• Matter and Radiation at Extremes
  • Vol. 5, Issue 2, 24401 (2020)
C. Martínez-Flores1 and R. Cabrera-Trujillo2、*
Author Affiliations
  • 1Departamento de Química, Universidad Autónoma Metropolitana Iztapalapa, San Rafael Atlixco 186, Col. Vicentina, Iztapalapa, C.P. 09340 México D.F., Mexico
  • 2Instituto de Ciencias Físicas, Universidad Nacional Autónoma de México, Ap. Postal 43-8, Cuernavaca, Morelos, 62251, Mexico
  • show less
    DOI: 10.1063/1.5139099 Cite this Article
    C. Martínez-Flores, R. Cabrera-Trujillo. High pressure effects on the excitation spectra and dipole properties of Li, Be+, and B2+ atoms under confinement[J]. Matter and Radiation at Extremes, 2020, 5(2): 24401 Copy Citation Text show less

    Abstract

    Properties of atoms and molecules undergo significant changes when subjected to spatial confinement. We study the excitation spectra of lithium-like atoms in the initial 1s22s electronic configuration when confined by an impenetrable spherical cavity. We implement Slater’s X-α method in Hartree–Fock theory to obtain the excitation spectrum. We verify that as the cavity size decreases, the total, 2s, 2p, and higher excited energy levels increase. Furthermore, we confirm the existence of crossing points between nsnp states for low values of the confinement radius such that the nsnp dipole transition becomes zero at that critical pressure. The crossing points of the sp states imply that instead of photon absorption, one observes photon emission for cavities with radius smaller than the critical radius. Hence, the dipole oscillator strength associated with the 2s → 2p transition becomes negative, and for higher pressures, the 2s → 3p dipole oscillator strength transition becomes larger than unity. We validate the completeness of the spectrum by calculating the Thomas–Reiche–Kuhn sum rule, as well as the static dipole polarizability and mean excitation energy of lithium-like atoms. We find that the static dipole polarizability decreases and exhibits a sudden change at the critical pressure for the absorption-to-emission transition. The mean excitation energy increases as the pressure rises. However, as a consequence of the critical transition from absorption to emission, the mean excitation energy becomes undetermined for higher pressures, with implications for material damage under extreme conditions. For unconfined systems, our results show good to excellent agreement with data found in the literature.

    I. INTRODUCTION

    Confined quantum systems exhibit significant changes to their structure, stability, reactivity, binding interactions, dynamics, and spectra as a consequence of modifications to the spatial boundary conditions in the presence of an extreme environment. Atoms or molecules within cavities, organic/inorganic host–guest complexes, quantum dots, fullerenes, and nanotubes are examples of real quantum confined systems. The main objective when studying confined quantum systems is to construct an accurate theoretical model that takes into account changes in the electronic wavefunction and energy levels due to the boundary conditions imposed by the surrounding environment. In this respect, theoretical calculations can be realized by adopting a suitable choice of boundary conditions. The pioneering work of Michels et al.1 provided the first model of a hydrogen atom confined in an impenetrable spherical cavity to simulate the effect of pressure. In recent years, the topic of confined atoms has attracted much attention and has become a very active field of research.2–5 Reviews with detailed discussion of the progress in this field can be found in Refs. 2–6 and references therein.

    To describe the electronic structure and interactions of a multi-electron system, a variety of theoretical methods have been developed, including, among others, Hartree–Fock (HF) theory and density functional theory (DFT), and these have had great success in different problems in atomic physics.6–11 Owing to the complexity of the N-body problem, some approximations have been implemented, in addition to having the system under confinement by spatial limitation of the electrons, which leads to complications in the one- and two-electron integrals. A widely used approach to the treatment of N-body systems is to consider a pseudopotential that describes the inner electronic structure of the atom. The basic idea of pseudopotentials is to take into account the multi-electron core interaction with a single valence electron by using a modified Coulomb potential.7,8,11 Such pseudopotential approaches have been used to describe the electronic spectra of confined systems.12,13 Lin and Ho12 used a pseudopotential for the lithium atom to simulate the core interaction with the single valence electron with optimized parameters. They calculated the photoionization cross section of the 2s shell electron under confinement by a power exponential potential due to an endohedral cavity and found that multiple Cooper resonances emerged.14 Their results show the importance of cage thickness and a smooth shell boundary in the photoionization cross section. Another example is due to Sarsa et al.,13 who studied the effects of confinement on the outer valence electrons for the ground state configurations of carbon and iron atoms. In standard HF theory, the main complication that arises when dealing with quantum confined systems is the treatment of the one- and two-electron integrals, particularly the electron exchange integral, as shown by Ludeña.15 Fortunately, Slater16 proposed a simplified approach to treat the electron exchange operator in the HF method by replacing it by a term proportional to the charge density of the inner electrons. This approach has been fruitful in treating problems in atomic and molecular structure with satisfactory results,17 particularly in the development of DFT theory. However, to the authors’ knowledge, there have not been any studies of the effects of confinement on the excitation spectra of a multi-electron system in the context of HF theory by means of Slater’s X-α approach.

