Second-order perturbative correlation energy functional in the ensemble density-functional theory
Abstract
We derive the second-order approximation (PT2) to the ensemble correlation energy functional by applying the Görling-Levy perturbation theory on the ensemble density-functional theory (EDFT). Its performance is checked by calculating excitation energies with the direct ensemble correction method in 1D model systems and 3D atoms using numerically exact Kohn-Sham orbitals and potentials. Comparing with the exchange-only approximation, the inclusion of the ensemble PT2 correlation improves the excitation energies in 1D model systems in most cases, including double excitations and charge-transfer excitations. However, the excitation energies for atoms are generally worse with PT2. We find that the failure of PT2 in atoms is due to the two contributions of an orbital-dependent functional to excitation energies being inconsistent in the calculations. We also analyze the convergence of PT2 excitation energies with respect to the number of unoccupied orbitals.
I Introduction
The ensemble density functional theory (EDFT)T79 ; GOKb88 ; GOK88 ; OGKb88 is a formally exact excited-state extension to the highly successful density-functional theory (DFT)HK64 ; KS65 ; FNM03 . EDFT is shown to be promising in solving some of the difficult problems in the widely used time-dependent density-functional theory (TDDFT)RG84 ; C96 ; MMNG12 ; U12 such as the double excitationTH00 ; LKQM06 ; EGCM11 ; M16 charge-transfer excitationTAHR99 ; DWH03 ; DH04 ; M16 ; M17 and so on, and recently there is a renewed interest in EDFTF15 ; ADKF17 ; GP17 ; YPBU17 ; GKP18 ; SB18 ; DF19 ; GP19 ; LFM19 ; F20 ; G20 ; GSP20 ; LF20 ; MSFL20 ; FLNC21 ; GKP21 ; HH21 .
One need to approximate the ensemble Hartree-exchange-correlation (Hxc) energy functional for practical calculations with EDFT. Unlike ground state DFT, it is more difficult to develop approximations as explicit density functionals, since ensemble functionals have an extra dependence on ensemble weights. Furthermore, the ensemble Hartree and exchange energies are not separated naturallyGSP20 and an involved process is required to utilize ground-state experiences in approximationsGKP21 , which is another difficulty in developing approximated functionals in EDFT. Only a few approximations are available in EDFTOGKb88 ; YPBU17 ; ADKF17 ; DMSF18 ; F20 ; LF20 ; MSFL20 ; GKP21 , and many are not built for general usage.
One can avoid most of the difficulties associated with developing explicit density functionals by using orbital-dependent approximations. The most simple one is the exact-exchange approximation known as symmetry-eigenstate Hartree-exchangePYTB14 ; YTPB14 ; YPBU17 or ensemble exact-exchange (EEXX)DMSF18 ; GKP18 ; SB18 ; SF18 ; DF19 ; SF20 in literature, the form of which is simply the definition of the ensemble Hartree-exchange energy. EEXX in general performs well, but the errors in the excitation energies can be large in some casesYPBU17 ; GKP18 ; SB18 due to the lack of correlation. An approximated ensemble correlation energy functional is needed to improve the accuracy.
In this paper, we derive an orbital-dependent second-order perturbative approximation (PT2) to the ensemble correlation energy functional by applying the Görling-Levy perturbation theory (GLPT)GL93 ; GL94 on EDFT. Since EEXX can be obtained from the first order correction to the non-interacting Kohn-Sham (KS) ensemble energy, the PT2 correlation is the most simple natural extension to EEXX. Since the optimized effective potential (OEP)KLI92 in EDFT is more complicated than in DFT due to the dependence on ensemble weights, we test the PT2 correlation with the direct ensemble correction (DEC)YPBU17 instead of doing self-consistent ensemble Kohn-Sham (EKS) calculations. We find that PT2 generally improves the accuracy of excitation energies in 1D model systems including double excitations and charge-transfer excitations. The performance of PT2 is unsatisfactory in 3D atoms, however, and we discuss the possible reasons. The convergence of PT2 excitation energies with respect to orbitals is discussed as well.
II Theory
II.1 Background
We use atomic units [] unless otherwise specified. The interacting system is described by the Schödinger equation:
(1) |
where the Hamiltonian is with , and being the kinetic, external potential and electron-electron interaction potential operators, and denotes a set of degenerate states (‘multiplet’) and an individual state in the multiplet. The multiplets are ordered by energy with being the ground state. The ordering of wavefunctions within a multiplet is arbitrary but fixed. An ensemble in EDFT contains consecutive multiplets from the ground state to a highest-energy multiplet . Each state in multiplet is assigned a weight satisfying for .GOKb88 We require for simplicity, where is the degeneracy of multiplet . We denote all weights as in the following.
The ensemble density and energy are defined as
(2) |
and
(3) |
where the ensemble density matrix is
(4) |
and .
EDFT proves the one-to-one correspondence betweeen and the external potential given GOKb88 , so is a functional of . One can define a non-interacting EKS system with the same as the interacting system, with the EKS orbitals satisfying
(5) |
The EKS density matrix is
(6) |
where is an EKS wavefunction corresponding to the state of the interacting system. The existence of an adiabatic connectionPY89 ; FNM03 between the interacting and EKS systems is assumed. if is exact.
