Introduction

Air and water pollution is a crucial problem in the modern world. Industrial chemicals, organic compounds, and fossil fuel use have attracted attention [1, 2]. The presence of metal ions (Ba2+, Pb2+, Cd2+, Cr3+, Cr6+, Co2+, Cu2+, Ni2+, Zn2+, and Hg2+) within the water is a problem that has caused environmental and health challenges in recent years [3].

Heavy metals exceed the density of 5 g/cm3. Heavy metal reactivity and toxicity depend upon concentration and the ability of the heavy metal to be accumulated or transformed in living organisms. Environmental and toxicological properties depend on speciation. Heavy metal speciation is the state in which the molecule or ion appears in the environment.

The heavy metals are accumulated surreptitiously and reach the toxicity thresholds. Heavy metal releases identify the problems of heavy metal contamination in industrialized countries. As a result of cadmium and mercury contamination accidents [3,4,5]. Heavy metals occur in aquatic environments through both natural and anthropogenic sources. The exposure may result from direct releases into freshwater and marine ecosystems or through indirect streams such as dry and wet landfills and agricultural runoff. Some of the most relevant natural heavy metal sources are volcanic activity, contaminated continents, and wildfires. Volcanoes may contribute through either voluminous and sporadic volcanic activity-induced emissions or low-volume continuous emissions caused by magma outgassing and geothermal processes [6]. However, the main source of atmospheric mercury is the degassing of the earth’s crust. Inconsiderate use and disposal of mercury in metallurgical and chemical industries lead to serious impacts; most of these heavy metals are ingested, stored, and even accumulated in the human body, which leads to chronic diseases [7,8,9,10,11]. Therefore, the rejected pollutants in the environment require attention to be removed from the wastewater. The processes of adsorption, precipitation, solvent extraction, ion exchange, reverse osmosis, phytoremediation, flotation, filtration, sedimentation, electrochemical treatment, etc., are applied and further improved to treat heavy toxic metals from wastewater [12, 13]. The treatment of contaminated wastewater uses adsorption as a low-cost, environmentally friendly method. Heavy metal adsorption capacity study requires sludge [13], clays [14], silica gel, alumina, activated carbon [15], biochar and biochar-based composites [16], carbohydrate biopolymers [17], chitosan [18], and zeolites [19] as adsorbents. However, zeolites, which are inorganic crystalline polymers, may be used instead of activated carbon since the adsorption properties of such materials are significant. Zeolites are easily controlled (pore size variation, hydrophobicity, acidity, etc.) and are easily regenerated. Several fields use zeolites, such as adsorption, catalysis, and ion exchange. Among the various types of zeolites, the Y zeolite, which is considered a powerful molecular sieve with a Si/Al fraction of 1.5 to 3 and a skeleton geometry that is similar to that of the X zeolites and the uncommon mineral zeolite Faujasite, was identified by D. W. Breck to be the first successful exploration in this range of compounds [20]. The properties of the zeolite structure are the basis for its use in molecular adsorption. The ability to adsorb types of molecules rather than others confers a range of applications to zeolite as molecular sieves. Parameters that control access to the zeolite are the size and shape of pores; in other cases, several types of molecules penetrate the zeolite, and some diffuse through the channels fast, while the back of the zeolite retains other molecules, as in the paraxylene removal by Y zeolite [21]. Mintova records the particle size of Y zeolite [22] to be 60–70 nm and the channel size to be 7.4 Å. Herein is the removal of heavy metals (Pb2+ and Cd2+) using a Y zeolite study. In the present work, the authors used a commercial zeolite labeled CBV 760 and tested it for the removal of inorganic pollutants, namely, lead and cadmium, and studied the sorption properties of CBV 760 through those pollutants and an actual effluent.

Materials and methods

All chemicals were analytical-grade reagents. For the preparation of aqueous solutions, use distilled water. Y zeolite (CBV 760) is a zeolyst brand. The NaOH and HCl were analytical Sigma-Aldrich products; the Pb (NO3)2 was from Riedel de Haën, of 99% purity; and the Cd (NO3)24H2O was from Acrōs Organics, with a purity of 99%. Solutions of lead and cadmium were from Ficher Chemical commercial solutions at concentrations of 10,000 mg L−1 and 1000 mg L−1, respectively. Successive dilutions of the stock solution are made until desired concentrations are obtained and analyzed by atomic absorption spectroscopy. The analysis of atomic absorption spectroscopy was made with a varian AA240FS flame analyzer with an acetylene flame and a hollow cathode lamp as a light source (λ = 217 nm for lead and λ = 228.8 nm for cadmium). The SpectrAAv.5 software processes the results to obtain the residual concentrations of Pb2+ and Cd2+.

