Abstract
We have performed calculations of adsorption energetics on the graphene surface using the state-of-the-art diffusion quantum Monte Carlo method. Two types of configurations are considered in this work: the adsorption of a single O, F, or H atom on the graphene surface and the H-saturated graphene system (graphane). The adsorption energies are compared with those obtained from density functional theory with various exchange–correlation functionals. The results indicate that the approximate exchange–correlation functionals significantly overestimate the binding of O and F atoms on graphene, although the preferred adsorption sites are consistent. The energy errors are much less for atomic hydrogen adsorbed on the surface. We also find that a single O or H atom on graphene has a higher energy than in the molecular state, while the adsorption of a single F atom is preferred over the gas phase. In addition, the energetics of graphane is reported. The calculated equilibrium lattice constant turns out to be larger than that of graphene, at variance with a recent experimental suggestion.
1. Introduction
Being a genuine two-dimensional system, graphene can readily serve as an ideal substrate for chemical modification. It has an extended, double-sided surface and high mechanical strength, making it particularly attractive to create adsorption patterns on graphene for various potential applications. Many possibilities exist to functionalize graphene with rich chemical properties to be explored [1].
To facilitate experimental investigations, computational studies of the energetics of these chemically modified graphene systems are highly desirable. For example, atomic adsorption on graphite or graphene has been an important test system in the history of band structure calculations [2]. Recently, many of the calculations of the adatom adsorption energies on graphene were carried out using density functional theory (DFT) [3–10]. As been realized in the past, DFT results for the adsorption energies on the surface can carry significant errors [11, 12]. Therefore, a systematic study and comparison of these calculated energies not only can provide the much needed guideline for the preferred adsorption process on graphene, but also will improve our understanding of the validity of different exchange–correlation functionals in addressing adsorption problems on the surfaces.
In addition, a puzzling piece of experimental data was reported for the measured in-plane lattice constant of graphane, a graphene sheet fully saturated by H atoms. The experimentally suggested value was 2.42 Å, as extracted from the transmission electron microscope (TEM) measurements [13], which is smaller than that of pure graphene, 2.46 Å. In contrast, a previous DFT study [14] predicted a lattice constant 4% larger than that of graphene. Since it is uncommon for the DFT calculations to exhibit such a large discrepancy in structural parameters, further computational investigations are needed.
In this work, we report calculations using the state-of-the-art diffusion quantum Monte Carlo (DMC) method to investigate the adsorption energetics on graphene. We have studied the interaction of a single atom (O, F, or H) with graphene and, in addition, determined the formation energy of graphane. Additional DFT calculations are also performed using various exchange–correlation functionals, including the local-density approximation (LDA) [15], PBE [16], PW91 [17], and revPBE [18]. The DMC results are then used to provide a definitive description of the adsorption energetics and to check the accuracy of these DFT results.
2. Methodology
2.1. Diffusion Monte Carlo method
For an N-electron system, the diffusion quantum Monte Carlo (DMC) method is, in principle, an exact approach for solving the imaginary-time Schrödinger equation
where τ is the imaginary time, , constant ET is an arbitrary energy offset, and V(R) is the Coulomb interaction potential. Equation (1) is written in atomic units, where e = me = ħ = 4π0 = 1. The ground state of the Hamiltonian in equation (1) can be obtained by taking the limit τ → ∞ [19]. Equation (1) can also be regarded as a 3N-dimensional diffusion equation with a drift term, and ψ(R,τ) can be considered as the density of walkers. However, two issues have to be addressed in practical calculations. First, the drift term (V(R) − ET), corresponding to the Coulomb interaction, may diverge, which will lead to large fluctuations in the population process and will slow down or hinder the simulation. Second, the wavefunction ψ(R,τ) could be either positive or negative, while the density of walkers has to be positive everywhere. This is the well-known fermion sign problem.
The introduction of importance sampling can help smooth the fluctuations and, at the same time, solve the fermion sign problem [20]. Specifically, a guiding wavefunction ψG(R) is introduced to guide the diffusion walkers and rewrite equation (1) in terms of the new distribution function f(R,τ) = ψG(R)ψ(R,τ):
where EL(R) is the local energy calculated with respect to ψG(R), , and the vector F(R) = ∇ψG(R)/ψG(R) is the drift term controlling the velocity of the walkers. If one can choose a reasonably accurate ψG(R), EL(R) will be a well-behaved function without large fluctuations.
