Elaheh K. Goharshadi*ab,
Golnoosh Akhlamadia and
Sayyed Jalil Mahdizadeha
aDept. of Chemistry, Ferdowsi University of Mashhad, Mashhad 91779, Iran. E-mail: gohari@um.ac.ir; Tel: +98-51-38805558
bCenter of Nano Research, Ferdowsi University of Mashhad, Iran
First published on 2nd December 2015
In this work, the dispersion behavior of graphene oxide (GO) nanosheets in water was studied by a series of MD simulations for water, GO, and the water/GO mixed systems. The simulation results showed that the water/GO system has a well-ordered structure with strong H-bond interactions. The initial value (4.0 Å) of the interlayer distance between two GO sheets at the beginning of the simulation clearly increased (∼7.2 Å) in the presence of water as compared to its value after simulation in the absence of water (5.7 Å). The solubility parameter of GO was calculated as a function of both temperature and number of layers. The solubility parameter of GO at 300 K reached a plateau at 44.9 MPa1/2 when the number of layers was five. Hence, water is a good solvent for dispersing GO since their solubility parameters are close to each other especially at temperatures close to the freezing point of water. The strong H-bond between water and oxygen-functional groups of GO makes the enthalpy favorable to form stable GO dispersions in water.
Many industrial and scientific applications benefit from the outstanding chemical, thermal, and electrical properties of GO, making its colloidal suspensions stable in various solvents, especially in water, very important.1–5 A number of studies have reported the dispersion of GO in polar solvents.3–5 The choice of an appropriate solvent for the dispersion of GO was based primarily on trial-and-error experiments.
Cohesive energy density is the net effect of all the interatomic/molecular interactions including van der Waals interactions, covalent bonds, ionic bonds, hydrogen bonds, and electrostatic interactions. An understanding of cohesive energy density is very important for predicting the stability of a colloid. The cohesive energy density of a material can be quantified in a number of ways. The most common approach is to use the solubility parameter, δ. The solubility parameter is one of the most useful concepts in the physical chemistry and thermodynamics of solutions and colloids. There is much interest in utilizing solubility parameters for rationally designing novel processes such as supercritical extraction, coatings, or new materials such as drugs and polymer alloys.6
In 1936, Hildebrand proposed a simple definition for a “solubility parameter” which provides a systematic description of the miscibility behavior of solvents.7 The solubility parameter is defined as the square root of the cohesive energy density (CED), i.e. the amount of energy needed to completely remove unit volume of molecules from their neighbors to infinite separation:8
![]() | (1) |
The dispersion behavior of GO at different organic solvents were investigated by Paredes et al.3 Graphite oxide was dispersed in N,N-dimethyl formamide, tetrahydrofuran, and ethylene glycol. They mentioned that in all of these solvents, full exfoliation of the graphite oxide into individual and single-layer GO sheets was achieved by sonication. Konios et al.5 investigated the dispersion behavior of GO and reduced graphene oxide (rGO) in eighteen solvents by estimating the solubility parameter using UV-Vis spectroscopy. Hadadian et al.9 investigated the stability of GO–distilled water dispersions by a number of experiments. Although these studies provide useful preliminary information, they do not permit molecular-level design of new solvents that are capable of efficiently dispersing GO. Hence, developing a molecular-level understanding of GO–solvent interactions represents a very important step toward optimizing the design of stable GO dispersions. Molecular dynamics (MD) simulations provide a promising computational tool for elucidating the nature of GO–solvent interactions at the molecular level. Simulations can provide important insights into the structural, electronic, and chemical properties of GO.
The aim of the paper is theoretical study of GO dispersion in water using MD simulation, calculating GO and water's solubility parameters by changing the effective parameter “temperature”, in order to find the optimum conditions for having stable GO's suspension in water. To the best of our knowledge, up to now, no research has reported for calculating the solubility parameter of GO using MD simulation. Although there are several scattered reports on different aspects of MD simulation for GO's dispersion in water, none of these reports directly address the solubility parameter. For example, Bagri et al.10 used MD simulation to study the atomistic structure of progressively reduced graphene oxide. Medhekar et al.11 elucidated the atomic level structure and mechanical properties of GO paper based on MD simulations using the reactive force field (ReaxFF). Shih et al.12 carried out a series of comparative experimental and MD simulation studies to understand the fundamentals of the surface activity and colloidal stability of GO aqueous solutions at different pH values. Dimiev et al.13 proposed an unconventional view of GO chemistry and develop the corresponding “dynamic structural model”. Nicolaï et al.14 used the quantum mechanical calculations to develop a full set of force field parameters in order to perform MD simulations to understand and optimize the molecular storage properties inside graphene oxide frameworks. Zokaie and Foroutan15 compared the structural and dynamical properties of water confined between two GO sheets through MD simulations. They showed that the structure and dynamics of water near the GO surfaces changes under confinement conditions. However, none of the recent MD simulation studies have been calculated the solubility parameter of GO.