Adsorption experiment

Adsorption kinetic

Adsorption kinetics of lead and cadmium were investigated in solutions with pH 4.7 and 5, respectively. In a controlled temperature bath (298 K), flasks were filled with 50 mL of a 100 mg L−1 solution of metal ions. Contact time ranged from 1 min to 24 h, with a stirring rate of 130 oscillations per minute. Then, the solution was centrifuged, and the supernatant was analyzed by atomic absorption spectroscopy (AAS). The empirical equations below were used to determine the adsorbed amount of metal ions onto the adsorbent and the Y (%) removal efficiency.

$${Q}_{t}=\frac{V\left({C}_{0}-{C}_{\mathrm{e}}\right)}{\mathrm{m}}$$
(1)
$$Y \left(\mathrm{\%}\right)=\frac{{C}_{0}-{C}_{\mathrm{e}}}{{C}_{0}}\times 100$$
(2)

where C0 and Ce are the initial and equilibrium metal ion concentrations in the solution (mg L−1), V is the volume of the solution (L), and m is the adsorbent weight (g).

Effect of adsorbent dose

The operational conditions were the mass of the zeolite (60, 180, and 300 mg), contact times ranging from 1 min to 24 h, the pH of the medium, and the initial concentration of 100 mg L−1, temperature of 298 K, and a stirring rate of 130 oscillations per minute were maintained.

Effect of metal ions initial concentration

The effect of pollutant initial concentration was investigated in this study. Therefore, the concentration was varied by 50, 100, and 150 mg L−1.

Adsorption isotherm

Adsorption isotherms, which plot the adsorbed amount of pollutant against residual concentrations, were determined under the same experimental conditions as the adsorption kinetics by using a range of concentrations (25, 50, 100, 150, 200, 300, 400, 500, 600, 700, and 800 mg L−1). Within each solution (50 mL), the solution was kept in contact with 60 mg CBV 760 at 298 K for 24 h to ensure that equilibrium was reached.

Effect of temperature

The temperature effect was studied to determine the thermodynamic parameters. The adsorbed amounts of the selected pollutants were calculated at 308 and 318 K temperatures under the same experimental conditions (solution pH). The solution concentrations used for the present study were the same as those used for the adsorption isotherm study.

Effect of pH

The pH effect study uses 60 mg of zeolite to disperse in 50 mL of 150 mg L−1 solution to 298 K in 24 h. The solution pH is in the range of 2 to 8. To adjust the solutions, 0.1 N NaCl and NaOH solutions were used. Shake the suspensions for 24 h and centrifuge them. The equilibrium pH is measured using a pH meter (HANNA HI 123).

Sample characterization

The pHpzc pH at the zero charge point of the CBV 760 zeolite was measured to understand its behavior toward metal ions. Therefore, a NaCl (0.1 N) solution was prepared and adjusted to the desired pH value in the range (2–12) with HCl and NaOH (0.1 N) solutions. The adjusted 50 mL solution was mixed with the sample for 24 h at 25 °C before being filtered to determine the equilibrium pH. A diffractometer D8 Advance (BRUKER-AXS) records the X-ray spectrum of CBV using copper radiation (λ = 0.15418 nm) at a scan range of 2θ = 1–70° with a step size of 0.02°. FT-IR spectra were recorded in the wavenumber range 4000–400 cm−1 and acquired on an infrared Fourier transform (ALPHA, BRUKER spectrometer). An ASAP 2010 instrument measures the adsorption-desorption isotherms of N2 at 77 K. To calculate the specific surface area, the Brunauer-Emmett-Teller (BET) method was used. [23]. The sample was characterized through a Quanta 250 scanning electron microscope (SEM) coupled to an energy-dispersive X-ray analyzer (EDX). An aluminum support coated with a self-adhesive graphite pellet powdered with sample.

