Open Access Article
Dina Hajjar*a,
M. Mohery
*b and
Gamal M. A. Mahran
c
aDepartment of Biochemistry, College of Science, University of Jeddah, Jeddah 80327, Saudi Arabia. E-mail: dhajjar@uj.edu.sa
bDepartment of Physical Sciences, College of Science, University of Jeddah, Jeddah, Saudi Arabia. E-mail: mmohery@uj.edu.sa
cMining Engineering Dept., Faculty of Engineering, King Abdulaziz University, Jeddah 21589, Saudi Arabia
First published on 15th October 2025
We present a new-generation, amine-decorated cellulose adsorbent (CDAM) engineered to overcome the inherent limitations of unmodified cellulose in the sequestration of heavy metals. By covalently grafting branched amine functionalities onto a crosslinked cellulose network, CDAM introduces a dense array of nitrogen-donor sites and enhanced porosity, as confirmed by FTIR (appearance of –NH– peaks), XPS (shifts in N 1s indicating quaternary and imine nitrogen), TGA (improved thermal stability), 1H/13C-NMR (chemical shifts consistent with successful grafting), and GC-MS (molecular fragmentation patterns). This tailored surface chemistry enables CDAM to achieve a record-high Cd2+ adsorption capacity of 483.7 mg g−1 under optimized conditions (pH 5.5, 30 min contact, 298 K), substantially outperforming benchmark bio-sorbents. Kinetic studies reveal a pseudo-second-order mechanism, indicative of chemisorption, while equilibrium data conform closely to the Langmuir isotherm, demonstrating monolayer coverage on a homogeneous set of active sites. Thermodynamic analysis (ΔH° = +10.4 kJ mol−1, ΔS° = +53 J mol−1 K−1, ΔG° < 0) confirms that Cd2+ uptake is endothermic, entropically driven, and spontaneous, with increased randomness at the solid–liquid interface due to desolvation effects. High-resolution XPS of Cd-loaded CDAM shows the emergence of Cd 3d peaks at 405.1 eV and 411.9 eV alongside shifted N 1s binding energies, directly validating metal coordination to imine and amine nitrogen. CDAM exhibits excellent reusability, retaining over 90% of its initial capacity after seven adsorption–desorption cycles with 0.25 M HCl elution, highlighting its practical viability. In a real-world demonstration, CDAM was deployed to treat acid leachates from cadmium-rich Wadi Um-Gheig rock samples. Quantitative Cd2+ recovery was achieved, and the purified metal ions were subsequently transformed into high-purity CdO nanoparticles. The resulting CdO exhibits a crystalline monoclinic structure (XRD), uniform nanorod morphology (TEM), and a surface area of 58.4 m2 g−1 (BET), illustrating the dual role of CDAM in environmental remediation and resource valorization. These findings position CDAM as a sustainable, high-performance platform for cadmium removal and value-added nanoparticle synthesis, with broad implications for water treatment and circular-economy strategies.
Its potential to bioaccumulate in trophic chains and its greater solubility in water compared to many other heavy metals make it particularly challenging to remove from contaminated environments. Therefore, developing environmentally benign, cost-effective, and efficient treatment technologies is essential.
Among existing remediation techniques, adsorption has gained prominence because of its simplicity, high effectiveness at low metal concentrations, cost-effectiveness, and scalability for industrial applications.9,10 Among the various remediation techniques available, adsorption has gained prominence due to its simplicity, high efficiency, cost-effectiveness, and potential for industrial scalability, even when metal concentrations are low.11–13 While effective, they often suffer from drawbacks such as high reagent consumption, sludge generation, membrane fouling, and high energy demands. In contrast, adsorption offers greater flexibility in adsorbent design and regeneration, making it particularly attractive for heavy metal removal.14–18
Many sorbent materials have been explored to remove cadmium from aqueous environments, including industrial byproducts such as red mud and fly ash, naturally occurring compounds like iron oxides, and advanced materials such as organoceramic composites. Bio-waste-derived adsorbents—such as those produced from rice husks, date pits, and other agricultural residues—have garnered considerable attention due to their cost-effectiveness and environmental sustainability.19–21 Despite these efforts, many of these materials present notable limitations: some display insufficient selectivity toward Cd2+ ions, while others suffer from relatively low adsorption capacities or poor reusability following desorption cycles, thereby limiting their practical application in long-term water treatment operations.22–25
In contrast, biopolymer-based sorbents, particularly those derived from cellulose and its chemical derivatives, have emerged as promising alternatives. Their natural abundance, biodegradability, and inherent capacity for chemical modification make them attractive candidates for developing customized adsorption systems. Introducing specific functional groups onto the cellulose backbone—such as amines, carboxyls, or phosphates—enhances their binding affinity toward heavy metal ions like Cd2+. Furthermore, the renewable nature of cellulose-based materials aligns well with the growing demand for green and sustainable technologies in water purification. As such, cellulose and its modified forms are increasingly recognized for their potential to address traditional adsorbents' drawbacks and provide scalable, efficient solutions for cadmium remediation.
Cellulose is the most abundant natural polymer on Earth, sourced from various organisms, including plants, marine life, algae, and bacteria. Chemically, it is a linear polymer comprised of D-glucose units connected by β(1 → 4)-linked D-glucose units. It is widely present in plant biomass, such as cotton, wood, and agricultural residues. Despite its abundance and biodegradable nature, raw cellulose exhibits limited effectiveness in adsorbing heavy metals due to its surface's lack of strong binding sites.26 Thus, the chemical modification of cellulose has become a vital strategy for enhancing its performance as an adsorbent material, particularly for removing heavy metal ions from solutions. Introducing functional groups like amine, carboxyl, thiol, and Schiff bases onto the cellulose backbone can significantly improve its affinity for various metal ions. These functional groups create active sites that facilitate stronger interactions with pollutants, enhancing adsorption efficiency. Additionally, cellulose materials that have been modified in this way demonstrate improved properties, such as greater water stability, increased mechanical strength, expanded surface area, and heightened chemical reactivity. These enhancements boost adsorption performance and position modified cellulose as a promising candidate for various environmental remediation applications, especially in industrial wastewater treatment.27–30 Modified cellulose materials also exhibit desirable characteristics, including improved water stability, mechanical strength, surface area, and reactivity, making them well-suited for environmental applications.
