DOI:
10.1039/D4FD00073K
(Paper)
Faraday Discuss., 2024,
254, 598-611
Introducing electron correlation in solid-state calculations for superconducting states†
Received
10th April 2024
, Accepted 11th April 2024
First published on 30th August 2024
Abstract
Analyzing the electronic localization of superconductors has been recently shown to be relevant for understanding their critical temperature [Nature Communications, 12, 5381, (2021)]. However, these relationships have only been shown at the Kohn–Sham density functional theory (DFT) level, where the onset of strong correlation linked to the superconducting state is missing. In this contribution, we approximate the superconducting gap in order to reconstruct the superconducting the one-reduced density matrix (1RDM) from a DFT calculation. This allows us to analyse the electron density and localization in the strong correlation regime. The method is applied to two well-known superconductors. Electron localization features along the electron–phonon coupling directions and hydrogen cluster formations are observed for different solids. However, in both cases we see that the overall localization channels are not affected by the onset of superconductivity, explaining the ability of DFT localization channels to characterize the superconducting ones.
Introduction
In the last few years, there has been growing interest in hydrogen-rich materials due to their potential as high-temperature superconductors. H3S,1 YH6,2 LaH10 (ref. 3) and YH9 (ref. 4) have been demonstrated to reach critical temperatures, Tc, all above 200 K at megabar pressures. However, one of the great challenges of modern condensed-matter theory is the prediction of the material-specific properties of superconductors, such as the critical temperature, Tc, and the gap at zero temperature, Δ. The model of Bardeen, Cooper, and Schrieffer (BCS) accounts for the universal behavior of all conventional superconductors, but it is not able to describe material-dependent properties.5 A major step in the first-principles approach to superconductivity is Eliashberg theory.6 This method describes the effects of electron–phonon coupling very accurately. However, the treatment of Coulomb correlations is much harder, so that the electron–electron repulsion is usually introduced by means of an adjustable parameter µ*, chosen to reproduce the experimental Tc.7,8
Combining good accuracy with moderate numerical effort, density functional theory (DFT) is the method of choice for ab initio calculations of macroscopic properties in the weak correlation regime. Following the discovery of high-temperature superconductors, the formal framework of DFT has been generalized to describe strong correlation in the superconducting state.9 Compared to traditional many-body approaches, superconducting DFT (SC-DFT) is absent of empirical parameters. Moreover, given its low computational cost, it has enabled the study of the full k-dependent anisotropy in the superconducting gap. In SC-DFT, the superconducting gap is estimated by making a connection with Eliashberg theory, formulated in terms of Green's functions, and solving a self-consistent equation.9–13 Unfortunately, this procedure is very complex and there are no simple computational tools developed for this purpose available to the public at this moment.
As a faster and alternative approach for predicting Tc, some of the authors have recently shown that the localization/delocalization pattern of a structure correlates with the critical temperature.14 Nevertheless, this quantity has only been analyzed for the normal state (NS) (and conclusions extended to the superconducting one). In order to better understand its ability to reveal Tc, it would be necessary to analyze these delocalization channels in the superconducting state (SC). A first step in this direction has been recently put forward for a model Hamiltonian (1D tight-binding hydrogen chain).15 It was then shown that the effect of the SC Hamiltonian is a temperature-resilient redistribution of the electrons over the base of the Kohn–Sham (KS) orbitals. Making use of this fractional occupation, the authors could propose a reformulation of the electron localization function for superconducting states.15
During this contribution we will overcome the full self-consistent solution of the SC-DFT equations by making use of an approximated BCS occupation number in order to approximate the electron density matrix and electron localization function from a Kohn–Sham calculation. This will allow us to grasp electron reorganization introduced from correlation in well-known superconductors, at DFT-level cost.
Theoretical background
The electron localization function (ELF) was first introduced by Edgecome and Becke to identify regions of localized same-spin electron pairs, or groups of them, in atomic and molecular systems.16 Its original definition is based on the same-spin pair probability as approximated at the Hartree–Fock level. However, a more useful definition was later introduced by Savin17, as the excess of local kinetic energy density (KED) due to Pauli repulsion: | | (1) |
where τ(r) is the positive definite KED of the system, and τvW(r)18 is its form in the von-Weizsacker approximation. For a 3D system, the Thomas–Fermi kinetic energy density takes the form .
