Open Access Article
Panagiota Bikaa,
Nadia Todorovaa,
Maria-Anna Gatoub,
Michael Pissasa,
Eamonn Devlina,
Elias Sakellisac,
Nikos Boukosa,
Nefeli Lagopatid,
Theopisti Lymperopouloue,
Lamprini-Areti Tsakanikae,
Evangelia A. Pavlatoub,
Vasileios K. Tzitzios
a and
Panagiotis Dallas
*af
aInstitute of Nanoscience and Nanotechnology, NCSR Demokritos, 15341, Athens, Greece. E-mail: pdallas@eie.gr; p.dallas@inn.demokritos.gr
bLaboratory of General Chemistry, School of Chemical Engineering, National Technical University of Athens, Zografou Campus, 15772, Athens, Greece
cDepartment of Physics, National and Kapodistrian University of Athens, Athens, Greece
dLaboratory of Biology, Department of Basic Medical Sciences, Medical School, Biomedical Research Foundation, Academy of Athens, National and Kapodistrian University of Athens, 11527 Athens, Greece
eLaboratory of Inorganic and Analytical Chemistry, School of Chemical Engineering, National Technical University of Athens, Polytechniopolis Zografou, Iroon Polytechniou 9, 15780 Athens, Greece
fTheoretical and Physical Chemistry Institute, National Hellenic Research Foundation, 11635 Athens, Greece
First published on 12th November 2025
In this work, we present three different pathways to render commercial melamine sponges, both magnetic and hydrophobic, thereby offering them the capacity to effectively and selectively remove crude oil and heavy metals from aqueous environments. The magnetic properties were endowed by the deposition of as-synthesized iron oxide nanoparticles, functionalized with hydrophobic oleic acid and oleylamine ligands, via a large-scale synthesis. Aiming for an even higher hydrophobicity, terminal vinyl groups were attached to the pristine and modified sponges through a sol–gel hydrolysis of VTES@SiO2. The magnetic materials were thoroughly characterized and evaluated for their efficiency in water purification applications, regarding the adsorption of crude oil and heavy metal pollutants. Each pathway of preparation was effectively practical for a different application. The initial coating of the sponges with a hydrophobic layer (Sp-h-m) enhanced the adsorption and the retention of iron oxide nanoparticles, resulting in materials with a maximum magnetization of 25 × 10−3 emu. This modified sponge also exhibited the highest removal of metal ions (As3+ and Hg2+) from aqueous solutions within 60 minutes, and its extraction efficiency was evaluated in systems with single metal ions (250 ppm) and in the removal of metal ions (As3+ and Hg2+) from aqueous solutions within the same time frame. It was evaluated for its extraction efficiency in systems with single metal ions (250 ppm) and under competitive conditions (500 ppm of toxic metals in total). The coating with a hydrophobic layer, following the deposition of the magnetic nanoparticles (Sp-m-h), led to further improvement of the sponge's hydrophobicity (from 129° to 140° water contact angle; WCA), excellent selectivity to crude oil, and water repellency. The magnetically modified sponges exhibited significant initial adsorption capacities (60–100 g g−1) and average adsorption capacities (35–65 g g−1) over 15 cycles. The high selectivity, adsorption/desorption efficiency (up to 99.8%), their adsorption capacity for As and Hg metal ions and their responses to external magnetic fields confirmed the suitability of the developed magnetic sponges for water purification systems. The versatility of the proposed modification route allows the preparation and optimization of specific magnetic sponges for targeted applications.