Unlike in ground-state DFT, is not necessarily a single Slater determinant of EKS orbitals since the EKS degenercies are in general greater than or equal to the corresponding interacting ones.PYTB14 We require to have the same spatial and spin symmetries as PYTB14 to distinguish the otherwise degenerate EKS states by linearly combine the EKS Slater determinants:
(7) |
where and denote the EKS multiplet and its degeneracy corresponding to interacting multiplet , denotes the EKS Slater determinant in EKS multiplet , and is the linear combination coefficient.
can be decomposed as
(8) |
The ensemble Hartree-exchange energy is
(9) |
which can be evaluated using the Slater-Condon rulesL14 . The ensemble correlation energy is defined by Eq. (8) and must be approximated in practice. The EKS potential is determined variationally by as
(10) |
EDFT calculates excitation energies either by subtracting ensemble energies or by taking derivative of with respect to the weightsGOK88 ; DF19 , and both requires solving Eq. (5) self-consistently. The DEC method of our previous workYPBU17 can be used to calculate excitation energies without extra self-consistent calculations other than the ground-state one. The DEC method uses a special type of ensemble defined as
(11) |
where is the total number of states in the ensemble, and is the weight parameter. The excitation energy can be written as a correction to the ground-state KS excitation energy:
(12) |
The ensemble density reduces to the ground-state density at , so Eq. (12) only requires a self-consistent ground-state KS calculation.
II.2 GLPT of EDFT
GLPTGL93 ; GL94 is developed for ground-state DFT as a systematic approach for developing orbital-dependent xc functionals. The perturbation is along the adiabatic connection with the KS Hamiltonian as the zeroth order and density fixed. The perturbation is the difference between the interacting and KS Hamiltonians , so one can extract approximations to and from the perturbative corrections. GLPT differs from the Rayleigh-Schrödinger perturbation theory since the perturbation is dependent on the coupling strength, and perturbative corrections to the density vanish at all orders. GLPT yields the exact exchange (EXX) approximation which is widely used in both DFT and TDDFT.
We apply GLPT on EDFT in this paper. GLPT has to be applied along a different adiabatic connection with the ensemble density fixed, with the EKS Hamiltonian as the zeroth-order Hamiltonian. The Schrödinger equation with the electron-electron interaction scaled by the coupling constant is
(13) |
where , and the existence of is assumed. The scaled ensemble energy is
(14) |
where . Treating as the perturbation and expanding as , we obtain the formula for the ensemble PT2 correlation energy:
(15) |
Details of the derivations is available in Appendix A.
Degenerate perturbation theoryD61 ; L14 requires that the zeroth-order wavefunctions of an EKS multiplet diagonalizes , and this is satisfied by the EKS wavefunctions in Eq. (7) since they are eigenfunctions of the spatial and spin symmetry operators that commute with . in Eq. (15) is then
(16) |
where is the EKS energy of EKS multiplet .
The OEP equations for the perturbative corrections of the potential can be derived by requiring the perturbative corrections of the density to vanishGL93 . However, these equations involve weighted sums over states and are more difficult to solve than their counterparts in ground-state DFT. Approximations to can be derived and are provided in the supplemental materialsupplemental . While approximations to can be derived similarly in principle, it is more cumbersome as it would contain functional derivatives of , so self-consistent EKS calculation with ensemble PT2 correlation can be unhandy. We also encounter divergences related to the ensemble derivative discontinuityL95 in the OEP calculations. We therefore choose to use the DEC method to assess the performance of the ensemble PT2 correlation in this paper to avoid solving the EKS equation self-consistently.
II.3 Ensemble PT2 correlation in the DEC method
The orbital-dependent Eqs. (9) and (15) are implicit density functionals. In the DEC method, these equations represent and instead of and required by Eq. (12). Since the -dependences of the functional and of the EKS density are inseparable in these equations, we calculate the -derivatives in Eq. (12) byYPBU17
(17) |
where is the ground-state Hxc potential, and the shorthand means
(18) | ||||
(19) |
with and being a complete set of one-electron orbitals and the ground-state KS orbitals respectively, and being the EKS density of the ensemble defined by Eq. (11) with being the highest energy multiplet constructed with orbitals .
The orbitals are held fixed for the derivatives in Eq. (17) to avoid the otherwise required self-consistent EKS calculations, since the -dependence of and originates from that of the EKS orbitals. is then the direct differentiation of Eqs. (9) and (15). Its PT2 part is
(20) |
where , , and the notation means
(21) | ||||
(22) |
We ignore the second term of Eq. (20) in the calculations in this paper since it is usually small (Appendix B). This approximation allows us to greatly simplify the calculation by using a bi-ensemble comprised of the ground state and the -th multiplet, which is equivalent to the full ensemble due to the functional form of the first term of Eq. (20).YPBU17 Eq. (20) apparently indicates that the number of orbitals required for convergence determines the scaling of the computational cost.