The guiding wavefunction ψG(R) also plays an important role in the so-called fixed-node approximation [21, 22]. The new distribution function f(R,τ) in equation (2) can be treated as a Boson ground-state wavefunction (without the sign problem) if ψG(R) has the nodes of the exact fermion ground state that we are looking for. The fixed-node approximation is designed to limit the walkers to moving within one nodal region, and moves across the nodal surface will be rejected [23]. The introduction of the guiding wavefunction ψG(R) sets up the nodal surface in the simulation. For complicated materials systems, the guiding wavefunction ψG thus determines the accuracy of DMC results.
2.2. Optimized trial wavefunctions
The guiding wavefunction ψG(R) can be obtained from optimized variational quantum Monte Carlo (VMC) calculations. The VMC method provides an upper bound for the exact ground-state energy. Given the Hamiltonian , the variational energy with respect to a trial wavefunction ψG is given by
The trial wavefunction has the usual Slater–Jastrow form
where D↑ and D↓ are the up- and down-spin Slater determinants of single-particle orbitals, respectively. These single-particle wavefunctions are usually generated from density functional theory (DFT) or Hartree–Fock calculations. J is the Jastrow factor that contains the electron–electron (u), nucleus–electron (χ), and nucleus–electron–electron (f) terms [24]:
where rij is the distance between electrons i and j, and riI denotes the distance between electron i and ion I. The u, χ and f terms in equation (5) are power expansions containing variational parameters and dependent on the spins of electrons. Since the electron–electron Coulomb potential exhibits discontinuous first derivatives with respect to rij when two electrons get close, the parameters in u are constrained in such a way that the Kato cusp condition [25] is satisfied. The Slater–Jastrow trial wavefunction in equation (4) can be optimized by minimizing either the variance of the energy [26] or the energy alone [27].
2.3. Computational details
For DMC calculations of an extended system, one needs to consider a finite number of electrons in one large supercell and construct the many-body wavefunction from one-particle orbitals associated with a single k-point in the new Brillouin zone (BZ) of this supercell [28]. To make these calculations feasible, we have used a 4 × 4 hexagonal unit cell including a total of 32 C atoms for both graphene and graphane studies. The one-particle orbitals are obtained from spin-polarized DFT calculations with norm-conserving pseudopotentials. For these systems, the quality of these one-particle orbitals turns out to be weakly dependent on the exchange–correlation functionals employed, therefore, most of our DMC calculations have started with LDA wavefunctions. We have also tested the effect of switching to the PBE wavefunctions in the graphane calculations, but found little difference.
For the single-atom adsorption on graphene, three possible high-symmetry adsorption sites are considered: the bridge (B) site between two C atoms, the top (T) sites on top of a single C atom, and the hollow (H) site at the center of the hexagon. The geometry is shown in figure 1, together with the 4 × 4 unit cell used in the calculations. The structure of graphane is assumed to have fully loaded hydrogen on both sides of the surface.
Figure 1. Three high-symmetry adsorption sites on graphene considered in the present work: bridge (B), top (T) and hollow (H) sites. A 4 × 4 unit cell used in the calculations is shown.
Download figure:
Standard imageWe carry out spin-polarized DFT calculations using norm-conserving pseudopotentials with the plane-wave basis as implemented in the CASTEP package [29]. We use the OPIUM code [30] to generate norm-conserving pseudopotentials and Casula's approach [31] is used in DMC calculations to minimize the nonlocal pseudopotential error. The plane-wave cutoff energy is 2000 eV, and the optimized force on each atom is less than 0.03 eV Å−1. In the single-atom adsorption study, a 4 × 4 supercell with a lattice constant of 4 × 2.464 Å is used, and the Brillouin zone is sampled by a 3 × 3 k-point grid to obtain the optimized adsorption structures. We then use a single special k point at this relaxed geometry to do the DMC calculations. For graphane, we have used a 1 × 1 unit cell and a 4 × 4 k-point grid in DFT calculations, and its accuracy is checked by denser grids. We then use this initial wavefunction (equivalent to one k point in a 4 × 4 unit cell) in DMC calculations. A layer separation of 10 Å is sufficient to effectively eliminate the interaction between neighboring layers. Calculations with the projector augmented wave (PAW) method as implemented in VASP [32, 33] are also carried out, and the results are very close to those using norm-conserving pseudopotentials. For isolated single atoms and diatomic molecules, we have repeated the calculations using the Quantum ESPRESSO package [34] and obtained consistent results. Different exchange–correlation functionals are compared, including the LDA [15], PBE [16], PW91 [17], and revPBE [18] functionals.