The modeling adsorption kinetics

To investigate the Pb2+ and Cd2+ adsorption on CBV 760, pseudo-first order, pseudo-second order, and intra-particle diffusion rate equations were used to interpret experimental data. The pseudo-first-order Lagergren rate expression [24] as defined by the

$$\mathrm{log}\left({q}_{e}-{q}_{t}\right)=\mathrm{log}{q}_{e}-\frac{{k}_{1}t}{2.303}$$
(3)

Here, k1 (g/ (min mg)) is the psedo-first-order model rate constant, and the plot of Ln (qeqt) against t is used to determine k1 and qe values. The equation below expresses the pseudo-second-order rate equation of Ho [25]:

$$\frac{t}{{q}_{t}}=\frac{1}{{k}_{2}{q}_{e}^{2}}+\frac{1}{{q}_{e}}t$$
(4)

The pseudo-second-order adsorption rate constant is k2 (g mg−1 min−1). The plot of t/qt as a function of t is used to determine the k2 and qe values. In the intraparticle diffusion model, the relationship between qt and t0.5 [26] is written as

$${q}_{t}={k}_{3}{t}^{0.5}+c$$
(5)

where k3 (mg/g g min0.5) is the rate of the intraparticle diffusion model and c is the line intercept proportional to the thickness of the boundary layer.

The modeling adsorption isotherms

Using the Freundlich isotherm model [27] to fit the experimental results of the Pb2+ and Cd2+ adsorption isotherms for all concentrations and the Langmuir model [28] by using Eqs. (6) and (7), respectively:

$$\mathrm{log}\;{Q}_{a}=\mathrm{log}\;{K}_{F}+\frac{1}{n}\mathrm{log}\;{C}_{e}$$
(6)
$$\frac{{C}_{e}}{{Q}_{a}}=\frac{1}{{Q}_{m}}{C}_{e}+\frac{1}{b{Q}_{m}}$$
(7)

where Qa is the amount of pollutant adsorbed onto solids in milligrams per gram, and Ce (mg L−1) is the residual concentration of Pb2+ and Cd2+.

KF and n were Freundlich constants related to the adsorption capacity and intensity, respectively, and KL (L mg−1) was a coefficient assigned to the affinity between the adsorbent and the adsorbate [29]. Another relevant parameter, RL, commonly called the separation factor, is obtained from the relation (8) [30].

$${R}_{L}=\frac{1}{1+{K}_{L}{C}_{0}}$$
(8)

where C0 is the concentration of the solute in the liquid phase. It was found that (i) 0 < RL < 1 indicates favorable adsorption, (ii) RL > 1 indicates unfavorable adsorption, (iii) RL = 1 indicates linear adsorption, and (iv) RL = 0 indicates irreversible adsorption.

The usage of the normalized standard deviation to assess the agreement between measured and calculated values is expressed as follows:

$$\Delta q \left(\mathrm{\%}\right)=\sqrt{\sum \frac{{\left[\frac{\left({q}_{\mathrm{exp}}-{q}_{\mathrm{cal}}\right)}{{q}_{\mathrm{exp}}}\right]}^{2}}{n-1}}\times 100$$
(9)

Thermodynamic parameters

Free energy (ΔG°) is used to determine the spontaneity of the system; this is obtained from the following equations [31]:

$$\Delta G^\circ =-RT\mathrm{ln}\;{K}_{d}$$
(10)

Thermodynamic parameters were calculated based on the below equation:

$$\begin{array}{c}{Ln \;K}_{d}\frac{-\Delta H^\circ }{RT}+\frac{\Delta S^\circ }{R}\\ {K}_{d=}\frac{{q}_{e}}{{C}_{e}}\end{array}$$
(11)

Thermodynamic parameters are determined by the plot of ln Kd = f (1 / T), slope, and intercept are \(-\frac{\Delta H^\circ }{R}\mathrm{ and }\frac{\Delta S^\circ }{R}, \mathrm{respectively}.\)

Results and discussion

Characterization

Analysis infrared spectroscopy

Figure 1 shows the CBV 760 zeolite infrared spectrum.