III Results and discussion
We carry out DEC calculations with EEXX and EEXX+PT2 functionals on 1D model systems for double excitation and charge-transfer excitation and on 3D He and Be atoms. TDDFT results are reported as well for comparison. We use the numerically exact KS ground state and the exact and in the calculations. For EEXX calculations, the errors in excitation energies would exclusively reflect the performances of the method; for EEXX+PT2, however, the results contain higher-order terms due to the exact , which is used to avoid solving for the OEP .
The orbital-dependent ensemble Hxc energy functional contributes to the excitation energy in two different ways corresponding to the two terms of Eq. (17), one from the -dependence of the orbital-dependent functional, and another one from the Hxc potential. This is relevant to self-consistent EKS calculations as well since the excitation energies also depend on GOK88 . For the EEXX functional, the two terms of Eq. (17) are found to have the same orders of magnitudeYPBU17 , but this is not true for the correlation part. We demonstrate this by reporting excitation energies of EEXX+ and EEXX+, where and denote the correlation contribution to the two terms of Eq. (17) respectively. The EEXX+PT2 result can be written as EEXX++ in this notation.
In ground-state DFT, a common approximation to the PT2 correlation energy is to neglect the ‘singly-excited’ matrix elements where the Slater determinants of the bra and ket differ by one orbital (such as in the B2PLYPG06 double hybrid functional). We test the performance of this approximation as well and denote it as PT2(no single).
III.1 1D model systems
We solve the interacting Schrödinger equation for 1D model systems by direct diagonalization on a grid. All 1D model systems have two electrons so that the exact can be obtained as . The second term of Eq. (20) vanishes for two-electron systems, so the 1D PT2 calculations are not approximated. Numerically exact KS ground state is obtained by inverting the ground-state KS equationYTPB14 .
We use an 1D Hooke’s atom as the model system for double excitation MZCB04 ; YPBU17 . The system has and contact interaction , where is the Dirac function. The doubly excited states in this system are close to singly excited states due to the weak electron-electron interaction. We use a large grid with grid-point spacing and to ensure that the effect of the grid boundary on all orbitals is negligible. Table 1 lists the errors in the excitation energies.
(mH) | |||||||
---|---|---|---|---|---|---|---|
DEC | TDDFT | TDDFT | |||||
EEXX | + | + | +PT2 | AEXX | DSPA | ||
1(1,2) | 1.389 | 2.240 | 1.350 | 2.201 | 2.401 | 1.041 | 1.400 |
2(2,2) | 17.24 | 4.565 | 17.16 | 4.487 | 5.001 | N/A | -1.900 |
3(1,3) | -16.65 | -1.929 | -18.27 | -3.550 | -3.554 | -16.94 | 2.200 |
4(2,3) | 28.34 | 19.85 | 26.68 | 18.19 | 18.15 | N/A | -1.800 |
5(1,4) | -26.60 | -15.78 | -28.40 | -17.58 | -17.05 | -26.55 | 1.600 |
shows the error in excitation energies due to the lack of correlation. With PT2 correlation, EEXX+PT2 in Table 1 improves the accuracy of excitation energies in most cases, including both doubly-excited states. Due to the perturbative nature of the ensemble PT2 correlation, however, improvement in accuracy is not guaranteed, as seen in the first excited state.
The correlation contribution of the two terms of Eq. (17) can be obtained from Table 1 by subtracting column 1(EEXX) from columns 2(EEXX+) and 3(EEXX+), which shows that the correlation effect in the 1D Hooke’s atom is dominated by the first term of Eq. (17). This dominance allows us to estimate the PT2 errors in excitation energies with this first term alone while ignoring the effect of using exact : comparing (the effect of this term of the exact functional) and (that of the PT2 approximation) yields , , , , for the 5 states in Table 1. These large errors for PT2 indicate that higher-order terms are probably needed to properly describe the ensemble correlation energy.
We use a system with two potential wells in a box as the model system for charge-transfer excitation. The external potential is
(23) |
and the electron-electron interaction is the soft-Coulomb potential
(24) |
We set in order to carry out adiabatic TDDFT calculations with the 1D local-density approximation (LDA)HFCV11 kernel for comparison. We orbtain the numerically exact KS system on a grid with grid-point spacing and . Excitations to the first (triplet) and second (singlet) excited states are charge-transfer excitations, both of which correspond to one electron transferring from the right potential well to the left one. The distance between the two wells lead to negligible overlap between the involved KS orbitals, so the TDDFT couplings between the initial and final states vanish for local or semi-local approximated xc kernels. The corresponding TDDFT excitation energies would be very close to the KS ones.
(mH) | ||||||
---|---|---|---|---|---|---|
DEC | TDDFT | |||||
(mH) | EEXX | + | + | +PT2 | ALDA | |
-53.38 | -53.38 | -53.18 | -0.1011 | 0.1027 | 0.2205 | -53.38 |
We list the error in the KS excitation energy in Table 2. Since the singlet-triplet splitting between the first two states is very small, Table 2 only lists the calculation result for the first excited state. The correlation effect dominates the correction to for the charge-transfer box as well. Both DEC/EEXX and TDDFT/ALDA fail to correct the KS excitation energy due to non-overlapping orbitals leading to vanishing matrix elements. Unlike the 1D Hooke’s atom, the correlation contribution from the -dependence of the orbital-dependent [first term of Eq. (17)] only changes the result slightly due to either vanishing matrix elements or large energy differences in Eq. (20). The contribution from dominates the correlation effect. This suggests that the two correlation contributions in Eq. (17) can be regarded as representing the local and the non-local correlation effects, respectively.