This innovative design exploits the synergistic effects of a crosslinked cellulose backbone for stability, Schiff base (C
N) moieties for strong chelation, and the introduction of branched, multi-dentate polyamine (triethylenetetramine) functionalities to create a high density of nitrogen-donor sites. This combination is engineered to overcome common limitations of earlier modifications, such as low site density or poor stability, resulting in a record-high adsorption capacity for Cd2+. Furthermore, this study extends beyond remediation by demonstrating a circular-economy approach: the subsequent valorization of the captured cadmium into high-purity CdO nanoparticles, a valuable functional material.
The present study introduces a novel cellulose-network-based adsorbent functionalized with Schiff base and branched amine moieties, creating a high-performance adsorbent for Cd(II) adsorption. This innovative design exploits the synergistic effects of the cellulose backbone, Schiff base coordination sites, and the introduction of quaternary amine functionalities to enhance metal-binding efficiency. The chemical and physical characteristics of the synthesized adsorbent were thoroughly analyzed using many tools such as FTIR, X-ray, XPS, 13C-NMR, H-NMR, and mass spectrometry to confirm successful functionalization. The adsorption behavior of the new material was systematically investigated against synthetic Cd(II) solutions and real environmental samples collected from the Wadi Um-Gheig region—a geologically significant area between El-Qusier and Ras Banas, known for its mineral-rich rock formations. Moreover, the adsorbed cadmium was successfully recovered and thermally converted into cadmium oxide (CdO) nanoparticles, with their structure, crystallinity, and morphology evaluated using XRD, scanning electron microscopy, and transmission electron microscopy. This dual-purpose strategy enables the effective remediation of toxic cadmium from natural samples. It offers a pathway for resource recovery and value-added nanoparticle synthesis, demonstrating environmental and technological innovation.
The modified cellulose CDAM was then obtained by reacting the cross-linked cellulose with triethylenetetramine in the presence of carbonyl diimidazole as a catalyst in dimethylformamide under reflux for 10 h at 80 °C. The obtained CDAM was then separated and rinsed many times with distilled water. Finally, the newly synthesized sorbent was dried by placing it in a laboratory oven set at 70 °C for four hours. This step was essential to remove residual moisture and stabilize the material, ensuring its structural integrity and readiness for subsequent characterization and adsorption experiments—Scheme 1.
The interaction mechanism is summarized as a flow. The reaction begins when carbonyl diimidazole (CDI) activates the cellulose hydroxyls by converting –OH groups into reactive imidazolyl carbonates, releasing imidazole as a byproduct. Next, the multi-amine molecule triethylenetetramine (TETA) performs a nucleophilic attack on the electrophilic carbonate carbon, displacing the imidazole and forming stable carbamate (–NH–CO–O–) linkages that tether TETA to the cellulose backbone. Because TETA contains several amine sites, this coupling can repeat at multiple points, introducing branching or crosslinking depending on the CDI-to-TETA ratio. The entire process is carried out in DMF at 80 °C for around 10 hours, ensuring efficient attachment of the amine groups to the polysaccharide network.
![]() | (1) |
In this equation, Cin and Ceq represent the initial and equilibrium concentrations of Cd2+ in the liquid phase (mg L−1), respectively. The equation was also utilized to calculate the adsorption capacity, qe (mg g−1), which indicates the amount of Cd2+ adsorbed per gram of CDAM sorbent at equilibrium conditions.
![]() | (2) |
amount of CDAM sorbent in grams, and V =
volume of solution in liters.
The desorption experiments were conducted under carefully controlled conditions to ensure accurate and reproducible results. These conditions were selected based on preliminary optimization studies to maximize desorption efficiency while preserving the structural integrity of the sorbent material. In these studies, 100 mL of a 250 ppm Cd2+ solution (initial pH 5.5) was placed in a conical flask, and 150 mg of adsorbent was added. The mixture was shaken at 150 rpm for 30 minutes at 298 K, then filtered through filter paper. The recovered adsorbent was thoroughly washed with HCl and ultrapure water until no Cd2+ was detected in the rinse solution, followed by drying in an oven. This dried adsorbent was then used for the next adsorption cycle. The solution concentration was adjusted to 1.5 mg mL−1 relative to the adsorbent amount, and the process was repeated for nine cycles. The durability efficiency of the improved cellulose was assessed by measuring the percentage of Cd2+ removal after each cycle. All adsorption tests were conducted in triplicate, with results reported as standard deviations. Experimental parameters remained consistent throughout all cycles: initial Cd2+ concentration of 250 mg L−1 at pH near 5.5, temperature at 298 K, dosage of CDAM is 0.08 g, and a contact time of 30 minutes.
N or C–N stretching from Schiff-base or crosslinking reactions. Additional new peaks at 1163 and 1058 cm−1 are assignable to C–N vibrations,35 and the band at 1732 cm−1 confirms the presence of carbonyl groups (C
O) introduced during the functionalization process.36 Together, these spectral changes unambiguously demonstrate that amine groups have been grafted onto the cellulose matrix without compromising its core structure, thus endowing the material with active sites for enhanced adsorption of pollutants or metal ions.
O and C–N bonds in the modified structure, which require more energy to break.38
![]() | ||
| Fig. 3 XPS results of CDAM: (A) before and after sorption (full spectra) and (B) after the adsorption of Cd 3d. | ||
High-resolution spectra provide further mechanistic insights:
(1) N 1s spectrum (Fig. 3a): The binding energy of the –NH2 group shifts from 399.8 eV (pristine CDAM) to 400.6 eV post-adsorption, indicative of nitrogen's coordination with Cd(II) through lone-pair electron donation.