Fig. 1
figure 1

The FTIR spectrum of CBV 760 zeolite

The spectrum depicts bands at 529, 610, 680, 793, 837, 975, 1048, 1220, 1630, 1869, and 3430 cm−1. The band at 975 cm−1 is due to the Si-O-Al anti-symmetric stretching vibrations in the T-O bonds (where T = Si or Al) [32,33,34]. The 793 cm−1 band is due to the symmetric T-O-T elongation of S4R (single 4-ring) [35], while the 529 cm−1 absorption is attributed to the T-O-T symmetric stretching of the double 6-ring (D6R) [36]. Both bands at 610 and 680 cm−1 are assigned to the Si-O-Al symmetric elongation vibration and S4R symmetric bending modes [37], respectively. The 1048 cm−1 band is associated to asymmetric elongation modes of external links [38, 39], while bands at 837 and 1220 cm−1 are attributed to the symmetric and asymmetric elongation modes of the inner tetrahedra, respectively [40,41,42,43]. The 3430 cm−1 band corresponds to the bridged bond between Bronsted acid sites and OH groups and is found in the case of a protonated zeolite (CBV 760) [38, 44].

pHpzc

Figure 2 represents the curve of pH at zero charge (pHpzc). The pHpzc was determined to understand the surface chemistry of CBV 760. If the solution pH is lower than pHpzc, the adsorbent surface is positively loaded, thus making the adsorption of metal ions (lead and cadmium) difficult [45]. Instead, the cationic adsorption occurs at a solution pH above pHpzc.

Fig. 2
figure 2

pH dependence of adsorbent charge: determination of the Point of Zero Charge (PZC)

Analysis by X-ray diffraction (XRD)

The X-ray diffraction (XRD) studies the high-crystallinity Y zeolite structure. Figure 3 displays the XRD pattern of CBV zeolite.

Fig. 3
figure 3

XRD pattern of CBV zeolite

It was noticed that the sample spectrum in Fig. 3 showed a crystallinity identical to that of faujasite. The indicated peak positions are similar to those of Na-Y zeolite (the HY zeolite was synthesized from Na-Y zeolite), as reported in the COD card files (COD 96-900-0125).

The sample belongs to the Faujasite-type Y zeolite class, and the zeolite is in the Fd 3 m cubic space group (cell parameter a = 24.24) [34, 46]. With a chemical formula of H5.54Na0.1Al5.64Si186.35O384, it is HY ( protonated shape).

N2 adsorption–desorption

Figure 4 illustrates the adsorption–desorption isotherms of N2 on CBV zeolite.

Fig. 4
figure 4

The nitrogen adsorption/desorption isotherm of CBV 760

Based on the IUPAC classification, the isotherm shape is of type IV [47], with a hysteresis loop above P/P0 = 0.6 of type H3, which indicates that the adsorption and desorption processes do not occur in the same way [48, 49], as a result of mesopore availability. In low pressure, the adsorption is sharp, demonstrating the presence of micropores. The plot of pore volume versus pore size of the sample allows to determine the pore distribution by the BJH method [50] in the relative pressure range 0 < P/P0 < 0.35, indicating that the pore distribution is focused between 10 and 15 Å (Fig. 4a), which corresponds to micropores (Dp < 2 nm). The pore distribution is shown in Fig. 4b and was for the pore diameter between 178 and 480 Å, which corresponds to mesopores (2 < Dp < 50 nm). To determine the microporous and mesoporous volumes, researchers employ the t-plot method. This involved plotting the volume of gas adsorbed (Vadsorbed) as a function of the thickness (t) of the adsorbate layer, resulting in the construction of a t-plot. By plotting the Vadsorbed for CBV versus thickness (t), thus obtaining the t-plot. Surface area (S) was calculated based on BET theory using the following equation:

$$\frac{\frac{P}{{P}_{0}}}{{V}_{ads}(1-\frac{P}{{P}_{0}})}=\frac{1}{{V}_{mono}\times C}+\frac{C-1}{{V}_{mono}\times C}\times \frac{P}{{P}_{0}}$$
(12)

Vads is the adsorbed volume of nitrogen in Cm3.g−1 at STP under pressure P.

Vmono is the volume of the monolayer of nitrogen adsorbed onto the solid in Cm3g−1 at STP.

C is a constant that is characteristic of the interactions between adsorbate and adsorbent. Vmono and C are calculated by plotting \(\frac{\frac{P}{{P}_{0}}}{{V}_{ads\left(1-\frac{P}{{P}_{0}}\right)}}\) versus P/P0 over a low-pressure range 0.05 < P/P0 < 0.30.

The SBET specific surface (m.2 / g) can therefore be calculated by the equation:

$$S_{BET} = \frac{{V_{mono} \times N \times \sigma }}{{V_{m} }} = 4,35 \times V_{mono}$$
(13)

Vm is the molar volume of a gas under normal temperature and pressure conditions (= 22415 Cm3. mole−1).σ is the cross-sectional area of the adsorbate at solid or liquid density, which is 16.2 Å for nitrogen.