The simulations of water were run at different temperatures for 1000 water molecules with the time step of 0.1 fs. The periodic boundary conditions were applied for all three directions. All water simulations were initially run for 0.1 ns in NPT ensemble (Nose–Hoover thermostat and barostat) at 1 bar and desired temperature to reach the equilibrium and then switched to NVE ensemble for 0.1 ns. The NVE ensemble was selected for production and averaging step because the energy fluctuations are minimum in this ensemble.
The REBO potential20 was used for carbon and hydrogen atoms. The bond stretching was represented by Morse potential (for C–O, O–H, and CO bonds). The parameters for the Morse potential were taken from elsewhere.21 The harmonic cosine form for bending potential was used.21 The dihedrals were described by cosine form.22 LJ (12, 6) potential was used for describing the vdW interactions. The cutoff distance for non-bonded interactions was considered to be 15 Å. The LJ parameters for H, C, and O atoms are listed in Table 1.
Atom | σ (Å) | ε (kcal mol−1) |
---|---|---|
C | 3.80 | 0.080 |
H | 2.60 | 0.008 |
O | 3.60 | 0.150 |
The columbic interactions were also considered using Coulomb law with the cutoff radius of 15 Å. The long range columbic interactions beyond the Coulomb cutoff distance were also calculated using the particle–particle particle-mesh (PPPM) method.24
To calculate the partial atomic charges, first the initial structure of GO was optimized in B3LYP/6-31G(d) level of theory. Then, the natural bond orbital (NBO) was employed to calculate the partial atomic charges using density functional theory at the same level of theory. The mean partial atomic charges for oxygen and hydrogen atoms of hydroxyl groups were calculated as −0.46 and 0.27, respectively. The mean partial atomic change for oxygen atoms of epoxy groups was calculated as −0.33. These values were 0.35 and −0.30 for carbon and oxygen atoms of quinone groups, respectively. Also, the mean atomic charges calculated for carbon and oxygen atoms due to carbonyl group and oxygen and hydrogen atoms of hydroxyl group of carboxyl groups were 0.61 and −0.40, −0.43 and 0.39, respectively. The partial atomic charges calculated using B3LYP/6-31G(d) level of theory are a preliminary and raw guess for atomic charges. Hence, we used the charge equilibration method,25 as implemented in LAMMPS, to minimize the electrostatic energy of the system by adjusting the partial charge on individual atoms based on interactions with their neighbors. All the quantum mechanics calculations were performed by Gaussian 09 software program package.26 After that, the simulations of GO were carried out in NPT ensemble with the total simulation time of 40 ns (time step of 1 fs). Three MD simulations for water, GO, and mixed water/GO systems were performed.
The oxygen–oxygen, gOO(r), oxygen–hydrogen, gOH(r), and hydrogen–hydrogen, gHH(r), RDFs were derived using MD simulation with considering TIP4P potential model at 300 K and 1 bar. gOO(r), gOH(r), and gHH(r) are given in Fig. 2a–c, respectively. As Fig. 2 shows, there is a good agreement between our work and previous experimental and theoretical studies.28–31
The mean square displacements (MSDs) of liquid water at different temperatures were computed using a series of MD simulations. The curve of MSD versus time calculated from averaging over for 1000 water molecules and the 0.1 ns trajectories in the temperature range from 280 to 360 K (Fig. 3). Fig. 3 shows that as time elapses, the MSD grows linearly. Also, the MSD increases with raising temperature since the mobility of molecules increases.
![]() | ||
Fig. 3 The MSD plot for liquid water molecules at different temperatures and 1 bar. The curves were calculated using the MD simulation with considering TIP4P potential model for liquid water. |
Table 2 gives the calculated internal energy change of vaporization (ΔEvap), molar volume (Vm), and solubility parameter of water (δwater). The solubility parameters were derived using eqn (1). As Table 2 shows there is a good agreement between the calculated and experimental values of solubility parameter for liquid water at different temperatures. Table 2 also indicates that as temperature rises, the solubility parameter decreases. With increasing the temperature, the interactions between the molecules in the system weaken as a result of an increase in kinetic energies of the molecules. Therefore, the amount of energy required to overcome the interactions and vaporize the molecules decreases, and the solubility parameter decreases.
