Assessment of the LFAs-PBE exchange–correlation potential for high-order harmonic generation of aligned H2+ molecules

Hsiao-Ling Sunab, Wei-Tao Penga and Jeng-Da Chai*abc
aDepartment of Physics, National Taiwan University, Taipei 10617, Taiwan. E-mail: jdchai@phys.ntu.edu.tw
bPhysics Division, National Center for Theoretical Sciences (North), National Taiwan University, Taipei 10617, Taiwan
cCenter for Theoretical Sciences and Center for Quantum Science and Engineering, National Taiwan University, Taipei 10617, Taiwan

Received 9th February 2016 , Accepted 17th March 2016

First published on 18th March 2016


Abstract

We examine the performance of our recently developed LFAs-PBE exchange-correlation (XC) potential [C.-R. Pan, P.-T. Fang, and J.-D. Chai, Phys. Rev. A, 2013, 87, 052510] for the high-order harmonic generation (HHG) spectra and related properties of H2+ molecules aligned parallel and perpendicular to the polarization of an intense linearly polarized laser pulse, employing the real-time formulation of time-dependent density functional theory (RT-TDDFT). The results are compared with the exact solutions of the time-dependent Schrödinger equation as well as those obtained with other XC potentials in RT-TDDFT. Owing to its correct (−1/r) asymptote, the LFAs-PBE potential significantly outperforms conventional XC potentials for the HHG spectra and the properties that are sensitive to the XC potential asymptote. Accordingly, the LFAs-PBE potential, which has a computational cost similar to that of the popular Perdew–Burke–Ernzerhof (PBE) potential, can be very promising for the study of the ground-state, excited-state, and time-dependent properties of large electronic systems, extending the applicability of density functional methods for a diverse range of applications.


I. Introduction

Due to its reasonable accuracy and computational efficiency, time-dependent density functional theory (TDDFT)1 has been one of the most popular methods for studying the excited-state and time-dependent properties of large electronic systems.2–4 Nevertheless, as the exact time-dependent exchange–correlation (XC) potential vXC(r,t) in TDDFT is not known, an approximate treatment for vXC(r,t) is necessary for practical applications.

For a system under the influence of a slowly varying external potential, the best known approximation for vXC(r,t) is the adiabatic approximation:

 
image file: c6ra03713e-t1.tif(1)
where vXC(r,t) is approximately given by the static XC potential (i.e., the functional derivative of the XC energy functional EXC[ρ]) evaluated at the instantaneous density ρ(r,t). Although the entire history of the density (i.e., memory effects) is ignored in the adiabatic approximation, the accuracy of the adiabatic approximation can be surprisingly high in many situations (even if the system is not in this slowly varying regime). However, since the exact EXC[ρ], the essential ingredient of both Kohn–Sham density functional theory (KS-DFT)5,6 (for ground-state properties) and adiabatic TDDFT (for excited-state and time-dependent properties), remains unknown, there has been considerable effort invested in developing accurate density functional approximations (DFAs) for EXC[ρ] to improve the accuracy of both KS-DFT and adiabatic TDDFT for a wide range of applications.2–4,7–9

Conventional density functionals (i.e., semilocal density functionals) are based on the local density approximation (LDA),10,11 generalized gradient approximations (GGAs), and meta-GGAs (MGGAs).12 They are computationally efficient for large systems, and reasonably accurate for properties governed by short-range XC effects, such as low-lying valence excitation energies. However, they can predict qualitatively incorrect results in situations where an accurate description of nonlocal XC effects is critical.2–4,7–9 For example, the LDA and most GGA XC potentials do not exhibit the correct asymptotic behavior in the asymptotic region (r → ∞) of a molecular system, yielding erroneous results for the highest occupied molecular orbital (HOMO) energies, high-lying Rydberg excitation energies,13–16 charge-transfer (CT) excitation energies,16–22 and excitation energies in completely symmetrical systems where no net CT occurs.23

To properly describe properties that are sensitive to the asymptote of the XC potential, two different density functional methods with correct asymptotic behavior have been actively developed over the past two decades. One is the long-range corrected (LC) hybrid scheme,24–37 and the other is the asymptotically corrected (AC) model potential scheme,38–46 both of which have recently become very popular. In the LC hybrid scheme, as 100% Hartree–Fock (HF) exchange is adopted for long-range interelectron interactions, an AC XC potential can be naturally generated by the optimized effective potential (OEP) method.7,47–49 However, for most practical calculations, the generalized Kohn–Sham (GKS) method, which uses orbital-specific XC potentials, has been frequently adopted in the LC hybrid scheme to avoid the computational complexity involved in solving the OEP equation, as the electron density, total energy, and HOMO energy obtained with the GKS method are very close to those obtained with the OEP method.7,49 Recently, we have examined the performance of various XC energy functionals for a diverse range of applications.50,51 In particular, we have shown that LC hybrid functionals can be reliably accurate for several types of excitation energies, involving valence, Rydberg, and CT excitation energies, using the frequency-domain formulation of linear-response TDDFT (LR-TDDFT).52,53 Nevertheless, due to the inclusion of long-range HF exchange, the LC hybrid scheme can be computationally expensive for large systems.