    The goal of this work is to apply Slater’s X-α approach to the ground and excited states of Li-like atoms confined in an impenetrable spherical cavity. To show the strength of Slater’s X-α approach to the calculation of the excitation spectrum of atoms, we take as a benchmark example the ground states of lithium-like atoms in the initial electronic configuration 1s22s, adopting a restricted HF approach. We focus on the dipole oscillator strength (DOS) and derived properties such as the static dipole polarizability and the mean excitation energy.

    The remainder of this paper is organized as follows. In Sec. II, we present the theoretical approach used to study the influence of a spherical confinement cavity on lithium-like atoms. In Sec. III, we discuss our results and findings. In Sec. IV, we give our conclusions and perspectives. Note that we use atomic units (a.u.) throughout, unless physical units are explicitly stated.

    II. THEORY

    A. Slater’s X-α approach in Hartree–Fock theory

    In this section, we present the HF method to obtain the 1s22s ground state energies of Li, Be+, and B2+ atoms, incorporating confinement conditions.

    In an HF approach, the total wavefunction is defined as a Slater determinant Ψ(r) = N det{ψi}, where N is a normalization constant. The general restricted HF approach18 considers the two electrons in the 1s2 core to be in the same orbital. To simplify the calculation of the excitation spectra, we assume that the inner electrons do not see the outer electrons. This approach is known as the frozen-core approximation, which is usually treated within a pseudopotential approach.7,8,11 However, here we do not use any specific pseudopotential formula to describe the interaction of the inner electrons, since this is calculated explicitly for every confinement configuration in a self-consistent manner. Then, for the ground state of a lithium-like atom, the HF equations are(ĥ1+Ĵ1)ψ1r=ϵ1ψ1r,(ĥ2+2Ĵ1K^1)ψ2(r)=ϵ2ψ2(r),where ϵj are the eigenvalues, ψj are the eigenfunctions, ĥj are the one-electron operators, and Ĵj and K^j are the two-electron Coulomb and exchange operators, respectively. These operators are defined asĥj(r)=12j2Zr+Vc(r),Ĵj(r)ψi(r)=dr2ψj(r2)ψj(r2)r12ψi(r),K^j(r)ψi(r)=dr2ψj(r2)ψi(r2)r12ψj(r),where r12 = |rr2|. The one-electron operator, Eq. (3), includes the confinement potential Vc(r) (see below). The total ground state energy of the confined system is given byEHF=2ĥ11+ĥ22+Ĵ11+2Ĵ12K^12.It can be seen that Eq. (1) is in eigenvalue form, so the excitation spectra of the core electron can be readily obtained. However, Eq. (2) is not, owing to the presence of the exchange operator, Eq. (5). To have an eigenvalue equation, we resort to Slater’s X-α approach,16 in which the exchange operator is replaced byK^1(r)ψ2(r)=αXρ11/3(r)ψ2(r),where αX is a parameter, and ρ(r) is the charge density due to the 1s2 inner electrons18 and is given byρ1(r)=2|ψ1(r)|2,with ψ1(r) the eigenfunction of the core electrons. To solve the HF equations (1) and (2), we implement a finite-difference numerical approach within a self-consistent field (SCF) procedure to obtain the eigenvalues, eigenfunctions, and dipole-dependent properties. For each iteration, we calculate Eq. (5) explicitly and determine Slater’s X-α constant as19,20αX=ψ2(r)|K^1(r)|ψ2(r)ψ2(r)|ρ11/3(r)|ψ2(r),finding the excitation spectrum of the 2s electron. The procedure is repeated until Eq. (6) has reached self-consistency within a 10−6 error difference for the total ground state energy.

    Owing to the spherical symmetry of the system, we assume ψ(r)=Rl(r)Ylm(θ,φ) and Rl(r) = ul(r)/r, where Ylm are spherical harmonics. This is done for both core and valence wavefunctions.

