Skip to main content
Erschienen in: Biomass Conversion and Biorefinery 3/2024

Open Access 06.04.2022 | Original Article

Humic acid and nano-zeolite NaX as low cost and eco-friendly adsorbents for removal of Pb (II) and Cd (II) from water: characterization, kinetics, isotherms and thermodynamic studies

verfasst von: Mamdouh S. Masoud, Alyaa A. Zidan, Gehan M. El Zokm, Rehab M. I. Elsamra, Mohamed A. Okbah

Erschienen in: Biomass Conversion and Biorefinery | Ausgabe 3/2024

Aktivieren Sie unsere intelligente Suche, um passende Fachinhalte oder Patente zu finden.

search-config
loading …

Abstract

Humic acid as a green-sorbent was synthesized from marine sediments. While kaolin was modified to nano-zeolite NaX. Different tools such as FT-IR, SEM, EDX and XRD were applied to confirm the characteristics of the generated green-sorbents. Different factors such as pH, contact time, sorbent dosage, initial metal ion concentration, temperature and interfering ions were carefully examined and used to optimize the batch adsorption process for Cd2+ and Pb2+. A small dose of nano-zeolite of 100 mg was required to attain the maximum adsorption of Pb2+ at pH about 7, shacking time at 60 min and Pb2+ concentration at 30 ppm. Also, the maximum sorption capacity of Cd2+ ions on nano-zeolite was achieved in a neutral medium and very short contact time implying the economic feasibility of the adsorption process. In the case of humic acid, the maximum removal capacity for Pb2+ and Cd2+ was operated at acidic medium and shacking time was 40 min. Metal ions remediation results were evaluated by some adsorption isotherm models at different temperatures. The kinetic and thermodynamic variables were also computed. The data fitted very well with the linear Langmuir and the pseudo-second-order model implying a favourable adsorption process. The sorption of Cd2+ and Pb2+ was regulated by both external mass transfer and intraparticle diffusion steps over the whole range of concentrations, as shown by the results. The metal ions removal percentage from four real water samples by green sorbents were applied and provides good evidence of two sorbents as promising eco-sorbent for removal of heavy metal ions.
Begleitmaterial
Hinweise

Supplementary Information

The online version contains supplementary material available at https://​doi.​org/​10.​1007/​s13399-022-02608-9.

Highlights

• Eco synthesized of humic acid and nano-zeolite NaX from modified marine sediments.
• Characteristics of the generated green-sorbents were confirmed using different tools.
• Different factors were examined to optimize the batch adsorption process for Cd2+ and Pb2+.
• Adsorption isotherm models were used to evaluate the metal removal efficiency.
• Good evidence of two sorbents as promising eco-sorbent for removal of heavy metal ions.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

1 Introduction

Heavy metals are one category of hazardous substance to aquatic ecosystems that keeps getting a considerable interest [13]. Even at low concentrations, they have significant environmental consequences, owing to their inability to be easily degraded. Cadmium and lead, for example, are highly toxic metals. Both metals may lead to serious health threats including renal damage, cancer and neurological destructions [4]. Cd (II) is a mass-produced waste coming from fertilizer, battery and alloy industries. On the other hand, Pb2+ often results from corrosion of plumbing and deposition of airborne aerosols [5]. Sources such as agriculture and sewage that contain a huge amount of organic matter (dissolved and particulate) disposal also contribute to the level of toxic heavy metals in the marine environment [6]. Both Cd2+ and Pb2+ have raised concern over their higher than the allowed concentration found in water [7]. Removal of heavy metal ions from wastewater is now a subject of major interest for industrial and environmental protection agencies. Besides, heavy metal removal is an important stage in landfill leachate treatment. Many physical-chemical treatment processes can be very costly because large quantities of chemicals are required through the progression steps. As such, the use of low-cost natural and effective adsorbents would be beneficial and hence should be extensively investigated as a target for the removal of heavy metals from landfill leachate [8, 9].
Marine sediments were used as a precursor to synthesize an effective green-sorbent for remediation of heavy metals from aquatic matrices. Humic acid (HA) is a portion of humic substances that exist naturally in rivers and lakes with black colour and high molecular weight. It can be naturally extracted from marine sediment at pH~2 but may be soluble at higher pH [10]. Humic acid possesses a large number and versatile functional groups such as carboxylic acid, phenolic hydroxyl, sulfonic acid, carbonyl and methoxy moieties that are connected with each other through hydrogen bonds and hydrophobic dispersion forces [11, 12]. In aqueous solutions, these groups can dissociate to units of negative charges that bind chemically to the cross-linked carbon network of humic macromolecules. Moreover, the carboxylic and phenolic groups of HA act as complexation sites to metal ions, therefore HA has high reactivity and can easily reduce metal ion concentrations in the environment by chelation [13, 14]. Furthermore, HA has many technological applications due to its non-toxic, inexpensive and biocompatible features [15].
Zeolite is a naturally occurring crystalline aluminosilicate mineral that consists of AlO4 and SiO4 tetrahedrons connected through oxygen atoms to form a porous aluminosilicate structure with the chemical formula xNa2O. Al2O3. ySiO2. zH2O where x, y and z are sodium oxide, silicon oxide and water molar ratios, respectively [16]. The intra-crystalline voids in the zeolite framework are occupied with water molecules and cations to neutralize the negative charge of AlO4 [17]. Zeolite is generally made from sodium silicate and hydrogel aluminate. Kaolin is preferentially used as a precursor in zeolite synthesis due to its high content of silica and alumina. The unique porous structure, thermal as well as chemical stabilities of zeolites make them appropriate for several practical applications, including gas and vapour separations, membrane reactors, chemically selective adsorption and slow-release fertilizers in agriculture [18]. The chemical composition of zeolite has a net negative charge on the crystal which allows it to combine with cations by ion exchange. Recently, nano-porous zeolites have received a lot of attentions [19, 20]. Reducing the particle size from micrometer to nanometer scale has led to a substantial change of material characteristics and their capabilities in adsorption [21]. For nano-zeolites, the number of atoms in the unit cell increases which cause the external surface area and surface energy to increase. Also, the diffusion path length in nano-zeolites is shortened as compared to that in the conventional micrometer zeolites.
The goal of our research was to develop natural sorbents that were effective, inexpensive and environmentally friendly. The materials chosen for their high metal ion adsorption properties are HA and nano-zeolites. Some modifications occur during the synthesis of nano-zeolites from kaolin using the hydrothermal gel process in order to achieve economic benefit. Different spectroscopic techniques were used to characterise the powder. In addition, the prepared material’s adsorption capacity was investigated. Marine sediments from the Egyptian coast were used as a precursor to create an effective HA for heavy metal remediation in aquatic matrices. Humic acid was investigated in this paper using equilibrium batch experiments and possible scenarios to determine the mechanism and effectiveness of HA adsorption.

2 Materials, preparations and methods of measurements

2.1 Chemicals and instrumentations

All reagents used in this study such as cadmium acetate, a source of cadmium ions (Cd2+), lead nitrate, a source of lead ions (Pb2+), sodium hydroxide, sodium chloride, potassium chloride, magnesium sulfate, potassium nitrate, sodium carbonate and calcium carbonate were purchased from Sigma-Aldrich and used without further purification. The pH meter (Jenway: model, 3505) used in the adjustment of the sorption solution was calibrated by three standard buffer solutions of pH 4, 7 and 12. Deionized double distilled water was used in all investigation steps such as cleaning, adsorbent synthesis and metal stock solution preparation. The green synthesized adsorbents were characterized by different techniques. Fourier transform infrared spectroscopy (FT-IR, Shimadzu spectrophotometer) of the dried samples was carried out by mixing 1 mg of the tested material with 300 mg of pure dry potassium bromide (KBr). The measurements were recorded between 400 and 4000 cm−1. X-ray diffraction (XRD) patterns for phase structure and material identification of different types of adsorbents were recorded by Peak Find-Memory-27 spectrophotometer using a generator of 30 mA with 40 kV and Cu Kα radiation (λ = 1.5406 Å). Scanning electron microscope (SEM, JEOL: JSMIT 200) was employed to visualize the size and morphology of the synthesized particles. The cracked surfaces of the samples were coated with a thin layer of gold before investigation. The elemental composition of the yielded sorbents was analyzed by an energy-dispersive x-ray spectrometer (EDX). The metal ions concentration (Cd2+ and Pb2+) were measured using atomic absorption spectrophotometer (AAS, Analytic Jena contra 300 Atomic Absorption, Germany). The experimental work of the current study was carried out at the National Institute of Oceanography and Fisheries (NIOF). Instrumental measurements were done at the central laboratory of the Faculty of Science, Alexandria University and the City of Scientific Research and Technological Applications, Egypt.