In contrast, in the AC model potential scheme, an AC XC potential is modeled explicitly with the electron density, retaining a computational cost comparable to that of the efficient semilocal density functional methods. In the asymptotic region of a molecule where the electron density decays exponentially, the popular LB94 (ref. 38) and LBα39 potentials exhibit the correct (−1/r) decay by construction. However, as most AC model potentials, including the LB94 and LBα potentials, are not functional derivatives,54,55 the associated XC energies and kernels (i.e., the second functional derivatives of the XC functionals) are not properly defined. Due to the lack of an analytical expression for EXC[ρ], the XC energy associated with an AC model potential is usually evaluated by the popular Levy–Perdew virial relation,56 which has, however, been shown to exhibit severe errors in calculated ground-state energies and related properties.43,50 In addition, due to the lack of a self-consistent adiabatic XC kernel, an adiabatic LDA or GGA XC kernel has been constantly employed for the LR-TDDFT calculations using the AC model potential scheme. Such combined approaches have been shown to be accurate for both valence and Rydberg excitations, but inaccurate for CT excitations,43,50,57–59 due to the lack of a space- and frequency-dependent discontinuity in the adiabatic LDA or GGA kernel adopted in LR-TDDFT.60 Besides, due to the lack of the step and peak structure in most adiabatic AC model potentials (e.g., LB94),61 we have recently shown that adiabatic AC model potentials can fail to describe CT-like excitations,22 even in the real-time formulation of TDDFT (RT-TDDFT),4 where the knowledge of the XC kernel is not needed.

Very recently, we have developed the localized Fermi–Amaldi (LFA) scheme,43 wherein an exchange energy functional whose functional derivative has the correct (−1/r) asymptote can be directly added to any semilocal functional. When the Perdew–Burke–Ernzerhof (PBE) functional62 (a very popular GGA) is adopted as the parent functional in the LFA scheme, the resulting LFA-PBE functional, whose functional derivative (i.e., the LFA-PBE XC potential) has the correct (−1/r) asymptote. Among existing AC model potentials, the LFA-PBE XC potential is one of the very few XC potentials that are functional derivatives.43–46

In addition, without loss of much accuracy, the LFAs-PBE XC potential (an approximation to the LFA-PBE XC potential)43 has been developed for the efficient treatment of large systems. For a molecular system, the LFAs-PBE XC potential can be expressed as

 
vLFAs-PBEXC(r) = vPBEXC(r) + vLFAsX(r). (2)
Here vPBEXC(r) is the PBE XC potential, and vLFAsX(r) is the LFAs exchange potential:
 
image file: c6ra03713e-t2.tif(3)
where the range-separation parameter ω = 0.15 bohr−1 was determined by fitting the minus HOMO energies of 18 atoms and 113 molecules in the IP131 database to the corresponding experimental ionization potentials.36 In eqn (3), the sum is over all the atoms in the molecular system, RA is the position of atom A, and wA(r) is the weight function associated with atom A:
 
image file: c6ra03713e-t3.tif(4)
ranging between 0 and 1. Here ρ0A(r) is the spherically averaged electron density computed for isolated atom A. Note that even with the aforementioned approximation, the LFAs-PBE XC potential remains a functional derivative (as discussed in ref. 43). Due to the sum rule of image file: c6ra03713e-t4.tif, the asymptote of vLFAsX(r) is shown to be correct:
 
image file: c6ra03713e-t5.tif(5)

As vPBEXC(r) decays much faster than (−1/r) in the asymptotic region, vLFAs-PBEXC(r) (see eqn (2)) retains the correct (−1/r) asymptote.

LFAs-PBE has been shown to yield accurate vertical ionization potentials and Rydberg excitation energies for a wide range of atoms and molecules, while performing similarly to their parent PBE semilocal functional for various properties that are insensitive to the XC potential asymptote. Relative to the existing AC model potentials that are not functional derivatives, LFAs-PBE is significantly superior in performance for ground-state energies and related properties.43

As LFAs-PBE is computationally efficient (e.g., similar to PBE), it may be interesting and perhaps important to understand the applicability and limitations of LFAs-PBE for a diverse range of applications. In this work, we examine the performance of LFAs-PBE for various time-dependent properties using adiabatic RT-TDDFT. The rest of this paper is organized as follows. In Section II, we describe our test sets and computational details. The time-dependent properties calculated using LFAs-PBE are compared with those obtained with other methods in Section III. Our conclusions are presented in Section IV.