    We consider the lithium-like atoms to be confined by an impenetrable spherical cavity, so the confinement potential is given by1–6Vc(r)=0,r<R0,,rR0.where R0 is the confinement radius of the cavity and is commensurate with the confinement pressure (see below).

    1. Finite-difference approach

    As the finite-difference approach has been reported previously by Cabrera-Trujillo and Cruz,21 here we just summarize our implementation to find the eigenvalues and eigenfunctions of Eqs. (1) and (2). The finite-difference approach consists in discretizing the function u(r) → uk and rrk, known at the kth point on a numerical grid, with k = 0 corresponding to u0 and k = N + 1 to uN+1, which are the boundary conditions of the system.22 In our case, for the impenetrable cavity, uN+1 = 0 and rN+1 = R0. We implement the finite-difference approach centered at the midpoint. With this, Eqs. (1) and (2) are rewritten asHϕ=Eϕ,where H is a tridiagonal symmetric matrix with N eigenvalues and eigenfunctions, and ϕ is related to u by a linear transformation.21

    We solve Eq. (11) in a grid box that extends from r = 0 to r = R0, with a total of N = 2000 points spaced logarithmically in this range as a function of the confinement radius R0. This logarithmic grid allows us to give a better description of the wavefunction cusp at the origin and a good number of continuum states. We have found that N = 2000 satisfies the Thomas–Reiche–Kuhn (TRK) sum rule up to five decimal digits. The accuracy of our finite-difference approach can be controlled by the number of points in the grid and their spacing; for example, for the free lithium atom, we obtain eigenvalues with precision up to the fifth decimal place. This approach gives a total of N excited states to describe the DOS electronic properties for each spherical cavity with radius R0, per electron. The values of R0 are chosen between 0.5 a.u. and 100 a.u. Our approach is implemented in a Fortran 95 code that calculates the eigenvalues, eigenfunctions, and physical properties of the system.

    B. Physical properties

    1. Dipole oscillator strengths

    The DOS accounts for the transition probability from an initial state to a final excited state and is defined asfn0=2(EHFnEHF0)|Ψn(r)|i=1Nriϵ^|Ψ0(r)|2,where ϵ^ is the direction of momentum transfer or the polarization vector of the electromagnetic radiation. Here, EHF0 and EHFn are respectively the initial and final total energies, with Ψ0 and Ψn respectively the initial and final total wavefunctions for a given transition. Owing to the presence of just a single determinant and for an electron operator, as in the case of the dipole operator Ô = Σiri, the DOS are reduced to single-electron transitions from either the core or the valence electron.18 Under the spherical symmetry of the system, Eq. (12) becomes23,24fn0i=23(ϵniϵ0i)Rni(r)|r|R0i(r)2,where the i stands for electron i = 1 (core electron transitions) or i = 2 (valence electron transitions). To confirm that our numerical approach has rendered a complete set of states, the TRK sum rule,25 Σnfn0 = Ne, must be satisfied, i.e., Ne = 3. Note that for absorption, Eq. (13) is positive, but for emission, i.e., when ϵni<ϵ0i, the DOS becomes negative.

    2. Static dipole polarizability

    The static dipole polarizability is defined through the DOS [Eq. (12)], and is given byαs=nfn0(EHFnEHF0)2=i,nfn0i(ϵniϵ0i)2,where it exhibits an explicit dependence on the single-electron dipole oscillator strengths. Consequently, it can be rewritten asαs=2αs1s+αs2s,where αsi is the contribution from the core (i = 1) or the valence (i = 2) electron. For photon absorption, the polarizability is positive, but for photon emission, it becomes negative.

    3. Mean excitation energy

    A parameter that characterizes the amount of energy loss when a swift heavy ion penetrates a target is provided by the mean excitation energy I0, as defined by Bethe:26lnI0=nfn0ln(EHFnEHF0)nfn0=i,nfn0iln(ϵniϵ0i)i,nfn0i.Using the TRK sum rule,25 Eq. (16) can be rewritten as3lnI0=2nfn01sln(ϵn1sϵ01s)+nfn02sln(ϵn2sϵ02s)=2lnI01s+lnI02s,which is the orbital decomposition or Bragg rule for the mean excitation energy, as used by Oddershede and Sabin27 in energy loss deposition studies.