2.2 Green synthesis of nano-zeolite-NaX

The nano-sized zeolite-NaX was synthesized by the hydrothermal alkaline activation method [21] using natural clay, kaolin, as a source of silica. The raw material was calcined at 600 °C for 3 h to obtain a metastable phase referred to metakaolin. The stable phase was mixed with freshly prepared aluminosilicate containing Na2O, NaOH, Al2O3, SiO2, EDTA, NaCl and H2O in the optimal molar ratio of 5Na2O.Al2O3.4SiO2.70H2O.1.2 EDTA.2NaCl. The mixture was set aside for ageing for 4 days at room temperature. The obtained nano-zeolites were repeatedly washed with deionized water until the pH of the supernatant was 9. Nanoparticles were then dried at 110 °C and calcined at 550 °C.

2.3 Extraction of humic acid

Humic substances were extracted by the fractionation method as shown by [22]. The soil used as the source of humic acid (HA) was taken from El-Umum drain, El-Mex bay, Egypt. The extraction scheme is shown in Fig. 1.

2.4 Metals adsorption by batch equilibrium technique

A stock solution of (1000 mg/L) of lead (Pb2+) and (1000 mg/L) of cadmium (Cd2+) were prepared using deionized water. The metal removal efficiency was evaluated by batch equilibrium technique in which a certain amount of each adsorbent was left in contact with adsorbate solutions under continuous stirring [23]. pH, adsorbent dosage, contact time, temperature, adsorbate concentration and interfering ions are the crucial parameters that were carefully considered and optimized in separate experiments to detect the metal sorption efficiency

2.4.1 Effect of pH

The effect of pH on the sorption capacity of metal ions was determined experimentally in the range of pH 2 to 7 at room temperature. A 10 ppm as an initial Pb2+ and Cd2+ concentration was prepared in a 50 mL measuring flask and the pH was adjusted using standard buffer solutions. The pH-adjusted metal ion solution was then added to 20 mg of green-synthesized sorbents and shacked automatically at a rate of 150 rpm for 30 min. After equilibrium was achieved, the samples were filtered and heavy metal concentrations were analyzed.

2.4.2 Optimal adsorbent dosage

The effect of the green-sorbent dosage on the metal sorption capacity has been investigated to determine the optimal dose. Various weights of 10, 20, 30, 40, 60, 80 and 100 mg of eco-sorbents for Pb2+ removal, and 20, 40, 60, 80, 100, 120, 140, 160, 180 and 200 mg of eco-sorbents for Cd2+ removal have been added to a different series of 25 mL of 10 ppm of selected metal ions solution while maintaining the pH value of selected metal ions and specific sorbent constant. The different series were shacked at 150 rpm for 30 min at room temperature.

2.4.3 Optimal contact time

The optimal contact time for sorption of Pb2+ and Cd2+ was assessed at room temperature and 150 rpm stirring rate over a range of time intervals; 10, 20, 30, 40, 50 and 60 min, on 25 mL of working metal ions solution with a concentration 10 ppm for each metal ion at certain green-sorbent dosage and pH.

2.5 Effect of temperature

Temperature variability experiments were conducted at 10, 20, 30, 40, 50 and 60 °C, with an initial concentration of each metal equal 10 ppm at optimum pH, contact time and sorbent doses.

2.5.1 Effect of metal ion concentrations

The impact of initial metal concentration (Cd2+ and Pb2+) on adsorption onto the sorbents at optimal pH, doses, contact time and metal concentrations of 5, 10, 15, 20, 25, 30, 40, 60, 80 and 100 ppm were examined in 25 mL deionized water. The remaining heavy metal concentrations at equilibrium were analyzed. The results were fitted to different isotherm and kinetic models.

2.6 Effect of interfering ions

A 25 mL of optimum concentration 30 ppm, 25 ppm of Pb2+ and 10 ppm, 5 ppm of Cd2+ on nano-zeolite nax and humic acid, respectively, were mixed with selected interfering ions such as NaCl, KCl, MgSO4, KNO3, NaCO3 and CaCO3 at optimum pH, contact time and sorbent doses to investigate metal ions sorption capacity when competing ions are present. The green-sorption capacity value was calculated from a metal mass balance equation:
$$q=\frac{\left({C}_o-C\right)}{m}{V}_L$$
(1)
where Cο and C (mol/L) is the initial and residual metal ion concentration, respectively. VL is the aqueous volume of the sorption reaction in a litre, m is the mass of dry sorbent in gram and q (mg/g) is the metal sorption capacity.

2.7 Metals adsorption by microcolumn technique

The microcolumn technique was performed for water treatment studies of environmental samples collected from Egyptian water bodies and wastewater discharge [24, 25]. The applicability and implementation of different sorbents for the removal of Cd2+ and Pb2+ were examined using a multistage column packed with each sorbent. Water samples were collected from different areas as Eastern Harbor (normal seawater), El-Mex Bay (polluted brackish-water), El-Mahmoudia Canal (freshwater) and West Dessert Operating Petroleum Company (wastewater). The physicochemical parameters of different water samples such as salinity, pH and oxidizable organic matter (OOM), as well as, dissolved nutrient salts (nitrate (NO3-N), nitrite (NO2-N), dissolved inorganic phosphate (DIP) and dissolved inorganic silicate (DSi), non-conservative constituents such as Ca2+, Mg2+, Na+, Li+ and K+ and heavy metals (Pb2+ and Cd2+) were evaluated. Removal of heavy metal by adsorption on green adsorbents was performed by passing 1 L of water sample (spiked with Cd2+ and Pb2+) through a packed column. The flow rate was adjusted to 0.5 mL/min. Three consecutive extractions have been performed and the filtrate was collected and analyzed for Cd2+ and Pb2+ ions using AAS.

3 Results and discussion

3.1 Spectral characterization of the synthesized sorbents

3.1.1 FT-IR spectroscopy

Detailed fundamental FT-IR bands of the synthesized sorbents are given in Figs. 2 and S1. The spectrum of free nano-zeolites NaX showed six main absorption bands (Fig. 2a). The stretching vibration of hydroxyl groups (Si-OH) which are present in the faujasite super and sodalite cages as the building blocks of the zeolite structures causes the band at 3449 cm−1 [26]. Also, the well-defined band at 1635 cm−1 is attributed to OH bending vibration (M-OH-M, where M = Si, Al) since it coexists with OH stretching vibration positioned at 3449 cm−1 [27]. Moreover, the sharp IR band at 1036 cm−1 and the less intense band at 789 cm−1 are assigned to the asymmetrical and symmetrical stretching of (M-O-M) group. Besides, the band in the spectral region 538–695 cm−1 is due to external linkage vibrations, which indicates the association of rings that connect the sodalite cages (Si-O-Na). The sharpness of this peak implies the crystallinity of zeolite X [27, 28]. Furthermore, the band at 469 cm−1 is characteristic of O-M bending vibration of the SiO4 and AlO4 in the internal tetrahedral structure [29]. Remarkably, the band at 1036 cm−1 of nano-zeolite (Fig. 2a) exhibited changes in the intensity and shifts in position in the spectra of the metal exchanged samples (to 915 and 917 cm−1 for NaX-Cd2+ and NaX-Pb2+, respectively, Fig. 2b, c). This is probably due to the Ḿ-O (Ḿ = Pb2+ or Cd2+) bond formation which alters the vibrational frequencies depending on the mass, charge, ion size and cation environment [30]. As well, the disappearance of some bands in the metal exchanged spectra can provide evidence for structural substitution of the Al atoms or dealumination. Similarly, the diverse functional groups present in the free HA were assessed by FT-IR (Fig. S1a). The spectrum of HA showed bands at 3425 cm−1 corresponds to the stretching vibration of the O–H group in –COOH as well as alcoholic and phenolic –OH groups [19]. The absorption bands in the range 2920–2851 cm−1 could be attributed to aliphatic C–H stretching vibrational mode in methyl and/or methylene groups [31]. Also, the band at 1583 cm−1 is ascribable to aromatic COO, C=O of carbonyl and quinone stretching vibration [32]. The sharpness of the bands approved the good extraction and purification of humic acid. Additionally, the FT-IR spectra of the M(II)–HA complex give information about the binding modes of HA with the metal ions. For instance, the band at 1583 cm−1 in HA was shifted to 1573 cm−1 and 1560 cm−1 after M(II) treatment by Cd2+ and Pb2+ ions, respectively, indicating the involvement of oxygen in COOH groups in the interaction with the M(II) ions (Fig. S1b, c). The new IR bands emerged at 1381 cm−1 and 1361 cm−1 in the Cd2+ and Pb2+ spectra are assigned to the M-OH stretching vibrations which indicated the interaction between M(II) and OH groups of the surface [33]. Also, the band at 1041 cm−1 is attributed to –CO stretch of polysaccharides and Si–O from silicate impurities [34]. Besides, the peak at 536 cm−1 (assigned to C–O–C bonding in the free HA) was slightly shifted to 539 cm−1 and 527 cm−1 and exhibited notable intensity change after loading the sorbent surface with Cd2+ and Pb2+, respectively. Therefore, the observations of the current FT-IR study indicated that -COOH, -OH and -CO functional groups present in HA are primarily involved in M(II) binding via ionic interactions on the sorbent's surface as reported by previous studies on HA [35].