Environmental significanceMelamine sponge is an excellent absorbing material, non-toxic and low cost, with chemical and mechanical stability. Its 3D porous structure possesses functional groups, allowing its effective modification with other nanomaterials to obtain targeted properties, such as magnetic response and hydrophobicity. We propose versatile routes for the modification of the sponge, aiming to develop composite materials that deal with the pollutants of oil and heavy metal ions in aqueous environments. The modifiers, based on magnetic iron oxide and silica nanoparticles, are not only environmentally friendly by composition, without toxic halogen components, but they are also attached to the sponges' matrix, thus avoiding any secondary pollution. Through remote navigation, the magnetic hydrophobic modified sponges offer safe decontamination of water under extreme or hazardous conditions. |
The melamine sponge is receiving great attention because of its availability, high elasticity and capacity. It should be noted that in the literature the performance of the sponge materials is expressed by both “absorption” and “adsorption” capacity due to the processes of holding oil within their volume (pores and channels) and the adherence of the oil to the sponge surface, especially after modification. In the present work, the term “adsorption” capacity was adopted for the general discussion and for the presentation of the results to emphasize the performed surface modification. It is known that surface functional groups such as hydroxyl (–OH), amino (–NH2), and secondary amine (–NH) groups are important features, which can be engaged either in chemical reactions (e.g. silanization) or in the attachment of various nanoparticles to endow specific properties.16 For oil removal applications, a high selectivity to oil and simultaneous water repellency is targeted by altering the sponges' roughness and surface energy as well as a high stability over multiple sorption/desorption cycles without any secondary pollution of the treated water. Therefore, melamine sponges possessing inherent hydrophilicity and oleophilicity are being modified with hydrophobic components and functional nanoparticles to become highly selective for oil compounds as well as to respond to external stimuli (e.g., magnetic field, light) when appropriate. Beyond hydrophobicity, additional magnetic, photothermic, anti-icing, etc. properties have been pursued. Modifiers such as silanes,14,17 hydrophobic polymers,18,19 nanoparticles of titania and iron oxide,20–23 two-dimensional MXenes and carbon nanomaterials (graphene, reduced graphene oxide)6,24,25 as well as combinations of them26–28 have been investigated. Adsorption capacities ranging from 4 to 240 g g−1 (g oil per g sponge) and hydrophobicity between 120° and 169° WCA) have been achieved, depending on the selected agents and modification methods. Synthetic procedures, such as dip-coating,17,18,29,30 in situ growth of nanoparticles,31 and polymerization32,33 as well as solvothermal treatment34 and sol–gel,35 have been employed. Among those, sol–gel and dip-coating impregnation with or without thermal curing appeared to be the most frequently used16 due not only to the relatively high selectivity (140°–150°) and adsorption capacities (50–140 g g−1) achieved but also to the mild conditions and cost-effective, scalable processing.
Multifunctional magnetic superhydrophobic melamine sponges are being developed for oil–water separation processes to enhance the adsorption capacity and selectivity towards oil as well as the capacity to operate in hazardous conditions and respond remotely. Iron oxide nanoparticles are prominent candidates for these applications when combined with hydrophobic compounds and incorporated into the desired matrix. The stability of these composites is an additional parameter that needs to be addressed not only to ensure the reusability of the sponge but also to avoid side pollution of the water with hazardous compounds and nanoparticles. In the literature, numerous reports exist on magnetic (super)hydrophobic oil-adsorbing melamine sponges. For example, magnetic nanoparticles have been synthesized in one-pot solvothermal reaction of iron precursor and various hydrophobic substances, although some of them are F-containing silane (e.g., 1H,1H,2H,2H-heptadecafluorodecyltriethoxysilane).34 In that work, bonding between the magnetic nanoparticles and the melamine sponge was achieved using a silicon adhesive, reaching maximum adsorption capacities ranged from 39.8 g g−1 (pump oil) to 78.7 g g−1 (silicone oil). Following a similar approach, a magnetically responsive and superhydrophobic sponge was developed,29 where a ferroferric oxide nanoparticle/graphene oxide hybrid was solvothermally prepared and then hydrophobically modified with a perfluorinated silane. The composite was anchored onto a melamine sponge, resulting in a superhydrophobic (163° WCA)/superoleophilic (0° OCA) material that exhibits selective adsorption of various oils between 63 and 156 g g−1 with good recyclability. In a more environmentally friendly approach, magnetic Fe3O4 nanoparticles were coupled with selected non-toxic deep eutectic solvents (DES@Fe3O4)30 for a 101.7–127.3 g g−1 adsorption capacity of real crude oils. A different route has been suggested26 with a two-step modification route, where Fe3O4 and WS2 nanoparticles were first deposited on the melamine sponge, and then a polydopamine/n-octadecanethiol film was built to obtain a rough hydrophobic (156.8 WCA) structure with low surface energy. As a result, an adsorption capacity of 40–100 g g−1 and good reusability were obtained. Polydopamine has also been used27 in combination with 3-aminopropyltriethoxysilane to achieve a 158.3° WCA and an adsorption capacity of 43.2–80.7 g g−1. Thus, it is evident that the efficiency and the environmental appropriateness of the magnetic hydrophobic sponges are strongly dependent on (i) the interplay between the two parts, i.e. selected hydrophobic substance and magnetic nanoparticles, (ii) the in situ or subsequent modification of the nanoparticles, and (iii) the selected sequence of nanoparticle deposition and stabilization processes.