    III. RESULTS

    A. Unconfined lithium-like atoms

    To show the reliability of our approach when applied to a multi-electron system, we present in Table I the results for the unconfined Li, Be+, and B2+ atoms. We show the core electron ground and excited orbital energies ϵ01s and ϵ02p, the total HF energy EHF, the DOS fis2pi for the first dipole transition, the polarizability αsi, and the mean excitation energy I0i. The same quantities are also reported for the valence (2s) electron. In the case of the ϵ01s and EHF energy values for the Li, Be+, and B2+ atoms, we observe good agreement up to four-decimal precision when compared with the results of Froese-Fischer.28 For the Li atom mean excitation energy I01s, we observe a difference of less than 5% with respect to the value reported by Oddershede and Sabin.27 For the valence (2s) electron, we obtain orbital energy values of ϵ02s,Li=0.20116 a.u., ϵ02s,Be+=0.67282 a.u., and ϵ02s,B2+=1.39722 a.u., for the Li, Be+, and B2+ atoms, respectively, with a difference of less than 3% in comparison with the results of Froese-Fischer.28 However, for the total HF energy, we obtain values with a difference of less than 1% with respect to the results of Froese-Fischer.28 For the dipole polarizability, we obtain αs2s,Li=171.188 a.u. and αs2s,Be+=27.3836 a.u., in comparison with values of 164.05 a.u. and 24.4966 a.u. reported by Schwerdtfeger and Nagle31 and Tang et al.,32 respectively.

    Valence (2s1)
    LiBe+B2+
    ϵ02s−0.201 16−0.672 82−1.397 22
    (−0.196 32)a(−0.666 15)a(−1.389 85)a
    ϵ02p−0.138 61−0.547 31−1.210 49
    EHF−7.437 49−14.283 8−23.38 26
    (−7.432 72)a(−14.277 4)a(−23.375 9)a
    (−7.419 23)c
    f2s2p2s0.651 270.413 150.299 63
    (0.767 1)d(0.510 9)d
    αs2s171.18827.383 68.994 40
    (164.05)e(24.496 6)f
    I02s3.567 8412.052 925.765 5
    (3.29)b

    Table 1. Unconfined ground state properties for free Li, Be+, and B2+ atoms. We report values for the core (i = 1) and valence (i = 2) electrons for the ground (ϵ0is) and excited (ϵ02p) orbital energies, the total HF energy EHF, the DOS fis2pi, the static dipole polarizability αsi, and the mean excitation energy I0i. Slater’s αX parameter, Eq. (9), takes the values αXLi=0.58002, αXBe+=0.52632, and αXB2+=0.49922.

    Thus, our approach using Slater’s X-α allows us to account for the ground state properties ϵ0is and EHF of the Li, Be+, and B2+ atoms, with good agreement with available theoretical results.

    In Fig. 1, we show the wavefunctions for the 1s and 2s orbitals as functions of position r for unconfined Li, Be+, and B2+ atoms. For comparison, we also show the results of Froese-Fischer.28 We observe excellent agreement for the 1s orbital for all ranges. For the 2s electron, we also obtain very good agreement for electron distances r > 1 a.u. However, we observe that small deviations appear for the inner part of the wavefunction, r < 1 a.u., compared with the Froese-Fischer results.

    Wavefunctions for the 1s and 2s states of Li, Be+, and B2+ atoms as functions of the radial coordinate r for the unconfined atoms. The curves are our results, while the symbols are from Ref. 28.

    Figure 1.Wavefunctions for the 1s and 2s states of Li, Be+, and B2+ atoms as functions of the radial coordinate r for the unconfined atoms. The curves are our results, while the symbols are from Ref. 28.

    B. Confined total and orbital HF energies

    In Fig. 2, we show the 1s, 2s, 2p, 3s, and 3p energy levels for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity as a function of the confinement radius R0. For comparison, we also show the results of Weiss33 for the free atom energy levels at R0 = 30 a.u. We can see that the energy levels increase, reaching the continuum, as the confinement radius decreases. Furthermore, we observe the appearance of crossing points for the 2s and 2p energy levels, as well as for the 3s and 3p levels, which are highlighted by circles for better visualization. Our results confirm the energy level behavior and crossing already reported from other approaches.34,35Figure 2(a) shows the 3s and 3p states of the Li atom. We find that the energy levels reach the continuum for cavities with R0 < 10 a.u. and R0 < 12 a.u., respectively. Here, the 3s energy level is deeper than the 3p state. As the confinement radius decreases, the 3s and 3p energy levels increase until a crossing point around R0 ∼ 6 a.u. For values of R0 < 6, the 3p energy is lower than the 3s energy. For the 3s and 3p states of the Be+ and B2+ atoms, we find a similar trend as for the Li atom. We can see that the crossing points between the 3s and 3p states of the Li, Be+, and B2+ atoms are in the positive spectrum. For the Be+ atom, we find that the crossing point occurs at R0 ∼ 4.6 a.u.

