Wei Yangabc,
Xiaohui Xiac,
Xueying Liuc and
Shaoqiu Zhang*c
aKey Laboratory of Building Safety and Energy Efficiency of the Ministry of Education, College of Civil Engineering, Hunan University, Changsha 410082, China
bNational Center for International Research Collaboration in Building Safety and Environment, Hunan University, Changsha 410082, China
cCollege of Civil Engineering, Hunan University, Changsha 410082, China. E-mail: zhangshaoqiu@hnu.edu.cn
First published on 2nd May 2023
The intercalation of cetyltrimethylammonium bromide (CTMAB) into montmorillonite will cause interlayer expansion and surface charge reversal. In this study, CTMAB–Mt is prepared by adding CTMAB with different multiples of montmorillonite cation exchange capacity (CEC), and the intercalated CTMAB structural arrangement, as well as the dynamics behavior, are investigated by combining molecular dynamics (MD) simulation with experimental characterization. According to RDF analysis of MD simulations, the interaction between CTMA+ and the surface of montmorillonite is mostly electrostatic interaction and hydrogen bond production. At low loading (≤1.00CEC), the XRD profile exhibits a peak value corresponding to one type of intercalation structure and interlayer spacing, but at high loading (>1.00CEC), two peaks are visible, each of which has a fixed value but a varied strength, corresponding to the existence of two types of expanded structures. The d-spacing (d001) values obtained from MD simulations are quite close to XRD values when CTMAB loading is lower than 1.00CEC. Density distribution profiles obtained from MD analysis reveal that as loading increases, CTMA+ is arranged in the interlayer from a monolayer to a bilayer and then to a pseudo-trilayer. At high loadings (>1.00CEC), due to the fact that the excess loading leads to inhomogenous intercalation, XRD demonstrates the existence of two different arrangements: bilayer and pseudo-trilayer. The self-diffusion coefficients of MD simulations show that the dynamic behavior of CTMA+ is influenced by both the interlayer space and the electrostatic interaction of the montmorillonite clay. The abrupt rise in interlayer spacing increases mobility, whereas the increased interaction between alkyl chains decreases mobility.
Among all, organic modification has attracted considerable attention for eliminating organic and inorganic contaminants from wastewater.9 The interlayer cations (usually Na+, K+, and Ca2+) of clay minerals can be easily exchanged by surfactants or polymers to synthesize organoclays with hydrophobic and organophilic characteristics. By selecting suitable organic surfactants and clay minerals and controlling the loading of organic modifiers, materials can be prepared with desired properties, such as increased retention capacity for anionic and non-polar compounds.3,10–12 The adsorption capacity and surface properties of montmorillonite were greatly affected by different types of modifiers and the loading of modifiers. The exchange of interlayer cations leads to changes in the interlayer space dimension, surface charge and wettability, and swelling pressure. The modification not only changed the surface properties of the clay but also affected the arrangement of the clay mineral layers, the microstructure of the particles, and thus the characteristics of the pores. The addition of organic cations to the clay structure can significantly alter its physical and chemical properties (such as adsorption capacity, wettability, porosity, or diffusivity).13 Quaternary ammonium cations (QACs) are the most widely used organic surfactants for clay modification.14 Structural change is the essential factor that affects the properties of clay, such as the diffusion transport process, which is affected by the surface composition and properties of materials in the clay barrier with low hydraulic conductivity.15 Therefore, it is an important issue to study the interfacial interaction and structural arrangement of organic clay at the nanoscale in the application of organic clay to environmental protection.
Due to the fine particle size and complex chemical composition of clays, it is difficult to obtain detailed information on the interlayer structure and the atomic local environment of organoclays from experimental measurements. Most of the experimental studies on organoclay have focused on its properties and structures, especially the changes in the interlayer structure. Advanced microscopy techniques such as scanning electron microscopy (SEM) and atomic force microscopy (AFM) are used for the detailed morphological characterization of organoclays.16 X-ray diffraction measurements have shown that the interlayer microstructure is affected by both the charge density of the clay layers and the alkyl chain length of the alkylammonium ions.17,18 Based on an all-trans configuration of surfactant after intercalation, monolayer, bilayer, pseudo-trilayer, and paraffin-monolayer and paraffin-bilayer19–21 are unrealistic. Vaia et al.20 showed on the basis of Fourier transform infrared spectroscopy (FTIR) that alkyl chains can vary from liquid like phase to solid like phase, while constrained amine chains are in varying degrees of ordered conformations. Furthermore, 13C magic-angle spinning nuclear magnetic resonance (13C MAS NMR) demonstrates that there is an extensive coexistence of gauche and trans conformers and that the complete liquidlike state does not exist.22 So the results of microscopic characterization only provide an average state of molecular arrangement and interlayer structure and cannot accurately describe the microscopic mechanism of interlayer structure and the interaction between the surfactant and the clay surface.