Apart from the extraction of oil, sponges are often suggested for the removal of heavy metal ions (Hg2+, Pb2+, As3+, Cd2+, etc.).36,37 A super-amphiphilic silicon sponge was reported to adsorb a high amount of Hg2+ from a water solution with concentrations between 63 and 6231 ppm. Nevertheless, an efficiency of 98% to 99% is achieved depending on the concentration. These results are related to the porosity of the sponge and the strong affinity between Hg2+ and the sulfur atoms in the network skeleton.14 A polyurethane sponge modified with reduced graphene oxide was also examined for both oil–water separation and the removal of hazardous metal ions (Cu2+) from wastewater.6 The PU@rGO sponge exhibited a ∼100% removal within 2 h at low Cu2+ concentrations (6 ppm). Most importantly, magnetic nanoparticles, when dispersed in water, can effectively remove Pb2+ and Cr6+ through electrostatic mechanisms, as reported by Wang et al.38
In the present work, a low-cost, commercially available melamine sponge was used for the preparation of ecologically friendly magnetic hydrophobic adsorbers for the removal of crude oil and heavy metal ions from water. A thermolytic one-step reaction was employed for the synthesis of hydrophobic iron oxide nanoparticles in organic solvents before the main dip-coating modification procedure of the sponges in the stable colloidal suspension Aiming for a greater water repellency and a high adsorption efficiency, an additional sol–gel silanization with hydrophobic vinyl ligands was applied before and after the deposition of the magnetic nanoparticles, thus altering the functional terminations on the sponge's surface. The prepared magnetic–hydrophobic melamine sponges were characterized and evaluated for both crude oil and heavy metal ion (Hg2+, As3+) removal from water. The influence of the attached magnetic (Fe3O4) and hydrophobic (VTES@SiO2) nanoparticles and their synergetic effects as well as the sequence of the modification processes of the modified sponges were evaluated on the crude oil and metal selectivity.
For the assessment of adsorption capacity and reusability, oil-rich systems composed of 60 g of water and 10 g of crude oil were prepared. In this case, a thick oil layer completely covering the water surface was formed. The sponges were added to the systems and remained under stirring for 30 min. Similarly, the sponges were removed, pressed to release the adsorbed oil and returned to the system. The sorption–desorption procedure was performed 15 times. For each cycle, the weight of the sponges before the adsorption (pressed) and after the adsorption (with the adsorbed oil) was recorded. The values were used to calculate the adsorption capacity (AC) using eqn (1) and (2):
![]() | (1) |
![]() | (2) |
The OA and OLE ligands employed in this work are widely used for the synthesis of colloidal nanoparticles.41 Oleylamine is used as a high-boiling-point solvent, thermolytic, and capping agent,39 while the addition of oleic acid provides higher stability to the final nanoparticles, specifically for non-polar solvents. The organic layer encapsulating the nanoparticles prevents their aggregation and offers them solubility in organic solvents through the growth of hydrophobic interactions between the aliphatic chains of OA and OLE. The magnetic behavior of the organophilic nanoparticles' dispersion was examined using a permanent magnet, as shown in Fig. S1, where their ferrofluid behavior is presented, confirming the successful synthesis.