3.1.2 XRD, EDX and SEM analyses

Qualitative mineralogy of the synthesized sorbents was determined with the standard interpretation procedures of XRD, where the position of a diffraction peak that represents a lattice plane can be identified by using a database of diffraction patterns. The XRD spectra measured in a range of 2θ from 5° to 70° (Fig. 3). The green synthesized nano-zeolite NaX structure has achieved a crystallinity percentage of about 88.12% after 4 days of ageing at room temperature, indicating amorphous aggregated cubic crystals. The XRD peaks at 2θ = 6.15°; 9.64°; 17.97°, 20.15°, 22.88°, 26.79°, 28.77° and 37.86° indicated the formation of nano-zeolite NaX crystals [17, 36]. These peaks may be indexed as the diffraction of the planes (111) (220) (511) (440) (620) (733) (822) (1022), respectively (JCPDF File No. 39-0218) [37] which indicate the cubic crystals of nano-zeolite. However, the peaks at 2θ = 20.76° and 35.57° were attributed to kaolin which suggests the incomplete dissolution of kaolin with sodium hydroxide during the preparation of gel mixture. As zeolites are thermodynamically metastable phases, the crystallization time allowed the formation of a mixture of Na-X and NaP types. The peaks at 2θ = 28.1°, 38.1° and 46.1° are the characteristic peaks of NaP zeolite [36, 38]. The formation of NaP in the synthesis of Na–X zeolite was also reported by [39]. On the other hand, the X-ray diffraction pattern of the extracted HA exhibited wide bands at 2θ = 20.5° which is attributed to the planes of aliphatic carbon chains and at 2θ = 25.5°, attributed to the structure of condensed aromatic carbon rings (Fig. S2). The diffractogram also showed many sharp peaks which represent the crystalline structures of some minerals such as silicon oxides [4042].
EDX is a semi-quantitative technique, where the elemental composition in weight and atomic percentage is obtained along with the spectra and peak area for all sample components. EDX analysis of green-synthesized NaX indicating the presence of Si, Al, Na, C and O (Fig. S3; Table S1) with silicon ion and oxygen as the main elements and sodium as stabilizer cation for zeolite framework [27]. Typically, nano-zeolite NaX is synthesized with Si/Al ratio of 1 [43]. The Si/Al and Na/Al ratios in the current study are 1.16 and 0.08, respectively. Thus, the formation of zeolites NaX was confirmed by the EDX techniques [27, 44].
Likewise, the elemental distribution in the extracted HA was obtained by EDX, as shown in Fig. 4 and Table S1. The major elements observed were C, O, Si, Al, K, Ca and Na, with some other minor elements such as S, Cl and Mg [45]. The presence of Si in HA may be the possible site for the interaction with the metal ions. The combination use of EDX with SEM techniques provides an accurate spatial resolution that is commonly used for surface analysis. In the present study, the morphologies of the synthesized nano-zeolite NaX were characterized by SEM analysis, as presented in Fig. 5. Nano-zeolite NaX appeared as cubic crystals with average sizes of 34–41 nm (Fig. 5a). Additionally, SEM images of HA demonstrated that Pb2+ and Cd2+ are adsorbed on the surface by replacing aluminium ions (Fig. 5e, f) [33]. Previous reports also suggested that metal ions adsorb to the surface of clay minerals by replacing Ca and Mg ions by ion exchange mechanism [46].

3.2 Effects of operating parameters on the performance of the synthesized sorbent

3.2.1 Effect of the solution pH

The pH is a critical criterion for determining the adsorption capacity of materials. Metal ion solutions were prepared at different pH levels ranging from pH 2 to 7 [16]. Table 1 shows the change in Cd2+ and Pb2+ uptake by nano-zeolite NaX and HA at different pH levels. Clearly, the adsorption capacity of nano-zeolite NaX to metal ions increased with the pH of the solution. Whereas, in the case of HA, the adsorption capacity of Cd2+ and Pb2+ ions increased to its maximum when the initial pH of the solution raised from pH 2 to 4 after that the adsorption performance of HA decreased significantly. Although the strong polarization of heavy metal cations could increase the sorption ability, this force cannot change the charge density of the sorbents [47] and hence, the strong polarization could not explain the large adsorption capacity of the heavy metal cations. Normally, variation of pH leads to physisorption and chemisorption on the adsorbent surface [48]. Nano-zeolite NaX has a great number of terminal silica tetrahedrons in their external surfaces and consistently a great number of negative silanol groups which enables it to have a substantial attraction with inorganic cationic contaminants [16, 49]. For nano-zeolite NaX, the lowest adsorption of all metals was obtained at the lowest pH values (Table 1). This may be due to the increase in the competition for adsorption sites by hydrogen (H+) and the existence of aluminium (Al3+) ions which is dissolved from the aluminosilicate layers [50]. Natural zeolite preferentially adsorbs H+ ions from solution forming a positively charged surface [51], and thus in more acidic conditions more H+ ions are adsorbed from the solution. The increase in positive charge on the surface sites of zeolite causes electrostatic repulsion with the metal ions. The process can be described by the following reactions [52].
$$\mathrm{M}-\mathrm{OH}+{\mathrm{H}}^{+}\to \mathrm{M}-{\mathrm{OH}}_2^{+}$$
$$\mathrm{M}-\mathrm{OH}+{\mathrm{H}}^{+}\to \mathrm{M}-{\mathrm{O}}^{-}+{\mathrm{H}}_2\mathrm{O}$$
Table 1
The effect of pH on the % removal of Pb (II) and Cd (II) using different adsorbents
The capacity of metal sorption (mg g−1)
pH
Nano-zeolite NaX
Humic acid
Pb2+
Cd2+
Pb2+
Cd2+
2
2.33
3.30
21.96
3.24
3
3.71
3.83
23.72
7.66
4
5.49
4.33
22.78
15.04
5
7.21
5.38
6.09
0.21
6
8.34
6.46
9.24
3.11
7
8.45
6.91
4.33
1.93
Humic acids are composed of combinations of weak aliphatic carbon chains and aromatic carbon rings that are not soluble in acidic medium but are soluble in alkaline medium [52]. The pH value of the solution is often the most significant parameter affecting metal adsorption on the HA surface. Data showed that Cd2+ and Pb2+ adsorption capacities on HA peaked at pH 4 and 3, respectively, and subsequently began to drop (Table 1). Excess H+ compete with heavy metals for superficial sites of HA when the pH of the solution is lower than the optimum value [53]. Deprotonation of humic acid particles causes the surface to be negatively charged and increases the electrostatic attraction with the surrounding metal ions [11]. The deprotonation of HA surface sites at low pH can be described by the reaction: HA(OH) + H+ → HA(O-) + H2O