Note that for representation reasons, τrel is usually inversed and rescaled, leading to the common expression of the ELF, which varies in the [0,1] range:
| | (2) |
Because the von-Weizsäcker kinetic energy is exact for a bosonic system of the same density
ρ(
r), the term
τrel is a local measure of the excess kinetic energy due to the fermionic nature of the electrons, or what we call the Pauli kinetic energy density. If this is high, it means that electrons in that regions are not paired/localized and the ELF will be small.
19 If the kinetic energy density is not locally increased as an effect of the exclusion principle, in that case we say that electrons are localized, which will be reflected in a high value of the ELF.
Fig. 1 provides an example for CH3OH, where the different regions of electron localization are highlighted by an ELF isosurface: the cores, the C–O, C–H and O–H bonds, and the oxygen lone pairs.
|
| Fig. 1 Isosurfaces of ELF = 0.85 for the CH3OH molecule. | |
This function allows for the description of electron localization within normal DFT. However, in order to recover the effects of the superconducting state, it is necessary to resort to alternative formulations that recover the strong correlation. Staying within the DFT framework, it is possible to resort to superconducting DFT (SC-DFT),9–13 which represents an adapted version of density functional theory (DFT) tailored specifically for superconductors. SC-DFT has demonstrated remarkable accuracy in predicting the critical temperatures (Tc) of traditional superconductors, all without the need for empirical parameters. In order to recover the nature of the superconducting state, the Hamiltonian in SC-DFT incorporates external potentials that account for crucial aspects of superconductors: electron–phonon interaction and Cooper pairs (see the ESI† for a more detailed description).
As recently introduced by some of the authors,15 it is possible to approximately reconstruct the electron density and the kinetic energy densities of the superconducting state, as described in SC-DFT, in terms of the (normal state) KS natural orbitals, their energies, occupancies, the gap, and the temperature (see ESI†):
| | (3) |
where
| | (4) |
Here, we have used the “SC” superscript to differentiate this density from the normal state (NS) electron density. Note that within this approach, the orbital energies
Enk take the form
, with
Δs(
nk) being the superconducting gap, and
φnk(
r) and
ξnk the Kohn–Sham orbitals and energies of the periodic system, respectively. Note that in the normal state limit,
Δs(
nk) → 0, so that we recover the normal state density,
| | (5) |
where
nNSnk = (1 + e
βEi)
−1 is the Fermi–Dirac distribution. Hence, within the SC-DFT framework, we can express the difference in the electron densities when going from the normal to the superconducting state as a difference in the occupation numbers. Fractional occupation numbers are a necessary condition for a SC state, as this is a superposition of states with different numbers of electrons (number symmetry breaking). However, it is not sufficient. Other correlation effects are introduced by the pairing potential. Nevertheless, a similar approach to this has been proved to be useful to describe the correlation effects on the ELF in both the ground
20 and excited states.
21
Similarly, it is possible to show that the positive definite KED of the superconducting state can be defined as,
| | (6) |
The von Weizsacker and Thomas–Fermi approximations to the KEDs are obtained using ρSC(r) from eqn (3),
| | (7) |
and
| | (8) |
respectively. Note that once again within this formalism, the differences between normal and superconducting states are expressed as changes in occupation numbers.
Finally, eqn (6)–(8) enable defining a superconducting ELF,
| | (9) |
This allows the analysis of the spatial distribution of the electrons in a superconductor by means of the electron localization function (ELF). In the same way as in the normal state, this function describes the localization patterns of the electrons in the superconducting state.
Hence, the superconducting ELF should help shine light on how the electrons arrange in the SC state, its similarities and differences with respect to the NS, and how this organization changes from one system to another. These equations have been applied to simple tight binding models;15 however, they have not been used to understand the electronic distribution in real solids so far. This is the aim of this contribution.
Computational approach
Within the SC-DFT approach, the difference between the spatial properties of the normal and the superconducting state lies exclusively in the occupation numbers of the KS orbitals (eqn (4) and (6)). An approximate calculation of these SC occupation numbers will thus allow us to study the electron localization in superconducting states.