In recent years, molecular simulations have been widely used in the study of organoclays to explore detailed information on molecular arrangements and layering structures. The simulation results not only provide complementary evidence to the experimental results but also provide new insights into the microstructure of organoclays, especially thermodynamic and dynamic information at the atomic level.23–25 In addition to the monolayer, bilayer, and pseudo-trilayer structures proposed by the structural model, a pseudo-tetralayer structure was observed by MD simulations in the double alkyl chain modified organoclays,26 which were not discussed in previous experimental results. This means that the real structure of the organoclay is probably a lot more complicated than the model that was made based on measurements.
In this paper, cetyltrimethylammonium bromide (CTMAB) is selected as one of the most easily available QACs surfactants to synthesize organoclay, and the intercalated CTMAB structural arrangement and dynamics behavior are investigated. Characterization analysis and MD simulations will be carried out to investigate the modification mechanism, interlayer structure arrangement, and diffusion characteristics of interlayer species. MD simulations consider the effects of different loading on interface interactions and interlayer structure arrangement, providing atomic-level structural and dynamic information for CTMAB–Mt and establishing its basic structure as an effective adsorbent for pollutants.
SiO2 | Al2O3 | Fe2O3 | CaO | Na2O | MgO | K2O | TiO2 | P2O5 | SO3 |
---|---|---|---|---|---|---|---|---|---|
62.839 | 16.337 | 5.817 | 5.399 | 3.614 | 3.224 | 1.331 | 1.331 | 0.255 | 0.136 |
Loading levels | Na0.75(Si7.75Al0.25)(Al3.5Mg0.5)O20(OH)4 | ||
---|---|---|---|
Na+ | CTMA+ | Br− | |
0.33CEC | 16 | 8 | 0 |
0.66CEC | 8 | 16 | 0 |
1.00CEC | 0 | 24 | 0 |
1.33CEC | 0 | 32 | 8 |
1.66CEC | 0 | 40 | 16 |
2.00CEC | 0 | 48 | 24 |
2.66CEC | 0 | 64 | 40 |
The total energy of the simulated system included coulombic interaction, van der Waals interaction, and bonded interaction. The total potential energy for the system can be calculated by eqn (1) (more details are in Table 3).
Etotal = EVDW + Ecoulombic + Ebond_stretch + Eangle_stretch + Etorsion | (1) |
Energy term | Potential term | Equation | Interacting atoms | Applied to |
---|---|---|---|---|
a Where: σij and εij are the Lennard-Jones size parameters and well-depth, respectively. Lorentz-Bertholet combining rule is used in the Lennard-Jones parameters calculation;31 e is the charge of electron; ε0 is dielectric permittivity of vacuum; qi and qj are the partial charges of atom i and j; rij is the distance between atoms i and j. r0 and θ0 are the equilibrium values of radius and angle; k1–k3 are the prefactors; d = ±1, n is an integer. | ||||
van der Waals | Lennard-Jones (12–6) | 2 | Any 2 atoms | |
Coulombic | Coulombic | 2 | Any 2 pairs | |
Bond stretch | Harmonic | Ebond-stretch = k1 (rij − r0)2 | 2 | O–H bond; CTMA+ |
Angle bend | Harmonic | Eangle-bond = k2 (θijk − θ0)2 | 3 | CTMA+ |
Torsional | Harmonic | Etorsion = k3[1 + dcos(nϕ)] | 4 | CTMA+ |
The first two terms of the total energy contributed were the Lennard-Jones potential (12–6 potential) and coulombic potential energy terms, and the sum of them represented the noncovalent interaction that is universal for two atoms. These parameters for all nonbonding and bonding are collected in Tables 4–6.
