Lucas
Hynes
a,
Gonzalo
Montiel
bc,
Allison
Jones
a,
Donna
Riel
a,
Muna
Abdulaziz
a,
Federico
Viva
d,
Dario
Bonetta
a,
Andrew
Vreugdenhil
e and
Liliana
Trevani
*a
aFaculty of Science, Ontario Tech. University, Oshawa, Ontario, Canada. E-mail: liliana.trevani@uoit.ca
bInstituto Nacional de Tecnologia Industrial, Buenos Aires, Argentina
cUniversidad de San Martin, San Martin, Buenos Aires, Argentina
dComision Nacional de Energia Atomica, San Martin, Buenos Aires, Argentina
eChemistry Department, Trent University, Peterborough, Ontario, Canada
First published on 29th April 2020
In this work, a comparative study between three carbon materials has been carried out to investigate the impact of the micro/mesoporous structure of the carbon substrate on their adsorption capabilities. The study included two commercial carbons: Darco KB-G (AC), and Vulcan XC-72R (VC). The third carbon material was a mesoporous material (MC), with tailored micro/mesoporous structure and surface area obtained by carbonization of a resorcinol–formaldehyde (RF) polymer gel using both soft and hard template materials. Melamine was used as a model adsorbate in both acid and alkaline solutions. For all carbons, melamine adsorption was found to be pH dependent with higher adsorption from alkaline solutions than from acidic solutions. To the best of our knowledge, these are the first reported values for the adsorption of melamine to these carbon materials. Adsorption data obtained using the Langmuir model were compared with theoretical studies involving melamine as a building block in the self-assembly of molecular structures on carbon substrates, and analyzed using the results of several characterization studies carried out as part of this research work, some of which include nitrogen and CO2 adsorption isotherms, Raman spectroscopy, powder X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and transmission electron microscopy (TEM).
Even though the adsorption of melamine on polymeric substrates, such as organic–inorganic hybrid melamine imprinted materials12–14 and monolithic polymeric cartridges with specific recognition sites for the extraction of melamine15 have been investigated in the past, the studies on carbon are scarce despite the fact that activated carbons have been extensively used as adsorbent material for organic species in aqueous media.5,16,17 A better understanding of the adsorption of melamine on micro and mesoporous carbon materials could contribute to developing new materials with high adsorptive capacity for melamine. As adsorbents, recent studies have shown the potential of carbon blacks (CBs), with mesoporous and microporous structure for the adsorption of humid acids,3 mercury,6 carbofuran,7L-histidine,18 and radioactive materials.9 These studies have also included Vulcan XC 72R (Cabot), a type of carbon commonly used as catalyst support on fuel cell applications,19 and investigated as adsorbent material for phenol and 1-naphthol.20
In this work, we present a comparative study on the adsorption of melamine on three carbon materials with distinct surface areas, pore size distributions, and oxygen contents: Darco KB-G (AC), Vulcan XC-72R (VC), and a synthetic carbon with micro/mesoporous structure (MC). Special efforts were dedicated to characterizing the carbon substrates using gas adsorption experiments, Raman spectroscopy, XRD, XPS, TEM, and other techniques to get a thorough insight into the melamine adsorption process and differences between carbon materials. Theoretical monolayer loadings for the adsorbate molecule were calculated based on the size of the molecule and the surface area of the carbon, and compared with experimental and literature values for model carbon surfaces.21–24
Nitrogen and carbon dioxide adsorption experiments were carried out to determine the surface area and pore size distribution of the carbon materials using a Micrometrics TriStar II Plus system. Typically 0.05–0.1 g of sample was vacuum degassed for 24 hours at 110 °C prior to adsorption. Data analysis was performed with Micromeretics MicroActive software.
Surface functional groups were identified by Fourier transform infrared spectroscopy (FTIR) on a PerkinElmer Spectrum 100 FT-IR Spectrometer using a Horizontal Attenuated Total Reflectance (HATR) sampling accessory and a DTGS detector. X-ray photoelectron spectroscopy (XPS) measurements were obtained on a ThermoFisher ESCALAB 250Xi with monochromatic Al Kα radiation (1486.6 eV) and a spot size of 900 microns. The functional groups on the surface of AC and VC were also analyzed by Boehm titration.
Thermogravimetric analysis (TGA) was used to evaluate the extent of removal of the SiO2 hard template. These were carried out on an SDT Q600 Simultaneous TGA/DSC from TA Instruments under air (50 mL min−1) at a scan rate of 10 °C min−1 from 30 to 1000 °C.