N is the number of Avogadro = 6.023 × 1023 mol−1.

The physicochemical characteristics of the studied CBV zeolite are summarized in Table 1.

Table 1 Physico-chemical properties of CBV zeolite

Analysis by scanning electron microscopy (SEM)

The SEM method allows us to analyze the physical appearance, crystallinity, and crystal morphology of zeolite. The SEM result in Fig. 5a shows octahedral crystals, as expected from Y zeolite’s high crystallinity. The EDX analysis (Fig. 5b) indicates that the Si/Al ratio of CBV zeolite is 4.53.

Fig. 5
figure 5

Scanning electron microscopy (a) and EDX analysis (b) of zeolite CBV760

Lead, cadmium adsorption

Adsorption kinetic

The Pb2+ and Cd2+ adsorption kinetics on CBV zeolite were studied over the contact time (Fig. 6) to determine the kinetic equilibrium time.

Fig. 6
figure 6

Adsorption kinetics of Pb2 + and Cd2 + on CBV zeolite at 298 K, m = 60 mg, [C] = 100 mg.L−1, pH 4.7 and 5 for Pb2 + and Cd2 + , respectively

The adsorption kinetics reaches the equilibrium time within 30 min for lead and 15 min for cadmium. This difference in equilibrium time is due to the ionic radius of the two metals and their degree of electronegativity. The ionic radii are 0.98 and 0.78 A° for lead and cadmium, respectively. The metal with a small ionic radius (cadmium) diffuses before Pb2+ in the internal surface of the HY zeolite. In terms of electronegativity, the one with high electronegativity has an attraction for the single electrons of the functional groups (OH) on the surface of zeolite, as in the case of lead (2.33) compared to cadmium (1.69), hence a high adsorption capacity for Pb2+.

The kinetic curve shows that the removal of metal ions was fast in the first few minutes due to surface adsorption. Curves are simple and lead to saturation because of the intraparticle diffusion process. Such a curve indicates the possibility of monolayer coverage of both metals on the CBV zeolite surface [51].

Effect of adsorbent dose

Pb2+ and Cd2+ adsorption kinetics onto the CBV 760 zeolite in a varied adsorbent dose range are displayed in Fig. 7.

Fig. 7
figure 7

The effect of adsorbent dose on retention capacity of Pb2 + , Cd2 + on CBV at 298 K, [C] = 100 mg.L−1, pH 4.7 and 5 for Pb2 + and Cd2 + , respectively

Figure 7 shows that the adsorbed amount decreased with the increase in adsorbent mass; this is due to

  1. (i)

    The aggregation of particles and the increase in the diffusion path length led to a decrease in the total surface area of the adsorbent and a reduction in the adsorbed amount [52].

  2. (ii)

    The high mass of adsorbent leads to saturated adsorption sites [53].

Effect of metal ions initial concentration

Figure 8 depicts how the adsorption capacity increases with increasing initial concentration.

Fig. 8
figure 8

Effect of metal ions initial concentration on retention of Pb2 + , Cd2 + on CBV at 25 °C, m = 60 mg, pH 4.7 and 5 for Pb2 + and Cd2 + , respectively

Variations in the initial metal concentration may affect the adsorption efficiency through different factors, such as the availability of the surface functional groups and the ability of the surface functional groups to bond to the metal ion [54]. It was also the result of the driving force of the metal ions to the adsorption sites increasing as the initial concentration of the metal ions increased. Thus, the increase in the metal ion adsorbed amount per unit mass. Results shown in Fig. 8 indicate that the Pb2+ adsorbed amount is more than that of Cd2+ due to the chemical and physical properties of the investigated metal ions, such as the atomic radius [55].

Effect of pH

The pH effect study assesses the retention behavior of heavy metals on the zeolite. The adsorbed amount (Qa) of Pb2+ is higher than that of Cd2+ at the same pH value as intended (Fig. 9).