This approximation is based on the fact that it is possible to represent the dependence of the gap at zero Kelvin with respect to the energies ξ as an isotropic Lorentzian function:22
| | (10) |
where
ω is a parameter that adjusts the width of the peak, and
N0 is a normalization such that the height of the peak at
ξ = 0 is the gap at
T = 0 K in BCS theory, taking into account the dependence of the gap with respect to the temperature:
| | (11) |
We therefore have a way to obtain the gap for a given
Tc at any temperature.
Fig. 2 offers the profile of the gap function for ω = 0.2 eV at different temperatures, T, for the H3S crystalline system assuming Tc = 200 K. These functions represent the width of the window around the Fermi energy where electrons will form Cooper pairs. One can see how the gap goes to zero everywhere when T = Tc. There, the properties of the normal and superconducting state become the same, as expected from SC-DFT.
|
| Fig. 2 Gap function around the Fermi energy for Tc = 200 K. The dependence on the temperature T is shown by different colored lines. | |
Taking into account the gap in Fig. 2, the occupation numbers in the normal and superconducting states are those shown in Fig. 4. With this, it is possible to obtain the electron density and ELF of the superconducting state from a NS solid-state calculation via a simple change in the fractional occupations. We have performed DFT calculations in well-known superconductor structures with the VASP package.23–26 Geometries were relaxed using the Perdew–Burke–Ernzerhof27 (PBE) functional. The same functional is used to compute the electronic bands. The rapid oscillations of the inner shells of atoms were addressed with the projector augmented-wave method28 with a kinetic energy cutoff of 600 eV. For integration on the reciprocal space, the first Brillouin zone was discretized with a 32 × 32 × 32 Monkhorst–Pack algorithm.
First we performed a self-consistent calculation of the NS state at zero Kelvin and stored the Kohn–Sham Bloch states (WAVECAR file in VASP). Then, we calculated the occupation numbers of the superconducting state for each band and each k-point with eqn (4). The gap is considered to depend on the energy and the temperature according to eqn (10) and (11). Here we assumed that the same equation applies for all k-points and we fixed ω at 0.2 eV in order to fit the results obtained with SC-DFT in ref. 29. Finally, the superconducting state was prepared by reading these occupancies and applying them to the previously stored Kohn–Sham states. An accuracy in the occupation numbers of 10−3 was used, such that only those differing from 0 or 2 (normal state) by more than that threshold value were changed.
Two options were potentially possible: (i) to use directly these Kohn–Sham states, or (ii) to perform a self-consistent calculation (SCF) keeping the occupancies unaltered. In the latter case, special care must be taken to ensure that during the self-consistent procedure the labels of the KS states are not altered, as this would result in an artificial change in the occupancies. For this purpose, we resorted to the VASP keywords ALGO = damped and FERWE. After a quick check, it was realized that although the ELF results were not affected by the convergence, the densities were. Hence, the SCF (option (ii)) was kept. Applications to H3S and LaH10 are analyzed in the coming section.
Applications
H3S: a superconductor characterized by H–S interactions
This compound was synthetized and characterized at high pressure in 2015.30 It has a H3S stoichiometry above 100 GPa, with a Tc of ca. 200 K at 150 GPa.30,31
The high-pressure phase shows a rhombohedral structure, which can be approximated by a body-centred cubic (bcc) structure (at 150 GPa, this means a change in the cell angle from 109.49° in the rhombohedral to 109.47° in the bcc).32,33 As shown in Fig. 3-left, the sulphur atoms occupy the corners of the cube as well as its centre, and the hydrogen atoms occupy the centre of the edges and faces of the cube.
|
| Fig. 3 Left: H3S with space group Imm and lattice constant acell = 3.07 Å. Sulfur atoms occupy Wyckoff position 2a and H atoms occupy position 6b. Right: LaH10 with space group Fmm, acell = 4.42 Å. La in position 4b and H in positions 8c and 32f (with x = 0.1205). | |
Fig. 4 allows comparison of the occupation numbers for the metallic and the superconducting state resulting from the previous gap approximation. For a fixed critical temperature, it can be seen that the occupancies of the normal state deviate from the step function as the temperature increases, softening the transition around the Fermi energy. On the other hand, the superconducting occupation numbers do not suffer great alterations with the temperature. In fact, they tend to resemble the occupations at the critical temperature, showing a larger correlation of the superconducting state in comparison with the normal state below Tc. These occupations are temperature resilient.