    Orbital energies for the 1s, 2s, 2p, 3s, and 3p states of Li, Be+, and B2+ atoms confined by a spherical impenetrable cavity as a function of the confinement radius R0: (a) 3s and 3p states; (b) 2s and 2p states; (c) 1s ground state. The crossing points between the ns–np levels are highlighted by circles for better visualization. The curves without symbols are for the ns states, while the curves with symbols are for the np states. For comparison, the HF results of Weiss33 for the unconfined atom energy levels are also shown at R0 = 30 a.u (▿).

    Figure 2.Orbital energies for the 1s, 2s, 2p, 3s, and 3p states of Li, Be+, and B2+ atoms confined by a spherical impenetrable cavity as a function of the confinement radius R0: (a) 3s and 3p states; (b) 2s and 2p states; (c) 1s ground state. The crossing points between the nsnp levels are highlighted by circles for better visualization. The curves without symbols are for the ns states, while the curves with symbols are for the np states. For comparison, the HF results of Weiss33 for the unconfined atom energy levels are also shown at R0 = 30 a.u (▿).

    In Fig. 2(b), we show the 2s and 2p energy levels, and we can again see an energy increase and the emergence of crossing points as R0 decreases. We find that the 2p energy level of the lithium atom becomes positive for R0 < 4.6 a.u. and the 2s energy level becomes positive for R0 < 4.4 a.u. In this case, the crossing point is present for R0 ∼ 3.4 a.u. This cavity size corresponds to a pressure of 85 GPa (see below), which is lower than the 210 GPa reported by Rahm et al.36 The discrepancy is attributed to the different confinement models. The same occurs for the Be+ atom at a confinement radius of R0 ∼ 2.5 a.u. and for B2+ at R0 ∼ 2.1 a.u. We obtain positive energy values for the 2s and 2p levels of the Be+ atom for R0 < 2.4 a.u. and R0 < 2.3 a.u., respectively. For the 2s and 2p states of the B2+ atom, we find positive energies for R0 < 1.71 a.u. and R0 < 1.55 a.u., respectively. Note that since our initial ground state electronic configuration occupies the 2s orbital level, for cavities with R0 lower than the crossing-point radius, one would have photon emission instead of photon absorption for the initial electronic configuration, 1s22s. We should stress that for R0 smaller than the critical cavity radius, the lowest-energy state of the Li-like systems becomes p-type, and hence the photon emission from the initial s to the final p states brings the excited electron to its ground state. Once this transition has taken place, as the cavity radius is reduced, the p-state evolution lies energetically below the corresponding s state, and hence the DOS become positive for excitations.

    In Fig. 2(c), we show the 1s energy levels as functions of R0. We find that the effect of the cavity on the 1s ground state energy is minimal for R0 > 2 a.u., R0 > 1.5 a.u., and R0 > 1 a.u. for Li, Be+, and B2+ atoms, respectively. For the Li atom, we observe a sudden change in the energy for R0 < 2 a.u. and it reaches a positive value for the 1s core state for R0 < 0.77 a.u. A similar situation occurs for the Be+ atom for R0 < 0.555 a.u. and for B2+ for R0 < 0.427 a.u.

    The increase in energy is explained as follows. For an impenetrable cavity, the electrons remain localized within the cavity. As the pressure increases, so does the electron kinetic energy (as a consequence of the Heisenberg uncertainty principle), and the total energy can become positive for a critical pressure,36 as we have just shown, but the system is still bounded.

    In Fig. 3, we show the results for the total HF energy for the Li, Be+, and B2+ atoms in the initial 1s22s configuration, confined by an impenetrable spherical cavity, as functions of the confinement radius. For comparison, we also show, in the case of Li, the theoretical results of Sañu-Ginarte et al.,29 Le Sech and Banerjee,37 Sarsa and Le Sech,38 and Sarsa et al.,13 and we can see that there is excellent agreement. In addition, the HF results for the unconfined atoms obtained by Weiss33 are shown at R0 = 5 a.u. For the free case, when R0 a.u., we obtain total energies of EHFLi=7.43749 a.u., EHFBe+=14.2838 a.u., and EHFB2+=23.3826 a.u., in good agreement with the values reported by Froese-Fischer28 and Sañu-Ginarte et al.29 As the confinement radius decreases, the HF energy increases, for all atoms, as previously reported by Connerade et al.39 For the Li, Be+, and B2+ atoms, we observe in Fig. 3 that the effect of the cavity is minimal for R0 > 7.3 a.u., R0 > 5.5 a.u., and R0 > 2.5 a.u., respectively. For the Li atom, for R0 < 1.3 a.u., the HF energy becomes positive. A similar situation occurs for the Be+ and B2+ atoms for R0 < 0.94 a.u. and R0 < 0.75 a.u., respectively.