In order to increase the efficiency of the DMC calculations, the one-particle wavefunctions are subsequently fitted by localized B-spline functions [35]. The scale of the B-spline grid spacing is given by the parameter b = π/Gmax, where Gmax is the length of the largest plane-wave component. A grid spacing of b/1.5 is used in the calculations for adatoms on graphene, while a grid spacing of b/2 is used for the study of graphane. The number of walkers in each DMC calculation is 1000 with a time step of 0.01 au for both graphene and graphane calculations. The trial wavefunctions with a Jastrow form in equation (4) are optimized by the energy minimization method [27]. All the DMC calculations described above are performed using the CASINO code [36].
3. Results and discussions
3.1. Single-atom adsorption on graphene
We first present energetics results for the adsorption of single O, F and H atoms on graphene. The adsorption energy of these atoms on graphene is evaluated with respect to that of the isolated atom: Ead = E(graphene + atom) − E(graphene) − E(atom). We will also discuss the results with respect to the corresponding molecules later. The calculated energies for the isolated single atoms and molecules are discussed in the appendix.
An important issue in DMC calculations is the choice of the special k-point, which effectively determines the imposed boundary condition for the many-body wavefunctions [28]. A careful selection of this special k point is desirable in order to reduce the finite-size error and to obtain a reasonably accurate energy using a finite supercell. The adequacy of a particular k-point choice can be assessed by DFT calculations that are much less computationally demanding. For the 4 × 4 supercell used in the DMC calculations in this work, we find the M point (1/2, 0) at the zone boundary of the supercell BZ to be a good choice for the special k-point. We use the DFT-LDA calculations to compare the energy results obtained from this special k-point with those using denser k-grids in the same BZ, including a 2 × 2 grid without offset, a 2 × 2 grid with an offset of (1/4, 1/4), and a 3 × 3 grid without offset. These results for a single H atom at three different adsorption sites calculated using the CASTEP package with norm-conserving pseudopotentials are shown in figure 2. It can be seen that the adsorption energies for the top and hollow sites obtained using the special k-point are quite close to those obtained using a 3 × 3 k-point mesh. As shown in figure 2, the LDA energy difference between the results using the special k-point and the 3 × 3 k-point set is less than 0.05 eV for H on the top site, the lowest-energy site. For the high-energy bridge site, the relatively larger energy difference is probably due to the reduction of the graphene C6v symmetry to C2v for the bridge site. This choice of the zone-boundary special k-point yields real-valued determinants, gives an excellent energy result for the lowest-energy H adsorption site, and results in a calculated DMC cohesive energy of 7.385 eV/atom for graphene, which is very close to the experimental value of the cohesive energy of graphite (7.374 eV/atom) [37].
Figure 2. Adsorption energies of a single H atom at different sites on the graphene surface calculated using a 4 × 4 unit cell and the local-density approximation. Results for different k-point meshes are compared with that using a special k-point (see text). Results using norm-conserving pseudopotentials (NCPP) and the projector augmented wave (PAW) method are also compared.
Download figure:
Standard imageFor the adsorption of a single O atom at the bridge site, the LDA adsorption energy difference between the special k-point and the 3 × 3 k-point set is also small, only about 0.16 eV. For this case we have further carried out more DMC calculations for each of the point on a 3 × 3 k-grid for the supercell and obtained a final average that is 0.29 eV higher than the energy value using the single special k-point. On the other hand, the LDA overbinding amount for this geometry is around 1.7–2.0 eV, much larger than the energy uncertainty resulting from the k-point choice.