The crystalline nature of the as-synthesized nanoparticles was evaluated by XRD analysis. This technique cannot distinguish between the oxide types of nanoparticles, specifically the formation of either magnetite (Fe3O4) or maghemite (γ-Fe2O3), as both have a cubic spinel structure. In Fig. 1a, we present the XRD patterns of the nanoparticles in comparison with the diffraction peaks of magnetite (JCPS: 96-900-5839) and maghemite (JCPS: 96-900-6317). The diffractogram reveals the reflections at 30.23° (202), 35.57° (311), 43.25° (400), 54.02° (422), 57.2° (511), and 62.93° (440). From the most intense reflection, the average crystallite size of the surface-modified nanoparticles was calculated using an XRD analysis software according to the Scherrer equation: D = Kλ/(β
cos
θ), where D is the crystallite size, K is the Scherrer constant, β is the full-width at half-maximum (FWHM), and θ is the Bragg diffraction angle.42 The calculated average crystallite size, as determined by the equation, is ∼8 nm for the iron oxide nanoparticles. In Fig. 1b, TEM images at a 100 nm and 50 nm scale present the various geometries of the synthesized nanoparticles. Their size fluctuates from 6 nm to 17 nm and a representative size distribution diagram can be seen in Fig. S2. The synthesis and the in situ functionalization of the iron oxide nanoparticles did not cause any aggregation in the suspension. Their shape is polyhedral, and it is a natural evolution of their nucleation and growth under the synthetic conditions used.43
It is well known that the ratio of Fe2+ and Fe3+ cations during the synthesis can influence the ratio of magnetite and maghemite phases in iron oxide nanoparticles.44,45 Mössbauer spectroscopy is a powerful tool for identifying the phase composition in as-synthesized nanoparticles. As magnetite is easily oxidized to maghemite with time, directly after their synthesis the nanoparticle solution was frozen to 125 K for the measurement. Thus, we expect minimal oxidation. The experimental spectrum along with theoretical simulations are plotted and presented in Fig. 1c. The sample can be fitted using three well-defined and one broadened sextet (Table 1). The spectrum, obtained at 125 K, has a characteristic asymmetry associated with magnetite, allowing the sextets to be assigned to magnetite A (one sextet) and B (three sextets) sites by their distinctive isomer shift values. The subspectral areas of the sextets yield an A
:
B area ratio of 1
:
1.85. This differs but only slightly from that expected from well-crystallised stoichiometric magnetite, 1
:
2. This observation combined with the broad range of the three subspectra associated with the B site indicates that the sample is composed of non-stoichiometric magnetite.
| Subspectrum | Site | Isomer shift (MM s−1) | Hyperfine field (kG) | Area (%) |
|---|---|---|---|---|
| 1 | A | 0.38(5) | 490(3) | 35(3) |
| 2 | B | 0.77(5) | 478(3) | 20(3) |
| 3 | B | 0.64(5) | 441(5) | 35(3) |
| 4 | B | 0.7(1) | 279(10) | 10(5) |
Regarding the chemical composition of the capped magnetic nanoparticles, FTIR analysis is provided in Fig. 1d. The intense vibrations at the high frequencies of 2956 cm−1 are assigned to the C–H mode, and those at 2920 cm−1 and 2852 cm−1 are ascribed to the symmetric and asymmetric –CH2– vibrations of the aliphatic chains of oleic acid and oleylamine.46 These peaks, in comparison with the pure oleic acid, are slightly shifted towards lower frequencies due to adsorption to the solid surface and the close packing of hydrocarbon molecules surrounding the nanoparticles.47 Meanwhile, the stretching vibration of the –N–H in oleylamine on the nanoparticles' surface is not evident at the expected wavenumber of 3322 cm−1 due to its low intensity. The adsorptions centered at 1558 cm−1 and 1457 cm−1 are attributed to the asymmetric and symmetric stretching vibrations of the bidentate –COO group, respectively.46 The presence of both carboxylate bands confirms the attachment of oleic acid on the surface of the nanoparticles. A low-intensity peak at 1296 cm−1 corresponds to the –C–O stretch, while the intense band of the characteristic –C
O asymmetric band of oleic acid is absent, verifying the bridged conjugation. At lower frequencies, the strong adsorption at 584 cm−1 is attributed to the Fe–O bond, which indicates the formation of an iron oxide network with oleate groups bridged on its surface.48
Afterwards, the NPs were incorporated into either a pristine sponge or one coated with a hydrophobic silane agent.49 FT-IR analysis was conducted to confirm the chemical modifications (Fig. 1e). The FTIR results reveal the characteristic adsorption bands of the pristine sponge at 809 cm−1 (bending vibration) and at 1546 cm−1 (C
N stretching vibration). The peaks at 1331 cm−1 and 1448 cm−1 correspond to the –C–H bending vibrations. The peaks at 1689 cm−1 are assigned to C
C mode, and at 2848 cm−1, 2916 cm−1, and 2958 cm−1 are attributed to the –CH2 symmetric asymmetric and C–H vibrations.49 In all magnetically modified sponges, the Fe–O vibrational mode appears at 562 cm−1. In contrast, in the hydrophobically modified sponges, vibrations assigned to the vinyl characteristic groups11 occur at 2980 cm−1 (–C–H), 3022 cm−1 (C–H), and 3061 cm−1 (
C–H). Additionally, others evolved at 1601 cm−1, indicative of the –CH
CH2 stretching, 1409 cm−1 for the Si–CH
CH2 bending, and 963 cm−1 for the C
C vibrations.49,50 The characteristic broad band of 1097–1043 cm−1 is further detected and attributed to the asymmetric stretching vibrations of the Si–O–Si siloxane bridge,51 while its symmetric stretching vibrations appear at 800 cm−1 and its out-of-plane deformations at 412 cm−1.52 The peak at 751 cm−1 is attributed to Si–CH3 bonding and Si–O–C,53 while C–H deformation modes at 1275 cm−1 and 1004 cm−1 are most likely related to remaining Si–O–CH3.53,54 Thus, both the iron oxide nanoparticles and the VTES@SiO2 have successfully modified the surface of the melamine sponges.