II. Test sets and computational details

High-order harmonic generation (HHG) from atoms and molecules is a nonlinear optical process driven by intense laser fields.63–91 Recently, HHG has attracted considerable attention, as it can serve as a probe for the structure and dynamics of atoms and molecules and chemical reactions on a femtosecond time scale. Besides, HHG may also be adopted for producing attosecond pulse trains and individual attosecond pulses.65,66 In the semiclassical three-step model of HHG,63,64 first, an electron from an atom or molecule is ionized by a strong laser field; secondly, the electron released by tunnel ionization is accelerated by the oscillating electric field of the laser pulse; thirdly, the electron returns and recombines with the parent ion, emitting a high-energy photon.

In RT-TDDFT, HHG spectra can be obtained by explicitly propagating the time-dependent Kohn–Sham (TDKS) equations. As the asymptote of the XC potential should be important to the ionization, acceleration, and recombination steps in the HHG process, the LFAs-PBE potential is expected to provide accurate results for HHG spectra and related properties. Here we examine the performance of various adiabatic XC potentials in RT-TDDFT for the HHG spectra and related properties of the one-electron H2+ system, as the exact results can be easily obtained with the time-dependent Schrödinger equation (TDSE) for direct comparison.

The TDSE calculations and the RT-TDDFT calculations employing the adiabatic LDA,10,11 PBE,62 LB94 (ref. 38), and LFAs-PBE43 XC potentials, are performed with the program package Octopus 4.0.1 (ref. 92). The H2+ system is described on a uniform real-space grid with a spacing of 0.19 bohr and a sphere of radius rm = 25 bohr around the center of mass of the system. The nuclei are positioned parallel (θ = 0°) or perpendicular (θ = 90°) to the x-axis with a frozen equilibrium bond length of 2.0 bohr.93 The electron–ion interaction is represented by norm-conserving Troullier–Martins pseudopotentials.94 To obtain the HHG spectrum, H2+, which starts from the ground state, experiences an intense linearly polarized laser pulse at time t = 0. The strong field interaction is generated by an oscillating electric field linearly polarized parallel to the x-axis:

 
image file: c6ra03713e-t6.tif(6)
Here the interaction with the electric field is treated in the dipole approximation and the length gauge.95 The frequency of the laser pulse ω0 = 1.55 eV (i.e., the wavelength λ = 800 nm), the peak intensity I0 = 1 × 1014 W cm−2, and a pulse duration of 13 optical cycles (i.e., T = 26π/ω0 = 34.7 fs) are adopted (see Fig. 1). To propagate the TDSE and TDKS equations, we adopt a time step of Δt = 0.02 a.u. (0.484 as) and run up to tf = 1500 a.u. (36.3 fs), which corresponds to 7.5 × 104 time steps. The approximated enforced time-reversal symmetry (AETRS) algorithm is employed to numerically represent the time evolution operator.96


image file: c6ra03713e-f1.tif
Fig. 1 Electric field of the laser pulse adopted.

In the HHG process, the electron released by tunnel ionization may travel far away from the simulation region (rrm). For computational efficiency, we separate the free-propagation space by employing the masking function:97,98

 
image file: c6ra03713e-t7.tif(7)
Here a = 5 bohr is the width of the masking function. At each time step, each Kohn–Sham (KS) orbital ψj(r,t) is multiplied by M(r):
 
ψj(r,t) → M(r)ψj(r,t). (8)

While the KS orbitals inside the simulation region are unaffected by the masking function, those at the boundary may gradually disappear and never return to the nuclei. Due to the masking function, the normalization of the KS orbitals may decrease in time. Therefore, the ionization probability of the KS orbital ψj(r,t) can be defined as

 
Pj(t) = 1 − ∫|ψj(r,t)|2dr. (9)

After the electron density image file: c6ra03713e-t8.tif is determined, the induced dipole moment is calculated by

 
d(t) = ∫(r,t)dr (10)
and the dipole acceleration is calculated using the Ehrenfest theorem:99
 
image file: c6ra03713e-t9.tif(11)
where vs(r,t) is the time-dependent KS potential. The HHG spectrum is calculated by taking the Fourier transform of the dipole acceleration:100
 
image file: c6ra03713e-t10.tif(12)

As the asymptotic behavior of the XC potential can be important for a wide variety of time-dependent properties, in this work, we examine the ionization probability of the HOMO (1σg) [eqn (9)], the induced dipole moment [eqn (10)], and the HHG spectrum [eqn (12)] for H2+ aligned parallel (θ = 0°) or perpendicular (θ = 90°) to the laser polarization, using the exact TDSE and RT-TDDFT with the adiabatic LDA, PBE, LB94, and LFAs-PBE XC potentials.