(2) O 1s spectrum (Fig. 3a): The binding energy of the –CO–NH2 moiety increases from 532.4 eV to 533.3 eV after Cd(II) adsorption, corroborating the involvement of both N and O atoms in Cd(II) chelation.39
The Cd3d high-resolution spectrum (Fig. 3b) exhibits two characteristic peaks at 405.23 eV (Cd3d5/2) and 411.97 eV (Cd3d3/2), consistent with the formation of Cd-containing species such as –OCdOH, CdCO3, or Cd(OH)2. These results collectively validate the effective adsorption of Cd(II) onto CDAM, with chemical interactions driven by coordination with N/O functional groups and potential surface precipitation.40,41
, J = 5.17 Hz), 3.5–3.59 (t, 1H, –C
, J = 7.24 Hz), 3.76 (d, 2H, –C
2–O, J = 3.68 Hz), 1.42–1.77 (m, 2H, –C
2, J = 7.47 Hz), 5.66 (t, 2H, –C
2–O, J = 4.49 Hz), 2.59–3.16 (t, 2H, –C
2–NH2, J = 5.3 Hz), 3.11 (s, 2H, –CH2–N
2, J = 6.33 Hz), 5.76 (s, 1H, O
C–N
–, J = 5.38 Hz).1H NMR analysis, performed at 400 MHz using DMSO-d6 as the solvent, proved to be an effective tool for providing detailed information about the proton environment in the synthesized composite, thereby aiding in structural elucidation. Two primary assignments represent the pyranose ring and the cross-linked branch containing amine groups. In the pyranose moiety, obvious assignments appear at 4.21–5.99, 3.5–3.59, and 3.76 ppm, which are related to the hydroxy, methine, and methylene protons attached to the oxygen of the pyranose ring. Also, clear assignments of the cross-linked branch containing amine groups appear at 1.42–1.77, 5.66, and 2.59–3.16 ppm, which are related to different methylene protons located upon the cross-linked branch containing amine groups. Moreover, two distinct assignments were observed, which are associated with the amino groups with higher coupling constants. The amino group attached to the methylene moiety appeared as a singlet at 3.11 ppm (less de-shielded) with a coupling constant of 6.38 Hz. In comparison, the amino group (–NH) attached to the carbonyl moiety appeared at 5.76 ppm (more de-shielded) with a coupling constant of 5.38 Hz. Characterization of the CDAM adsorbent using 1H-NMR is illustrated in Fig. 4.
H–OH, J = 3.31 Hz), 75.14 (s, 1C, –
H–O–, J = 2.3 Hz), 66.42 (s, 1C, –CH–
H2–O–), 23–33.43 (s, 4C,
H2–), 99.4 (s, 1C, –
H–O–C
O), 39.69–51.8 (s, 6C, –
H2–NH–), 155.61 (s, 1C, –C
O).13C NMR analysis is a powerful tool for explaining the carbon structure of the synthesized composite. In the case of the synthesized CDAM adsorbent, the spectrum confirms two main structural features: the presence of the pyranose ring and the cross-linked branches containing amine groups. In the pyranose moiety, clear assignments appear at 70.99–95, 75.14, and 66.42 ppm, which are correlated to the methine attached to the hydroxyl group and the oxygen atom, respectively, besides a methylene carbon attached to the oxygen. Also, obvious assignments for the cross-linked branch containing amine groups appear at 23–33.43, 39.69–51.8, 99.4, and 155.61 ppm, which are related to different methylene carbon locations through the cross-linked branch containing amine groups. Moreover, two distinct assignments were observed, which are related to the methylene carbon connected to the –NH group and the methine carbon attached to the –O–C
O group. Finally, the more deshielded carbonyl carbon was detected at a chemical shift of 155.61 ppm. The structural specification of the synthesized CDAM adsorbent, as determined by 13C NMR analysis, is shown in Fig. 5.
Important fragments corresponding to the synthesized composite were identified. A quasi-molecular ion peak at m/z 816, attributed to [C32H64N8O16]˙ with a relative abundance of 17%, represents the molecular formula of the composite. Moreover, other fragments which were pointed to the successful synthesis of the composite such as pyranose ring [C6H11O6]˙ with m/z of 179 and relative abundance of 71, and [H2O]˙ with m/z of 18 and relative abundance of 34, [CO2]˙ with m/z of 44 and relative abundance of 28, and [NH3]˙ with m/z of 17 and relative abundance of 12.
Additionally, other fragments were associated to polyamine moiety were observed such as [CH5N]˙ with m/z of 31 and relative abundance of 18 (methyl amine), [C2H7N]˙ with m/z of 45 and relative abundance of 39 (ethyl amine), [C4H13N3]˙ with m/z of 103 and relative abundance of 13 (diethyl triamine), [C6H18N4]˙ with m/z of 146 and relative abundance of 18 (triethyl tetramine) and [C7H18N4O]˙ with m/z of 190 and relative abundance of 23 (triethyl amine carboxamide). The explanation of the fragmentation pattern of the CDAM adsorbent via MALDI-TOF/MS is shown in Fig. 6.
![]() | ||
| Fig. 7 (A) Impact of pH on the sorption of Cd+2 using CDAM, (B) pHZPC of the CDAM adsorbent. Cd+2 concentration, 250 mg L; CDAM dose 0.1 g; equilibration time, 20 min, and temperature, 30 °C. | ||
At low pH values (1–3), the high concentration of (H3O+) in the solution results in intense competition with Cd2+ ions for available sorption sites. Simultaneously, the surface of the CDAM material is highly protonated under acidic conditions, resulting in a positively charged surface that repels the cationic Cd2+ species. Consequently, the adsorption efficiency remains limited in this pH range.