To confirm the speciation diagram for melamine accurately describes the speciation in solution as a function of pH, mainly in acid media, the changes in the UV-visible absorption spectrum of melamine were also investigated. As shown in Fig. 2, the absorbance at λ = 234 nm increases as the pH decreases and remains constant in the pH region where MH+ is the predominant species in solution (Fig. S1, ESI†). Consequently, to ensure that the adsorption experiments were carried out with only one species in solution (M or MH+), the adsorption studies were performed in 0.01 M NaOH (pH ∼ 12) and 0.01 M HCl (pH ∼ 2–3) at 30 °C. The Lambert–Beer law is obeyed in the concentration ranged adopted in this study as shown in Fig. 2.
For the adsorption studies, ∼40 mg of carbon was suspended in 100 g of melamine solution (2 × 10−4 to 7 × 10−3 mol kg−1 of solution) and left to equilibrate at a constant temperature in an Excella E25 Incubator Shaker for at least 4 days. After equilibration, samples were filtered through a syringe filter (0.45 μm Teflon filters) to remove adsorbent. The concentration of melamine in the solutions before and after adsorption was determined using UV-visible spectroscopy (Cary 50 Bio UV-Visible Spectrometer, 300–190 nm, slow scan rate). Prior to acquiring the UV-visible spectrum, the pH of all solutions was adjusted to pH 3 to ensure that all of the melamine was present as a single species.
The melamine adsorption was calculated for each carbon material with eqn (1):
(1) |
The Langmuir28 and Freundlich29 models were both used to fit the experimental data to be able to compare with other studies. In both cases, Origin Pro 2018 software was used to determine the best fit and regression parameters. According to the Langmuir model, the equilibrium adsorbate uptake, qe, as a function of the equilibrium concentration of adsorbate in solution, Ce is given by eqn (2):
(2) |
(3) |
To confirm that the addition of carbon did not introduce a significant change in the pH of the solutions, control experiments under the same conditions but without melamine were performed in which 40 mg of carbon was mixed with 100 g of solution, with the pH tested before and after stirring for 24 hours in a closed container to minimize the contact with air. The carbon was shown to not change the pH of the solutions.
The texture of the carbon materials was investigated by gas adsorption/desorption experiments. The nitrogen adsorption isotherms for the VC and AC commercial carbons are compared with that obtained for the MC sample in Fig. 4. The presence of mesopores is evidenced by the hysteresis loop characteristic of capillary condensation at high pressure (type IV with H1 hysteresis, IUPAC classification isotherm).30 As expected, this feature is more pronounced in MC than in VC and AC, due to the fact that the synthesis was carried out in the presence of both a polyelectrolyte and a hard-template. BET surface areas were calculated as 1315 m2 g−1 for AC, 592 m2 g−1 for MC and 246 m2 g−1 for VC. A comparison with published data and available information for the commercial materials from the manufacturers indicates that the BET surface area for AC is lower than the value reported by the company for this material (1700 m2 g−1), but consistent with an independent determination performed in an alternative system. The value obtained for VC is in quite good agreement with the BET areas reported in other studies31,32 and by the manufacturer (∼240 m2 g−1).
Fig. 5a illustrates typical BJH incremental pore size distribution curves obtained from the N2 desorption isotherms at 77 K for the carbon materials, while the NLDFT incremental and cumulative pore size distribution plots for AC, MC, and VC from CO2 adsorption isotherms at 273 K are shown in Fig. 5b. The pore volumes obtained from the N2 desorption isotherms were 0.83 cm3 g−1 for AC, 2.17 cm3 g−1 for MC and 0.45 cm3 g−1 for VC, with maximum pore radii at 62 Å, 189 Å and 175 Å, respectively. In the case of MC, the distribution is bimodal, a narrow peak at ∼60 Å, and a broad distribution in the mesopore region (100 to 200 Å) that it is likely due to the large size of the SiO2 nanoparticle templates and the destabilization of the structure upon its dissolution. These results are in good agreement with the TEM images obtained for the MC and shown in Fig. 6, where a distribution of mesopores with pore sizes between 100 and 200 Å is clearly visible. The volume in micropores (size <5.59 Å) obtained using the NLDFT model for the adsorption of CO2 at 273 K for AC, MC, and VC, were found to be 0.020 cm3 g−1, 0.036 cm3 g−1, and 0.0039 cm3 g−1, while the total volume for pores with sizes less than ∼10 Å were 0.075 cm3 g−1, 0.083 cm3 g−1, and 0.014 cm3 g−1, respectively. The results show AC, the material with the larger surface area, has the larger content of micropores, while VC, and mainly MC have a more developed mesopore structure.