Before delving into the crystalline structure of the composite sponges, as mentioned in a previous work,55 the pristine Sp sponge exhibits a broad band around 24°, indicating its amorphous structure. Some additional peaks observed in its XRD pattern were related to the sample holder. The XRD patterns of the magnetic sponges are presented in Fig. 1e. The modified composites exhibit characteristic reflections of iron oxide nanoparticles at 30.23° (202), 35.57° (311), 43.25° (400), 54.02° (422), 57.2° (511), and 62.93° (440), which are observed in all samples. In the Sp-h-m and Sp-m-h samples, the additional diffraction peak at 9.08° (110) signals the coexistence of the magnetic nanoparticles with the a-phase of VTES@SiO2.
The 3D porous network of the sponge is successfully maintained in all the modified sponges, as evidenced by SEM analysis (Fig. 2). For the Sp-m sample (Fig. 2a), a thick layer at the branches of the sponge is observed, consisting of accumulated polyhedral nanoparticles, along with some agglomerated entities. Energy-dispersive X-ray spectroscopy detects the three elements, C, N, O, originating from the sponge and the ligands, in addition to Fe and O of the iron oxide nanoparticles.
![]() | ||
| Fig. 2 SEM images and the corresponding EDX spectra of the (a) Sp-m, (b) Sp-h-m and (c) Sp-m-h samples. | ||
In Fig. 2b, where Sp-h-m is depicted, the deposition of the iron oxide nanoparticles on the superhydrophobic sponge significantly increases the roughness of the fully covered sponge. Some compact assemblies are present at the nodes of the sponges' 3D network, and upon higher magnification, additional irregular and spherical cavities can be observed. These cavities are probably formed by the introduction of iron oxide nanoparticles and the evaporation of the dispersion solvent, offering various micro- and nanostructures on the three-dimensional sponge, which are crucial for their superhydrophobicity56 and beneficial to adsorption applications.57 The sample's EDX spectrum revealed all the previously identified elements, in addition to silicon, which was found in a higher percentage compared to iron.
For Sp-m-h (Fig. 2c), where VTES@SiO2 adsorption was performed after the deposition of the iron oxide nanoparticles, a thick smooth layer grew on the surface, which could hinder the access to the iron oxide nanoparticles, protecting them. This could be useful for certain applications. Its EDX spectrum showed all the elements of the modified sponges, with a prevalence of Si over Fe. The modifications performed on the sponges were successful, allowing for the maintenance of the open pores and elasticity of the initial 3D network, which is crucial for the adsorption processes. Moreover, their surface was homogeneously covered on the interior and the exterior. The high surface roughness achieved is attributed to the hydrophobic interface of some composite sponges, resulting from their low surface energy.34,58 These multi-hierarchical micro- and primarily nano-structures57 may significantly benefit environmental applications for the purification of aqueous environments.
The magnetically modified sponges were submitted to magnetization measurements at room temperature to evaluate their magnetic properties. Fig. 3 shows the hysteresis magnetization loops of a sample consisting only of nanoparticles and representative samples of sponges loaded with magnetic nanoparticles. The magnetite NPs exhibit a saturation magnetization of 60 emu g−1 and negligible coercivity and remanence, indicating superparamagnetic behavior.59 The saturation magnetization is higher than that reported in some other studies,60,61 but lower than the bulk value which is to be expected for nanoparticles.62,63
To evaluate the magnetic behavior of sponges loaded with magnetic nanoparticles quantitatively, the reference sponges (without magnetic nanoparticles) and the sample holder were first measured to estimate the background magnetization contribution. These measurements indicate that the background magnetization varied linearly with the applied magnetic field. As a result, an appropriate linear variation of magnetization due to the background was subtracted from the magnetic measurements of the samples containing magnetic nanoparticles. As one can see in Fig. 3b–d, the magnetization curves are typical for samples containing ferrimagnetic nanoparticles. This is experimental proof that magnetic nanoparticles are loaded into the sponges. Their saturation magnetization was found to be smaller than that of the as-synthesized nanoparticles due to mass dilution of the sponges, but nonetheless, they were still magnetically controllable (Fig. S3).