|
| Fig. 4 Occupation numbers at different temperatures for Tc = 200 K. To the left, the occupation numbers of the normal state, to the right, the same displayed for the superconducting state. The normal state at T = 0 K is depicted in black. | |
If we analyze the electron density for the SC system at 10 K (Fig. 5-left), we can see that non-nuclear maxima appear in the direction of the S–H bonds.34 The electron delocalization along these channels is also highlighted by the electron localization function (Fig. 5-right). We can see how the region of electron localization around the hydrogens stretches towards the sulfur atoms and leads to a localization region along the S–H direction in a “candy shape”, as expected for an anionic hydrogen. Given the electron density build-up in the S–H direction, the S–H interaction is strengthened and it is expected to behave anharmonically. This was already observed by Errea et al., who pinpointed that anharmonicity hardens the S–H stretching mode.35
|
| Fig. 5 H3S isosurfaces at T = 10 K. Left: ρSC = 0.18 a.u.; right: ELFSC = 0.8. | |
This picture is rather stable upon increasing the temperature (see T = 200 K in the ESI†). In order to better visualize this difference, we have carried out a 3D grid difference analysis of both scalar fields (Fig. 3 in the ESI†). It can be seen that the differences appear mainly at the nuclei and S–H bonding regions. However, interpretation of these data should be carried out with care. The differences are so small that they might be affected by numerical issues. Since the main differences are related to changes in population around the Fermi level (FL), we have thus decided to promote this information by the use of adapted formulae, which only focus on the FL region:
| | (12) |
where FL makes reference to the fact that we are only calculating the quantity around the Fermi level, using thereto a criterion based on the fractional occupations (see Fig. 4 in the ESI
† for a visual depiction of the occupations). In order to verify that our implementation is working, we have carried out the integration of
ρFL over the unit cell and compared it to the integration of
ρ. Whereas 18 electrons are obtained in the unit cell for the superconducting state at
T = 200 K (6.137 electrons on S and 0.94–0.96 on hydrogens with a Bader partition), only 0.185 electrons are obtained when only the levels around the Fermi level are included (0.036 electrons on S and 0.018–0.019 electrons on hydrogen – note that compared to S, a bigger percentage of hydrogen charge is included in this integration, as expected from the contribution of hydrogen atoms to the DOS
14).
A similar FL criterion as in eqn (12) is applied to the local quantities (kinetic energy densities) involved in the ELF definition. This allows us to focus on the correlated region of the superconductor. Results are shown for T = 10 K in Fig. 6. These new expressions allow confirmation of the S–H direction as the one where correlation sets in, both from the ρFL and from the ELFFL viewpoints. It is noteworthy that important electron localization regions appear in the S–H direction (small “tear-like” features in Fig. 6-right).
|
| Fig. 6 Electronic organization around the Fermi level at T = 10 K for H3S. Left: ρFL = 0.0005 a.u.; right: ELFFL = 0.01. | |
Now, we can identify the main electronic reorganizations in the superconductor by plotting the same isosurface as we cool the superconductor from the critical temperature (200 K) to very low temperatures (10 K) (Fig. 7). Notice that, because in SC-DFT the superconducting state converges to the normal state at the critical temperature (where the gap goes to zero), ρ200KFL and ELF200KFL, which have been defined for the superconducting state, also correspond to the normal state quantities around the Fermi level at 200 K.
|
| Fig. 7 H3S isosurfaces at 10 K (blue), 100 K (orange) and 200 K (red). Left: ρ = 0.08 a.u.; right: ρFL = 0.0015 a.u. | |
Although the overall density values (ρ) do not change considerably with temperature (Fig. 7-left), those of ρFL do, as can be seen in Fig. 7-right.‡ Decreasing the temperature below Tc induces changes in the localization of electrons near the Fermi level, which is favoured around the S–H bond stretching direction. Similarly, the isosurfaces of the total ELF do not show any difference with the temperature, whereas those of ELFFL do suffer some changes, while keeping a similar topology (see Fig. 5 in the ESI†).