Typical Raman spectra are summarized in Fig. 7. The spectra do not show significant differences between the samples. The expected D and G bands of carbon at ∼1360 cm−1 and 1600 cm−1, respectively, can be clearly seen, where the D band corresponds to the disordered or defect content of the carbon and G bands corresponds to the graphitic, ordered component of the material. The ratio of the intensity of these bands, ID/IG, is a useful measure of the degree of graphitization,33,34 The lower the ID/IG ratio, the higher the content of sp2 hybridized carbon correlating to a more graphitic structure, and less sp3 hybridized carbon correlating to a more amorphous structure. Integration of the D and G bands after deconvolution allowed the calculation of the ID/IG ratios for the three carbon samples: 1.8 (AC), 1.6 (MC) and 1.7 (VC). Based on these results, AC should present the higher content of carbon defects, while MC and VC have a more graphitic structure.33 However, the differences are small, and it is difficult to make an assessment. Nevertheless, the higher ID/IG ratios for AC is in good agreement with the surface oxygen contents calculated from the XPS spectra: 13.0 atomic% for AC, 4.9 atomic% for MC, and 2.3 atomic% for VC. The Boehm titration method also used to quantify the surface functional groups on AC and VC are summarized in Table 1. AC, which contained 13 atomic% O according to XPS data, had by far the greatest number of acidic groups by Boehm titration, while VC, which had an O content of only 2.3 atomic%, contained only a negligible number of acidic groups. A similar study was not performed on MC because of the amount of carbon required for this determination.
Carbon | O content (atomic%) | Total acidic groups | Phenolic groups | Lactonic groups | Carboxylic groups |
---|---|---|---|---|---|
AC | 13.0 | 342 μmol | 158 μmol | 17 μmol | 169 μmol |
VC | 2.3 | Negligible content – unable to be determined with this method |
The X-ray diffraction patterns for VC, AC and MC are shown in the ESI† (Fig. S2). Similar bands are present in the three cases, but the lines corresponding to the 002, 100 or 101 and 110 planes of graphite are more pronounced in VC, indicating a more graphitic structure. The VC spectrum is also very similar to spectra for the same material reported in the literature.34
Fig. 8 Adsorption isotherms for melamine on AC, MC, and VC in 0.01 M NaOH (top) and 0.01 M HCl (bottom) aqueous solutions at 30 °C. |
Optimized structures for melamine on different adsorption sites on graphene obtained using density functional theory (DFT) have been recently reported by Quesne-Turin et al.22 The study showed the adsorption and 2D self-organization process are the result of intermolecular interaction, and molecule–surface interactions to a lesser extent. The authors reported the adsorption energy increases from the monomer to the 2D-network, and the total stabilization energy on the surface decreases in the same range. Based on these findings, the authors speculate that melamine molecules are adsorbed alone or through dimers before forming a hexagonal porous supramolecular network. These results explain why the adsorption in acid and alkaline media are significantly different. In acid media, the interaction between the protonated melamine molecules, and between melamine and the carbon ring, is likely less favorable, resulting in very low adsorption values. Even electrostatic interactions will be absent at low pH values because of the protonation of the oxygen functional groups on carbon.
On the other hand, in alkaline media, MC and VC show a higher melamine uptake at low concentrations (H type isotherm based on Giles categorization35). Although this cannot be used to definitively analyze the isotherms, it could indicate that adsorption on AC is less favorable at low concentrations, perhaps due to the microporous structure of this material.
The Langmuir and Freundlich model fitting parameters are summarized in Table 2. The obtained melamine uptake for each carbon material expressed in mg melamine per g C are 208.7, 135.0, and 56.4 for AC, MC, and VC respectively. Following the expected trend, the amount of melamine required for the formation of a monolayer on carbon per gram of sorbent decreases as the BET surface area of the carbon adsorbent decreases (1315 m2 g−1, 592 m2 g−1, and 222 m2 g−1 for AC, MC, and VC, respectively). Due to the significant difference in surface area between materials, it is not surprising that normalization by surface area significantly changes this trend (Fig. 9), with monolayer saturation values per m2 of adsorbent material in alkaline media equal to 0.15, 0.23, and 0.23 mg melamine per m2 C, for AC, MC, and VC, respectively. AC, with a significantly higher oxygen content and larger micropores is able to adsorb less melamine than the other studied carbon materials. In terms of the affinity of each type of carbon for melamine, the higher the Kads value in the Langmuir model, the higher the affinity of that particular carbon for melamine. These Kads values are 232.6, 1682.9 and 1497.6 kJ mol−1 for AC, MC and VC, respectively. The results match the shape of the isotherms, as MC and VC both have very large Kads values, indicating a higher affinity for melamine than AC. The adsorption constant values were used to calculate the Gibbs energy of adsorption (ΔGads) using eqn (4).