Fig. 9
figure 9

pH Effect on adsorption of Pb2 + and Cd2 + onto CBV at 298 K, m = 60 mg, [C] = 150 mg.L1

Regardless of the considered ion, the adsorbed amount (Qa) is increased with pH and reaches a maximum in the pH range of 6–8. The increase in Qa is relatively fast at pH values of 4–6. Indeed, the Pb2+ ion adsorption capacity increased as the pH increased from 4 to 6, from 24.83 to 106.66 mg. g−1. For the Cd2+, the adsorbed amount rises from 6.5 to 44.91 mg. g−1 with the increase in the Cd2+ pH solution from 4 to 6. At high-acidic pH values, the H+ ions are reported that compete with Pb2+ and Cd2+ ions for the adsorption sites upon the CBV zeolite. However, the electrostatic attraction of the dissolved ions to the surface charge controls the surface complexation reactions. Indeed, as cadmium and lead have ionic radii (0.97 and 1.20 Å), respectively, thus possess a low charge density and are hence more affected by surface group protonation [56] that reduces the adsorption sites on the CBV zeolite. As a result, fewer sites are available to uptake the studied metal ions. The metal ion adsorbed amount increases with the pH solution, reducing the competition of H+ ions. For further explanation, surface hydroxyl sites can be yielded through functional group hydrolysis on the zeolite surface at alkaline pH values, which are beneficial for metal ion removal. Furthermore, the increase in the adsorbed amount of metal ions is due to Pb2+ and Cd2+ ions hydration. Indeed, at pH > 8, Pb2+ and Cd2+ hydrate, forming the PbOH+, Pb(OH), and Cd(OH)2 species [57] (precipitation of ions occurs within this pH range).

Adsorption isotherm

Figure 10 depicts the adsorption isotherms of Pb2+ and Cd2+ on CBV zeolite.

Fig. 10
figure 10

Adsorption isotherm of Pb2 + (a) and Cd2 + (b) on CBV zeolite at 298 K, pH 4.5 and 5 (the solution pH of Pb2 + and Cd2 + , respectively)

The maximum CBV adsorption amounts are 178 and 176 mg. g−1, with residual concentrations of 586.45 mg. L−1 and 616 mg. L−1, for lead and cadmium, respectively, as shown in Fig. 10. The Pb2+ and Cd2+ adsorption isotherms onto CBV zeolite are of S type by the classification of Giles [58]. Such isotherms indicate an intense competition for the adsorption sites between the solvent molecules and the studied ions. S-shaped curves characterize the wide pore sizes of the adsorbents. This kind of curve with an inflection point in Fig. 10a is associated with the successive adsorption process that occurs in the case of metal cation adsorption through the functional groups on the adsorbent surface with various mechanisms of interactions between the adsorbate and the functional groups, as reported by Neskoromnaya et al. [59]. Adsorption of lead and cadmium on zeolite occurs in two steps: the first one is a monolayer (Langmuir) in the low concentration range, and the second one is a multilayer (Freundlich) in the high concentration range. As the solution concentration increases, the adsorption becomes easier, reaching saturation when adsorbed ions occupy all adsorbent receptor sites. Thus, a single layer of adsorbate (a monolayer) forms.

Effect of temperature on adsorption of Pb2+ and Cd2+ onto CBV zeolite

Temperature is a relevant thermodynamic parameter that can impact the adsorption on a solid surface. Indeed, adsorption is a physical process related to the molecular interactions between a liquid or gas phase and the adsorbent surface. The temperature effect may modify these interactions [60, 61].

The result in Fig. 11 demonstrates that for all investigated temperatures (298, 308, and 318 K), the Pb2+ adsorbed amount (Qa) on the CBV is more than that for Cd2+. Indeed, considering the temperature of 308 K and at the same initial concentration (C0) of Pb2+ and Cd2+ of 500 mg. L−1 Qa is 175 mg. g−1 and 148 mg. g−1. These adsorbed amounts correspond to the respective residual concentrations (Ce) of 316 and 322 mg. L−1. However, the driving force of metal ions to the active sites of the CBV zeolite may accelerate with the increase in temperature [62, 63].

Fig. 11
figure 11

Temperature effect on adsorption of Pb2 + (a) and Cd2 + (b) onto CBV zeolite, m = 60 mg, pH 4.7 and 5 for Pb2 + and Cd2 + , respectively

The modeling adsorption kinetics of Pb2+and Cd2+ onto CBV

The results (the rate constants and R2 correlation coefficients) in Tables 2 and 3 are from modeling the adsorption kinetics of Pb2+ and Cd2+ using the pseudo-first-order (PFO), pseudo-second-order (PSO), and intraparticle models.