LaH10
In 2017, a systematic high-pressure ab initio structural search revealed the stability of LaH10 at high pressure in the fcc phase36 and a very high coordination of La by hydrogen atoms (see Fig. 3-right). This structure was subsequently shown to have a critical temperature above 250 K.3 It is also worth noting that this compound is known to show very strong quantum effects.
Given the results for H3S, we will directly concentrate on the density and ELF around the Fermi level (eqn (12)). Fig. 8 shows how the electron density where correlation takes place (ρFL) is mainly located around the hydrogen atoms instead of the bonds. Electron localization (ELFFL) is also mainly centered in H atoms, with important delocalization channels in the H4 units.
|
| Fig. 8 LaH10 isosurfaces around the Fermi surface at 10 K. Left: ρFL = 0.0008 a.u.; right: ELFFL = 0.001. | |
The isosurfaces of the overall density show, once again, a temperature-independent character (see Fig. 9-left). On the other hand, while the isosurfaces of ρFL in Fig. 9-right have a similar shape, they do evidence some small changes in their values, leading to a square pattern. Those differences become more important with the decrease in temperature at the atomic sites inside the H4 units, revealing a change of character of the electrons in that region below Tc, when the system is in the superconducting regime.
|
| Fig. 9 LaH10 isosurfaces at 10 K (blue), 100 K (orange) and 250 K (red). Left: ρ = 0.08 a.u.; right: ρFL = 0.0015 a.u. | |
Localization channel topologies
Overall, no big changes have been observed in the full electron density and ELF. Hence, we have also analyzed their critical points both for H3S and LaH10 at different temperatures. We are especially interested in saddle points, since the values of the ELF at some of these points have been correlated to Tc in ref. 14. Looking at Fig. 9 in the ESI,† it can be seen that the saddle point position and ELF values are extremely stable with the changes in the fractional occupation given by eqn (4). This confirms the possibility of using the normal state to characterize the delocalization of the SC state and hence the equation that correlates Tc with delocalization channels.
Conclusions
Making use of the recently introduced SC-ELF, which allows calculation of the electron localization function for a superconducting state as provided by SC-DFT,15 we have introduced an approximation to calculate it from NS-DFT. By approaching the SC gap via a Lorentzian function, it was possible to calculate the expected occupations for a given gap, Tc and T. Hence, the SC electron localization can be obtained from a common solid-state calculation, where correlation is introduced as a redistribution of electrons around the Fermi level. We have applied this approach to two typical superconductors, H3S and LaH10. Given the fact that the electronic reorganization only happens around the Fermi energy, very little difference appears in the electron density and the electron localization function. Hence, we have focused in the region where the fractional occupation changes in order to understand the electronic structure. With this trick, we have been able to visualize the changes upon the set up of the superconducting state, as well as with respect to temperature. Whereas in H3S the main differences appear along the stretching and bending directions, the main differences in LaH10 show up in a degeneracy in the H4 region, which is boosted as temperature decreases. Since these directions could be related to the strong anharmonicity, it will be the aim of future work to analyze deformations along these directions. Overall, no important differences were observed in the values and positions of the critical points (unless directly focussed at the Fermi level). This explains the ability of DFT (de)localization channels to reflect (de)localization in the superconducting state.
Conflicts of interest
There are no conflicts to declare.
Acknowledgements
We would like to acknowledge support by ECOS-Sud C17E09 and C21E06, and the Agence Nationale de la Recherche under grant ANR-22-CE50-0014. This research was supported by the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Programme (grant agreement no. 810367), project EMC2. CC acknowledges ANID for the grants ECOS210019 and FONDECYT 1220366, and the Center for the Development of Nanosciences and Nanotechnology, CEDENNA AFB 220001. W. M. acknowledges the support of ANID Chile through the Doctoral National Scholarship no. 21211501. This research was supported by the High-Performance supercomputing Center (HPC) of Irene/TGCC and Jean Zay/Idris of GENCI (France) under projects A0160915069, and A0160815101 as well as the supercomputing infra-structure of the NLHPC (ECM-02).
References
- A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Ksenofontov and S. I. Shylin, Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system, Nature, 2015, 525, 73 CrossRef CAS PubMed.