The contribution of the sample holder is subtracted as follows.
The saturation magnetic moment of the sponges shown in Fig. 3b–d at 10 kOe is estimated to be Sp-m, ≈11 × 10−3 emu, Sp-h-m, ≈25 × 10−3 emu, and Sp-m-h ≈11 × 10−3 emu. The small saturation mass magnetization (total mass of the sponge and magnetic nanoparticles) is reasonable because of the small paramagnetic contribution of the sponge.64,65 Although their magnetic responses vary, all magnetically modified sponges were submersed into the same volume of iron oxide nanoparticles in hexane. Different quantities may have been functionalized on the sponge in each sample, which is dependent on the preparation method of the modified sponges. Thus, the mass of the magnetic nanoparticles was determined by dividing the saturation magnetic moment of each sample by the saturation magnetization measured for the pure magnetic nanoparticles. The calculated quantity of iron oxide nanoparticles explains the differences in the magnetization of the sponges. For Sp-m and Sp-m-h, in which the deposition of the nanoparticles was the first modification step, ≈0.2 mg is attached, whereas at Sp-h-m, more sites were added owing to the VTES@SiO2, and thus a doubled quantity of ≈0.4 mg is attached.
The Sp-h-m sample exhibits the highest saturation magnetization among the presented composites. Thus, the nanoparticles on the sponge's surface evidenced in Fig. 2b are efficiently immobilized, owing to the oleophilic nature of the superhydrophobic sponge beneath. Interestingly, the saturation magnetizations of the modified sponges presented in this work are higher than those previously reported for sponges employed in similar applications.66
After the second cycle, the adsorbed and the desorbed quantities decreased and remained stable for the next 4 cycles. This decrease for the modified sponges could be explained by (i) the low quantity of oil remaining available on the water surface after the initial cycles, and (ii) their significant water repellency as shown by the contact angle measurements. For these samples, the total quantities Qads and Qdes were close to the 500 mg of oil added to the oil/water system. Importantly, after the desorption process, pure oil without any water was obtained due to their hydrophobic properties. In contrast, water was clearly observed for the pristine Sp (Fig. S4c and d), attributed to its natural hydrophilicity. During the first cycle, the Sp sample adsorbed oil too, which is not unusual given its oleophilic properties.16 At the same time, the level of water purification (Fig. S5) was the lowest, and oil was present on the water surface even after the 6th cycle, thus confirming its low selectivity. In contrast, the modified sponges removed the entire oil quantity added to the system during the first two cycles, and after the 6th cycle, the water was clear without oil traces (Fig. S5).
The ACm/m and ACm/V of the sponges for the 1st cycle (initial) and for all 15 cycles (average) are compared in Fig. 4b and c. In this experiment, all the modified samples exhibited relatively high adsorption capacities per unit mass and volume of the sponge. The ACm/m of the Sp-m was the highest, owing to the high surface area of the magnetite NPs for oil adsorption, and reaching an initial ACm/m of ∼100 mg mg−1 and average ACm/m of ∼65 mg mg−1, which are very close to the values for the pristine Sp (∼117 mg mg−1 and ∼66 mg mg−1, respectively). The sponges with additional hydrophobic modification, i.e. Sp-h-m and Sp-m-h, exhibited lower initial ACm/m (∼70 mg mg−1 and ∼60 mg mg−1) and average ACm/m of ∼35 mg mg−1.
All the modified sponges exhibited excellent gravimetric/volumetric AC. For the Sp-m sponge, the initial ACm/V was found to be ∼1300 mg cm−3 and the average ACm/V was ∼745 mg cm−3. The Sp-h-m and the Sp-m-h sponges also showed similar or higher AC than the pristine Sp, the values of which were ∼1175 mg cm−3 as initial and ∼657 mg cm−3 as average ACm/V. These findings suggest that modifying the melamine sponge with F3O4 through the selected route enhances the overall oil adsorption capacity of the sponge. The additional hydrophobic modification with VTES@SiO2 nanoparticles decreases the initial oil uptake but does not significantly affect the average oil adsorbance capacity within the frame of 15 cycles. The regeneration of the sponges by washing with hexane and ethanol and drying overnight resulted in a return to ∼70% of the initial AC and to ∼95% of the hydrophobicity (water contact angle; WCA). It should be noted that despite the large thickness of the oil layer on the water surface, the pristine sponge Sp exhibited water release after the 5th cycle, which became more prominent after the 9th cycle. Some water release was also observed for the sponges with final magnetic modification, Sp-m and Sp-h-m, after the 7th and the 12th cycle, respectively. Such a phenomenon was not observed for the sponge with final hydrophobic modification, Sp-m-h, which makes it highly effective in oil separation from the water surface. Importantly, the sponges with both magnetic and hydrophobic properties can be maneuvered at a distance on the oil/water surface in a direction defined by an external magnetic field (Fig. S7 and S8).