To check the quality of the calculations with norm-conserving pseudopotentials, we have also performed a calculation using the VASP package with the PAW method. As shown in figure 2, the adsorption energy results for a single H in a 4 × 4 supercell with the 3 × 3 k-mesh agree with those using the CASTEP package with norm-conserving pseudopotentials to within 4–17 meV for the three H adsorption sites. The calculations are spin-polarized, resulting in an unpaired electron centered at the H site and a net magnetic moment of 1 μB per supercell.
The DMC adsorption energies for a single O, F, and H atom at different adsorption sites on graphene are plotted in figure 3 and listed in table 1, in comparison with DFT results using different exchange–correlation functionals, including the LDA [15], PBE [16], PW91 [17], and revPBE [18] functionals. The DMC calculations are performed using structures optimized by the LDA, while the structures in DFT calculations are optimized individually. For H, O and F adatoms on the preferred adsorption sites, no significant differences in the structural parameters are found using different functionals. The C–H, C–O and C–F bond length values obtained using the LDA are 1.13 Å (top), 1.44 Å (bridge), and 1.50 Å (top), respectively, while the bond length values obtained using the PBE functional are 1.13 Å, 1.47 Å, and 1.56 Å, respectively. The difference between LDA and PBE bond lengths is therefore smaller than 0.01, 0.03 and 0.06 Å for C–H, C–O and C–F, respectively. The numbers of the spin-up and spin-down electrons per cell differ by one for F and H, giving rise to a magnetic ground state. In all three cases, the most stable sites predicted by the DMC and DFT calculations are in agreement, with O favoring the bridge site and F and H favoring the top site. However, the DMC results indicate that DFT calculations vastly overestimate the binding of O and F atoms on graphene. The PBE, PW91, and revPBE results for the adsorption energies are generally better than the LDA results, but they significantly underestimate the energy difference between various sites of F on graphene, giving rise to questionable results for the diffusion barrier. The situations are much better for H on graphene, for which PBE, PW91, and revPBE calculations give quite accurate adsorption energies for the top and hollow sites. Overall, the revPBE adsorption energies are closest to the DMC results, although the relative energy order at different adsorption sites obtained by different functionals is generally consistent. The absolute error of the LDA energies for adsorbed O and F is of the order of 2 eV and 1.5 eV, respectively. Most of the energy errors of the PBE results come from the solid calculation, since PBE is able to give quite accurate atomic energies, as shown in the appendix. The revPBE errors are noticeably smaller compared with LDA results, about 0.2 eV and 0.5 eV for O and F at their preferred sites, respectively. In all three cases, the adsorption energy is negative at their lowest-energy site compared with the atomic state, with O having the strongest energy gain by forming bonds with two C atoms at the bridge site.
Figure 3. Adsorption energies of a single O, F and H atom at different adsorption sites on graphene are compared. Results from DFT calculations using various exchange–correlation functionals are compared with the DMC energies. The adsorption sites considered include B (bridge), T (top), and H (hollow) sites, and energies are with respect to that of an isolated atom. The DMC error bars are about 70 meV.
Download figure:
Standard imageTable 1. Energies (in eV) of a single oxygen, fluorine, and hydrogen atom at different adsorption sites on graphene with respect to its atomic energy. Results using different exchange–correlation energy functionals (see text) calculated with the projector augmented wave (PAW) method are compared with the diffusion quantum Monte Carlo (DMC) results. B, T, and H stand for the bridge, top, and hollow site, respectively. The values in the parentheses are results with respect to a half of the molecular energy.