We also calculate the excitation energies of an 1D flat box, where both terms of Eq. (17) contribute significantly to the correlation effect (Appendix C). We find that the inclusion of ensemble PT2 correlation improves the exchange-only EEXX excitation energies for most of the states of the tested 1D systems. However, the PT2 correction to EEXX is not always in the correct direction. The magnitude of improvement also varies a lot for different states.
III.2 3D atoms
We carry out DEC calculations on He and Be atoms using numerically exact KS potentialsUG93 ; UG94 , and compare the excitation energies to the experimental valuesNIST_ASD . Since we solve the ground-state KS orbitals on a radial grid, the higher-energy orbitals are not well represented due to tails being truncated at the grid boundary. This leads to the ordering of higher-energy orbitals being different from that of the exact KS system (see He PT2 convergece curve in supplemental materialsupplemental ). Nevertheless, this should not affect the PT2 results since the orbitals form a complete basis set with respect to the grid, and convergence is achieved fairly quickly for PT2 (Sec. III.3).
(mH) | ||||||
DEC | TDDFT | |||||
EEXX | + | + | +PT2 | ALDA | ||
He atom | ||||||
3S(1s2s) | -4.996 | -23.93 | 9.092 | -9.841 | 9.318 | -8.724 |
1S(1s2s) | 11.25 | 2.476 | 25.34 | 16.56 | 23.16 | 10.61 |
3P(1s2p) | -0.9324 | -22.22 | 12.91 | -8.376 | 13.48 | -16.41 |
1P(1s2p) | 5.355 | -21.59 | 19.20 | -7.748 | 19.45 | -3.247 |
3S(1s3s) | -1.075 | -24.14 | 13.48 | -9.581 | 14.65 | -0.3482 |
1S(1s3s) | 2.574 | -17.12 | 17.13 | -2.566 | 18.57 | 4.217 |
3P(1s3p) | -0.1077 | -24.22 | 14.33 | -9.776 | 15.21 | -2.276 |
3D(1s3d) | 0.3006 | -24.97 | 15.02 | -10.24 | 15.92 | -1.036 |
1D(1s3d) | 0.3715 | -25.09 | 15.09 | -10.37 | 15.95 | -0.6598 |
1P(1s3p) | 1.712 | -23.70 | 16.15 | -9.257 | 17.17 | -0.0629 |
3S(1s4s) | -0.2530 | -24.47 | 14.39 | -9.822 | 15.50 | 0.2161 |
1S(1s4s) | 1.140 | -21.70 | 15.79 | -7.049 | 17.20 | 2.354 |
Be atom | ||||||
3P(1s22s2p) | -38.06 | 15.46 | -38.76 | 14.76 | 17.59 | -46.89 |
1P(1s22s2p) | 4.067 | 52.91 | 3.368 | 52.22 | 52.11 | 6.473 |
3S(1s22s3s) | -4.407 | 42.58 | -36.84 | 10.14 | 17.30 | -13.64 |
1S(1s22s3s) | 6.179 | 41.58 | -26.26 | 9.145 | 11.86 | 1.836 |
1D(1s22p2) | 10.77 | 17.78 | 9.374 | 16.38 | 17.55 | N/A |
3P(1s22s3p) | -4.778 | 45.19 | -37.64 | 12.33 | 20.00 | -13.32 |
3P(1s22p2) | -35.49 | 19.19 | -36.88 | 17.79 | 20.51 | N/A |
Table 3 lists the errors in the excitation energies for He and Be. The performance of the ensemble PT2 correlation energy functional is disappointing in atoms. Unlike 1D systems, EEXX+PT2 excitation energies are in general worse than those of EEXX despite a few exceptions. This includes the He atom as well, which is a two-electron system as the 1D model systems. The ground-state PT2 correlation is also found to be problematic when used directlyMWY05 , and the failure is attributed to the algebraic structure of KS orbitals. However, this is unlikely the reason for the failure seen in Table 3 since the DEC method does not involve self-consistent ensemble PT2 calculations.
The reason for larger error of EEXX+PT2 might be related to the two terms of Eq. (17) being incompatible. For EEXX, these two terms are close in magnitude and have different signsYPBU17 . Due to their cancellation, the resulting corrections to KS excitation energies have much smaller magnitudes. If these two terms are incompatible, the error can be large since Eq. (17) subtracts two large numbers, which can be seen from the EEXX+ results. This might be the case for EEXX+PT2 shown in Table 3, where we use the exact in Eq. (17) instead of the OEP . As an example to demonstrate the effect of incompatible and , using the LDA correlation potentialPW92 in an EEXX calculation of the first excitation energy of He yields an error of -55.98 mH, which is much greater than the one in Table 3. Since the correlation is dominated by one of the two terms for the 1D Hooke’s atom and charge-transfer box, the error due to this incompatibility is small (but the error due to PT2 itself can still be large), so the overall accuracy of EEXX+PT2 only depends on the dominant term. This problem is a disadvantage of orbital-dependent Hxc energy functionals and does not affect explicit density functionals. This failure also indicates that the OEP is not close to the exact for atoms, which agrees with the large relative errors due to PT2 in the 1D Hooke’s atom, showing that higher-order terms are probably needed for correlation.