As the pH increases from 3 to approximately 5.5, the concentration of H3O+ diminishes, reducing competition and decreasing surface protonation. This transition enhances the electrostatic attraction between the adsorbent surface and Cd2+ ions. The significant enhancement in sorption observed in this region is further explained by the material's point of zero charge (pHZPC), which was experimentally determined to be about 4.6, as shown in Fig. 7b. Below this pH, the surface bears a net positive charge. In contrast, above pHZPC, the surface becomes increasingly hostile. Thus, at pH values above 4.6, the negatively charged surface favors the uptake of Cd2+ ions via electrostatic interactions and possibly through complexation with surface functional groups such as carboxylate, phosphate, or hydroxyl moieties.
The maximum adsorption efficiency observed near pH 5.5 corresponds to conditions where the surface of the adsorbent is sufficiently deprotonated and negatively charged, and cadmium ions predominantly exist as free Cd2+ species in solution. Above this pH, the gradual decline in adsorption capacity is likely due to the formation and precipitation of cadmium hydroxide as the concentration of hydroxide ions (OH−) increases. This precipitation removes Cd2+ ions from the solution through a process other than adsorption, consequently lowering the observed adsorption capacity.
The best pH range for Cd(II) adsorption onto CDAM lies above the material's pHpzc, where the surface charge transitions to negative and electrostatic attraction toward Cd2+ ions is maximized. These findings underscore the significant influence of the adsorbent's surface characteristics and the chemical forms of metal ions in solution on adsorption. Consequently, optimizing the pH is essential for successfully using this material to remove heavy metals from its solutions.
Beyond the 0.06 g threshold, however, further increases in sorbent mass yielded only marginal gains in cadmium uptake, indicating that the system had approached site-saturation under the given conditions. At this plateau, nearly all accessible functional groups on the CDAM surface were occupied, and additional material contributed redundant binding capacity. Based on these observations, 0.06 g per 100 mL (equivalent to 0.6 g L−1) was selected as the optimal dose for subsequent kinetic, isotherm, and thermodynamic experiments, balancing maximal removal performance with material efficiency. This optimized dose ensures that adsorption trials proceed under conditions where capacity is sufficiently exploited without excess sorbent material.
To elucidate the underlying mechanism of Cd2+ sorption, the kinetic data were investigated using four widely applied models: pseudo-first-order (PFO), pseudo-second-order (PSO), intra-particle diffusion (IPD), and Elovich. The kinetic parameters obtained from the linearized forms of these models are summarized in Table 1.
| 1st ordered kinetic | 2nd order kinetic | ||||||
| K1 | q(max)cal | R2 | qmax(exp) | K2 | q(max)cal | R2 | |
| 0.0643 | 97.656 | 0.961 | 83.04 | 0.001 | 86.956 | 0.9938 | |
| Elovich model | Intra-particle diffusion rate model | ||||||
| α | β | R2 | Kd (mg g−1 min−0.5) | C | R2 | ||
| 0.0356 | 29.155 | 0.9875 | 1st stage | 0.0283 | 3.2194 | 0.9985 | |
| 2nd stage | 0.0648 | 0.5894 | 0.9961 | ||||
The PFO model,43 which operates under the assumption that the rate at which sorption sites are filled is proportional to the number of sites still available, did not accurately reflect the observed experimental results. As shown in Fig. 9b, the linear plot of log(qe − qt) versus time displayed a low correlation coefficient (R2), and the theoretical equilibrium adsorption capacity (qe, cal) calculated from this model significantly deviated from the experimentally observed value (qe, exp). These discrepancies suggest that the model was unsuitable for describing how Cd+2 was adsorbed onto CDAM, likely because the interaction between the adsorbent and the adsorbate involves chemical bonding.
![]() | (3) |
In contrast, the PSO model, which posits that the adsorption rate is limited by chemical bonding, closely matched the experimental data (Fig. 9c). The high R2 value and the strong agreement between the calculated (qecal) and experimental (qeexp) equilibrium adsorption capacities confirm the suitability of the PSO model. The rate constant (k2) and the sorption capacity derived from this model suggest that the adsorption process is mainly governed by the formation of chemical bonds, such as those between the amine groups on the CDAM and the Cd2+ ions, rather than by simple physical attraction.44,45
![]() | (4) |
To further probe the mass transfer mechanisms and rate-limiting steps, the intra-particle diffusion (IPD) model was employed.46 According to the Weber–Morris equation:
| qt = kdt0.5 + C | (5) |
The Elovich model, which describes chemisorption on heterogeneous surfaces, was also applied to assess the kinetics of surface interaction. The equation expresses the model:49
![]() | (6) |
In summary, the sorption kinetics of Cd2+ onto CDAM is best defined by the PSO model, confirming that chemisorption is the dominant mechanism, likely involving electron sharing or exchange between Cd2+ and amine functional groups on the sorbent surface. However, the intra-particle diffusion and Elovich models also provided meaningful insights. The IPD model demonstrated that film diffusion and pore diffusion play roles in the rate-limiting steps of the process. In contrast, the Elovich model confirmed that chemisorption occurs on a heterogeneous surface. Therefore, while the PSO model offers the most accurate prediction of kinetic behavior, the IPD and Elovich models complement the analysis by revealing the complexity of diffusion and surface interaction phenomena involved in cadmium uptake by CDAM. These findings provide a robust foundation for understanding the adsorption mechanism and optimizing the adsorbent for practical applications in wastewater treatment.
In summary, the sorption kinetics are best defined by the pseudo-second-order model. The high correlation coefficient (R2 > 0.99) and the agreement between theoretical and experimental qe values confirm that chemisorption is the dominant, rate-limiting mechanism. This likely involves electron sharing or exchange between the Cd2+ ions and the nitrogen-based functional groups (amines, imines) on the CDAM surface.