In addition to establishing the magnetic response at the sponges, sufficient hydrophobicity needed to be achieved and further retained through the repeated oil adsorption/desorption cycles and their purification. The wettability of the composite's surface was evaluated through contact angle analysis, and the measured values are presented in Fig. 5. When the droplet of deionized water was shed upon the melamine sponge, it was immediately absorbed and no contact angle could be measured (not shown). This is related to its superhydrophilicity and the presence of the abundant polar groups –NH2, N–H and –OH on the melamine's surface.67 After the modifications of the sponges, the contact angle increases, and the surfaces exhibit superhydrophobic properties. The droplet of deionized water remained stable upon the surface. Generally, SiO2 nanoparticles have a hydrophilic character in nature; however, when they are functionalized with a hydrophobic silane agent, they obtain superhydrophobic properties.11 Similarly, since the magnetic nanoparticles are coated with ligands consisting of large aliphatic chains, they offer a hydrophobic interface,68 which was expected as the dispersion is stable only in non-polar solvents. The average contact angle measured for the Sp-m sample is 129°. The contact angle of the hydrophobic and magnetic composite sponges is dependent on the sequence of their preparation (Sp-m-h and Sp-h-m). Both demonstrate higher contact angles than Sp-m, but lower than Sp-h. When the final step of the modification is the adsorption and silanization process (Sp-m-h), they have a CA of 140°, higher than of Sp-h-m with 132° CA, when the final step is the deposition of the magnetic nanoparticles. After their application in the oil removal, the sponges were rinsed with n-hexane in order to remove their adsorbed quantity. Their form was slightly changed after the manual desorption cycles, but their hydrophobic properties remained. In Fig. 5b, the contact angles of the sponges after the oil experiments are gathered. There was a slight decrease in the range of 5°–10°, indicating the preservation of superhydrophobicity for all modified sponges and their stability after repetitive adsorption/desorption cycles.
Overall, the recorded adsorption of ∼100 g g−1 and repeatability of ∼65 g g−1 for 15 cycles are among the highest published in the literature for different types of sponges.26,27,30,69,70 In our work, the performance is attributed to the synergy of the selected modifiers of the sponge's surface. Specifically, the silicon oxide mediator (SiO2) is readily attached on the sponge and is subsequently modified with the hydrophobic silane (VTES), thus guaranteeing high stability and hydrophobicity. In addition, the use of magnetic nanoparticles modified with large aliphatic chains allows their immobilization on the sponge's skeleton, providing magnetic response and hydrophobicity. Moreover, the deposition of VTES@SiO2 as a final coating onto Sp-m (see Fig. 2c) further stabilizes the nanoparticles and increases the water repellence, thus leading to the superior performance of Sp-m-h, which demonstrated the highest number of adsorption cycles without uptake of water.