III. Results and discussion

The vertical ionization potential of a neutral molecule, defined as the energy difference between the cationic and neutral charge states, is identical to the minus HOMO energy of the neutral molecule obtained from the exact KS potential.101–106 Accordingly, the minus HOMO energy calculated using approximate KS-DFT has been one of the most common measures of the quality of the underlying XC potential.43,50,51,107 As shown in Table 1, the minus HOMO (1σg) energy of the one-electron H2+ system, obtained with the exact time-independent Schrödinger equation (TISE) is 29.96 eV. However, the minus HOMO energies of H2+, calculated using the LDA and PBE XC potentials are about 6 eV smaller than the exact value, as the LDA and PBE XC potentials exhibit an exponential decay (not the correct (−1/r) decay) in the asymptotic region (r → ∞) of a molecule. By contrast, the minus HOMO energies of H2+, calculated using the LB94 and LFAs-PBE XC potentials are in good agreement with the exact value (within an error of 1.8 eV), due to the correct asymptotic behavior of the LB94 and LFAs-PBE XC potentials.
Table 1 Minus HOMO (1σg) energy (in eV) of H2+, calculated using the exact TISE and KS-DFT with various XC potentials
Orbital Exact LDA PBE LB94 LFAs-PBE
1σg 29.96 23.24 23.72 28.24 28.25


Fig. 2 plots the ionization probability of the HOMO (1σg) for H2+ aligned parallel (θ = 0°) or perpendicular (θ = 90°) to the laser polarization, calculated using the exact TDSE. After 5 optical cycles, the ionization probabilities of the HOMO calculated using the LB94 and LFAs-PBE XC potentials become much more accurate than those calculated using the LDA and PBE XC potentials for both the parallel (Fig. 3) and perpendicular (Fig. 4) alignments. Owing to the correct (−1/r) asymptote, the LB94 and LFAs-PBE XC potentials are more difficult to ionize as compared with the LDA and PBE XC potentials (which decay much faster in the asymptotic region).


image file: c6ra03713e-f2.tif
Fig. 2 Ionization probability of the HOMO (1σg) for H2+ aligned parallel (θ = 0°) or perpendicular (θ = 90°) to the laser polarization, calculated using the exact TDSE.

image file: c6ra03713e-f3.tif
Fig. 3 Ionization probability of the HOMO (1σg) for H2+ aligned parallel (θ = 0°) to the laser polarization, calculated using the exact TDSE and RT-TDDFT with various adiabatic XC potentials. Inset shows the TDSE, LB94, and LFAs-PBE results on a logarithmic scale.

image file: c6ra03713e-f4.tif
Fig. 4 Same as Fig. 3, but for H2+ aligned perpendicular (θ = 90°) to the laser polarization.

In addition, we examine the induced dipole moment for H2+ aligned parallel (Fig. 5) or perpendicular (Fig. 6) to the laser polarization, calculated using the exact TDSE and RT-TDDFT with various adiabatic XC potentials. The induced dipole moments obtained with the adiabatic XC potentials in RT-TDDFT are very similar, and are all close to those obtained with the exact TDSE, revealing that the choice of the XC potential has only a minor effect on the induced dipole moments.


image file: c6ra03713e-f5.tif
Fig. 5 Induced dipole moment for H2+ aligned parallel (θ = 0°) to the laser polarization, calculated using the exact TDSE and RT-TDDFT with various adiabatic XC potentials.

image file: c6ra03713e-f6.tif
Fig. 6 Same as Fig. 5, but for H2+ aligned perpendicular (θ = 90°) to the laser polarization.

In comparison to the HHG spectra obtained with the exact TDSE (Fig. 7), we plot the HHG spectrum for H2+ aligned parallel (Fig. 8) or perpendicular (Fig. 9) to the laser polarization, calculated using various adiabatic XC potentials in RT-TDDFT. In contrast to the work of Wasserman and co-workers,85 the LFAs-PBE potential, which is an AC model potential, performs significantly better than the LDA, PBE, and LB94 XC potentials, especially for the higher harmonics (from the 50th to 80th harmonics). As a result, the fine details of the XC potential (not just the XC potential asymptote) are expected to be important for the calculated HHG spectra.


image file: c6ra03713e-f7.tif
Fig. 7 HHG spectrum for H2+ aligned parallel (θ = 0°) or perpendicular (θ = 90°) to the laser polarization, calculated using the exact TDSE. Here the unit of H(ω) is Å (ħeV2)−1.

image file: c6ra03713e-f8.tif
Fig. 8 HHG spectrum for H2+ aligned parallel (θ = 0°) to the laser polarization, calculated using the exact TDSE (black curve) and RT-TDDFT with various adiabatic XC potentials. Insets show the differences between the RT-TDDFT and TDSE results. Here the unit of H(ω) is Å (ħeV2)−1.

image file: c6ra03713e-f9.tif
Fig. 9 Same as Fig. 8, but for H2+ aligned perpendicular (θ = 90°) to the laser polarization.