Oxygen | LDA | PBE | PW91 | revPBE | DMC |
---|---|---|---|---|---|
B | − 3.48(0.27) | − 2.40(0.87) | − 2.50(0.82) | − 2.04(1.07) | − 1.79(0.64) |
T | − 2.55(1.20) | − 1.66(1.62) | − 1.79(1.54) | − 1.34(1.77) | − 0.43(1.99) |
H | − 0.82(2.93) | − 0.56(2.68) | − 0.60(2.72) | − 0.56(2.54) | 0.39(2.82) |
Fluorine | |||||
B | − 2.26(−0.64) | − 1.86(−0.50) | − 1.92(−0.53) | − 1.71(−0.48) | − 0.55(0.05) |
T | − 2.67(−1.06) | − 2.07(−0.72) | − 2.13(−0.74) | − 1.83(−0.61) | − 1.31(−0.71) |
H | − 2.01(−0.39) | − 1.73(−0.37) | − 1.78(−0.39) | − 1.64(−0.41) | 0.26(0.34) |
Hydrogen | |||||
B | − 0.44(2.01) | − 0.10(2.37) | 0.11(2.39) | 0.23(2.51) | − 0.70(3.07) |
T | − 1.26(1.19) | − 0.90(1.36) | − 0.88(1.39) | − 0.84(1.44) | − 0.91(1.46) |
H | − 0.09(2.36) | − 0.02(2.28) | 0.05(2.32) | 0.01(2.30) | − 0.05(2.42) |
An alternative way to examine the adsorption energetics is to compare the adatom energies on the surface with that of the molecular state. The DFT results may become better due to error cancelation. These comparisons are shown in figure 4 in the same format as in figure 3. Indeed, the energy errors are reduced, but are less consistent. For example, the PBE and revPBE energy results for O at the lowest-energy bridge sites are higher than the DMC value, while the LDA value is lower. In addition, the results show that single O and H atoms on graphene have a higher energy than their molecular state by 0.6 eV and 1.5 eV, respectively. Only a single F atom on graphene is more favorable than in the molecule, with an energy gain of 0.7 eV/atom. The detailed energy values are given in table 1.
Figure 4. Adsorption energies (with respect to the molecular state) of a single O, F and H atom at different adsorption sites on graphene are compared. Results from DFT calculations using various exchange–correlation functionals are compared with the DMC energies. The adsorption sites considered include B (bridge), T (top), and H (hollow) sites, and energies are with respect to a half of the energy of corresponding molecules. The DMC error bars are about 70 meV.
Download figure:
Standard image3.2. Graphane
We next investigate the graphene sheet fully loaded by H. At the saturation concentration, H atoms are adsorbed on both sides of the graphene plane. The structure of graphane, which was first named by Sofo et al [14], is illustrated in figure 5. Similar to the previous case of single-atom adsorption on graphene, we need to first determine the special k-point to be used in our DMC calculations. The unit cell of graphane contains two C atoms and two H atoms. The results of our LDA calculations using norm-conserving pseudopotentials indicate that the total energy obtained with a 4 × 4 k-point mesh with an offset of (0.5, 0.5) is very close to that obtained by using a 12 × 12 mesh. The former is then used in the DMC calculations, which is equivalent to using a 4 × 4 supercell with a special k-point of (0.5, 0.5). The total number of electrons included in the DMC calculations for graphane is 160.
Figure 5. Structure of graphane: top view (left) and side view (right).
Download figure:
Standard imageFigure 6 shows the DFT and DMC results for the formation energy of graphane defined as Ef = E(graphane) − [E(graphene) + 2E(H atom)] as a function of the in-plane lattice constant a. The equilibrium lattice constants from LDA, PBE and DMC calculations determined by the fitted curves are 2.50, 2.52 and 2.50 Å, respectively. The DMC energies have a statistical error of 50 meV, which will not affect the lattice constant result. We have also examined the effect of using different one-particle wavefunctions in constructing the Slater determinant, and the difference in the DMC energies using LDA and PBE orbitals is completely negligible. The DMC calculation gives a lattice constant of 2.50 Å, which is similar to the results from our DFT calculations and previous DFT work [14]. However, this value is larger than the reported experimental value of 2.42 Å in a TEM measurement [13]. The DFT calculations with the LDA or PBE exchange–correlation functional give similar lattice constants, with the PBE result slightly larger. Our DMC and DFT results indicate that the lattice constant of graphane is greater than that of graphene (2.46 Å). Therefore, the small lattice constant reported previously [13] remains unexplained, and could arise from the substrate interaction. It is noted that DFT calculations do not provide an accurate value for the formation energy. The errors are about 1 eV for LDA and 0.5 eV for PBE.