EEXX+PT2(no single) results for the 1D Hooke’s atom and 1D charge-transfer box are close to the DEC/EEXX+PT2 results, but the error is larger for atoms. Despite this, the error introduced by this approximation is still small comparing with the excitation energies (see supplemental materialsupplemental and Ref. NIST_ASD ), suggesting that the ground-state approximation of neglecting singly-excited matrix elements is also applicable in EDFT.
III.3 Convergence for PT2
Although the - and -sums in Eq. (20) sums over all KS states, these sums can be truncated by only including a finite number of KS orbitals, since the contributions from higher-energy orbitals vanish due to the KS energy differences in the denominators. We check the convergence of the EEXX+PT2 excitation energies with respect to the number of KS orbitals, which are plotted in Fig. 1 for the 1D Hooke’s atom.

Fig. 1 shows that convergence is achieved for all EEXX+PT2 excitation energies of the 1D Hooke’s atom with only a few KS orbitals. Higher-energy orbitals have a small impact on excitation energies as the KS energy differences become large. Due to the near-degeneracies between KS states and and between and , the changes of the and excitation energies are big when orbitals 3 and 4 are included, respectively. The excitation energies of the 1D charge-transfer box and 1D flat box (see supplemental materialsupplemental ) also converge quickly with respect to the number of orbitals, since the orbital energies of these systems increase rapidly.

Fig. 2 plots the EEXX+PT2 excitation energies of the Be atom with different numbers of KS orbitals included. The He atom have similar trends (see supplemental materialsupplemental ). Comparing with 1D cases, higher-energy orbitals can affect the excitation energies significantly since the orbital energies form a Rydberg series. More orbitals are needed to achieve convergence in 3D atoms.
The excitation energies in Fig. 2 only change notably when orbitals with certain angular momentum quantum numbers are included, since otherwise many of the newly included matrix elements in Eq. (20) would vanish due to symmetry. Therefore one can achieve convergence with fewer orbitals for atoms by ignoring the orbitals whose angular momentum quantum numbers differ too much from those of the orbitals in the KS ensemble. As an example, the EEXX+PT2 excitation energy of the 1S(1s22s3s) state of Be calculated with all orbitals (upto 6g) only differ from that calculated with only s and p orbitals (upto 6p) by 0.007 mH.
IV Conclusion
We apply GLPT on EDFT and derive the orbital-dependent ensemble PT2 correlation energy functional. We test the performance of the ensemble PT2 functional on 1D model systems and 3D atoms using the DEC method in our previous work. For 1D model systems, inclusion of improves the exchange-only EEXX results in general, but the errors of EEXX+PT2 are at the same orders of magnitude as those of EEXX. We show that the ensemble correlation effect on excitation energies of an orbital-dependent functional consists of contributions from the -dependence of and from . The ratio of these two can differ a lot for different systems, and double excitation and charge-transfer excitation represent two extremes, which are dominated by the -dependence of and by respectively. Calculation results suggest that PT2 may not be enough to properly approximate the ensemble correlation effect, which may need higher order terms.
For 3D He and Be atoms, the EEXX+PT2 results are generally worse than EEXX. The failure of EEXX+PT2 is probably related to the inconsistency between the functional and the exact used in the calculation. This problem only affects orbital-dependent functionals, and the OEP of is needed to completely assess the performance of the ensemble PT2 correlation. However, we find the ensemble PT2 OEP more difficult to evaluate than in ground-state due to ensemble sums and divergences, and it is beyond the scope of this paper to develop proper approximations for it. We also find that the commonly used ground-state approximation of neglecting singly-excited matrix elements is applicable in EDFT with slightly larger error. Because of the problems associated with the ensemble PT2, a double hybrid functional mixing in a part of PT2 in the generalized EKS schemeGK21 might be a more viable approach for developing approximated ensemble Hxc energy functionals.
contains sums over all EKS Slater determinants, and we studied the convergence of EEXX+PT2 excitation energies with respect to the number of KS orbitals. We find that in 1D systems convergence is achieved with only one or two extra orbitals than the minimal case, and the excitation energies only change slightly when higher-energy orbitals are included. Although more orbitals are needed for convergence of 3D atoms due to the orbital energies resemble the Rydberg series, convergence is still achieved fairly fast, allowing the use of in practical calculations.
Acknowledgements
This work is supported by the National Natural Science Foundation of China Grant No. 11804314.
Appendix A Details in GLPT for EDFT
The scaled Hamiltonian is equal to in GLPT, and the zeroth-order and perturbative Hamiltonians are
(25) |
and
(26) |
Eq. (26) differs from the Rayleigh-Schrödinger perturbation theory since the perturbation is dependent on .