Species | Symbol | Charge (e) | D0 (kcal mol−1) | σ0 (Å) |
---|---|---|---|---|
Hydroxyl hydrogen | Ho | 0.425 | ||
Hydroxyl oxygen | Oh | −0.95 | 0.1554 | 3.1655 |
Bridging oxygen | Os | −1.05 | 0.1554 | 3.1655 |
Bridging oxygen with octahedral substitution | Obos | −1.1808 | 0.1554 | 3.1655 |
Bridging oxygen with tetrahedral substitution | Obts | −1.1688 | 0.1554 | 3.1655 |
Bridging oxygen with double substitution | Obss | −1.2996 | 0.1554 | 3.1655 |
Hydroxyl oxygen with substitution | Ohs | −1.0808 | 0.1554 | 3.1655 |
Tetrahedral silicon | St | 2.1 | 1.8405 × 10−6 | 3.3020 |
Octahedral aluminum | Ao | 1.575 | 1.3298 × 10−6 | 4.2712 |
Tetrahedral aluminum | At | 1.575 | 1.8405 × 10−6 | 3.3020 |
Octahedral magnesium | Mgo | 1.36 | 9.0298 × 10−7 | 5.2643 |
Sodium ion | Na | 1 | 0.1301 | 2.3500 |
Ammonium N | N | −0.68 | 0.1670 | 3.5012 |
Methyl carbon–CH3 | C3 | (−0.3, 0.12) | 0.0390 | 3.8754 |
Organic hydrogen | H | 0.1 | 0.0380 | 2.4500 |
Alkyl carbon–CH2 | C2 | (−0.2, 0.22) | 0.0390 | 3.8754 |
Bromide | Br | −1 | 0.09 | 4.6238 |
Bond stretch | |||
---|---|---|---|
Species i | Species j | k1 (kcal mol−1 Å−2) | r0 (Å) |
Oh | Ho | 554.1349 | 1 |
Ohs | Ho | 554.1349 | 1 |
N | C3 | 356.5988 | 1.47 |
N | C2 | 356.8988 | 1.47 |
C3 | H | 340.6175 | 1.105 |
C2 | C2 | 322.7158 | 1.526 |
C2 | H | 340.6175 | 1.105 |
C3 | C2 | 322.7158 | 1.526 |
Angle bend | ||||
---|---|---|---|---|
Species i | Species j | Species k | k2 (kcal mol−1 rad−2) | θ0 (deg) |
C3 | N | C3 | 86.3 | 112 |
C3 | N | C2 | 86.3 | 112 |
N | C3 | H | 57.3 | 109.5 |
H | C3 | H | 39.5 | 106.4 |
N | C2 | C2 | 50 | 109.5 |
N | C2 | H | 57.3 | 109.5 |
C2 | C2 | H | 44.4 | 110 |
H | C2 | H | 39.5 | 106.4 |
C2 | C2 | C2 | 46.6 | 110.5 |
C3 | C2 | C2 | 46.6 | 110.5 |
C3 | C2 | H | 44.4 | 110 |
C2 | C3 | H | 44.4 | 110 |
Torsional | ||||||
---|---|---|---|---|---|---|
Species i | Species j | Species k | Species l | K (kcal mol−1) | d | n |
C3 | N | C3 | H | 0.0889 | 1 | 3 |
C2 | N | C3 | H | 0.0889 | 1 | 3 |
C3 | N | C2 | C2 | 0.0889 | 1 | 3 |
C3 | N | C2 | H | 0.0889 | 1 | 3 |
N | C2 | C2 | C2 | 0.1581 | 1 | 3 |
N | C2 | C2 | H | 0.1581 | 1 | 3 |
C2 | C2 | C2 | H | 0.1581 | 1 | 3 |
H | C2 | C2 | H | 0.1581 | 1 | 3 |
C2 | C2 | C2 | C2 | 0.1581 | 1 | 3 |
C3 | C2 | C2 | C2 | 0.1581 | 1 | 3 |
C3 | C2 | C2 | H | 0.1581 | 1 | 3 |
H | C3 | C2 | C2 | 0.1581 | 1 | 3 |
H | C3 | C2 | H | 0.1581 | 1 | 3 |
In this study, all MD simulations were performed with the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) software.32 Periodic boundary conditions were imposed on the three dimensions so that clay minerals basal surfaces were in contact with the organic cations. To obtain the trajectories, the Newton equations of motion were integrated numerically using the velocity Verlet algorithm with a time step of 1.0 fs. The short-range van der Waals and coulombic interactions were truncated using 10.0 Å and 12.0 Å cutoffs, respectively. Long-range coulombic electrostatic interactions were computed using Ewald33 summation. During the simulations, the montmorillonite slab was kept rigid and only the inner hydroxyl groups were allowed to move to reduce the errors caused by the “frozen” approach. After the initial configurations, energy minimization was performed using the conjugate gradient method to eliminate unreasonable contacts. The temperature and pressure were controlled using the Nose–Hoover34 thermostat and Hoover35 barostat method, respectively. First, the microcanonical ensemble (NVE) was performed for 0.1 ns at 298 K to relax the system. Then isothermal-isobaric ensemble (NPT) simulations were performed to achieve equilibrium in 2.5 ns and the last 0.5 ns to record the basal spacing at 298 K and 1 atm. Finally, a further 1 ns canonical ensemble (NVT) simulation was performed following the previous 2.5 ns NPT simulation to obtain the interlayer space structures and dynamics properties at 298 K.