IV. Conclusions

In summary, we have assessed the performance of our recently proposed LFAs-PBE XC potential for the HHG spectra and related properties of H2+ aligned parallel and perpendicular to the polarization of an intense linearly polarized laser pulse, using adiabatic RT-TDDFT. The results have been compared with those obtained with the exact TDSE as well as those calculated using adiabatic RT-TDDFT with the LDA, PBE, and LB94 XC potentials. Due to the correct (−1/r) asymptote, both the LB94 and LFAs-PBE XC potentials significantly outperform the conventional LDA and PBE XC potentials for the properties sensitive to the XC potential asymptote. For the HHG spectra, the LFAs-PBE potential performs significantly better than the LDA, PBE, and LB94 potentials, implying that the fine details of the XC potential (not just the XC potential asymptote) could also be important for HHG spectra. Relative to the LB94 potential and other AC model potentials that are not functional derivatives, the LFAs-PBE potential has been shown to be significantly superior in performance for ground-state energies and related properties.43 Therefore, the LFAs-PBE potential, which has a computational cost similar to that of the popular PBE potential, can be very promising for studying the ground-state, excited-state, and time-dependent properties of large electronic systems, extending the applicability of density functional methods for a diverse range of applications.

Acknowledgements

This work was supported by the Ministry of Science and Technology of Taiwan (Grant No. MOST104-2628-M-002-011-MY3), National Taiwan University (Grant No. NTU-CDP-105R7818), the Center for Quantum Science and Engineering at NTU (Subproject No. NTU-ERP-105R891401 and NTU-ERP-105R891403), and the National Center for Theoretical Sciences of Taiwan.