![]() | (7) |
The Freundlich isotherm model describes sorption on heterogeneous surfaces, where the energy of the binding sites varies across the surface (Fig. 10C).54,55 In contrast, the Temkin isotherm operates on the principle that the heat of adsorption decreases linearly as the surface becomes more covered due to interactions between the adsorbed molecules (Fig. 10D).56
![]() | (8) |
![]() | (9) |
In the Langmuir model, KL represents the association constant of the binding sites (L mg−1), while qm denotes the theoretical maximum monolayer sorption capacity (mg g−1). Ce is the equilibrium concentration of Cd+2 ions in solution (mg L−1), and qe indicates the equilibrium amount of Cd+2 adsorbed per unit mass of CDAM (mg g−1). For the Temkin model, AT is the binding constant (L g−1), R is the universal gas constant, T is the absolute temperature (K), and bT reflects the variation of adsorption heat (kJ mol−1).
The Dubinin–Radushkevich (D–R) isotherm is particularly effective in distinguishing between different adsorption mechanisms, such as ion exchange, physical adsorption, and chemisorption, by calculating the mean free energy of adsorption (E). eqn (10),57,58 Fig. 10E
Ln qe = ln qD − BD(ε)2
| (10) |
Additionally, eqn (11) defines the dimensionless separation factor RL, which helps predict the nature of adsorption based on the initial Cd2+ concentration Co. The parameter RL indicates the favorability of adsorption: Values greater than 1 (i.e., 1/n < 1) indicate favorable adsorption. Values between 0 and 1 (i.e., 1/n > 1) suggest unfavorable adsorption.59 The RL values of this study, ranging from 0.01 to 0.04 for CDAM adsorbents, confirm a highly favorable adsorption process.
![]() | (11) |
From Table 2, according to Langmuir factors controlling, the extreme theoretical sorption capacity of CDAM reached 476.19 mg g−1 at room temperature and pH 5.5. This closely matches the experimental results 483.67, reinforcing that the Langmuir model accurately defines the observed behaviour and that functional group interactions play a vital role in improving sorption capacity.
| Kinetic models | Parameters | |
|---|---|---|
| a The strong fit to the Langmuir isotherm (R2 = 0.9933) indicates that adsorption occurs through monolayer coverage onto a surface composed of homogeneous binding sites. The high theoretical maximum capacity (qm = 476.19 mg g−1), which aligns closely with the experimental value (483.67 mg g−1), underscores the efficacy of the functionalization strategy in creating a high density of uniform sites for Cd2+ sequestration. | ||
| Langmuir isotherm | Equation | y = 0.0021x + 0.0602 |
| qmax (mg g−1) | 476.19 | |
| Kl | 0.0349 | |
| R2 | 0. 9933 | |
| Freundlich isotherm | Equation | y = 0.2976x + 1.935 |
| Kf (mg g−1) | 86.099 | |
| 1/n (mg min g−1) | 0.2976 | |
| R2 | 0.9483 | |
| Tempkin isotherm | Equation | y = 74.2x + 0.9186 |
| AT (L m−1) | 1.0124 | |
| bT | 33.39 | |
| R2 | 0.9594 | |
| D–R isotherm | Equation | y = −0.0042x + 6.1716 |
| qD (mg g−1) | 478.95 | |
| BD (Mo12 kJ−2) | 0.0042 | |
| ED (kJ mo1−1) | 10.911 | |
| R2 | 0.9844 | |
| Practical capacity | qexp | 483.67 |
The Temkin model was also considered; it was effective in fitting the experimental data, likely due to its assumption of heterogeneous surfaces, and indicated by the higher correlation coefficient. In this model, the constant b, related to adsorption heat, was calculated as 33.39 kJ mol−1, while the binding energy (AT) was determined to be 1.0124 L g−1, as shown in Fig. 10E. Furthermore, the D–R isotherm model provided strong evidence for chemisorption, supported by the calculated mean adsorption energy E of 10.911 kJ mol−1 and a high correlation coefficient (R2 = 0.9844). These findings validate the suitability of the D–R model in describing the uptake of Cd+2 ions onto CDAM.
![]() | (12) |
| ΔG° = ΔH° − TΔS° | (13) |
The evaluated thermodynamic constants (ΔG°, ΔH°, and ΔS°) were systematically obtained using the relationships provided in eqn (12) and (13).
The thermodynamic isotherm study examined how temperature (ranging from 298 K to 343 K) affects the adsorption of Cd2+ by the CDAM adsorbent. Table 3 presents the characteristics of Cd2+ adsorption at these different temperatures. The results showed an apparent increase in sorption capacity as the temperature rose, from 483.70 mg g−1 to 536.30 mg g−1 (as detailed in Table 3), suggesting that the adsorption process is endothermic, meaning higher temperatures promote the uptake of Cd2+ ion.60,61
| ΔH° (kJ mol−1) | ΔS° (J mol−1 K−1) | ΔG° (kJ mol−1) | ||||
|---|---|---|---|---|---|---|
| 10.482 | 53.922 | 298 K | 313 K | 323 K | 333 K | 343 K |
| −16.058 | −16.867 | −17.406 | −17.946 | −18.485 | ||
The thermodynamic parameters summarized in Table 3 support the conclusion that chemical interactions are the primary mechanism governing the adsorption process. The enthalpy change (ΔH0) was 10.482 kJ mol−1, falling within the typical range for chemisorption.62 The Gibbs free energy values (ΔG0), ranging from −16.058 to −18.485 kJ mol−1, indicate a spontaneous process, further supporting the presence of chemical adsorption mechanisms.63 Additionally, the positive (ΔS0 = 53.922 J mol−1 K−1) reflects increased randomness at the solid–liquid interface, which enhances the interaction between the sorbent and Cd2+ ions.64
Overall, thermodynamic analysis confirms that CDAM exhibits strong chemisorption characteristics, as evidenced by the highest ΔH0 (10.482 kJ mol−1), most negative ΔG0 (−18.485 kJ mol−1), and largest ΔS0 (53.922 J mol−1 K−1). These indicate dominant chemical interactions, likely due to coordination with functional groups on the CDAM surface.