(2) |
Radial distribution function (RDF) can provide structural information about ions between interlayers, which is beneficial to understanding the aggregation state of substances. The bonding behaviors of the interlayer species in the CTMAB–Mt can be explored based on their RDF and coordination numbers (CN). It was calculated according to eqn (3).
(3) |
As an important parameter for kinetic analysis, the motion characteristics of the ions and solvent molecules are usually evaluated by the mean square displacement (MSD) evolution, which is the statistical mean of the particle trajectory with time. Using the mean square displacement, the self-diffusion coefficient can be calculated from the following Einstein relation:36
(4) |
Zeta potential can show the charges at the interface of montmorillonite particles and their corresponding aqueous solution. It can be seen from Fig. 3 that Na–Mt has a zero electrical potential (Pzc) at pH = 2 and remains negatively charged at other pH values due to the large number of isomorphic substitutions in octahedral and tetrahedral sheets. Due to the CTMAB modification, the surface charge is reversed to positive at acidic conditions for CTMAB–Mt when compared to Na–Mt. While at alkaline condition, the CTMAB–Mt pose a negatively charge surface, but with a lower charge density than that for Na–Mt. The Pzc of CTMAB–Mt is between pH = 6 and 8, and the zeta potential of CTMAB–Mt changes from a negative value (−11.43 mV) to a positive value (6.56 mV) at pH = 6. That is because the surface and interlayer cations of montmorillonite are exchanged by CTMA+ cations, and the large potential difference is mainly dependent on the cation exchange between layers. The positive zeta potential of CTMAB–Mt increased the adsorption capacity of anions. The zeta potential of montmorillonite is determined by both the permanent negative surface charge and the pH-dependent edge charge. The zeta potential of Na–Mt and CTMA–Mt gradually decreases with the increase in pH value due to the edge charge of clay changes from positive to negative as the pH increases. The hydroxyl group (Al/Si–OH) present at the edge of the montmorillonite particle can gain or lose a proton, which creates an additional pH-dependent contribution to the surface charge.37,38
The FTIR spectra of Na–Mt and CTMAB–Mt clearly confirm the successful intercalation of CTMAB chains into the interlayer space of montmorillonite (Fig. 4). Characteristic absorption bands appear in the Na–Mt and CTMAB–Mt sample at 3627 cm−1 (–OH stretching of Al, Mg(OH)), 1033 cm−1 (stretching vibration of Si–O–Si), and 518 cm−1 (bending vibration band of Si–O). The existence of the above three peaks indicates that the crystalline structure of Na–Mt has not changed after modification. The broad absorption bands at 1644 cm−1 and 3420 cm−1 are assigned to the –OH bending vibration and stretching vibration of water absorbed by the Na–Mt interlayer. Peak weakening at 3420 cm−1 is observed in the FTIR spectra (Fig. 4a) of CTMAB–Mt, suggesting the surface properties of Na–Mt had been changed from hydrophilic to hydrophobic by modifying it with CTMAB. All CTMAB–Mt samples show additional absorption bands at 2926 and 2853 cm−1 (asymmetric and symmetric stretching vibrations of –CH2–) and at 1473 cm−1 (bending vibrations of C–H of the methyl group of the ammonium groups), evidencing the incorporation of CTMAB into the structure of the clay. With increasing the CTMAB loading, the peak intensity and peak area of the CTMAB–Mt related absorption bands at 2926 and 2853 cm−1 shift toward lower frequencies (Fig. 4b), suggesting a transition from a disordered liquid like conformation of CTMAB alkyl chains at low loadings to a more ordered state at high loadings where the modifier alkyl chains exhibit well-aligned trans conformations.39