Table 2 Kinetic model parameters attributed to lead and cadmium
Table 3 Determination of intraparticular model rate constants to lead and cadmium onto CBV

The pseudo-second-order model (PSO) fits the experimental adsorption kinetics of Pb2+ and Cd2+. The R2 values are close to one, and the PSO qe values proximate to the experimental data support the obtained result. A comparison of the standard deviations of the two models finds that ∆q (%) is low for the PSO model compared to the PFO model. This statement is for both studied metal ions. The qe value determined by the PSO model is 68.49 mg. g−1, close to the experimental value of 69.80 mg. g−1. Concerning the cadmium, the value of qe calculated by the PSO model is 30.47 mg. g−1, close to the experimental value, which is 30.30 mg. g−1, in contrast to the PFO model. (Table 2). That result suggests that Pb2+ and Cd2+ adsorption might be a chemisorption [64]. The modeling of the adsorption kinetics uses the intraparticle model to further understand the adsorption by plotting qt versus t0.5. If the pattern is linear and passes through the origin, the adsorption follows the intraparticle diffusion model. However, if an external mass transfer or chemical reaction is involved, the straight line does not pass through the origin [65, 66].

The multilinearity obtained in the graph (figure not shown) implies that two or more reactions occur [67, 68]. The rate constants express the different stages of adsorption, and the variation of these constants can be due to the surface adsorption stages (Fig. 12). Heavy metals (Pb2+, Cd2+) are adsorbed fast by the outer adsorbent surface. Nevertheless, after the external surface is saturated, the heavy metals penetrate the adsorbent particles and adsorb onto the inner surface of the adsorbent. Therefore, the heavy metals diffuse into the adsorbent pores. This stage represents the final stage and reaches equilibrium, indicating that the intraparticle diffusion becomes slow due to the low solute concentrations in the solution and thus leads to a decrease in the rate constants [69, 70] (Table 3).

Fig. 12
figure 12

Variation of Pb2 + and Cd2 + adsorption constant onto CBV as a function of temperature

Adsorption mechanism of metal ions (Pb2+ and Cd2+) onto CBV zeolite

This study showed that heavy metals diffuse into the adsorbent pores, which indicates slow diffusion of ions into the internal channels of the zeolite. Then, the ions move to the position of exchangeable ions within the crystal structure [71,72,73]. Ion exchange occurs between the exchangeable cations (Ca2+, Na+, and K+) in the crystal structure of the zeolite and the metal ions (Pb2+ and Cd2+). The HY zeolite is porous, characterized by SiO44− and AlO45− tetrahedra linked by their oxygen. The coexistence of aluminol and silanol groups leads to a Bronsted and Lewis acidity. The adsorption of metal ions takes place at the pH of the solution (pH 4.5 for cadmium and pH 5 for lead) without adjustment. Within the acidic medium, the oxygen of aluminol and silanol groups releases a proton and forms a negative charge. The functional groups that contain oxygen are mainly involved in metal ion adsorption. By electrostatic attraction, the negative charge attracts the adsorbed metal ions to create a complex on the inner sphere adsorbent surface. That leads to the formation of the Pb-(Cd)-O bond, and the covalent bonds Pb-O-Si and Cd-O-Si promote the adsorption due to the presence of high adsorbent-adsorbate interactions [74]. Thus, the ion exchange, electrostatic, and surface complexes mechanisms govern the adsorption process [74, 75].

Modeling adsorption isotherms

The modeling of adsorption isotherms uses the Freundlich and Langmuir models for the studied temperatures (298, 308, and 318 K). Table 4 shows the obtained results.

Table 4 Model parameters Langmuir and Freundlich of adsorption of metal ions (Pb2+, Cd2+) onto CBV at different temperatures