The thermodynamic parameters collectively confirm that the adsorption process is endothermic (ΔH° = +10.482 kJ mol−1), spontaneous (ΔG° < 0), and results in an increase in randomness at the solid–liquid interface (ΔS° = +53.922 J mol−1 K−1). The positive ΔH° value falls within the typical range for chemisorption, further supporting the mechanism proposed from the kinetic and isotherm analyses.
| Adsorbent | Adsorption capacity, mg g−1 | Ref. |
|---|---|---|
| Silicon and β-cyclodextrin (β-CD) comodified rice husk biochar (β-CD@SiBC) | 137.6 | 65 |
| Mg-modified biocar (Cys@MgBC) | 175.9–223.7 | 66 |
| Canna indica-derived biochar | 188.8 | 67 |
| Carboxymethyl cellulose/polyacrylamide | 256.4 | 68 |
| Ca–Mg phosphate based on dolomite | 241.7 | 69 |
| Methylisothiocyanate decorated PAMAM dendrimer/mesoporous silica | 97.8 | 70 |
| Chelating polyacrylonitrile | 146.0 | 71 |
| TiO2/glutaraldehyde/carboxymethyl cellulose | 274.28 | 72 |
| CTPO | 307 | 21 |
| Silicate-modified biochar derived from Sawdust | 178.58 | 73 |
| Ultrasonic magnesium-modified biochar (UMBC) | 138.5 | 74 |
| Montmorillonite microwave-assisted acid treatment (4MAT-Mt) | 388.32 | 75 |
| NTW | 42.35 | 76 |
| WHHC-CA1 | 166.6 | 77 |
| CDAM | 483.7 | This study |
Even when compared to high-capacity sorbents such as CTPO (307 mg g−1) and Mg-modified biocarbon (up to 223.7 mg g−1), CDAM demonstrates significantly superior efficiency. This exceptional performance of CDAM can be attributed to its abundant surface functionalities, especially amine groups, which provide strong coordination sites for Cd2+ ions. Furthermore, the sorbent's well-developed porous structure enhances mass transfer by offering a large surface area and accessible diffusion pathways, which together enable rapid and efficient metal uptake.
The remarkable enhancement in sorption capacity is directly attributed to the tailored surface functionalization strategy employed in the preparation of CDAM. This strategic modification not only increases the number of effective adsorption sites but also enhances the material's affinity for cadmium ions. Combined, these structural and chemical advantages position CDAM as a highly effective and competitive candidate for real-world wastewater treatment applications, particularly those targeting cadmium contamination in industrial effluents. The data strongly support CDAM's potential to surpass traditional sorbents in efficiency and practical utility. Even when compared to these recent and high-performing adsorbents, CDAM demonstrates a significantly superior adsorption capacity, nearly doubling the performance of many materials reported in the last two years.
| Single system | Adsorption% | Binary system | Adsorption% |
|---|---|---|---|
| Cd2+ | 96.8 | Cd2+ | 96.8 |
| Mg2+ | 1.0 | Cd2+ + Mg2+ | 96.6 |
| Fe3+ | 0.8 | Cd2+ + Fe3+ | 96.7 |
| Al3+ | 1.5 | Cd2+ + Al3+ | 96.3 |
| Cr3+ | 1.3 | Cd2+ + Cr3+ | 96.5 |
| V5+ | 1.1 | Cd2+ + V5+ | 96.1 |
| Mn2+ | 1.2 | Cd2+ + M2+ | 96.4 |
| Ni2+ | 13.6 | Cd2+ + Ni2+ | 88.4 |
| Cu2+ | 14.9 | Cd2+ + Cu2+ | 86.7 |
| Pb2+ | 3.5 | Cd2+ + Pb2+ | 93.2 |
| Zn2+ | 11.8 | Cd2+ + Zn2+ | 97.1 |
| Zr4+ | 0.5 | Cd2+ + Zr4+ | 96.3 |
| Ca2+ | 1.3 | Cd2+ + Ca2+ | 96.5 |
The data indicate that CDAM strongly prefers cadmium ions, achieving a removal efficiency of 96.8% when no competing ions are present. However, introducing additional metal species in binary systems revealed varying degrees of interference in Cd2+ sorption. While some ions, such as Mg2+, Fe3+, Al3+, Cr3+, V5+, Ca2+, and Zr4+, showed minimal impact, leading to only slight reductions in cadmium removal, other metals significantly affected performance. In particular, Cu2+, Ni2+, and Zn2+ demonstrated the most pronounced competitive interactions, decreasing Cd2+ adsorption to 86.7%, 88.4%, and 97.1%, respectively.
These more significant competitive effects likely stem from the physicochemical similarities of these ions to Cd2+, particularly in ionic radius, charge density, and coordination chemistry. For example, Ni2+ and Cu2+ have comparable hydrated radii and a strong tendency to form stable complexes with functional groups such as amines and carboxyls, which are abundant on the surface of CDAM. Such competition suggests these ions may directly compete with Cd2+ for the same active sites on the sorbent surface.
This behavior underscores the importance of evaluating adsorbent selectivity in the presence of multiple coexisting ions, as is typical in real-world wastewater treatment scenarios. Although CDAM maintains a high selectivity for cadmium under most conditions, a reduction in performance has been observed. The presence of certain metals highlights potential challenges that could arise in practical applications. These findings emphasize the need for further optimization or functional modification to enhance the specificity of CDAM in environments where competitive adsorption from chemically similar ions may occur.
N) and primary/secondary amine (–NH–) nitrogens—creating stable five- or six-membered chelate rings. Hydroxyl oxygen atoms on the cellulose backbone may also contribute weaker coordination interactions, further stabilizing the bound metal. This multi-dentate binding motif concentrates Cd2+ at the adsorbent interface and secures it through strong covalent-like bonds, yielding the high uptake capacity and rapid kinetics observed (Scheme 2).