GLPT expands at as
(27) |
and and can be expanded similarly. Since and , we can derive the expression for the ensemble Hxc potential from Eqs. (10) and (27) as
(28) |
For the ensemble energy, we have and . The ensemble Hxc energy is derived similarly as in Eq. (28):
(29) |
Inserting the -expansions into Eq. (13) and collecting different orders of , the equations from to are
(30) | ||||
(31) | ||||
(32) |
where and .
The first and second order corrections to the EKS energy can be derived from Eq. (32) as
(33) |
and
(34) |
Eq. (33) indicates that since due to variational principle. Inserting Eq. (33) and (34) into the ensemble energy decomposition Eq. (14), and noticing that
(35) |
we arrive at the formula for the ensemble PT2 correlation energy Eq. (15).
Appendix B Justification of neglecting the fixed-orbital -derivative of in DEC/PT2
The second term of Eq. (20) is neglected the calculations, and we provide a justification here. Since the DEC method is evaluated at , this term reduces to
(36) |
where we assume that the ground state is non-degenerate for simplicity.
is a one-body potential, so is its fixed-orbital -derivative. Therefore the summations over and in Eq. (36) only involve ‘singly-excited’ that differs from by one orbital according to the Slater-Condon rules. The in Eq. (36) is truncated at the first order. We have since , so the magnitude of is approximately that of , which vanishes for Hartree-Fock (HF) single-Slater-determinant wavefunctions according to Brillouin’s theorem. Although this matrix element does not vanish for Kohn-Sham (KS) Slater determinants, it is small in most cases since the KS Slater determinants resemble the HF onesSH99 . Therefore the second term of Eq. (20) can be neglected in EEXX+PT2 calculations in most cases.
We carry out EEXX+PT2 calculations with this term for the Be atom to demonstrate the validity of this approximation. The fixed-orbital -derivative of term is calculated with the Slater approximation (see supplemental materialsupplemental ). These calculations must use the full ensemble containing all the states from the ground state to the interested excited multiplet, so the computational cost is much larger than the bi-ensemble calculations in the main text. Table 4 lists the differences of excitation energies with and without this term, and these differences are at least two orders of magnitudes smaller than the error in EEXX+PT2 shown in the main text. This demonstrates that the approximation of neglecting this term is viable.
I | (mH) |
---|---|
3P(1s22s2p) | 0.02396 |
1P(1s22s2p) | 0.03065 |
3S(1s22s3s) | 0.1580 |
1S(1s22s3s) | 0.1310 |
1D(1s22p2) | 0.04902 |
3P(1s22s3p) | 0.1510 |
3P(1s22p2) | 0.03527 |
Appendix C 1D flat box
We carry out DEC and TDDFT calculations of a two-electron 1D flat box whose external potential is
(37) |
and the soft-Coulomb interaction Eq. (24). We use to ensure that the differences between exact and KS excitation energies are large, so that the improvement of EEXX+PT2 results over EEXX can be seen clearly. We use a grid with 1001 points with grid-point spacing 0.001 for . Table 5 lists the calculation results.
(mH) | |||||||
(H) | (H) | EEXX | + | + | +PT2 | ||
Singlet | |||||||
2(1,2) | 15.62 | 13.88 | -78.40 | 28.41 | 31.76 | 138.6 | 104.2 |
3(2,2) | 28.86 | 27.76 | -145.2 | -220.0 | 75.16 | 0.3752 | 17.04 |
5(1,3) | 39.93 | 38.60 | -302.0 | 13.59 | -261.9 | 53.66 | 51.76 |
7(2,3) | 54.49 | 52.48 | -153.9 | -212.9 | -3.650 | -62.67 | -40.72 |
9(1,4) | 74.05 | 73.12 | -281.3 | -32.25 | -236.9 | 12.21 | 20.92 |
10(3,3) | 77.93 | 77.20 | -18.99 | -107.4 | 61.15 | -27.21 | -38.43 |
Triplet | |||||||
1(1,2) | 12.44 | 13.88 | -219.7 | -144.7 | -109.5 | -34.57 | -2.608 |
4(1,3) | 37.70 | 38.60 | -132.9 | -42.36 | -92.82 | -2.292 | 8.062 |
6(2,3) | 52.08 | 52.48 | -246.8 | -123.0 | -96.53 | 27.20 | 18.81 |
8(1,4) | 72.61 | 73.12 | -136.3 | -34.64 | -91.81 | 9.829 | 20.89 |
References
- [1] A.K. Theophilou. The energy density functional formalism for excited states. J. Phys. C, 12:5419, 1979.
- [2] E. K. U. Gross, L. N. Oliveira, and W. Kohn. Rayleigh-Ritz variational principle for ensemble of fractionally occupied states. Phys. Rev. A, 37:2805, 1988.
- [3] E. K. U. Gross, L. N. Oliveira, and W. Kohn. Density-functional theory for ensembles of fractionally occupied states. I. Basic formalism. Phys. Rev. A, 37:2809, 1988.
- [4] L. N. Oliveira, E. K. U. Gross, and W. Kohn. Density-functional theory for ensembles of fractionally occupied states. II. Application to the He atom. Phys. Rev. A, 37:2821, 1988.
- [5] P. Hohenberg and W. Kohn. Inhomogeneous electron gas. Phys. Rev., 136:B864, 1964.
