Introduction
The two-dimensional carbon allotrope graphene has been a material of extreme interest in recent years for its unique electronic and structural properties for experimental, theoretical, and computational study alike. Its phenomenological landscape can be broadened further when we consider its stackability into bilayer or other multiple-layered structures. (In the many-layer limit, it is the very familiar mineral graphite.) For example, bilayer graphene in particular has been of interest for the study of its eightfold-degenerate Landau levels.[1] Determination of its structure with density functional theory is complicated by the fact that the layers are bound together not by covalent bonds but by the van der Waals interaction, a correlation effect whose manifestation depends on the chosen functional. This project explores the predicted interlayer distance of bilayer graphene for several different functionals.
For our calculations, we use the DMol³[2,3] framework, using atomic orbitals as a basis (specifically, the DNP 3.5 basis set) and treating all electrons, requiring self-consistent field convergence within 10⁻⁶. We used PBE[4] and PW91[5] from the GGA family of functionals and the PWC[6] functional from the LDA family, all “vanilla” funcitonals without dispersion correction, plus we reran calculations with the PBE functional using the TS[7] method for DFT-D (DFT dispersion) corrections. We also considered only the Bernal stacking pattern, well-known to be the lowest in energy. We used periodic boundary conditions with a 7×7×2 k-point set and with a 18.4Å lattice constant in the direction normal to the plane. (This yields a vacuum gap of 15Å at an interlayer separation of 3.4Å; all equilibrium interlayer distances were between 3Å and 4Å.)
Figure 1 – (Above) In Bernal-stacked bilayer graphene, the lattices of the two layers are exactly aligned with one another, but shifted slightly so that the A site of one lattice is directly above the B site of the other. (Below) The A and B lattice sites are differentiated by the relative orientation of the nearest neighbors. For example, one could distinguish them by whether an angle opens leftwards (for A) or rightwards (for B).
Methods
First, we should clarify why the van der Waals interaction is exclusively a correlation effect, i.e. in Kohn-Sham DFT it arises solely from the exchange-correlation functional. In DFT we solve a one-electron problem using the total electron density of the solved problem. However, van der Waals interactions are the interactions of instantaneous dipoles, which are not captured by the electron density. See Figure 2 for a classical picture which demonstrates intuitively why van der Waals interactions can only be resolved when one considers the simultaneous motion of at least two electrons.
Figure 2 – If the small yellow negative charges orbiting the large red positive charges happen to be on opposite sides of the positive charges they orbit, the “atoms” will have instantaneous dipole moments and therefore be attracted.
The van der Waals interaction arises from the interaction of the electron density in one location with the electron density in another, so it is a fundamentally nonlocal effect. In contrast, LDA is by name the “local density approximation”, and GGA is only semi-local as the “generalized gradient approximation”, depending on the local density and its derivative and loosely speaking being nonlocal only for infinitesimal distances.
Several methods exist to account for van der Waals interactions. One is to add to the DFT energy the pairwise (i.e. for each pair of atoms) van der Waals interaction energy C₆/R⁶ where R is the interatomic distance, but this approach adds a degree of empirical dependence. The TS method[7] used in this project effectively applies this method, but uses the results of DFT calculations to determine the C₆ van der Waals interaction coefficient, circumventing much of the empirical nature of the approximation. (The TS method still has has two adjustable parameters used in a damping function to remove the singularity for small R. They were chosen by comparison and fitting to data on noble gases, van der Waals-bonded organic dimers, and a database of organic molecule interaction energies.) Another possible method, not used here, is to employ special supplementary van der Waals interaction functionals.[8,9]
To find the equilibrium interlayer distance, we searched for a minimum energy by adjusting that parameter only, since we already know that Bernal stacking is the equilibrium relative orientation of the graphene sheets as mentioned before.