Among the tested eluents, hydrochloric acid (HCl) demonstrated the most effective desorption performance for Cd+2 recovery from CDAM, as evidenced by the comparative data shown in Fig. 12A. The elution efficiency of HCl was further optimized by investigating its concentration range from 0.1 M to 2 M while maintaining constant desorption parameters. The results, depicted in Fig. 12B, indicate that 0.25 M HCl achieves the highest cadmium desorption efficiency, suggesting that higher acid concentrations do not necessarily enhance recovery and may instead compromise the adsorbent's structural integrity or functional activity.
Additionally, the influence of the desorption contact period was systematically considered over 30 to 180 minutes to determine the optimal recovery period. As illustrated in Fig. 12C, the Cd+2 recovery increased steadily with time and reached near-complete desorption at 30 minutes, beyond which no significant improvement was observed. These findings establish that a 30-minute desorption cycle using 0.25 M HCl at room temperature provides optimal conditions for effective regeneration of the CDAM adsorbent. In contrast, the CDAM material developed in this study demonstrated considerable ease in adsorption and regeneration operations. As shown in Fig. 12D, the batch reusability tests revealed that CDOP maintained nearly constant removal efficiency over the first 9 consecutive cycles. A slight decline to 82.4% was observed after the eighth cycle, which can be attributed to incomplete desorption or residual Cd2+ on the adsorbent surface.
| Element | Concentration, (%) | Element | Concentration (ppm) | |||||||||
|---|---|---|---|---|---|---|---|---|---|---|---|---|
| S1 | S2 | S1 | S2 | |||||||||
| SiO2 | 3.1 | 4.34 | Cd | 1100 | 1877 | |||||||
| Al2O3 | 1.3 | 0.86 | Ni | 27 | 44 | |||||||
| CaO | 28.4 | 8.01 | Ti | 240 | 180 | |||||||
| MgO | 4.2 | 0.56 | P | 390 | 416 | |||||||
| Fe2O3 | 12.1 | 4.07 | V | 23 | 16 | |||||||
| Na2O | 0.38 | 0.36 | Cu | 81 | 93 | |||||||
| K2O | 0.17 | 0.13 | Cr | 24 | 33 | |||||||
| ZnO | 20.1 | 37.5 | Li | 39 | 51 | |||||||
| MnO | 0.55 | 0.38 | Ag | 34 | 28 | |||||||
| PbO | 2.5 | 1.55 | B | 417 | 439 | |||||||
| L.O.I | 25.8 | 29.33 | ||||||||||
| Sample | pH | Ca2+ | Mg2+ | Na+ | K+ | Cl− | CO32- | SO42- | Th4+ | UO2+ | Cd2+ | Pb2+ |
| S3 | 3 | 2548 | 1035 | 34 812 |
1567 | 36 313 |
2045 | 2322 | 31 | 33.6 | 196.5 | 15.4 |
| S4 | 3.8 | 3381 | 2001 | 35 794 |
1318 | 36 424 |
2793 | 2938 | 461 | 21.5 | 101.3 | 22.3 |
It is important to mention herein that, in the Cd(II) sorption method from the samples S1 and S2, we observed a serious problem with the concentration of Zn2+ content, which is a serious problem in the adsorption process of Cd2+. Therefore, this content must be removed through a separation process, which can be summarized as follows: next to leaching with 40% H2SO4, the separation of zinc occurred using ammonium hydroxide by adjusting the pH to 8.5, after that CO2 gas was flowed through the leach liquor solution with stirring tell the Zinc precipitation was formed.78 The obtained zinc carbonate was converted to zinc oxide through thermal decomposition at 350 °C, which was sufficient for the complete decomposition of zinc carbonate for three hours. The obtained zinc oxide was confirmed using different tools, XRD and SEM.
Fig. 13 shows the XRD pattern of ZnO in nanoparticle form, recorded over a 2θ range of 10° to 80°. The presence of distinct peaks in the pattern confirms that the ZnO nanoparticles are semi-crystalline. The observed diffraction peaks at 2θ values of 31.9°, 34.56°, 36.42°, 47.62°, 56.8°, 63.04°, and 68.14° correspond to the characteristic planes of the wurtzite structure of ZnO, as referenced by the JCPDS data card No. 36–1451. These data are consistent with previously reported XRD patterns for ZnO nanoparticles.79,80
Energy-dispersive X-ray (EDX) analysis confirmed the high purity of the ZnO nanoparticles, with the elemental composition consisting primarily of zinc and oxygen. Scanning electron microscopy (SEM) images revealed a flaky morphology, characterized by flat and irregularly shaped nanocrystalline flakes.
500 mL of each sample were subjected to adsorption process using CDAM adsorbent using the best controlling factors reached (pH 5.5, 1.5 g CDAM dose, 45 min time at 70 °C) after that the pregnant CDAM with Cd2+ were separated and the Cd2+ was eluted using 0.25 M HCl for 30 min, finally the CDAM was separated and washed several time to be ready to used again.
The recovered cadmium solution was subjected to precipitation as a flow. First, the pH was adjusted to 8–9 using NH4OH solution added dropwise with continuous stirring to form Cd(OH)2 precipitate. After stirring for 30 minutes, allow the precipitate to settle for two to three hours, then filter using vacuum filtration with Whatman filter paper and wash thoroughly with deionized water until chloride-free (test with AgNO3). Dry the precipitate at 80–100 °C for 4–6 hours. The obtained precipitate is divided into two portions: one is converted directly to CdO product, and the other is converted to cadmium nanoparticles. The first portions are transferred to a ceramic crucible and calcined in a muffle furnace at 300–350 °C for 120–180 min. to convert them to CdO according to the reaction eqn (14):
| Cd(OH)2 → CdO + H2O | (14) |
The resulting brown to red-brown CdO powder should be cooled in a desiccator and kept in an airtight container. The obtained product was confirmed using XRD analysis and SEM.