- [6] W. Kohn and L. J. Sham. Self-consistent equations including exchange and correlation effects. Phys. Rev., 140:A1133, 1965.
- [7] C. Fiolhais, F. Nogueira, and M. Marques, editors. A primer in density functional theory. Lecture notes in physics. Springer, Berlin, 2003.
- [8] E. Runge and E. K. U. Gross. Density-functional theory for time-dependent systems. Phys. Rev. Lett., 52:997, 1984.
- [9] M. E. Casida. Time-dependent density functional response theory of molecular systems: theory, computational methods, and functionals. In J. M. Seminario, editor, Recent developments and applications in density functional theory. Elsevier, Amsterdam, 1996.
- [10] M. A. L. Marques, N. T. Maitra, F. M. S. Nogueira, E. K. U. Gross, and A. Rubio, editors. Fundamentals of time-dependent density functional theory. Lecture Notes in Physics. Springer, Berlin, 2012.
- [11] C. A. Ullrich. Time-Dependent Density-Functional Theory: Concepts and Applications. Oxford University Press, Oxford, 2012.
- [12] D. J. Tozer and N. C. Handy. On the determination of excitation energies using density functional theory. Phys. Chem. Chem. Phys., 2:2117, 2000.
- [13] B. G. Levine, C. Ko, J. Quenneville, and T. Z. Martínez. Conical intersections and double excitations in time-dependent density functional theory. Mol. Phys., 104:1039, 2006.
- [14] P. Elliott, S. Goldson, Canahui C., and N. T. Maitra. Perspectives on double-excitations in tddft. Chem. Phys., 391:110, 2011.
- [15] N. Maitra. Perspective: Fundamental aspects of time-dependent density functional theory. J. Chem. Phys., 144:220901, 2016.
- [16] D. J. Tozer, R. D. Amos, N. C. Handy, B. O. Roos, and L. Serrano-Andres. Dose density functional theory contribute to the understanding of excited states of unsaturated organic compounds? Mol. Phys., 97:859, 1999.
- [17] A. Dreuw, J. L. Weisman, and M. Head-Gordon. Long-range charge-transfer excited states in time-dependent density functional theory require non-local exchange. J. Chem. Phys., 119:2943, 2003.
- [18] A. Dreuw and M. Head-Gordon. Failure of time-dependent density functional theory for long-range charge-transfer excited states: the zincbacteriochlorin-bacteriochlorin and bacteriochlorophyll-spheroidene complexes. J. Am. Chem. Soc., 126:4007, 2004.
- [19] N. T. Maitra. Charge transfer in time-dependent density functional theory. J. Phys. Condens. Matter, 29:423001, 2017.
- [20] M. Filatov. Ensemble dft approach to excited states of strongly correlated molecular systems. Top. Curr. Chem., 368:97, 2015.
- [21] M. Alam, K. Deur, S. Knecht, and E. Fromager. Combining extrapolation with ghost interaction correction in range-separated ensemble density functional theory for excited states. J. Chem. Phys., 147:204105, 2017.
- [22] T. Gould and S. Pittalis. Hartree and exchange in ensemble density functional theory: avoiding the nonuniqueness disaster. Phys. Rev. Lett., 119:243001, 2017.
- [23] Z.-H. Yang, A. Pribram-Jones, K. Burke, and C. A. Ullrich. Direct extraction of excitation energies from ensemble density-functional theory. Phys. Rev. Lett., 119:033003, 2017.
- [24] T. Gould, L. Kronik, and S. Pittalis. Charge transfer excitations from exact and approximate ensemble kohn-sham theory. J. Chem. Phys., 148:174101, 2018.
- [25] F. Sagredo and K. Burke. Accurate double excitations from ensemble density functional calculations. J. Chem. Phys., 149:134103, 2018.
- [26] K. Deur and E. Fromager. Ground and excited energy levels can be extracted exactly from a single ensemble density-functional theory calculation. J. Chem. Phys., 150:094106, 2019.
- [27] T. Gould and S. Pittalis. Density-driven correlations in many-electron ensembles: theory and application for excited states. Phys. Rev. Lett., 123:016401, 2019.
- [28] I. S. Lee, M. Filatov, and S. K. Min. Formulation and implementation of the spin-restricted ensemble-referenced kohn-sham method in the context of the density functional tight binding approach. J. Chem. Theory Comput., 15:3021, 2019.
- [29] E. Fromager. Individual correlations in ensemble density functional theory: state- and density-driven decompositions without additional kohn-sham systems. Phys. Rev. Lett., 124:243001, 2020.
- [30] T. Gould. Approximately self-consistent ensemble density functional theory: toward inclusion of all correlations. J. Phys. Chem. Lett., 11:9907, 2020.
- [31] T. Gould, G. Stefanucci, and S. Pittalis. Ensemble density functional theory: insight from the fluctuation-dissipation theorem. Phys. Rev. Lett., 125:233001, 2020.
- [32] P.-F. Loos and E. Fromager. A weight-dependent local correlation density-functional approximation for ensembles. J. Chem. Phys., 152:214101, 2020.