Fig. 4 FTIR spectra of Na–Mt and CTMAB–Mt: (a) the region of 400–4000 cm−1; (b) the region of 2750–3000 cm−1. |
The interactions between CTMAB and Na–Mt surfaces in the CTMAB–Mt can be quantified based on their radial distribution functions (RDF) and coordination numbers (CN). Fig. 5 illustrates the curves of the RDF of the surface Os of the clay silicate layer around ammonium N, surface Os around alkyl C2, surface Os around methyl C3 and surface Os around hydrogen. The gN–Os(r) curves show the first peak at near 4.1–4.62 Å in all loadings and the average CN of N–Os are about 12 in 0.33CEC and 6 in other loadings, indicating that ammonium N all locate above the surface six-member rings and coordinate with surface Os (Fig. 6). Because methyl C3 is attached to ammonium N as the entire head group, the RDF for C3–Os is similar to that of N–Os, with the first peak located near 3.54–3.66 Å in all loadings. The g(r)C2–Os curves only show one peak at 0.33CEC and no peaks at other loading levels. It can be inferred that CTMA+ organic cations are strongly affected by the upper and lower silicate surfaces, and the migration between layers is mainly horizontal. The limitation of longitudinal displacement allows the coordination to peak at 4.86 Å. The RDF for H–Os in CTMAB has no distinct peaks at 0.33 and 0.66CEC, while the first peak appears at 3.06 Å at other loadings. At 0.33 and 0.66CEC, CTMA+ is strongly attracted by the upper and lower silicate surfaces, which makes the CTMA+ farther away from the surface of montmorillonite than other loadings. It is consistent with the gN–Os(r) results which the first peak appears at 4.62 Å at 0.33 and 0.66CEC, and at about 4.14 Å for other loadings. In general, CTMA+ and montmorillonite are mainly combined by electrostatic interaction at 0.33 and 0.66CEC, and stabilized in the interlayer by both electrostatic interaction and hydrogen bond when the loading exceeds 0.66CEC.
Fig. 5 Radial distribution functions: (a) g(r)N–Os; (b) g(r)C3–Os (–CH3); (c) g(r)C2–Os (–CH2); (d) g(r)H–Os. |
Fig. 6 Top view of the interlayer snapshot at 1.00CEC. On the left, H atoms have been removed for a better view of the head group position. |
The d001 of the experimental and simulated CTMAB–Mt are shown in Table 7. The simulation results are basically the same as the experimental results at the loading (≤1.00CEC). When the loading exceeds 1.00CEC, the d001 of the experimental results corresponds to the simulation results under 1.00 and 2.66CEC system, which shows that CTMAB–Mt mainly exists as the interlayer loading of 1.00 and 2.66CEC. It can be seen that with the increase of the loading, d001 is a process of approximately linear growth from the simulation results, which is different from the experimental results at the loading (≥1.00CEC). That is because in the simulations, the intercalation of CTMAB was set in advance uniformly into the interlayer space of the montmorillonite. However, the amount of CTMAB intercalation between layers may not be uniform in the real intercalation process, which makes the simulation results slightly different.