Figure 6. Formation energies of graphane as a function of the lattice constant obtained using density functional theory (DFT) with different exchange–correlation functionals (LDA and PBE) and the diffusion quantum Monte Carlo (DMC) method with Slater determinants constructed from two different one-particle wavefunctions (wf). The DMC errors are about 50 meV.
Download figure:
Standard image4. Conclusion
To summarize, we have performed DMC calculations to investigate the adsorption energetics on the graphene surface. The results are used to check the accuracy of various DFT exchange–correlation functionals. We first consider the adsorption of O, F, or H adatoms on the graphene surface. For O and F adatoms, we find that LDA, PW91, PBE and revPBE all yield over-bound results, while for the H adatom all GGA-based functionals are able to produce reasonable adsorption energies compared with the DMC results for the top (lowest-energy) and hollow sites. Regarding the ordering of the adsorption energies at different sites, DMC and DFT give consistent results. In addition, we have investigated the formation energy of graphane (the H-saturated graphene system) to find the equilibrium lattice constant. The DMC calculation predicts a larger lattice constant than that of graphene, similar to the DFT prediction. However, a recent experiment reported a smaller lattice constant than graphene, which we believe could arise from the substrate interaction.
Acknowledgments
CMW acknowledges support from the National Science Council of Taiwan under Grant No. 99-2112-M001-034-MY3. CRH and CMW acknowledges support from the National Center for Theoretical Sciences (NCTS) in Taiwan. MYC acknowledges support from the US Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering under Award No. DE-FG02-97ER45632.
Appendix.: Energies of single atoms and molecules
In this appendix, we tabulate the DMC energetics results obtained for free single atoms and molecules. We have calculated the energy of an isolated atom using norm-conserving pseudopotentials, and the results for the O, F, and H are listed in table A.1, including DFT values with different exchange–correlation energy functionals. In general, the LDA results tend to have an error of a few per cent compared with the DMC results, which could be as large as 3 eV for O and F. In contrast, other more advanced functionals can give quite accurate energies for a single atom; the error is of the order of 0.1%. For example, the energy difference between PBE and DMC results is only about 0.02 eV for atomic H and 0.2 eV for atomic F and O.
The calculated binding energies for O2, F2 and H2 molecules are shown in table A.2. Our DMC atomization energy of F2 is 1.35 eV, which is consistent with that of 1.38 eV from a previous QMC calculation [39]. For O2 and F2, all DFT binding energies are larger than experimental values by 1–2 eV, while the DFT errors for H2 are much less significant. The binding energy difference between DMC and experiment for H2 is only 12 meV. Overall, the DMC method is able to provide the best molecular binding energies compared with experiment.
Table A.1. Atomic energies for O, F and H obtained using density functional theory with different exchange–correlation functionals and the diffusion quantum Monte Carlo (DMC) method. Energies are in eV. DFT calculations were performed using the Quantum ESPRESSO package with norm-conserving pseudopotentials.
LDA | PBE | PW91 | revPBE | DMC | |
---|---|---|---|---|---|
H atom | − 13.028 | − 13.603 | − 13.644 | − 13.725 | − 13.625 |
F atom | − 652.264 | − 655.636 | − 656.294 | − 656.364 | − 655.794 |
O atom | − 429.816 | − 432.323 | − 432.803 | − 432.916 | − 432.617 |
Table A.2. Molecular binding energies for O2, F2 and H2 obtained using density functional theory with different exchange–correlation functionals and the diffusion quantum Monte Carlo (DMC) method. Energies are in eV. DFT calculations were performed using the Quantum ESPRESSO package with norm-conserving pseudopotentials. The experimental values are also shown for comparison.
LDA | PBE | PW91 | revPBE | DMC | Expt | |
---|---|---|---|---|---|---|
H2 | 4.908 | 4.536 | 4.560 | 4.576 | 4.738 | 4.75a |
F2 | 3.274 | 2.675 | 2.741 | 2.467 | 1.352 | 1.60b |
O2 | 7.516 | 6.850 | 6.862 | 6.622 | 5.009 | 5.11b |