T (K) | Vma (cm3 mol−1) | ΔEvapa (kJ mol−1) | δwater (MPa1/2) | % Error (δwater) | |
---|---|---|---|---|---|
Calc. | Exp.b | ||||
a Calculated using MD simulation.b Experimental data was calculated based on enthalpies of vaporization extracted from ref. 32.c Ref. 6. | |||||
280 | 17.908 | 40.368 | 47.480 | 48.591 | 2.3 |
300 | 18.124 | 39.106 | 46.454 | 47.994 | 3.2 |
47.820c | 2.8 | ||||
320 | 18.160 | 37.872 | 45.663 | 47.391 | 3.6 |
340 | 18.630 | 36.564 | 44.304 | 46.779 | 5.3 |
360 | 18.963 | 35.271 | 43.128 | 46.159 | 6.5 |
W(r) = −kT![]() ![]() | (2) |
The first peak of RDF and PMF for the O1–H1 appears at 1.98 Å (Fig. 4a). The negative value of PMF at this distance indicates there is a strong attraction between O1 and H1. The ratio of height of second peak is much less than the first one. It means there is a weak correlation between O1–H1 at this distance (second peak). In other words at distances greater than 2.2 Å, there is no correlation between O1 and H1.
Fig. 4b shows that the height of the prominent peak, first peak, is less than that of Fig. 4a. This means there is a weaker interaction between oxygen of the first layer and hydrogen of the second layer of GO compared to that of the first layer. Also, the value of PMF of O1–H2 (−0.365 kJ mol−1 nm−2), is more positive than that of O1–H1 (−0.428 kJ mol−1 nm−2) at the position of the first peak. As Fig. 4c to e show there is no interaction between O1–H3, O1–H4, and O1–H5.
Fig. 5a and b show the RDF and PMF plots for O1–H2 at different temperatures, respectively. Because of thermal motions, the intensity of the first peak of the RDFs decreases as temperature rises.
![]() | ||
Fig. 5 (a) The RDF and (b) the PMF plots for O1–H2 of GO for different temperatures at 1 bar. The insets are a close view of (a) the first peak of RDFs (b) the well depth of the PMF curves. |
ΔEvap = ENL − NESL | (3) |
T (K) | EGO (eV) | ||
---|---|---|---|
N = 1 | N = 4 | N = 5 (bulk) | |
280 | −3278 | −13![]() |
−16![]() |
300 | −3289 | −13![]() |
−16![]() |
320 | −3304 | −13![]() |
−16![]() |
340 | −3310 | −13![]() |
−16![]() |
360 | −3312 | −13![]() |
−16![]() |
The volumes and internal energy of vaporizations of GO for different number of layers at 300 K was computed using MD simulations and the results are shown in Table 4. As this table shows the internal energy of vaporization rises with the increase of the number of layers because the number of interactions increases.
N | V (Å3) × 103 | ΔEvap (eV) |
---|---|---|
2 | 10.800 | 90 |
3 | 14.852 | 155 |
4 | 19.531 | 245 |
5 | 24.281 | 306 |
The average number of hydrogen bonds (H-bond) per single layer of GO can be computed by integrating gO1–H2(r) plot up to the first minimum. Table 5 gives the average number of H-bond per single layer of GO and the number of H-bond per unit area (DHB). As temperature increases both number of H-bond and DHB decreases because of increasing of thermal motions.
T (K) | Average no. of H-bonds per single layer of GO | DHB (no. H-bond per nm2) |
---|---|---|
280 | 71.6 | 7.99 |
300 | 71.3 | 7.92 |
320 | 69.5 | 7.72 |
340 | 68.0 | 7.56 |
360 | 63.0 | 7.00 |
Although a large number of experimental work on solubility parameter for liquids can be found in literature, for solids only a few publications reported the solubility parameter. This is probably due to the fact that experimental measurements of solubility parameters for liquids are easier to obtain than for solids. In this work, the solubility parameter of GO was calculated as a function of both temperature and number of layers using volume and ΔEvap reported in Table 4. Fig. 6 shows the solubility parameter of GO versus the number of layers at 300 K and 1 bar. The solubility parameter of GO at 300 K reaches a plateau at 44.9 MPa1/2 when the number of layers is five. Therefore, the five layer GO was considered as bulk state.
Table 6 shows the computed values of volume and solubility parameter of the bulk GO (N = 5) at different temperatures. The computed values of solubility parameters were compared to the corresponding computed data for liquid water in Table 6. This table shows both values are very close to each other especially at lower temperatures. Hence, water is a good solvent for dispersing GO especially at temperatures close to freezing point of water. This observation may be associated with the fact that hydrogen bonds between water molecules and oxygen groups of GO are stronger at lower temperatures. As mentioned before, the solubility parameter is the net effect of all interatomic/molecular interactions that here are the electrostatic (H-bond) and van der Waals. Table 6 reveals that solubility parameter of the both GO and water increases as temperature falls from 360 to 280 K but the slope of this enhancement is steeper for GO. Finally, the minimum difference between these two solubility parameters occur at temperatures close to the freezing point of water. In the other words, the interactions between water molecules and between GO layers become more similar at lower temperatures. One of the basic concepts in solution chemistry is that materials with the same interactions mix more easily with each other.