Table 4 shows that the KL parameter (which gives information about the adsorbate-adsorbent bond) of the Langmuir model and the adsorbed amount (Qm) increase with the increase in temperature. Therefore, the bond between metallic ions and the CBV strengthens as the temperature rises. The KL constant of Pb2+ is greater than that of Cd2+ for the investigated isotherms. Hence, the CBV-Pb2+ bond is high than that of CBV-Cd2+. The Qm values obtained by the Langmuir model are close to experimental values at the same temperature. Indeed, at T = 318 K, Qm = 303.03 mg. g−1 for Pb2+, which is close to that obtained by the experiment of 304 mg. g−1. The Qm = 290.70 mg. g−1 (from the Langmuir model) for Cd2+ is close to 291 mg. g−1 (from the experiment). However, the RL parameter is in the 0 < RL <1 range, which indicates that the Langmuir isotherm is favorable, and the standard deviations ∆q (%) are less than those of the Freundlich model. Therefore, the metal ion adsorption is monolayer, which means the surface is homogeneous. The interactions between the adjacent adsorbed ions are non-lateral when a single ion occupies a single surface site [76]. Except for the Cd2+ adsorption isotherm over CBV at T = 318 K, the Freundlich model is almost not fitted to the obtained results due to the R2 correlation coefficient, which is R2 < 1 despite n > 1 (favorable isotherm) [77, 78]. The Freundlich model can be used for non-homogeneous surfaces and considers that the concentration of the adsorbed ionic species on the solid surface increases as the concentration of certain species in the liquid phase is enhanced [79, 80].

Thermodynamic study

The thermodynamic of the adsorption requires the knowledge of two thermodynamic parameters: the free enthalpy (∆G°) and the entropy (∆S°) [81, 82].

The thermodynamic parameters ∆H° (kJ. mol−1), ∆S° (kJ K−1. mol−1), ∆and G° (kJ. mol−1) of Pb2+ and Cd2+ adsorption on CBV at different temperatures are in Table 5.

Table 5 Thermodynamic parameters of Pb2+ and Cd2+ adsorption onto CBV

The ∆H° value is lower than 40 kJ. mol−1 (which indicates the interaction energy between adsorbate-adsorbent) and positive for Pb2+ and Cd2+ adsorption on CBV. It reveals a strong bond between the metallic ions and the CBV zeolite and proves that the Pb2+ and Cd2+ ions adsorption process is endothermic. [83, 84]. Adsorption metal ion kinetic modeling suggested that the adsorption may be a chemisorption. The thermodynamic study determined that the adsorption process is physical adsorption since the free enthalpy < 40 kJ. mol−1 [85].

During the adsorption of metal ions, structural changes at the adsorbate-adsorbent lead to a positive entropy. That increases the disorder during the adsorption of Pb2+ and Cd2+ on the CBV 760 zeolite [86, 87]. The result in Table 5 shows that the standard free enthalpy ∆G° decreases with the increase in temperature, indicating a spontaneous adsorption onto CBV zeolite [88, 89].

Adsorption of an actual contaminated battery effluent on CBV zeolite

The adsorption study of lead and cadmium from a synthetic solution onto CBV zeolite is efficient. The aim is to investigate the performance of CBV zeolite through heavy metals in an actual contaminated effluent, which is an R0 effluent issued from the battery manufacturing company from Oued S’mar (Algiers, Algeria). The experimental conditions are the same as those of the Pb2+ and Cd2+ adsorption kinetics. Table 6 displays the various concentrations of metal ions in R0 before CBV contact.

Table 6 Concentration of R0, R1 before and after contact with CBV

After the CBV zeolite contact, the supernatant is recovered after centrifugation and analyzed by AAS. Table 6 shows the results.

Table 6 displays that the removal capacity of heavy metals onto CBV follows the order: Ni2+<Cu2+<Cd2+<Pb2+ in R1. Due to the selectivity of CBV for lead and cadmium and their ion exchange capacities during the adsorption, the adsorbent exhibited high removal efficiency for these heavy metals. [90,91,92]. Results reported in Table 6 indicate a high capacity of the CBV for heavy metal removal from an actual contaminated battery effluent.

Conclusion

The present study uses a commercial faujasite Y zeolite (CBV 760) to remove heavy metal (Pb2+ and Cd2+) from synthetic solution and actual contaminated battery effluent. The adsorption efficiency of CBV 760 zeolite towards heavy metals such as Pb2+ and Cd2+ was tested. The adsorption kinetic of Pb2+ is fast compared to that of Cd2+. The adsorption kinetics reveals that the adsorption of Pb2+ and Cd2+ depends on the contact time, the adsorbent mass, the pH of the solution, the initial concentration, and the temperature. The comparative study of adsorption isotherms of (Pb2+, Cd2+) on Y zeolite (CBV 760) shows that the lead was better adsorbed than cadmium in synthetic solution and real contaminated effluent due to the physicochemical properties of both pollutants. Therefore, the CBV 760 can be used to deal with environmental problems due to the cages and supercages that characterize the zeolite, which confer it the highly porous texture, and also due to the high specific surface area, which provides the capacity to remove pollutants.