3.2.2 Effect of contact time

The contact time effect on the metal sorption capacity of the synthesized green sorbents was examined under the optimum buffering condition for each metal ion at different time intervals of 10, 20, 30, 40, 50 and 60 min. The results are illustrated in Fig. 6a. The data showed that nano-zeolite NaX exhibited the maximum sorption capacity of Cd2+ (8.13 mg/g), and Pb2+ (24.91 mg/g) upon reaching a contact time of 10 and 60 min, respectively. As for humic acid, the maximum adsorption capacities at 40 min were 12. mg/g and 23.11 mg/g for Cd2+ and Pb2+, respectively. The adsorption of Pb2+ on nano-zeolite proceeded in two stages (Fig. 6a). The first stage is characterized by a rapid adsorption behaviour due to the large available number of active sites on the sorbents and the possibility of the good orientation of the functional groups which enables reasonable interactions with the ions in solution. This stage is followed by a slower one due to sorbent surface saturation. Similar behaviour was discerned for the adsorption of Cd2+ on humic acid with a slight downward shift after 40 min probably due to repulsion interaction [37]. It is worthy to mention that the saturation of binding sites on nano-zeolite with Cd2+ was determined early, at 10 min of contact time. Achieving maximum efficiency of adsorption in a short contact time is an important factor in practical wastewater treatment systems which indicate less energy consumption in the shaking process and hence more economic feasibility [54].

3.2.3 Effect of the sorbents dosage

The amount of adsorbent is an important factor, which determines the sorption capacity at a given initial concentration of adsorbate [50]. The influence of nano-zeolite NaX and humic acid amount on the removal of Pb2+and Cd2+ ions from aqueous solutions for initial metal concentrations of 20 ppm, pH of 7, contact time of 30 min and with different adsorbent doses from 10–100 mg for Pb2+ and 10–200 mg for Cd2+ is shown in Fig. 6b. As shown, the adsorption capacities of Pb2+and Cd2+ ions were increased by increasing the eco-sorbents amounts. The maximum adsorption capacities were achieved using 10 mg and 200 mg of nano-zeolite NaX dose for Pb2+ and Cd2+ ions, respectively. However, in the case of humic acid, the maximum adsorption capacities were determined at 40 mg and 160 mg for Pb2+ and Cd+2 ions, respectively. This increase is due to the availability of more binding sites as the adsorbent dosage increased [55]. The observed decrease in Cd+2 adsorption with increasing humic acid dosage may be attributed to the possibility of particle aggregations, which causes a decrease in the total surface area of the sorbents and an increase in the divisional path length. Also, a large adsorbent amount effectively reduces the saturation of the adsorption sites per unit mass, resulting in a reduction in capacity [37, 56]. The order of removal efficiency was Pb2+- nano-zeolite NaX (99.89%) > Pb2+- HA (97.12%) > Cd2+- HA (91.22%) > Cd2+- nano-zeolite NaX (44.80%). For both green sorbents, the removal of lead from aqueous solutions at any value of dosage was greater than that of cadmium. This may be attributed to the affinity values of each element toward the reactive sorbent material [10].

3.2.4 Effect of temperature

Variation of temperature greatly affects the adsorption of metal ions on sorbents. The thermal effect was studied over a temperature range of 10–60 °C under optimum experimental conditions. The data revealed that the adsorption capacities of Pb2+ on both sorbents and the adsorption of Cd2+ on humic acid decreased with the increase in temperature, indicating that the adsorption process in these cases was exothermic in nature. In contrast, the adsorption capacity increased with the elevated temperature in the case of Cd2+ on nano-zeolite NaX, indicating endothermic adsorption. Increased adsorption with temperature can also be ascribed to an increase in the number of active adsorption sites, which is caused by the thermal breakdown of internal bonds at the adsorbent's surface [57]. The effect of temperature is further discussed in the modelling section.

3.2.5 Effect of interfering ions

Ions that are co-existing in solutions with the determined heavy metals significantly interfere with their adsorption by competing on the active sites of the sorbents. The effect of selected ions, Na(I), K(I), Mg (II) and Ca (II) as main coexisting ions in seawater, on the metal sorption capacity of the synthesized sorbents, was studied under the current optimal experimental conditions of Cd2+ and Pb2+. As a representative example, the percentage removal of Cd2+ by the HA sorbent in the presence of interfering ions is displayed in Fig. S4. There is a clear decline in the adsorption capacity of nano-zeolite and HA in the presence of the selected interfering ions, as expected. This observation is pointing to the competitive adsorption mechanism on the binding sites which depends on the ionic size, solvation layers and effective nuclear charge of the ions in solution. Moreover, the type of the surface loaded, as well as the functional groups present on the green sorbents, may also affect this behaviour [58].

3.2.6 Effect of metal ions concentration

The influence of initial metal ion concentration on adsorption was explored in the concentration range of 5 to 100 ppm (Fig. 6c). The results showed a gradual increase in the adsorption capacity (qe) of sorbents with the metal ion in the studied concentration range. It is noticed that at an equal concentration of Cd2+ and Pb2+ in solution, 100 ppm, nano-zeolite NaX exhibited identical adsorption capacity toward both metals. This could be explained by the effect of the nanoscale sorbents with high surface energy in quick attainment of the maximum adsorption capacity. Nevertheless, dissimilar metal adsorption behaviour was observed on the surface of HA. Obviously, adsorption is independent of the initial concentration at low metal ions concentrations where the ratio of metal ions to available adsorption sites is small. However, the case was different when the metal concentration raised up (Fig. 6c). At this point, many factors may govern the driving force of metal ions toward the active sites on the adsorbents.

3.3 Adsorption modellings

The influence of metal ion concentration on the extent of adsorption of Cd2+ and Pb2+ by nano-zeolite NaX and humic acid was assessed by different isotherms. Also, kinetic and thermodynamic models were utilized to figure out the correlation between metal ions and the surface of the green sorbents.

3.3.1 Adsorption isotherm models

The adsorption results were evaluated by a number of linear isotherm models including, Langmuir, Freundlich, Tempkin, Brunauer-Emmett-Teller (BET) and Dubinin Radushkevich. For all isotherms, the correlation coefficients (R2) were used to index the extent of agreement between the experimental data and the isotherm model. Best fitting is achieved when the correlation coefficient equals ~ 1.
The Langmuir adsorption quantifies the development of a monolayer adsorbate on the surface of the adsorbents with an assumption that no further adsorption occurs after this layer [58, 59]. The equation of the model with definitions of all parameters is stated in Table 2. Langmuir constants KL (Eq. 2) was extracted from the plot of 1/qe versus 1/Ce (Fig.7a), where qe is the adsorption capacity at equilibrium (mg/g) and Ce is the concentration of ions at equilibrium (mg/L). According to the calculated qe, the order of adsorption appears to be Pb2+ > Cd2+ for the two green-sorbents (Table 3). This trend refers to differences in binding energies and the stability of the resulted metal complexes [37, 64]. Langmuir constant, KL, indicates the amount of the bonding energy. The multilayer adsorption isotherm for heterogeneous (rough and multisite) surfaces is expressed using the Freundlich model [65]. In Freundlich isotherm (Eq. 3; Table 2), the constants Kf and nf denote the adsorption efficiency and the surface dissimilarity, respectively. Kf values (Table 3) pointed to a nearly similar removal feature for the studied ions on the green sorbents with much less efficiency for the removal of Cd2+ by nano-zeolite NaX under the same experimental conditions. Also, the nf values for all the studied cases are greater than 1 which infers heterogeneity of the adsorbent surface and easy separation of ions from solution and hence capable adsorption process [24, 37].
Table 2
Equations of the linear adsorption isotherm models
Isotherm model
Adsorption equation
Eq. no.
Constants
References
Langmuir
  
qe: the adsorption capacity at equilibrium (mg/g), qmax: the maximum adsorption capacity at monolayer coverage (mg/g)
Ce: the concentration of ions at equilibrium (mg/L), KL: the intensity of adsorption (L/mg).
 