T (K) | V (Å3) × 103 | δGO (MPa1/2) | δwater (MPa1/2) | % Error |
---|---|---|---|---|
280 | 24.246 | 48.8 | 47.5 | 2.7 |
300 | 24.281 | 44.9 | 46.4 | −3.3 |
320 | 24.292 | 42.5 | 45.7 | −7.5 |
340 | 24.317 | 41.6 | 44.3 | −6.5 |
360 | 24.322 | 41.2 | 43.1 | −4.6 |
It is worthy to mention that along with simulation studies, we investigated the dispersion of GO in water experimentally.9 Our experimental data are in acceptable agreement with the simulation results, i.e., the aqueous suspensions of the prepared GO were stable for more than 5 months without any sedimentation.
To study the influence of GO in the structure of water, we first computed the oxygen–oxygen RDF(gOO(r)) and oxygen–hydrogen RDF (gOH(r)) of water in the presence of double layer GO and compared to that of pure water (Fig. 2a and b). The main features of Fig. 8 are:
(i) The number of peaks of gOO(r) in the water/GO system with respect to liquid water increases, meaning that the structure of water in the presence of GO is more ordered than that of pure water. The gOO(r) plot of water in the water/GO system approaches to unity at longer distances (14 Å) compared with that of pure water (8 Å). This is due to the strong H-bond between oxygen-containing functional groups of GO and water molecules. Hence, the H-bond between GO and water has a profound effect on the structure of water.
(ii) The height of the first peak of gOO(r) for water in the water/GO system is greater than that of pure water. It means that the number of water molecules at the nearest neighboring shell increases because of more ordered structure of water at the presence of GO.
(iii) The position of the first peak of gOO(r) remains almost unchanged whereas the second peak shifts to longer distance in the water/GO system. The same results were obtained for water molecules confined between GO sheets.15 The shift observed for the second peak can be assigned to the disruption of hydrogen bond between water molecules in the presence of GO.15
(iv) Fig. 8b reveals an interesting fact. As this figure shows, in the bulk water, the first peak of gOH(r) located at 1.84 Å (assigned as mean of H-bond length) was disappeared in the presence of GO sheets. Instead, some overlapped broad peaks appear in the region 2.2–3.0 Å. This finding approves that H-bond interactions between water molecules in the water/GO system were strongly disrupted and a longer (so weaker), and non-uniform H-bond forms. Medhekar et al.11 shows that optimum H-bond length between water molecules in the similar water/GO system is 2.55 Å.
To get further knowledge on the structure of water/GO system, the most interesting RDFs, gOH(r) (oxygen atom of GO and hydrogen atom of water) was computed (Fig. 9). The first peak, which indicates the hydrogen bond interactions, located at about 1.20 Å. It was shifted to shorter distances compared to H-bond distance in water (1.84 Å) (Fig. 2b) which clearly shows a strong H-bond between GO oxygen functional groups and water molecules was formed.
The MSD plots in the X, Y, and Z-directions as well as total MSD of water for water/GO system (T = 300 K and 1 bar) were also computed using MD simulations (Fig. 10). A comparison between the MSD plots of bulk water (Fig. 3) and water in water/GO system (Fig. 10) indicates that strong H-bond network between GO and water molecules makes the water to behave as a pseudo-solid material with a saturated MSD plot. In fact, the mobility of water molecules diminishes because of strong H-bonds between GO and water. The MSD values in X and Z directions are greater than that of Y direction because Y axes is perpendicular direction to the GO surface (XZ plane). The similar results were observed in ref. 17.
The main conclusions of this work are:
(i) The internal energy change of vaporization of GO increases with the increase of layers.
(ii) Water/GO system has a well-ordered structure with strong H-bond interactions. Water molecules penetrate into the GO interlayer space and exfoliate it easily (thanks to H-bond interactions).
(iii) MD simulation results showed that water is a good solvent for dispersing GO since the values of solubility parameter of GO is close to those of water especially at temperatures close to freezing point of water. The strong H-bond between water and oxygen-functional groups of GO makes the enthalpy favorable to form stable GO dispersions in the water.
(iv) The strong hydrogen bonding between GO and water molecules is responsible to form stable colloidal suspension of GO in water. It is an enthalpy driven process. The value of solubility parameter for GO supports this tendency since it was calculated based on enthalpy of vaporizations. Hence, some sort of enthalpy driven feature of GO dispersion in water can be concluded.
Footnote |
† Electronic supplementary information (ESI) available: Table S1 (justification of validity of potential used in this work). See DOI: 10.1039/c5ra19932h |
This journal is © The Royal Society of Chemistry 2015 |