Subsequently, cadmium sorption was carried out using CDAM under optimized conditions: pH 5.5, 0.06 g of CDAM, 45 min. of reaction period, and a temperature of 70 °C. The Cd(II)-loaded CDAM was then eluted using 0.25 M HCl. Finally, cadmium was precipitated using Na2S, and the resulting product was investigated by X-ray diffraction (XRD) analysis (Fig. 14).
The XRD pattern (Fig. 14) confirms the formation of the CdO phase with a lattice constant of 4.695 Å and a cubic crystal structure belonging to the space group Fm3m. The diffraction peaks observed at 2θ values of 32.90°, 38.20°, 55.20°, 65.80°, and 69.20° correspond to the (111), (200), (220), (311), and (222) planes, respectively, which match the standard pattern for cubic CdO (JCPDS card No. 05-0640). These results indicate the successful synthesis of crystalline CdO.
SEM images of the CdO sample reveal a distinctive cauliflower-like morphology, characterized by clustered and aggregated nanostructures.
The 2nd portion was then used to preparation of cadmium nano particle as follow, The obtained cadmium hydroxide was converted to cadmium nitrate, thoroughly wash the Cd(OH)2 precipitate with deionized water 3–4 times and filter using vacuum filtration with Whatman filter paper, then place the washed precipitate in a clean beaker and add 2 M HNO3 dropwise with continuous stirring until complete dissolution occurs according to the reaction, eqn (15):
| Cd(OH)2 + 2HNO3 → Cd(NO3)2 + 2H2O | (15) |
Using minimal acid for dissolution, filter the resulting solution via a 0.45 μm membrane filter to eliminate any undissolved particles. Then, heat the filtrate on a hot plate at 70–80 °C with stable stirring to evaporate excess water, obtaining a concentrated Cd(NO3)2 solution. Avoid complete drying. Finally, the solution should be allowed to cool to room temperature before further processing.81,82
To synthesize CdO nanoparticles, dissolve 0.12 M of the prepared cadmium nitrate tetrahydrate (Cd(NO3)2) in 150 mL of deionized water in a 250 mL flask with continuous stirring. Adjust the pH to 11 by adding 25% w/w ammonia solution dropwise, which causes the formation of cadmium hydroxide intermediate, and stir the solution for 12 hours to ensure complete reaction. Transfer the mixture to a 250 mL autoclave and heat at 80 °C for 12 hours for hydrothermal treatment, promoting controlled nucleation and uniform nanostructure growth. Collect the white precipitate (cadmium hydroxide precursor) by centrifugation, wash several times with distilled water to remove residual ions, and dry overnight at 60 °C. Finally, calcine the dried material at 600 °C for 3 hours in a furnace to convert the hydroxide precursor to brown CdO nanoparticles through thermal decomposition, yielding crystalline nanoparticles with controlled size distribution and high purity suitable for various applications in catalysis, sensing, and electronic devices.
The high-resolution transmission electron microscopy image (Fig. 15a) of CdO nanoparticles reveals particles with a predominantly cubic morphology. The presence of larger particles is attributed to the aggregation of smaller particles. A log-normal fit of the particle size distribution attained from the TEM image yielded an average particle diameter of approximately 22.7 nm. The corresponding histogram and average particle size are presented in Fig. 15b. This average particle size is consistent with that calculated from X-ray diffraction (XRD) data.
![]() | ||
| Fig. 15 (A) high-resolution transmission electron microscopy image, (B) average particle size (C) XRD analysis, (D)BET analysis, (E) SEM image, of CdO nanoparticles. | ||
Fig. 15 displays the powder XRD pattern of CdO nanoparticles, indicating a cubic crystal structure with unit cell parameters: a = b = c = 4.694 Å and α = β = γ = 90°. The average particle size, estimated using Scherrer's equation, was approximately 31 nm. The sharp and well-defined diffraction peaks confirm the crystalline nature of the CdO nanoparticles with random orientation.83 The diffraction peaks corresponding to the (111), (200), and (220) planes align well with the standard JCPDS card No. 75-0593, confirming the phase purity of CdO.
The surface area and pore structure of the CdO nanoparticles were evaluated by nitrogen physisorption analysis. The nitrogen adsorption–desorption isotherms (Fig. 15b) correspond to a type IV isotherm with an H3 hysteresis loop, as classified by IUPAC.84 This type of hysteresis suggests the presence of slit-like mesopores, consistent with the mesoporous nature of the CdO particles (Gregg & Sing, 1982). The Brunauer–Emmett–Teller (BET) surface area was measured at 58.4 m2 g−1, with a total pore volume of 0.059 cm3 g−1. The Barrett–Joyner–Halenda (BJH) desorption pore size distribution revealed an average pore size of 4.7 nm, confirming the mesoporous structure and indicating high surface contact potential.
The shape and morphology of the CdO nanoparticles were further examined using FE-SEM. As shown in Fig. 15, the CdO nanoparticles exhibit a polygonal structure with an almost uniform size distribution.
The synthesis of high-purity CdO nanoparticles from recovered waste streams demonstrates a successful transition from environmental pollutant to functional nanomaterial. The produced CdO, with its crystalline structure and morphology, is not a waste product but a valuable material with several potential applications. CdO nanoparticles are known for their use in optical devices due to their transparent conductivity and high carrier mobility, in gas sensors for their sensitivity to various gases, as a photocatalyst for the degradation of organic pollutants, and as an electrode material in batteries and solar cells.74,75 This valorization pathway significantly enhances the economic feasibility and environmental benefits of the adsorption process, aligning perfectly with the principles of a circular economy.
| This journal is © The Royal Society of Chemistry 2025 |