\(\frac{1}{q_e}=\frac{1}{q_{\mathrm{max}}}+\left(\frac{1}{q_{\mathrm{max}}{K}_L}\right)\left(\frac{1}{Ce}\right)\)
(2)
[59]
Freundlich
\(\ln {q}_e={lnK}_f+\frac{1}{n_f} Ln{C}_e\)
(3)
Kf: adsorption capacity and of the solid adsorbent (mg/g),
nf: adsorption intensity of the solid adsorbent
[60]
Tempkin
qe = BTlnAT + BTLnCe
(4)
AT: the equilibrium binding constant (L/min)
[61]
Brunauer-Emmett-Teller (BET)
\(\frac{\frac{C_e}{C_0}}{q_e\ \left({C}_0-{C}_e\ \right)}=\frac{1}{K_b{q}_m}+\frac{K_b-1}{K_b{q}_m}\left(\frac{C_e}{C_0}\right)\)
(5)
qe: an adsorbate amount (mg/g), Ce: adsorbate content (mg/L),
Kb: adsorption isotherm constant related to energy of surface interaction (L/mg).
[62]
Dubinin–Radushkevich (D-R)
\(Ln{q}_e={lnq}_D+{B}_D^2 RTLn\left(1+\frac{1}{C_e}\right)\)
(6)
Be: free energy of adsorption per mole of the sorbate as it migrates to the surface of the adsorbent from infinite distance in the solution,
qm: D–R isotherm constant related to the degree of sorbate sorption by the sorbent surface.
[63]
\({E}_D=\frac{1}{\sqrt{2\ {B}_D}}\)
(7)
Table 3
Calculated parameters of the isotherm model for the adsorption of Cd2+ and Pb2+ ions on nano-zeolite NaX and humic acid
Equilibrium models
Parameters
Nano-zeolite NaX
Humic acid
Pb2+
Cd2+
Pb2+
Cd2+
Langmuir
qmax (mg/g)
72.992
33.223
44.642
41.152
KL (L/mg)
0.230
0.029
0.306
0.093
R2
0.9823
0.9925
0.9715
0.9913
Freundlich
Kf (L/mg)
4.056
1.060
4.040
1.832
nf (g/L)
2.333
1.241
2.079
1.471
1/nf (L/g)
0.429
0.806
0.481
0.680
R2
0.9475
0.9863
0.9863
0.9862
Tempkin
At (L/mg)
4.333
0.680
3.208
0.816
Bt (KJ/mol)
18.067
3.774
19.787
10.390
R2
0.9461
0.9656
0.9623
0.9485
BET
Kb
24.500
6.156
21.995
5.704
qm (mg/g)
68.027
2.465
2.155
5.963
R2
0.9833
0.9948
0.9159
0.9767
Dubinin-Radushkevich (D-R)
qD (mg/g)
81.337
14.682
84.032
40.585
ED (kJ/mol)
4.237
3.174
4.174
3.077
R2
0.9958
0.9525
0.9808
0.9635
BD
0.0278
0.0496
0.028
0.0528
Moreover, the data were fitted to the linear Tempkin’s isotherm which accounts for the adsorbent–adsorbate interactions (Eq. 4; Table 2). AT in Tempkin’s equation is the equilibrium binding constant (L/min) that corresponds to the maximum binding energy [66]. The graphs of qe versus ln Ce showed the viability of Tempkin’s isotherm (Fig. S5). The experimental data were also fitted to Brunauer–Emmett–Teller’s isotherm (BET, Eq. 5 in Table 2) which is a model used for the multilayer adsorption assuming that a layer does not need to be completed before the next layer starts to seal up [67]. The calculated R2 values of adsorbents confirm the applicability of the BET isotherm. Dubinin–Radushkevich’s isotherm (D-R) can explain the physicochemical characteristics of the adsorption process. This model was applied to estimate the porosity characteristics of the sorbent and the apparent energy of adsorption. The liner form of this model is presented by Eq. 6 in Table 2. In addition, the apparent energy ED in kJ/mol of adsorption from the Dubinin–Radushkevich isotherm can be computed using Eq. 7. The ED is a parameter used to predict the type of adsorption. The physical adsorption is represented by ED in the range (1–8) kJ/mol [68], whereas the chemical ion-exchange process is indicated by ED > 8 kJ/mol [69]. The computed ED values (Table 3) reflected chemical adsorption on the investigated adsorbents.

3.3.2 Adsorption kinetic models

Kinetic models are used to examine adsorption efficiency, adsorption mechanism and rate control step that includes mass transfer of the process. Pseudo-first-order, second-order, intraparticle diffusion and film diffusion models were fitted to the data at different contact time (t) [70]. Lagergren proposed a pseudo-first-order kinetic model that was successfully applied to explain the kinetics of many adsorption systems. The integral form of the model (Eq. 8) is expressed in Table 4, where k1 and qe are the slope and intercept of the graph of ln(qe-qt) vs. (t). The kinetic parameters for pseudo-first and second-orders for nano-zeolite NaX and HA with Pb2+ and Cd2+ were calculated and presented in Table 5. The low correlation coefficient (R2) values for the pseudo-first-order model (Table 5) imply less linearity and that the model did not fit the whole range of contact time [73]. On the other hand, the pseudo-second-order kinetic model (Eq. 9; Table 4) exhibited superior linearity over the whole range of contact time with correlation coefficients values higher than 0.99 (Fig. 8 and Table 5). The comparison of the experimentally obtained qe and the estimated qe is the true assessment of the kinetic model’s validity [37]. It can be noticed that in the pseudo-second-order kinetic model, the calculated qe are in excellent agreement with the experimental values. Moreover, a trend was observed in the rate constants of the pseudo-second-order model, such that k2 generally decreased due to the occupation of active sites as the initial concentration increased. No trend was observed for the rate constant k1 [74]. Usually, the pseudo-second-order model provides information about the overall reaction kinetics. However, it does not give evidence about the rate-limiting step. Metal ions adsorption kinetics from a solution into a solid always constructed on the basis of three consecutive steps: (1) film diffusion (i.e. mass is transferred from bulk liquid phase through boundary film to solid external surface), (2) intraparticle diffusion (i.e. diffusion in the liquid contained in the pores and/or along the pore walls) and (3) mass action (i.e. physical adsorption and desorption between the adsorbate and active sites). Since the adsorption step is very rapid, it is assumed that it does not influence the overall kinetics. The overall rate of adsorption process will be controlled by either surface diffusion or intraparticle diffusion. In many cases, intraparticle diffusion can be considered much slower step. Weber and Morris (1963) had cited some researches to establish intraparticle diffusion model, that assumes that the metal ions are transported from the solution through an interface between the solution and the adsorbent which called film diffusion, followed by a rate-limiting intraparticle diffusion step which bring them into the pores of the particles in the adsorbent. This model is correlated between qt as a function of the square root of time t (Eq. 10; Table 4). The slope and intercept values are represented by kid and C. The constant C value gives an idea about the thickness of the boundary layer; and the larger the constant C value is, the greater is the boundary layer effect [75]. According to these data (Table S2), the constant C values are positive at the period of experiment, suggesting that surface diffusion has a large role in the rate-limiting step. Increasing the intercept values with increase metal ions concentration, due to agitation used in batch adsorption experiments, high enough to minimize external mass transfer limitations. Figure S6 pointed out the weakness of the linearity of the curve and all initial parts of different curves present an approximately linear behaviour, which influenced by the ratio between adsorbent mass and liquid volume. The weak linear plot did not pass through the origin, indicating that intraparticle diffusion kinetic model was not the only rate-limiting step in the adsorption mechanism. Adsorption kinetics may thus be controlled by liquid film diffusion and intraparticle diffusion simultaneously. The liquid film diffusion kinetic model can be expressed as shown in Eq. 11 (Table 4). For all adsorbents, the plot of –ln (1 – F) as a function of time (t) (Fig. S7) gave straight lines that did not pass through the origin. It is clearly known that several steps are involved in the sorption of sorbate by a sorbent. However, when the transport of the solute molecules from the liquid phase up to the solid phase takes place, the boundary plays the most significant role in the adsorption process.
Table 4
Kinetic and thermodynamic models
Model
Adsorption equation
Eq. no.
Constants
References
Kinetic model
  
qe and qt are the values of amount adsorbed per unit mass at equilibrium and at any time t.
k1: equilibrium rate constant for pseudo-first-order adsorption.
 