References

  1. E. Runge and E. K. U. Gross, Phys. Rev. Lett., 1984, 52, 997 CrossRef CAS.
  2. E. K. U. Gross, J. F. Dobson and M. Petersilka, in Density Functional Theory II, Springer, New York, 1996, vol. 181, pp. 81–172 Search PubMed.
  3. M. A. L. Marques and E. K. U. Gross, Annu. Rev. Phys. Chem., 2004, 55, 427 CrossRef CAS PubMed.
  4. C. A. Ullrich, Time-Dependent Density-Functional Theory: Concepts and Applications, Oxford University, New York, 2012 Search PubMed.
  5. P. Hohenberg and W. Kohn, Phys. Rev., 1964, 136, B864 CrossRef.
  6. W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133 CrossRef.
  7. S. Kümmel and L. Kronik, Rev. Mod. Phys., 2008, 80, 3 CrossRef.
  8. A. J. Cohen, P. Mori-Sánchez and W. Yang, Chem. Rev., 2011, 112, 289 CrossRef PubMed.
  9. E. Engel and R. M. Dreizler, Density Functional Theory: An Advanced Course, Springer, Heidelberg, 2011 Search PubMed.
  10. P. A. M. Dirac, Proc. Cambridge Philos. Soc., 1930, 26, 376 CrossRef CAS.
  11. J. P. Perdew and A. Zunger, Phys. Rev. B: Condens. Matter Mater. Phys., 1981, 23, 5048 CrossRef CAS.
  12. J. P. Perdew, A. Ruzsinszky, J. Tao, V. N. Staroverov, G. E. Scuseria and G. I. Csonka, J. Chem. Phys., 2005, 123, 062201 CrossRef PubMed.
  13. M. E. Casida, C. Jamorski, K. C. Casida and D. R. Salahub, J. Chem. Phys., 1998, 108, 4439 CrossRef CAS.
  14. M. E. Casida and D. R. Salahub, J. Chem. Phys., 2000, 113, 8918 CrossRef CAS.
  15. N. N. Matsuzawa, A. Ishitani, D. A. Dixon and T. Uda, J. Phys. Chem. A, 2001, 105, 4953 CrossRef CAS.
  16. M. J. G. Peach, P. Benfield, T. Helgaker and D. J. Tozer, J. Chem. Phys., 2008, 128, 044118 CrossRef PubMed.
  17. D. J. Tozer, J. Chem. Phys., 2003, 119, 12697 CrossRef CAS.
  18. A. Dreuw, J. L. Weisman and M. Head-Gordon, J. Chem. Phys., 2003, 119, 2943 CrossRef CAS.
  19. A. Dreuw and M. Head-Gordon, J. Am. Chem. Soc., 2004, 126, 4007 CrossRef CAS PubMed.
  20. B. M. Wong and J. G. Cordaro, J. Chem. Phys., 2008, 129, 214703 CrossRef PubMed.
  21. B. M. Wong, M. Piacenza and F. D. Sala, Phys. Chem. Chem. Phys., 2009, 11, 4498 RSC.
  22. W.-T. Peng and J.-D. Chai, Phys. Chem. Chem. Phys., 2014, 16, 21564 RSC.
  23. W. Hieringer and A. Görling, Chem. Phys. Lett., 2006, 419, 557 CrossRef CAS.
  24. A. Savin, in Recent Developments and Applications of Modern Density Functional Theory, ed. J. M. Seminario, Elsevier, Amsterdam, 1996, pp. 327–357 Search PubMed.
  25. H. Iikura, T. Tsuneda, T. Yanai and K. Hirao, J. Chem. Phys., 2001, 115, 3540 CrossRef CAS.
  26. Y. Tawada, T. Tsuneda, S. Yanagisawa, T. Yanai and K. Hirao, J. Chem. Phys., 2004, 120, 8425 CrossRef CAS PubMed.
  27. R. Baer and D. Neuhauser, Phys. Rev. Lett., 2005, 94, 043002 CrossRef PubMed.
  28. O. A. Vydrov, J. Heyd, A. V. Krukau and G. E. Scuseria, J. Chem. Phys., 2006, 125, 074106 CrossRef PubMed.
  29. J.-W. Song, S. Tokura, T. Sato, M. A. Watson and K. Hirao, J. Chem. Phys., 2007, 127, 154109 CrossRef PubMed.
  30. J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2008, 128, 084106 CrossRef PubMed.
  31. J.-D. Chai and M. Head-Gordon, Phys. Chem. Chem. Phys., 2008, 10, 6615 RSC.
  32. J.-D. Chai and M. Head-Gordon, Chem. Phys. Lett., 2008, 467, 176 CrossRef CAS.
  33. J. A. Parkhill, J.-D. Chai, A. D. Dutoi and M. Head-Gordon, Chem. Phys. Lett., 2009, 478, 283 CrossRef CAS.
  34. J.-D. Chai and M. Head-Gordon, J. Chem. Phys., 2009, 131, 174105 CrossRef PubMed.
  35. T. Stein, L. Kronik and R. Baer, J. Am. Chem. Soc., 2009, 131, 2818 CrossRef CAS PubMed.
  36. Y.-S. Lin, C.-W. Tsai, G.-D. Li and J.-D. Chai, J. Chem. Phys., 2012, 136, 154109 CrossRef PubMed.