ΔGads = −RTln(Kads) | (4) |
Carbon | Adsorption in 0.1 M HCl | Adsorption in 0.1 M NaOH | ||||
---|---|---|---|---|---|---|
Langmuir model | ||||||
Q 0 (mg per g C) | K ads (kg mol−1) | r 2 | Q 0 (mg per g C) | K ads (kg mol−1) | r 2 | |
AC | 71.2 ± 11.3 | 200.8 ± 56.4 | 0.94 | 208.7 ± 21.5 | 232.6 ± 44.9 | 0.97 |
MC | 20.2 ± 3.0 | 1268.3 ± 852.8 | 0.90 | 135.0 ± 4.0 | 1682.9 ± 201.1 | 0.98 |
VC | 39.9 ± 5.4 | 257.7 ± 68.4 | 0.94 | 56.4 ± 4.9 | 1497.6 ± 549.1 | 0.64 |
Carbon | Adsorption in 0.1 M HCl | Adsorption in 0.1 M NaOH | ||||
---|---|---|---|---|---|---|
Freundlich model | ||||||
K F (mg per g C) | n | r 2 | K F (mg per g C) | n | r 2 | |
AC | 924.2 ± 231.0 | 0.616 ± 0.045 | 0.96 | 2620.7 ± 439.7 | 0.596 ± 0.030 | 0.98 |
MC | 108.6 ± 79.0 | 0.350 ± 0.129 | 0.77 | 685.2 ± 96.8 | 0.321 ± 0.024 | 0.97 |
VC | 477.4 ± 146.8 | 0.578 ± 0.055 | 0.94 | 216.5 ± 45.5 | 0.273 ± 0.036 | 0.86 |
Fig. 9 Adsorption isotherms for melamine on AC, MC and VC at 30 °C, normalized by carbon BET surface area. |
(5) |
As shown in Table 4, the calculated monolayer values obtained by adopting the simplifying assumption that melamine forms a densely packed adlayer on carbon, that high-resolution STM images showed is not the case, and are significantly higher than those obtained in this study. Clearly, the assumption that the carbon surface is a continuous graphite plane is not a sufficiently sophisticated model for high surface area carbons, highlighting the need for accurate experimental adsorption data. In acid media, a similar comparison cannot be made because melamine will be protonated (MH+) and it is impossible to make any further analysis. Despite the limitations, the results would indicate that the adsorption of melamine on mesoporous carbons such as MC and VC is more favorable than in AC, perhaps due to the size of the solvated melamine molecule.
Carbon | Surface area (m2 g−1) | Langmuir monolayer (in NaOH) (mg g−1) | Calculated melamine monolayer (eqn (5)) (mg g−1) | % difference between exp. and calc. values |
---|---|---|---|---|
AC | 1315 | 209 | 526 | 40 |
MC | 592 | 135 | 237 | 57 |
VC | 222 | 56 | 89 | 63 |
When comparing the adsorption isotherms for the three carbons in alkaline media, it was verified that the Langmuir monolayer adsorption follows the trend of surface area; AC has the largest surface area, and a melamine uptake of 209 mg g−1, followed by MC with 135 mg g−1 and finally VC with 56 mg g−1. This trend is as expected, as materials with a greater surface area usually contain more adsorption sites, increasing the amount of material that can be adsorbed. However, the surface normalized adsorption values showed that VC and MC have a greater affinity for melamine, but only MC is likely to find application as an adsorbent due to its significantly higher surface area. The MC solids produced for this study has a high surface area and porous structure, with one of the reasons for this being that the silica hard template is not removed until after the carbonization procedure to prevent the pores from collapsing.
This work seeks to serve as the first steps in the study of the adsorption of melamine on carbon materials with the expectation that the reported findings can be extended to other carbon materials such as graphene or CNTs, as a stepping stone toward surface modification for applications in the fields of self-assembled and electrochemical materials.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/d0ma00097c |
This journal is © The Royal Society of Chemistry 2020 |