Pseudo-first-order kinetics
ln(qe − qt) = lnqe − k1t
(8)
[71]
Pseudo-second-order kinetics
\(\frac{t}{q_t}=\frac{1}{k_2{q}_e^2}+\frac{t}{q_e}\)
(9)
k2: equilibrium rate constant for pseudo second-order adsorption.
 
Intraparticle diffusion
qt = kid t0.5 + C
(10)
qt: is adsorption capacity at any time t,
kid: intraparticle diffusion rate constant (mg/g min1/2), C: film thickness.
[72]
Liquid film diffusion
Ln(1 − F) =  − Kfd (t)
(11)
F: function attainment of equilibrium which equals qt /qe
Kfd: liquid film diffusivity ((L/min)
Thermodynamic model
Kd = qe/Ce
(12)
T: absolute temperature (K),
R: universal gas constant (8.314 J mol−1 K−1)
Kd (L/g): thermodynamic distribution coefficient for the adsorption.
qe: the sorption capacity (mg g−1) at the equilibrium,
Ce: the equilibrium concentration (mg L−1)
[71]
Van’t Hoff equations
Gο =  − RTLnKd
(13)
\(Ln{K}_d=\frac{\Delta {S}^o}{R}-\frac{\Delta {H}^o}{\mathrm{RT}}\)
(14)
Table 5
Values of kinetic parameters for pseudo-first-order and pseudo-second-order models of Cd (II) and Pb (II) ions adsorption on humic acid and nano-zeolite NaX
Adsorbent
Metal ions
Cο (mg/L)
qe (exp.) (mg/g)
Pseudo first-order kinetics
Pseudo second order kinetics
k1
qe (calc.)
R2
k2
qe (calc.)
R2
Nano-zeolite NaX
Pb2+
5
1.247
0.069
6.082
0.847
0.965
1.263
0.999
10
2.498
0.064
37.136
0.626
4.332
2.501
1
15
3.748
0.115
4.996
0.874
1.417
3.761
1
20
4.978
0.083
78.023
0.531
6.300
4.98
1
25
6.237
0.035
271.972
0.648
28.551
6.238
1
30
7.471
0.031
776.114
0.733
89.646
7.468
1
Cd2+
5
0.543
0.063
21.266
0.681
3.2977
0.5471
0.999
10
1.139
0.022
27.791
0.905
1.5679
1.1477
0.999
15
1.711
0.061
16.81
0.602
2.4021
1.7159
1
20
2.315
0.062
15.735
0.754
2.2462
2.3202
1
25
2.921
0.051
28.05
0.801
3.2928
2.9248
1
30
3.527
0.049
16.79
0.814
1.7959
3.5373
1
Humic Acid
Pb2+
5
1.2161
0.0337
36.2703
0.8589
3.8692
1.2198
1
10
2.4565
0.034
13.4167
0.6889
1.3502
2.4649
1
15
3.6932
0.0332
11.1551
0.7558
1.1182
3.7064
1
20
4.9097
0.0332
16.9658
0.9386
1.7588
4.9188
1
25
6.1615
0.035
12.6619
0.7243
1.339
6.1728
1
30
7.4464
0.0364
10.2707
0.5682
1.0276
7.4571
1
Cd2+
5
0.6738
0.0561
21.6694
0.8165
2.6876
0.6789
1
10
1.4243
0.0652
9.1422
0.7822
1.345
1.4339
1
15
2.1386
0.053
13.7097
0.9158
1.6193
2.1492
1
20
2.8931
0.0598
17.3467
0.8598
2.3602
2.8994
1
25
3.6513
0.0515
20.4401
0.9205
2.4427
3.6576
1
30
4.3914
0.049
25.873
0.9393
3.2348
4.3956
1
As a result, it is concluded that the adsorption in the present contribution is a complex mechanism where both surface adsorption and intraparticle diffusion have a role in determining the rate.

3.3.3 Thermodynamics of adsorption

The temperature dependence of the adsorption of metal ions on adsorbent was predicted by estimating the thermodynamic parameters such as free energy (ΔG°), enthalpy (ΔH°) and entropy (ΔS°) using the Van’t Hoff equations (Table 4) [76, 77]. The thermodynamic distribution coefficient Kd for the adsorption was calculated from Eq. 12, Table 4, at different temperatures of 283, 293, 303, 313, 323 and 333 K. ∆G° was then calculated from Eq. 13. The values of ∆H° and ∆S° were calculated from the slope and intercept of the plot of ln Kd against 1/T (Fig. 9) using Eq. 14 [78]. The degree of the spontaneity of the adsorption process is indicated by the free energy change (ΔG°), where a greater negative value implies more energetically favourable adsorption (Table S3).
The ΔG° values are negative at all temperatures indicating that the adsorption process is spontaneous and signifying physisorption rather than chemisorption [79]. In the case of Pb2+ adsorption on nano-zeolite NaX, ΔG° becomes more negative as the temperature increase which demonstrates the increase in the adsorption capability of Pb2+ at higher temperatures and the feasibility of the adsorption process [80]. On the other hand, the positive ΔG° values for the Cd2+ adsorption on nano-zeolite Nax indicates a non-spontaneous adsorption process that required a small amount of energy and hence is favourable at higher temperatures [56]. The positive values of ΔH° reveal the endothermic nature of adsorption of Cd2+ on nano-zeolite in the studied temperature range. One reason for the positive enthalpy is that metal ions dissolve well in water and must lose a portion of their hydration sheaths before being adsorbed, where the endothermicity of ion de-solvation requires more energy than the exothermicity of ion attachment to the sorbent surface [81]. Nevertheless, the negative ΔH° values of metal ion’s adsorption on humic acid point to exothermic adsorption. Also, the positive values of ΔS° (Table S3) confirm the randomness at the solid/liquid interface during the adsorption process on nano-zeolite [82].
Based on the present measurements, the affinity of adsorption of Pb2+ on eco-sorbents is greater than that of Cd2+ as reported by previous studies [83]. Several reasons control adsorption processes including the type and distribution of active groups on the adsorbent, the way metal ions interact with the adsorbent and the metal ionic radii. Pb2+, with the largest ionic size (the Pauling ionic radii is 1.20 Å), has the maximum adsorption capacity. However, smaller metal ions as in the case of Cd2+ (the Pauling ionic radii is 0.97 Å) interact less with the adsorbent because they have a higher hydration enthalpy and hence energy is required to separate H2O molecules from the metal ions surface [84]. Also, the bigger the unhydrated ions, the lower the charge density, and because the charge is more distributed, the hydration water is less retained, resulting in a weaker metal ion-water phase interaction [85].
The comparison from the literature between the performance of nano-zeolite NaX and HA adsorbents with other clays adsorbents for cadmium and lead ions sorption is shown in Table 6.
Table 6
Comparison of adsorption capacity (mg g−1) of prepared sorbents
Sorbents
Pb2+
Cd2+
References
Nano-Zeolite NaX
7.12
1.12
Present study
Low silica nano-zeolite X
909.09*
[44]
Iranian natural zeolite
5.9
4.01
[16]
Fariman sepiolite
31.5
[50]
Semnan zeolite
25.3
Natural zeolites (pH=5-7)
6.5
1.2
[86]
HA
6.02
0.66
Present study
Humic acid–treated coconut (Cocos nucifera) Husk
66.06
47.28
[87]
Humic Acid
185.9**
[88]
Chitosan grafted poly acrylic acid bentonite composites
51.6
[89]
Chitosan immobilized on bentonite
26.39
[90]
Alginate–montmorillonite nanocomposite
244.7
[91, 92]
Cystene–montmorillonite nanocomposite
0.18
[93]
Attapulgite/poly (acrylic acid)
38
[94]
Date seed biochar
74.60***
[95]
*Time = 100 min. **C0 = 0–200 mg/l, pH = 2.9–3.1. ***PH = 6