  37. Y.-S. Lin, G.-D. Li, S.-P. Mao and J.-D. Chai, J. Chem. Theory Comput., 2013, 9, 263 CrossRef CAS PubMed.
  38. R. van Leeuwen and E. J. Baerends, Phys. Rev. A, 1994, 49, 2421 CrossRef CAS.
  39. P. R. T. Schipper, O. V. Gritsenko, S. J. A. van Gisbergen and E. J. Baerends, J. Chem. Phys., 2000, 112, 1344 CrossRef CAS.
  40. D. J. Tozer, J. Chem. Phys., 2000, 112, 3507 CrossRef CAS.
  41. X. Andrade and A. Aspuru-Guzik, Phys. Rev. Lett., 2011, 107, 183002 CrossRef PubMed.
  42. A. P. Gaiduk, D. S. Firaha and V. N. Staroverov, Phys. Rev. Lett., 2012, 108, 253005 CrossRef PubMed.
  43. C.-R. Pan, P.-T. Fang and J.-D. Chai, Phys. Rev. A, 2013, 87, 052510 CrossRef.
  44. R. Armiento and S. Kümmel, Phys. Rev. Lett., 2013, 111, 036402 CrossRef CAS PubMed.
  45. J. Carmona-Espíndola, J. L. Gázquez, A. Vela and S. B. Trickey, J. Chem. Phys., 2015, 142, 054105 CrossRef PubMed.
  46. S. L. Li and D. G. Truhlar, J. Chem. Theory Comput., 2015, 11, 3123 CrossRef CAS PubMed.
  47. R. T. Sharp and G. K. Horton, Phys. Rev., 1953, 90, 317 CrossRef.
  48. J. D. Talman and W. F. Shadwick, Phys. Rev. A, 1976, 14, 36 CrossRef CAS.
  49. S. Kümmel and J. P. Perdew, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 035103 CrossRef.
  50. C.-W. Tsai, Y.-C. Su, G.-D. Li and J.-D. Chai, Phys. Chem. Chem. Phys., 2013, 15, 8352 RSC.
  51. J.-C. Lee, J.-D. Chai and S.-T. Lin, RSC Adv., 2015, 5, 101370 RSC.
  52. M. E. Casida, Recent Advances in Density Functional Methods, Part I, World Scientific, Singapore, 1995 Search PubMed.
  53. G. Onida, L. Reining and A. Rubio, Rev. Mod. Phys., 2002, 74, 601 CrossRef CAS.
  54. A. P. Gaiduk and V. N. Staroverov, J. Chem. Phys., 2009, 131, 044107 CrossRef PubMed.
  55. A. P. Gaiduk and V. N. Staroverov, Phys. Rev. A, 2011, 83, 012509 CrossRef.
  56. M. Levy and J. P. Perdew, Phys. Rev. A, 1985, 32, 2010 CrossRef CAS.
  57. D. J. Tozer, R. D. Amos, N. C. Handy, B. O. Roos and L. Serrano-Andres, Mol. Phys., 1999, 97, 859 CrossRef CAS.
  58. O. Gritsenko and E. J. Baerends, J. Chem. Phys., 2004, 121, 655 CrossRef CAS PubMed.
  59. J. Neugebauer, O. Gritsenko and E. J. Baerends, J. Chem. Phys., 2006, 124, 214102 CrossRef PubMed.
  60. M. Hellgren and E. K. U. Gross, Phys. Rev. A, 2012, 85, 022514 CrossRef.
  61. J. I. Fuks, P. Elliott, A. Rubio and N. T. Maitra, J. Phys. Chem. Lett., 2013, 4, 735 CrossRef CAS PubMed.
  62. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  63. P. B. Corkum, Phys. Rev. Lett., 1993, 71, 1994 CrossRef CAS PubMed.
  64. M. Lewenstein, P. Balcou, M. Y. Ivanov, A. L'Huillier and P. B. Corkum, Phys. Rev. A, 1994, 49, 2117 CrossRef.
  65. P. M. Paul, E. S. Toma, P. Berger, G. Mullot, F. Augé, P. Balcou, H. G. Muller and P. Agostini, Science, 2001, 292, 1689 CrossRef CAS PubMed.
  66. R. Kienberger, E. Goulielmakis, M. Uiberacker, A. Baltuska, V. Yakovlev, F. Bammer, A. Scrinzi, T. Westerwalbesloh, U. Kleineberg and U. Heinzmann, Nature, 2004, 427, 817 CrossRef CAS PubMed.
  67. X. Chu and S.-I. Chu, Phys. Rev. A, 2001, 63, 023411 CrossRef.
  68. X. Chu and S.-I. Chu, Phys. Rev. A, 2001, 64, 063404 CrossRef.
  69. I. V. Litvinyuk, K. F. Lee, P. W. Dooley, D. M. Rayner, D. M. Villeneuve and P. B. Corkum, Phys. Rev. Lett., 2003, 90, 233003 CrossRef CAS PubMed.
  70. X. Chu and S.-I. Chu, Phys. Rev. A, 2004, 70, 061402(R) CrossRef.
  71. T. Kanai, S. Minemoto and H. Sakai, Nature, 2005, 435, 470 CrossRef CAS PubMed.
  72. X. X. Guan, X. M. Tong and S.-I. Chu, Phys. Rev. A, 2006, 73, 023403 CrossRef.
  73. S. Baker, J. S. Robinson, C. A. Haworth, H. Teng, R. A. Smith, C. C. Chirilă, M. Lein, J. W. G. Tisch and J. P. Marangos, Science, 2006, 312, 424 CrossRef CAS PubMed.