- I. A. Troyan,
et al., Anomalous High-Temperature Superconductivity in YH6, Adv. Mater., 2021, 33, 2006832 CrossRef CAS PubMed.
- A. Drozdov, P. Kong, V. Minkov, S. Besedin, M. Kuzovnikov, S. Mozaffari, L. Balicas, F. Balakirev, D. Graf and V. Prakapenka, others Superconductivity at 250 K in lanthanum hydride under high pressures, Nature, 2019, 569, 528–531 CrossRef CAS PubMed.
- P. Kong, V. S. Minkov, M. A. Kuzovnikov, A. P. Drozdov, S. P. Besedin, S. Mozaffari, L. Balicas, F. F. Balakirev, V. B. Prakapenka, E. Greenberg and M. I. Eremets, Superconductivity up to 243 K in yttrium hydrides under high pressure, Nat. Commun., 2021, 12, 5075 CrossRef CAS PubMed.
- J. Bardeen, L. N. Cooper and J. R. Schrieffer, Theory of Superconductivity, Phys. Rev., 1957, 108, 1175–1204 CrossRef CAS.
- G. M. Eliashberg, Interactions between electrons and lattice vibrations in a superconductor, Sov. Phys.–JETP, 1960, 11, 696–702 Search PubMed.
- P. Morel and P. W. Anderson, Calculation of the Superconducting State Parameters with Retarded Electron-Phonon Interaction, Phys. Rev., 1962, 125, 1263–1271 CrossRef.
- D. J. Scalapino, J. R. Schrieffer and J. W. Wilkins, Strong-Coupling Superconductivity. I, Phys. Rev., 1966, 148, 263–279 CrossRef.
- L. N. Oliveira, E. K. U. Gross and W. Kohn, Density-Functional Theory for Superconductors, Phys. Rev. Lett., 1988, 60, 2430–2433 CrossRef PubMed.
- L. N. Oliveira, E. K. U. Gross and W. Kohn, Density-Functional Theory for Superconductors, Phys. Rev. Lett., 1988, 60, 2430–2433 CrossRef PubMed.
- M. Lüders, M. A. L. Marques, N. N. Lathiotakis, A. Floris, G. Profeta, L. Fast, A. Continenza, S. Massidda and E. K. U. Gross, Ab initio theory of superconductivity. I. Density functional formalism and approximate functionals, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 024545 CrossRef.
-
A. Sanna, Introduction to Superconducting Density Functional Theory, in The Physics of Correlated Insulators, Metals, and Superconductors, ed. Pavarini, E., Koch, E., Scalettar, R., and Martin, R., Forschungszentrum Jülich, 2017, vol. 7 Search PubMed.
- A. Sanna, C. Pellegrini and E. K. U. Gross, Combining Eliashberg Theory with Density Functional Theory for the Accurate Prediction of Superconducting Transition Temperatures and Gap Functions, Phys. Rev. Lett., 2020, 125, 057001 CrossRef CAS PubMed.
- F. Belli, T. Novoa, J. Contreras-García and I. Errea, Strong correlation between electronic bonding network and critical temperature in hydrogen-based superconductors, Nat. Commun., 2021, 12, 5381 CrossRef CAS PubMed.
-
M. di Mauro, I. Errea, B. Braïda, T. Novoa and J. Contreras-García, Molecularity: a fast and efficient criterion for probing superconductivity, arXiv, 2024, preprint, arXiv:2403.07584, DOI:10.48550/arXiv.2403.07584.
- A. D. Becke and K. E. Edgecombe, A simple measure of electron localization in atomic and molecular systems, J. Chem. Phys., 1990, 92, 5397 CrossRef CAS.
- A. Savin, R. Nesper, S. Wengert and T. F. Fässler, ELF: The Electron Localization Function, Angew Chem. Int. Ed. Engl., 1997, 36, 1808–1832 CrossRef CAS.
- C. F. von Weizsäcker, Zur Theorie de Kernmassen, Z. Phys., 1935, 96, 431 CrossRef.
- T. Novoa, J. Contreras-García, P. Fuentealba and C. Cárdenas, The Pauli principle and the confinement of electron pairs in a double well:Aspects of electronic bonding under pressure, J. Chem. Phys., 2019, 150, 204304 CrossRef PubMed.