Loading level (this work) | Simulation (Å) (this work) | Experiment (Å) (this work) | Loading level40 | Experiment40 (Å) |
---|---|---|---|---|
0.33CEC | 14.38 | 14.71 | 0.2CEC | 14 |
0.66CEC | 16.39 | 18.16 | 0.4CEC | 16 |
1.00CEC | 19.41 | 19.61 | 0.6CEC | 18 |
1.33CEC | 23.86 | 19.87/40.11 | 1.0CEC | 21 |
1.66CEC | 27.21 | 19.87/40.11 | 1.5CEC | 19/38 |
2.00CEC | 30.82 | 19.87/40.11 | 2.0CEC | 19/38 |
2.66CEC | 38.55 | 19.87/40.11 | 2.5CEC | 19/38 |
On the basis of the measured basal spacing and the length of the alkyl chains, various arrangement models have been proposed for the intercalated surfactants, including monolayer, bilayer, pseudo-trilayer, paraffin-monolayer, and paraffin-bilayer.24 These idealized models assume that the intercalated surfactants adopted an all-trans conformation after intercalation. To further clearly show the intercalated CTMAB structure arrangement at different loading levels, the density distribution profiles for CTMA+ chains as a function of surface distance (Z–Z0) and the snapshots at different loading levels are shown in Fig. 8. The density distribution profiles along the z direction of CTMA+ are statistically averaged under the last 0.5 ns NVT ensemble after the simulation is stabilized. There is only one clear peak for all CTMA+ cations from the density distribution at 0.33CEC (Fig. 8a1), and the thickness between the occupied layers is about 5 Å, which is comparable to the thickness of the CTMA+ (Fig. 1b), indicating that the CTMA+ is monolayer arrangement and parallel to the silicate surface (Fig. 8a2). The CTMA+ organic cations have two split peaks at 1.00CEC (Fig. 8c1), each with a width of about 5 Å, indicating that a bilayer arrangement has been formed. In addition to the CTMA+ monolayer parallel to the silicate surface, it can also be observed that based on the monolayer, part of the long chain is distributed in the other layer (Fig. 8c2). The CTMA+ starts to split slightly (Fig. 8b1), and the thickness increases to about 8 Å indicating that there is a transition state between monolayer and bilayer (Fig. 8b2). When the loading exceeds 1.00CEC, the CTMA+ is arranged as a pseudo-trilayer. Three peaks can be clearly seen on the density distribution profiles (Fig. 8d1–g1), and the methylene group is likely to jump to the middle layer because of the existence of the intermediate peak. It can also be observed that CTMA+ cations begin to tilt to the silicate surface, and the tilt angle becomes larger and larger (Fig. 8d2–g2). In fact, there are also monolayer and paraffin type CTMA+ cations in the pseudo-trilayer arrangement. With increasing addition loading, the proportion of monolayer CTMA+ decreases and the paraffin type arrangement increases. Paraffin-monolayer (Fig. 8d2) and paraffin-bilayer (Fig. 8d2–g2) appear in the pseudo-trilayer arrangement, so the pseudo-trilayer can also be considered a transition state from bilayer to paraffin type.
Fig. 9 Self-diffusion coefficients of CTMA+ cations at different loading levels and variation of interlayer spacing of adjacent loading. |
Fig. 10 Trajectories of amino N (*1) and alkyl C2 (*2) in CTMA+ cations at 1.00CTMAB–Mt: (a) side view; (b) top view, entire trajectories trace the motion in the last 1 ns NVT ensemble. |
The successful loading of CTMAB will change the structure of montmorillonite, forming layered curls, folds, and rough particles, and the surface potential will reverse. The interaction between CTMA+ and the surface of montmorillonite is mainly electrostatic interaction and hydrogen bond formation. CTMAB is successfully intercalated into the interlayer, and it is evenly intercalated into only one type of intercalated structure at low loading (≤1.00CEC), while it has the existence of both bilayer and pseudo-trilayer structures at high loading (>1.00CEC), with d001 of 1.98 nm and 4.01 nm, respectively. The change in arrangement of CTMA+ between layers includes monolayer, bilayer and pseudo-trilayer with increasing loading. Paraffin type CTMA+ cations are found in the pseudo-trilayer arrangement. It can be considered that the pseudo-trilayer is a transition state from the bilayer to the paraffin type. The size of the diffusion space and the interaction of the alkyl chains affect the mobility of CTMAB.
This journal is © The Royal Society of Chemistry 2023 |