Results
We found that the equilibrium interlayer distances for the four approaches discussed were between 3.78Å and 3.8Å for PBE, between 3.7Å and 3.75Å for PW91, 3.25Å for PWC, and 3.33Å for PBE with TS. This wide spread in results emphasizes the importance of knowing what capabilities each functional does and does not have, and warns against taking results blindly. For comparison, numerous experiments and calculations throughout the literature report interlayer spacings in the range between 3.3Å and 3.4Å. This applies just as well to corrected calculations, such as ours with PBE and TS. Regardless, the disparity between our TS-corrected and uncorrected PBE results emphasize the significance of the van der Waals in certain systems. Layered structures such as bilayer graphene are an excellent example of such systems and are a topic of dedicated research efforts.[9,10]
Table 1: Energy as a function of distance for the PBE GGA functional
Layer separation (Å) | 3.35 | 3.4 | 3.45 | 3.5 | 3.6 | 3.75 | 3.78 | 3.79 | 3.8 | 3.82 | 3.83 | 3.85 |
PBE energy (Ha) | -152.337163 | -152.337298 | -152.337405 | -152.337483 | -152.337584 | -152.337638 | -152.337639 | -152.337639 | -152.337639 | -152.337638 | -152.337637 | -152.337636 |
Table 2: Energy as a function of distance for the PW91 GGA functional
Layer separation (Å) | 3.4 | 3.5 | 3.7 | 3.73 | 3.75 | 3.8 | 3.85 | 3.9 | 4 |
PW91 energy (Ha) | -152.470189 | -152.470360 | -152.470483 | -152.470482 | -152.470483 | -152.470482 | -152.470476 | -152.470468 | -152.470457 |
Table 3: Energy as a function of distance for the PWC LDA functional
Layer separation (Å) | 3 | 3.2 | 3.24 | 3.25 | 3.27 | 3.3 | 3.4 | 3.65 | 3.75 | 3.85 |
PWC energy (Ha) | -151.173444 | -151.174213 | -151.174237 | -151.174238 | -151.174235 | -151.174218 | -151.174083 | -151.173520 | -151.173270 | -151.173027 |
Table 4: Energy as a function of distance for the PBE GGA functional with the TS DFT-D correction
Layer separation (Å) | 3.1 | 3.3 | 3.32 | 3.33 | 3.34 | 3.35 | 3.4 | 3.5 | 3.75 | 3.85 |
PBE w/ TS energy (Ha) | -152.350268 | -152.351181 | -152.351192 | -152.351193 | -152.351192 | -152.351188 | -152.351140 | -152.350905 | -151.173270 | -151.173027 |
Figure 3 – The shape, not only the minimum, of the energy-versus distance curve depends strongly on the choice of functional. Note for PBE and PW91 how slowly the energy increases past its minimum, in contrast to the PBE with TS and the PWC results. Since the layers have zero net charge, this distance can be qualitatively interpreted as a limit to the range of covalent interactions, beyond which only van der Waals interactions (if they were properly present in a functional) would noticeably affect energy. It should be cautioned, however, that the fact that PWC does not behave similarly does not imply that it accounts for the van der Waals interaction.
References
[1]R. Cote, J. Lambert, Y. Barlas, and A. H. MacDonald, “Orbital order in bilayer graphene at filling factor ν=−1” Phys. Rev. B 82, 035445 (2010).
[2] B. Delley, “An all‐electron numerical method for solving the local density functional for polyatomic molecules”, J. Chem. Phys. 92, 508 (1990).
[3] B. Delley, “From molecules to solids with the DMol³ approach”, J. Chem. Phys. 113, 7756 (2000).
[4] J. P. Perdew, K. Burke and M. Ernzerhof, “Generalized gradient approximation made simple”, Phys. Rev. Lett. 77, 3865 (1996).
[5] “Generalized gradient approximations for exchange and correlation: A look backward and forward”,
[6]”Accurate and simple analytic representation of the electron-gas correlation energy”,
[7] A. Tkatchenko, M. Scheffler, “Accurate molecular van der waals interactions from ground-state electron density and free-atom reference data”, Phys. Rev. Lett. 102, 073005 (2009).
[8] K. Berland et al., “van der Waals forces in density functional theory: a review of the vdW-DF method”, Rep. Prog. Phys. 78, 066501 (2015).
[9] H. Rydberg et al., “van der Waals Density Functional for Layered Structures”, Phys. Rev. Lett. 91, 126402 (2003).
[10] M. S. Alam, J. Lin, and M. Saito, “First-Principles Calculation of the Interlayer Distance of the Two-Layer Graphene”, Jpn. J. Appl. Phys. 50, 080213 (2011).
Rappoport D, Crawford NRM, Furche F, Burke K. 2009. “Approximate density functionals: Which should I choose?” In Computational Inorganic and Bioinorganic Chemistry, ed. E Solomon, R King, R Scott, pp. 159–72. New York: Wiley.