- F. Feixas, E. Matito, M. Duran, M. Solà and B. Silvi, Electron localization function at the correlated level: a natural orbital formulation, J. Chem. Theory Comput., 2010, 6, 2736–2742 CrossRef CAS PubMed.
- A. Echeverri, M. Gallegos, T. Gómez, A. M. Pendás and C. Cárdenas, Calculation of the ELF in the excited state with single-determinant methods, J. Chem. Phys., 2023, 158, 174101 CrossRef CAS PubMed.
-
M. Tinkham, Introduction to Superconductivity; International Series in Pure and Applied Physics, McGraw Hill, 1996 Search PubMed.
- G. Kresse and J. Hafner, Ab initio molecular dynamics for liquid metals, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 47, 558(R) CrossRef PubMed.
- G. Kresse and J. Hafner, Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 49, 14251 CrossRef CAS PubMed.
- G. Kresse and J. Furthmüller, Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set, Comput. Mater. Sci., 1996, 6, 15 CrossRef CAS.
- G. Kresse and J. Furthmüller, Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set, Phys. Rev. B: Condens. Matter Mater. Phys., 1996, 54, 11169 CrossRef CAS PubMed.
- J. P. Perdew, K. Burke and M. Ernzerhof, Generalized Gradient Approximation Made Simple, Phys. Rev. Lett., 1996, 77, 3865–3868 CrossRef CAS PubMed.
- P. E. Blöchl, Projector augmented-wave method, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed.
- A. Linscheid, A. Sanna, A. Floris and E. K. U. Gross, First-Principles Calculation of the Real-Space Order Parameter and Condensation Energy Density in Phonon-Mediated Superconductors, Phys. Rev. Lett., 2015, 115, 097002 CrossRef CAS PubMed.
- A. P. Drozdov, M. I. Eremets, I. A. Troyan, V. Ksenofontov and S. I. Shylin, Conventional superconductivity at 203 kelvin at high pressures in the sulfur hydride system, Nature, 2015, 525, 73–76 CrossRef CAS PubMed.
- M. I. Eremets, V. S. Minkov, A. P. Drozdov, P. P. Kong, V. Ksenofontov, S. I. Shylin, S. L. Budko, R. Prozorov, F. F. Balakirev, D. Sun, S. Mozaffari and L. Balicas, High-Temperature Superconductivity in Hydrides: Experimental Evidence and Details, J. Supercond. Novel Magn., 2022, 35, 965–977 CrossRef CAS.
- I. Errea, M. Calandra, C. J. Pickard, J. R. Nelson, R. J. Needs, Y. Li, H. Liu, Y. Zhang, Y. Ma and F. Mauri, Quantum hydrogen-bond symmetrization in the superconducting hydrogen sulfide system, Nature, 2016, 532, 81–84 CrossRef CAS PubMed.
- R. Bianco, I. Errea, M. Calandra and F. Mauri, High-pressure phase diagram of hydrogen and deuterium sulfides from first principles: Structural and vibrational properties including quantum and anharmonic effects, Phys. Rev. B, 2018, 97, 214101 CrossRef CAS.
- A. M. Pendás, M. A. Blanco, A. Costales, P. M. Sánchez and V. Luaña, Non-nuclear Maxima of the Electron Density, Phys. Rev. Lett., 1999, 83, 1930–1933 CrossRef.
- I. Errea, M. Calandra, C. J. Pickard, J. Nelson, R. J. Needs, Y. Li, H. Liu, Y. Zhang, Y. Ma and F. Mauri, High-Pressure Hydrogen Sulfide from First Principles: A Strongly Anharmonic Phonon-Mediated Superconductor, Phys. Rev. Lett., 2015, 114, 157004 CrossRef PubMed.
- H. Liu, I. I. Naumov, R. Hoffmann, N. W. Ashcroft and R. J. Hemley, Potential high-T(c) superconducting lanthanum and yttrium hydrides at high pressure, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 6990–6995 CrossRef CAS PubMed.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4fd00073k |
‡ To understand why the isovalues of ρFL are taken to be as low as 0.01 a.u., one must take into account the total number of electrons in the energy window, which is only 0.185. |
|
This journal is © The Royal Society of Chemistry 2024 |
Click here to see how this site uses Cookies. View our privacy policy here.