3.4 Removal of Cd2+ and Pb2+ ions from environmental samples

The existence of heavy metal contamination in the Mediterranean Sea and its vicinities has been verified by several investigations [37, 58]. A multistage column packed with eco-sorbent was used to study the usability and application of the synthesized low cost and eco-friendly sorbents in water treatment (removal of Cd2+ and Pb2+). The physicochemical parameters such as pH, salinity, oxidizable organic matter (OOM), the major cations such as K+, Na+, Li+, Ca2+ and Mg2+ and nutrient salts of different water samples were measured and listed in Table 7. The measured parameters differed from one region to another but remained within the same range as in prior investigations [95, 96]. The Cd2+ and Pb2+ removal percentage from different water salinities are stated in Table 8. The percentage of Cd2+ removal using nano-zeolite NaX was in the range of 40.82–70.18 % in El-Mex Bay. The extraction reached 99.81% in the third run of El-Mahmoudia Canal samples. While the adsorption of Pb2+ on nano-zeolite NaX was 93.14% (Petroleum Co.) and 98.63% in the third run of extraction of Eastern harbor samples. The removal efficiency of humic acid on metals in the environmental samples was 36.36–99.40% and 92.62–98.93% for Cd2+ and Pb2+, respectively. The following is the order of percentage removal of Cd2+ by green sorbents: Freshwater > brackish water > saline water, suggesting that interfering ions have a significant impact on adsorption processes. Pb2+ removal percentages are nearly identical in varied water salinities. This referral to a competitive adsorption mechanism on binding sites that is influenced by the ionic size, solvation layers and effective nuclear charge of ions in solution. The identified percentage extraction of Cd2+ and Pb2+ by the two eco-sorbents from the four real water samples, based on three passage steps over multi-stages micro-column systems, provides good evidence for the potential applications of nano-zeolite NaX and humic acid as promising eco-sorbent for removal of heavy metal ions from real environmental samples.
Table 7
Physicochemical parameters, major constituents and nutrient salts of different water samples
 
Physico-chemical parameters
Major constituents (mg/L)
Nutrient salts μM
Water samples
S%
pH
OOM (mg/L)
Na+
K+
Li+
Ca2+
Mg2+
SO42- (g/L)
NO3-N
NO2-N
DIP
DSi
Petroleum company
127.26
8.14
48.96
1063
700
0.12
2004
14583
8.64
5.63
11.00
1.39
122.66
El-Mahmoudia canal
1.25
7.35
2.24
980
7800
0.10
100.20
212.7
0.23
40.91
0.80
1.97
30.38
El-Mex bay
13.92
7.18
11.84
8600
147
0.09
240.48
522.6
1.35
12.74
68.98
13.97
223.50
Eastern harbor
41.76
7.52
0.32
11200
356
0.09
440.88
1859.3
0.38
4.04
0.15
0.91
2.22
Table 8
Percentage removal of Cd2+ and Pb2+ by green sorbents from environmental samples in the presence of interfering ions
 
Cd 2+ (% removal)
Pb2+ (% removal)
Nano-zeolite NaX
Humic acid
Nano-zeolite NaX
Humic acid
Water samples
Run 1
Run 2
Run 3
Run 1
Run 2
Run 3
Run 1
Run 2
Run 3
Run 1
Run 2
Run 3
Petroleum company
80.08
81.02
81.21
90.67
91.62
98.80
87.10
87.90
93.14
97.29
98.23
98.27
El-Mahmoudia canal
62.23
92.42
99.81
90.72
95.91
99.40
91.00
96.88
97.71
92.62
93.89
98.93
El-Mex bay
40.82
67.76
70.18
48.76
58.23
75.39
95.72
96.52
96.99
96.17
96.33
98.31
Eastern harbor
53.86
56.15
58.76
36.36
47.42
75.84
97.19
98.08
98.63
94.77
95.09
97.69

4 Conclusion

Nano-zeolite NaX and humic acid are eco-adsorbents with versatile active sites for removing a lot of metal ions. To describe the retention behaviour of Pb2+ and Cd2+ onto the eco-sorbents under acidic and neutral conditions, several sorption batch experiments were conducted. Humic acid and nano-zeolite NaX were analyzed using XRD to verify the synthetic sorbents’ qualitative mineralogy, EDX to validate the elemental composition in weight and atomic percentage and SEM images to demonstrate substantial changes in surface shape following adsorption. Operating parameters such as pH, contact duration, adsorbent mass and metal concentrations were all used in the batch process. For Cd2+ and Pb2+ on humic acid, the maximum adsorption capacity was achieved in acidic pH, whereas the pH was modified to a neutral value to reach the peak adsorption on nano-zeolite NaX, Metal adsorption was quite fast, taking only a few minutes. The Langmuir and Freundlich isotherm models well match the equilibrium adsorption data, confirming the hetero adsorption and physical-adsorption of Pb2+ and Cd2+ onto humic acid and nano-zeolite NaX. The pseudo-second-order model was the best fit for the kinetic data. Surface adsorption and intra-particle diffusion models that govern the adsorption process were used to study the mechanism. Pb2+ and Cd2+ adsorption on humic acid and Pb2+ adsorption on nano-zeolite NaX was both practicable and spontaneous, while Cd2+ adsorption on nano-zeolite NaX was non-spontaneous according to Van't Hoff thermodynamic parameters. The adsorption was exothermic, whereas the Cd2+ adsorption process on nano-zeolite NaX was temperature-dependent. Eco-sorbents demonstrated perfect effectiveness in the application with high removal from various water types. The findings show that eco-sorbents may be used as an alternative to chemical adsorbents in treating wastewater containing Pb2+ and Cd2+ contaminants.

Declarations

Conflict of interest

The authors declare no conflict of interest.
Open AccessThis article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://​creativecommons.​org/​licenses/​by/​4.​0/​.