The sorbed metal content (q, mg g−1) by the modified sponges (eqn (3)) and the removal percentage (%) (eqn (4)) of the examined metals from the aqueous solutions were defined as follows:
![]() | (3) |
![]() | (4) |
The sponges (Sp-h-m, Sp-m and Sp-m-h) were applied to adsorb metal ions. Fig. 6a indicates the experimental data for the adsorption of As3+ and Hg2+ using single metal ion aqueous solutions (the concentration of each metal ion was fixed at 250 ppm) on the examined sponges. Based on the acquired results, the Sp-h-m sponge exhibited the highest adsorption capacity for both heavy metals. More specifically, its adsorption capacity towards As3+ and Hg2+ was 20.5 and 20.3 mg g−1, respectively, which decreased to 19.9 (As3+) and 20.1 (Hg2+) mg g−1 for Sp-m-h and further to 19.4 (As3+) and 19.2 (Hg2+) mg g−1 in the case of the Sp-m sponge. It is possible that Sp-h-m demonstrates the highest adsorption capacity, as a synergetic effect exists between the magnetic and the hydrophobic modifications of the sponge. As seen from SEM, lots of cavities are present that can permit the diffusion of metal ions to both iron oxide nanoparticles and VTES@SiO2, which further enhances the accessible surface area for the adsorption. Moreover, its higher saturation magnetization among the modified sponges can influence the attraction and simultaneous extraction of metal ions from the aqueous solution. The Sp-m sponge previously presented the best oil removal due to the large surface of iron oxide nanoparticles, but in the metal removal application, it seems to have the lowest capacity. On the Sp-m-h sponge, VTES@SiO2 covered the NPs and blocked the available surface. Phase changes due to further oxidation of magnetite could have also altered the properties, in addition to the higher oxygen content. The interactions driven by the cooperation of these functional groups enrich the mechanism of adsorption, from physical adsorption and electrostatic interactions to hydrogen bonding and surface complexation.67,68
In addition to the singe metal ion experiment (non-competitive system) the Sp-h-m sponge, which indicated the most enhanced capacity, was submitted to an additional experiment. A mixed metal ion aqueous solution was prepared with 250 ppm concentration of each As3+ and Hg2+ ion (total 500 ppm of toxic metals). The results are compared and presented in Fig. 6b. The adsorption capacity of the Sp-h-m sponge towards As3+ and Hg2+ in the competitive conditions decreased in the mixed solution, in relation to the adsorption values for the single metal ion solutions. With these competitive conditions, the adsorption values were 18.9 and 18.4 mg g−1 against As3+ and Hg2+, respectively. Compared with the adsorption values of the single metal ion solutions, the adsorption capacity is slightly affected due to the competition of one metal ion with the other to occupy the modified sponge's sites, decreasing by about 8% for As3+ and by 9.4% for Hg2+. It is thus concluded that close to 500 ppm As3+ and Hg2+ together were adsorbed overall effectively. Iron oxide nanoparticles are excellent platforms for sorption of heavy metal cations, while both the pristine sponge and the silica layer can facilitate water purification processes. The efficiency of the magnetite crystals that are employed in our work is proven by the fact that the Sp-m sample that does not contain silica layers already adsorbs both Hg(II) and As(III). In previous studies, Faulconer et al. utilized iron oxide nanomaterials which achieved excellent Hg(II) removal efficiency. The proposed adsorption process involved a strong binding affinity of the iron oxide surface with divalent mercury ions through coordination with oxygen atoms.71 Arsenic removal poses a more complicated picture in the literature due to the stability of both As3+ and As5+ cations, two forms that may coexist due to redox reactions with the nanoparticles, as in the case of zero-valent iron NPs. In a comprehensive paper by Wei et al., it was revealed that the ionic strength of the solution had a negligible effect on the adsorption of both As3+ and As5+ in iron oxide nanoneedles.72 Furthermore, our work indicates that there is no competition between the two heavy metals in terms of adsorption capacity (see Fig. 6b). The latter, coupled with the fact the adsorption capacity is rather similar, indicates an adsorption mechanism through interaction with under-coordinated oxygen sites. The pH of the solution is acidic; hence we exclude the formation of hydroxide species with reduced solubility and removal of the heavy metals is due to adsorption by the modified sponges. By this straightforward preparation method of these multifunctional magnetic hydrophobic sponges compared to others,73,74 both ions present in the solution are adsorbed in high quantities in only 60 min. Typical regeneration processes involve mild treatment with 2 M NaOH solution75 for arsenic adsorption and electrolysis for mercury.76 To that end, even small pieces of the magnetically controllable sponge can efficiently lead to the purification of water systems to the levels required by the World Health Organization (0.01 ppm for As and 0.006 ppm for Hg) and the United States Environmental Protection Agency (0.01 ppm for As and 0.002 ppm for Hg). Notably, according to our cost and scalability evaluation (not presented here), the large-scale synthesis and commercialization of the developed magnetic hydrophobic sponges is totally feasible due to the low-cost raw materials and processing.
Supplementary information (SI): captures of the ferrofluid and the magnetic modified sponges with the application of an external magnetic field out and in an oil/water system; table of the data derived from the selectivity and oil capacitance experiments along with their photos and videos, size distribution diagram for the iron oxide nanoparticles and optical microscopy images for the Sp-h-m sample. See DOI: https://doi.org/10.1039/d5en00549c.
| This journal is © The Royal Society of Chemistry 2025 |