- [33] C. Marut, B. Senjean, E. Fromager, and P.-F. Loos. Weight dependence of local exchange-correlation functionals in ensemble density-functional theory: double excitations in two-electron systems. Faraday Discuss., 224:402, 2020.
- [34] M. Filatov, S. Lee, H. Nakata, and C.-H. Choi. Signatures of conical intersection dynamics in the time-resolved photoelectrons spectrum of furan: theoretical modeling with an ensemble density functional theory method. Int. J. Mol. Sci., 22:4276, 2021.
- [35] T. Gould, L. Kronik, and S. Pittalis. Double excitations in molecules from ensemble density functionals: theory and approximations. Phys. Rev. A, 104:022803, 2021.
- [36] D. Hait and Head-Gordon M. Orbital optimized density functional theory for electronic excited states. J. Phys. Chem. Lett., 12:4517, 2021.
- [37] K. Deur, L. Mazouin, B. Senjean, and E. Fromager. Exploring weight-dependent density-functional approximations for ensembles in the hubbard dimer. Eur. Phys. J. B, 91:162, 2018.
- [38] A. Pribram-Jones, Z.-H. Yang, J. R. Trail, K. Burke, R. J. Needs, and C. A. Ullrich. Excitations and benchmark ensemble density functional theory for two electrons. J. Chem. Phys., 140:18A541, 2014.
- [39] Z.-H. Yang, J. R. Trail, A. Pribram-Jones, K. Burke, R. J. Needs, and C. A. Ullrich. Exact and approximate kohn-sham potentials in ensemble density-functional theory. Phys. Rev. A, 90:042501, 2014.
- [40] B. Senjean and E. Fromager. Unified formulation of fundamental and optical gap problems in density-functional theory for ensembles. Phys. Rev. A, 98:022513, 2018.
- [41] B. Senjean and E. Fromager. N-centered ensemble density-functional theory for open systems. Int. J. Quantum Chem., 120:e26190, 2020.
- [42] A. Görling and M. Levy. Correlation-energy functional and its high-density limit obtained from a coupling-constant perturbation expansion. Phys. Rev. B, 47:13105, 1993.
- [43] A. Görling and M. Levy. Exact kohn-sham scheme based on perturbation theory. Phys. Rev. A, 50:196, 1994.
- [44] J. B. Krieger, Y. Li, and G. J. Iafrate. Construction and application of an accurate local spin-polarized kohn-sham potential with integer discontinuity: exchange-only theory. Phys. Rev. A, 45:101, 1992.
- [45] R. G. Parr and W. Yang. Density functional theory of atoms and molecules. Oxford University Press, 1989.
- [46] I. N. Levine. Quantum Chemistry. Pearson, Upper Saddle River, NJ, 2014.
- [47] A. Dalgarno. Stationary perturbation theory. In D. R. Bates, editor, Quantum Theory. Academic Press, New York, 1961.
- [48] See supplemental material at [url].
- [49] M Levy. Excitation energies from density-functional orbital energies. Phys. Rev. A, 52:R4313, 1995.
- [50] S. Grimme. Semiempirical hybrid density functional with perturbative second-order correlation. J. Chem. Phys., 124:034108, 2006.
- [51] N. T. Maitra, F. Zhang, R. J. Cave, and K. Burke. Double excitations within time-dependent density functional theory linear response. J. Chem. Phys., 120:5932, 2004.
- [52] N. Helbig, J. I. Fuks, M. Casula, M. J. Verstraete, M. A. L. Marques, I. V. Tokatly, and A. Rubio. Density functional theory beyond the linear regime: validating an adiabatic local density approximation. Phys. Rev. A, 83:032503, 2011.
- [53] C. J. Umrigar and X. Gonze. High performance computing and its application to the physical sciences. In D. A. Browne, editor, Proceedings of the Mardi Gras ’93 Conference, Singapore, 1994. World Scientific.
- [54] C. J. Umrigar and X. Gonze. Accurate exchange-correlation potentials and total-energy components for the hellium isoelectronic series. Phys. Rev. A, 50:3827, 1994.
- [55] A. Kramida, Yu. Ralchenko, J. Reader, and and NIST ASD Team. NIST Atomic Spectra Database (ver. 5.8), [Online]. Available: https://physics.nist.gov/asd [2021, July 13]. National Institute of Standards and Technology, Gaithersburg, MD., 2020.
- [56] S. Hirata and M. Head-Gordon. Time-dependent density functional theory within the tamm-dancoff approximation. Chem. Phys. Lett., 314:291, 1999.
- [57] P. Mori-Sánchez, Q. Wu, and W. Yang. Orbital-dependent correlation energy in density-functional theory based on a second-order perturbation approach: Success and failure. J. Chem. Phys., 123:062204, 2005.
- [58] J. P. Perdew and Y. Wang. Accurate and simple analytic representation of the electron-gas correlation energy. Phys. Rev. B, 45:13244, 1992.
- [59] T. Gould and L. Kronik. Ensemble generalized kohn-sham theory: the good, the bad, and the ugly. J. Chem. Phys., 154:094125, 2021.
- [60] Stowasser R. and Hoffmann R. What do the kohn-sham orbitals and eigenvalues mean? J. Am. Chem. Soc., 121:3414, 1999.