Publisher’s Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.
Anhänge

Supplementary Information

Literatur
7.
Zurück zum Zitat Medellin-Castillo NA, Padilla-Ortega E, Regules-Martínez MC, Leyva-Ramos R, Ocampo-Pérez R, Carranza-Alvarez C (2017) Single and competitive adsorption of Cd(II) and Pb(II) ions from aqueous solutions onto industrial chili seeds (C apsicum annuum ) waste. Sustain Environ Res 27(2):61–69. https://doi.org/10.1016/j.serj.2017.01.004CrossRef Medellin-Castillo NA, Padilla-Ortega E, Regules-Martínez MC, Leyva-Ramos R, Ocampo-Pérez R, Carranza-Alvarez C (2017) Single and competitive adsorption of Cd(II) and Pb(II) ions from aqueous solutions onto industrial chili seeds (C apsicum annuum ) waste. Sustain Environ Res 27(2):61–69. https://​doi.​org/​10.​1016/​j.​serj.​2017.​01.​004CrossRef
10.
22.
Zurück zum Zitat Trofimova, E.S., Zykova, M.V., Ligacheva, A.A., Sherstoboev, E.Y. Zhdanov,V.V. Belousov, M.V. Yusubov, M.S. Krivoshchekov, S.V. Danilets, M.G., Dygai. A.M. Influence of humic acids extracted from peat by different methods on functional activity of macrophages in vitro. Bull Exp Biol Med 162 (2017), https://doi.org/10.1007/s10517-017-3702-5. Trofimova, E.S., Zykova, M.V., Ligacheva, A.A., Sherstoboev, E.Y. Zhdanov,V.V. Belousov, M.V. Yusubov, M.S. Krivoshchekov, S.V. Danilets, M.G., Dygai. A.M. Influence of humic acids extracted from peat by different methods on functional activity of macrophages in vitro. Bull Exp Biol Med 162 (2017), https://​doi.​org/​10.​1007/​s10517-017-3702-5.
40.
43.
Zurück zum Zitat Treacy MMJ, Higgins JB (2007) Collection of simulated XRD powder patterns for zeolites fifth (5th) revised edition, Fourth Revised edn. Elsevier, Amsterdam, p 2001 Treacy MMJ, Higgins JB (2007) Collection of simulated XRD powder patterns for zeolites fifth (5th) revised edition, Fourth Revised edn. Elsevier, Amsterdam, p 2001
45.
Zurück zum Zitat Jamal A, Fatima N, Huang Z, Liaquat R, Ahmad B, Haider R, Ali MI, Shoukat T, Alothman ZA, Ouladsmane M (2021) Extraction and chemical characterization of humic acid from nitric acid treated lignite and bituminous coal samples. Sustainability 13:8969. https://doi.org/10.3390/su13168969CrossRef Jamal A, Fatima N, Huang Z, Liaquat R, Ahmad B, Haider R, Ali MI, Shoukat T, Alothman ZA, Ouladsmane M (2021) Extraction and chemical characterization of humic acid from nitric acid treated lignite and bituminous coal samples. Sustainability 13:8969. https://​doi.​org/​10.​3390/​su13168969CrossRef
50.
Zurück zum Zitat Sharifipour, F., Hojati, S., Landi, A. and Faz Cano, A. Kinetics and thermodynamics of lead adsorption from aqueous solutions onto Iranian sepiolite and zeolite. Int J Environ Res, 9(3) (2015)1001-1010, 10.22059/IJER.2015.988. Sharifipour, F., Hojati, S., Landi, A. and Faz Cano, A. Kinetics and thermodynamics of lead adsorption from aqueous solutions onto Iranian sepiolite and zeolite. Int J Environ Res, 9(3) (2015)1001-1010, 10.22059/IJER.2015.988.
56.
Zurück zum Zitat Taha AA, Shreadah MA, Hany FH, Ahmed AM (2017) Validity of Egyptian Na-montmorillonite for adsorption of Pb2+, Cd2+ and Ni2+ under acidic conditions: characterization, isotherm, kinetics, thermodynamics and application study. Asia Pac J Chem Eng. https://doi.org/10.1002/apj.2072 Taha AA, Shreadah MA, Hany FH, Ahmed AM (2017) Validity of Egyptian Na-montmorillonite for adsorption of Pb2+, Cd2+ and Ni2+ under acidic conditions: characterization, isotherm, kinetics, thermodynamics and application study. Asia Pac J Chem Eng. https://​doi.​org/​10.​1002/​apj.​2072
59.
Zurück zum Zitat Langmuir I (1916) The constitution and fundamental properties of solids and liquids. J Am Chem Soc 38:2221–2295CrossRef Langmuir I (1916) The constitution and fundamental properties of solids and liquids. J Am Chem Soc 38:2221–2295CrossRef
60.
Zurück zum Zitat Freundlich HMF (1906) Over the adsorption in solution. J Phys Chem 57:385–471 Freundlich HMF (1906) Over the adsorption in solution. J Phys Chem 57:385–471
61.
Zurück zum Zitat Tempkin MJ, Pyzhev V (1940) Kinetics of ammonia synthesis on promoted iron catalysts. Acta Physicochim URSS 12:217–222 Tempkin MJ, Pyzhev V (1940) Kinetics of ammonia synthesis on promoted iron catalysts. Acta Physicochim URSS 12:217–222
62.
Zurück zum Zitat Brunauer S, Emmett PH, Teller E (1938) Adsorption of gases in multimolecular layers. J Am Chem Soc 60:309–319CrossRef Brunauer S, Emmett PH, Teller E (1938) Adsorption of gases in multimolecular layers. J Am Chem Soc 60:309–319CrossRef
68.
Zurück zum Zitat Woodard A (2001) Industrial waste treatment handbook Woburn. Butterworth-Heinemann Inc., pp 260–261 Woodard A (2001) Industrial waste treatment handbook Woburn. Butterworth-Heinemann Inc., pp 260–261
70.
71.
Zurück zum Zitat Lagergren S (1898) Zurtheorie der sogenannten adsorption gelösterstoffe. Kungliga Svenska Vetenskapsa-kadem Lagergren S (1898) Zurtheorie der sogenannten adsorption gelösterstoffe. Kungliga Svenska Vetenskapsa-kadem
72.
Zurück zum Zitat Weber WJ, Morris JC (1963) Kinetics of adsorption on carbon from solution. J Sanit Eng Div 89:31–59CrossRef Weber WJ, Morris JC (1963) Kinetics of adsorption on carbon from solution. J Sanit Eng Div 89:31–59CrossRef
76.
Zurück zum Zitat El Nemr A, El-Sikaily A, Khaled A (2010) Modeling of adsorption isotherms of Methylene Blue onto rice husk activated carbon. Egypt J Aquat Res 36(3):403–425 El Nemr A, El-Sikaily A, Khaled A (2010) Modeling of adsorption isotherms of Methylene Blue onto rice husk activated carbon. Egypt J Aquat Res 36(3):403–425
79.
Zurück zum Zitat Dawodu FA, Akpomie GK, Ejikeme PCN (2012) Equilibrium, thermodynamic and kinetic studies on the adsorption of lead(II) from solution by “Agbani clay”. Res J Eng Sci 1:9–17 Dawodu FA, Akpomie GK, Ejikeme PCN (2012) Equilibrium, thermodynamic and kinetic studies on the adsorption of lead(II) from solution by “Agbani clay”. Res J Eng Sci 1:9–17
89.
Zurück zum Zitat Zou X, Pan J, Ou H, Wang X, Guan W, Li C, Yan Y, Duan Y (2011) Adsorptive removal of Cr(III) and Fe(III) from aqueous solution by chitosan/attapulgite composites: Equilibrium, thermodynamics and kinetics. Chem Eng J 167:112–121CrossRef Zou X, Pan J, Ou H, Wang X, Guan W, Li C, Yan Y, Duan Y (2011) Adsorptive removal of Cr(III) and Fe(III) from aqueous solution by chitosan/attapulgite composites: Equilibrium, thermodynamics and kinetics. Chem Eng J 167:112–121CrossRef
90.
Zurück zum Zitat Futalan CM, Kan CC, Dalida ML, Hsien KJ, Pascua C, Wan MW (2011) Comparative and competitive adsorption of copper, lead, and nickel using chitosan immobilized on bentonite. Carbohydr Polym 83:528–536CrossRef Futalan CM, Kan CC, Dalida ML, Hsien KJ, Pascua C, Wan MW (2011) Comparative and competitive adsorption of copper, lead, and nickel using chitosan immobilized on bentonite. Carbohydr Polym 83:528–536CrossRef
91.
Zurück zum Zitat Shawky HA (2011) Improvement of water quality using alginate/montmorillonite composite beads. J Appl Polym Sci 119:2371–2378CrossRef Shawky HA (2011) Improvement of water quality using alginate/montmorillonite composite beads. J Appl Polym Sci 119:2371–2378CrossRef
92.
Zurück zum Zitat Yadav VB, Gadi R, Kalra S (2019) Clay based nanocomposites for removal of heavy metals from water: A review. J Environ Manag 232:803–817CrossRef Yadav VB, Gadi R, Kalra S (2019) Clay based nanocomposites for removal of heavy metals from water: A review. J Environ Manag 232:803–817CrossRef
93.
Zurück zum Zitat Liu P, Jiang L, Zhu L, Wang A (2014) Novel approach for attapulgite/poly(acrylic acid) (ATP/PAA) nanocomposite microgels as selective adsorbent for Pb(II) Ion. React Funct Polym 741:72–80CrossRef Liu P, Jiang L, Zhu L, Wang A (2014) Novel approach for attapulgite/poly(acrylic acid) (ATP/PAA) nanocomposite microgels as selective adsorbent for Pb(II) Ion. React Funct Polym 741:72–80CrossRef
Metadaten
Titel
Humic acid and nano-zeolite NaX as low cost and eco-friendly adsorbents for removal of Pb (II) and Cd (II) from water: characterization, kinetics, isotherms and thermodynamic studies
verfasst von
Mamdouh S. Masoud
Alyaa A. Zidan
Gehan M. El Zokm
Rehab M. I. Elsamra
Mohamed A. Okbah
Publikationsdatum
06.04.2022
Verlag
Springer Berlin Heidelberg
Erschienen in
Biomass Conversion and Biorefinery / Ausgabe 3/2024
Print ISSN: 2190-6815
Elektronische ISSN: 2190-6823
DOI
https://doi.org/10.1007/s13399-022-02608-9

Weitere Artikel der Ausgabe 3/2024

Biomass Conversion and Biorefinery 3/2024 Zur Ausgabe