  74. L. Cui, J. Zhao, Y. J. Hu, Y. Y. Teng, X. H. Zeng and B. Gu, Appl. Phys. Lett., 2006, 89, 211103 CrossRef.
  75. P. B. Corkum and F. Krausz, Nat. Phys., 2007, 3, 381 CrossRef CAS.
  76. S. Baker, J. S. Robinson, M. Lein, C. C. Chirilă, R. Torres, H. C. Bandulet, D. Comtois, J. C. Kieffer, D. M. Villeneuve, J. W. G. Tisch and J. P. Marangos, Phys. Rev. Lett., 2008, 101, 053901 CrossRef CAS PubMed.
  77. A.-T. Le, R. Della Picca, P. D. Fainstein, D. A. Telnov, M. Lein and C. D. Lin, J. Phys. B: At., Mol. Opt. Phys., 2008, 41, 081002 CrossRef.
  78. F. Krausz and M. Ivanov, Rev. Mod. Phys., 2009, 81, 163 CrossRef.
  79. D. A. Telnov and S.-I. Chu, Phys. Rev. A, 2009, 80, 043412 CrossRef.
  80. X. Chu, Phys. Rev. A, 2010, 82, 023407 CrossRef.
  81. E. F. Penka and A. D. Bandrauk, Phys. Rev. A, 2010, 81, 023411 CrossRef.
  82. X. Chu and M. McIntyre, Phys. Rev. A, 2011, 83, 013409 CrossRef.
  83. J. Heslar, D. Telnov and S.-I. Chu, Phys. Rev. A, 2011, 83, 043414 CrossRef.
  84. E. F. Penka and A. D. Bandrauk, Phys. Rev. A, 2011, 84, 035402 CrossRef.
  85. M. R. Mack, D. Whitenack and A. Wasserman, Chem. Phys. Lett., 2013, 558, 15 CrossRef CAS.
  86. K. N. Avanaki, D. A. Telnov and S.-I. Chu, Phys. Rev. A, 2014, 90, 033425 CrossRef.
  87. D. A. Telnov, J. Heslar and S.-I. Chu, Phys. Rev. A, 2014, 90, 063412 CrossRef.
  88. M. Chini, X. Wang, Y. Cheng, H. Wang, Y. Wu, E. Cunningham, P.-C. Li, J. Heslar, D. A. Telnov, S.-I. Chu and Z. Chang, Nat. Photonics, 2014, 8, 437 CrossRef CAS.
  89. P.-C. Li, Y.-L. Sheu, C. Laughlin and S.-I. Chu, Nat. Commun., 2015, 6, 8178 CrossRef PubMed.
  90. J. Heslar, D. A. Telnov and S.-I. Chu, Phys. Rev. A, 2015, 91, 023420 CrossRef.
  91. Y. Chou, P.-C. Li, T.-S. Ho and S.-I. Chu, Phys. Rev. A, 2015, 91, 063408 CrossRef.
  92. A. Castro, H. Appel, M. Oliveira, C. A. Rozzi, X. Andrade, F. Lorenzen, M. A. L. Marques, E. K. U. Gross and A. Rubio, Phys. Status Solidi B, 2006, 243, 2465 CrossRef CAS.
  93. D. J. Willock, Molecular Symmetry, John Wiley & Sons, Chichester, UK, 2009, p. 227 Search PubMed.
  94. N. Troullier and J. L. Martins, Phys. Rev. B: Condens. Matter Mater. Phys., 1991, 43, 1993 CrossRef CAS.
  95. E. Cormier and P. Lambropoulos, J. Phys. B: At., Mol. Opt. Phys., 1996, 29, 1667 CrossRef CAS.
  96. A. Castro, M. A. L. Marques and A. Rubio, J. Chem. Phys., 2004, 121, 3425 CrossRef CAS PubMed.
  97. U. De Giovannini, D. Varsano, M. A. L. Marques, H. Appel, E. K. U. Gross and A. Rubio, Phys. Rev. A, 2012, 85, 062515 CrossRef.
  98. U. De Giovannini, G. Brunetto, A. Castro, J. Walkenhorst and A. Rubio, ChemPhysChem, 2013, 14, 1363 CrossRef CAS PubMed.
  99. C. Fiolhais, F. Nogueira and M. Marques, Lect. Notes Phys., 2003, 144 Search PubMed.
  100. L. D. Landau and E. M. Lifshitz, The Classical Theory of Fields, Pergamon, Oxford, 1975 Search PubMed.
  101. J. F. Janak, Phys. Rev. B: Condens. Matter Mater. Phys., 1978, 18, 7165 CrossRef CAS.
  102. J. P. Perdew, R. G. Parr, M. Levy and J. L. Balduz Jr, Phys. Rev. Lett., 1982, 49, 1691 CrossRef CAS.
  103. M. Levy, J. P. Perdew and V. Sahni, Phys. Rev. A, 1984, 30, 2745 CrossRef.
  104. C.-O. Almbladh and U. von Barth, Phys. Rev. B: Condens. Matter Mater. Phys., 1985, 31, 3231 CrossRef CAS.
  105. J. P. Perdew and M. Levy, Phys. Rev. B: Condens. Matter Mater. Phys., 1997, 56, 16021 CrossRef CAS.
  106. M. E. Casida, Phys. Rev. B: Condens. Matter Mater. Phys., 1999, 59, 4694 CrossRef CAS.
  107. J.-D. Chai and P.-T. Chen, Phys. Rev. Lett., 2013, 110, 033002 CrossRef PubMed.

This journal is © The Royal Society of Chemistry 2016