|
5
|
-1.031
|
17.669
|
62.571
|
6
|
-3.771
|
35.137
|
130.679
|
Zn(II)
|
4
|
2.843
|
44.246
|
139.359
|
5
|
0.698
|
19.076
|
61.376
|
6
|
-0.641
|
14.827
|
51.817
|
Cd(II)
|
4
|
4.495
|
83.461
|
265.307
|
5
|
1.132
|
53.298
|
175.959
|
6
|
-4.271
|
5.574
|
33.049
|
Column Experiments
Metal ion uptake by INaTAHA
Column experiments involving Pb(II), Zn(II), and Cd(II) were used to study metal ion uptake by the sodium form of INaTAHA. These experiments were carried out using an initial concentration of 1000 ppm, at room temperature (RT), and a flow rate of 0.50 mL/min, at the corresponding optimal pH levels for each metal ion. The results are shown in Table 6 and are expressed as the column's percentage of metal uptake. The observed uptake capacities for the metal ions follow the order: Pb(II) > Zn(II) > Cd(II). These results align with those obtained from batch experiments. However, it's noteworthy that the percentage uptake values in the column experiments are lower compared to those obtained in batch experiments conducted at pH 6 and 25°C. This disparity can be attributed to the need for an extended contact time to achieve complete saturation in the column experiments. Moreover, the absence of mechanical shaking in the column technique results in decreased metal ion uptake compared to the batch experiments.
Table 6
Metal ion uptake using column experiment at pH 6 and at RT.
Metal Ion
|
Initial concentration Ci (mg/L)
|
Final concentration
CF (mg/L)
|
Loaded Concentration
(mg/L)
|
Uptake (mg metal ion /g INaTAHA)
|
% uptake
|
Pb(II)
|
1000
|
510.2
|
489.8
|
12.2
|
48.9
|
Zn(II)
|
1000
|
568.4
|
431.6
|
10.8
|
43.2
|
Cd(II)
|
1000
|
859.6
|
140.4
|
3.51
|
14.0
|
Desorption Studies
The present investigation utilized three distinct concentrations of nitric acid (HNO3) as eluting agents, namely 1M HNO3, 0.5M HNO3, and 0.1M HNO3, with the aim of eliminating metal ions. A steady flow rate of 0.50 mL/min was maintained throughout the elution procedure. Five portions totaling 10.0 mL each were used to collect the eluate, and the results are expressed as a percentage of recovery, as detailed in Table 7. This experimental procedure was conducted to assess the regenerative capability of INaTAHA for subsequent use. Based on the values of the cumulative recovery percentage, the following trend was observed regarding the effectiveness of eluting agents for removing metal ions from INaTAHA:
1M HNO3 > 0.5M HNO3 > 0.1M HNO3.
The increased concentration of HNO3 corresponded to a higher desorption of Pb(II), Zn(II), and Cd(II) ions. This phenomenon can be attributed to the role of nitric acid as a proton-exchanging agent, which promotes physical adsorption rather than chemical adsorption (Salameh and Khalili 2010).
Table 7
Desorption of Pb(II), Zn(II), and Cd(II) ions from INaTAHA.
Eluting agent: HNO3
|
Pb(II)
|
Zn(II)
|
Cd(II)
|
1M
|
0.5M
|
0.1M
|
1M
|
0.5M
|
0.1M
|
1M
|
0.5M
|
0.1M
|
% Recovery 1st portion
|
35.2
|
26.8
|
24.4
|
38.1
|
28.5
|
26.5
|
39.9
|
30.5
|
28.6
|
% Recovery 2nd portion
|
20.6
|
17.5
|
8.3
|
20.2
|
19.9
|
8.9
|
21.6
|
21.3
|
10.2
|
% Recovery 3rd portion
|
13.0
|
9.3
|
6.6
|
13.8
|
10.5
|
7.5
|
14.7
|
12.1
|
9.2
|
% Recovery 4th portion
|
9.9
|
8.2
|
4.9
|
11.1
|
9.8
|
5.3
|
12.1
|
10.9
|
7.3
|
% Recovery 5th portion
|
5.8
|
6.5
|
3.8
|
7.4
|
6.7
|
4.5
|
8.4
|
8.2
|
5.8
|
% Cumulative Recovery
|
84.5
|
68.3
|
48
|
90.6
|
75.4
|
52.7
|
96.7
|
83
|
61.1
|