    Total HF energy, Eq. (6), as a function of cavity radius R0 for Li, Be+, and B2+ atoms. In the case of Li, the symbols show the theoretical results from Sañu-Ginarte et al.29 (×), Le Sech and Banerjee37 (□), Sarsa and Le Sech38 (○), and Sarsa et al.13 (▵). The HF results for unconfined atoms as reported by Weiss33 are shown at R0 = 5 a.u (▿).

    Figure 3.Total HF energy, Eq. (6), as a function of cavity radius R0 for Li, Be+, and B2+ atoms. In the case of Li, the symbols show the theoretical results from Sañu-Ginarte et al.29 (×), Le Sech and Banerjee37 (□), Sarsa and Le Sech38 (○), and Sarsa et al.13 (▵). The HF results for unconfined atoms as reported by Weiss33 are shown at R0 = 5 a.u (▿).

    From Figs. 2 and 3, we conclude that the effect of the confinement cavity is stronger on the valence electrons that on the core electrons.

    C. Pressure

    The order of magnitude of the pressure that the cavity exerts on the atomic system as R0 is shrunk is given by the static pressureP=EHFv=14πR02EHFR0,where v is the volume of the spherical cavity. In Fig. 4, we show the results for the pressure as a function of the spherical cavity radius R0 for Li, Be+, and B2+ atoms. For comparison, we show some characteristic pressures found in nature. We first note that for the same cavity radius, the pressure is lowest for B2+, increases for Be+, and it is highest for Li. This is a consequence of the ionic character of the system. The lithium atom is more diffuse in its 2s orbital, so the same cavity radius induces a higher pressure, while, owing to the high nuclear charge, the boron ion has already compacted its 2s electron, so the same cavity radius induces a smaller pressure on the ionic system. The figure also shows the cavity size and pressure for which the 2s → 2p transition occurs in our approach. For Li, we find it at 85 GPa (R0 = 3.4 a.u.), for Be+ at 350 GPa (R0 = 2.5 a.u.), and for B2+ at 690 GPa (R0 = 2.1 a.u.). These results are within an order of magnitude of those reported by Rahm et al.,36 which were obtained using a different confinement model, thus confirming the suitability of our approach. Rahm et al. reported a higher pressure, probably because their model considers penetrable confinement conditions.

    Static pressure induced by the cavity as a function of cavity size R0 for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity. The open square symbols (□) indicate the cavity size and pressure at which the 2s → 2p transition occurs. Some naturally occurring pressures are also shown.

    Figure 4.Static pressure induced by the cavity as a function of cavity size R0 for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity. The open square symbols (□) indicate the cavity size and pressure at which the 2s → 2p transition occurs. Some naturally occurring pressures are also shown.

    D. Dipole oscillator strength

    In Fig. 5(a), we show the DOS for the electronic transitions 1s → 2p (core excitations) and 2s → 2p (valence excitations) for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity as a function of the confinement radius R0. In the case of the Li atom, for R0 < 15 a.u., f2s2p2s begins to decrease, showing a change near R0 ∼ 9 a.u., where we find a value of f2s2p2s=0.58895 a.u. At R0 = 6.5 a.u., we obtain a DOS of 0.460 46 a.u., and at R0 = 3.5 a.u. a value of f2s2p2s=0.02320 a.u., near the radius for which the crossing point occurs. For the Be+ atom, we observe similar behavior. The DOS for the transition decreases rapidly near R0 ∼ 5 a.u., with f2s2p2s=0.35781 a.u., and then reaches a value of f2s2p2s=0.01523 a.u. at R0 = 2.6 a.u., which is near the crossing point. For the B2+ atom, we observe the same DOS reduction as R0 is decreased until the crossing point at R0 = 2.1 a.u. When R0 is decreased, the DOS for the valence electron excitation, f2s2p2s, is reduced until the 2s electron reaches the crossing point between the 2s and 2p energy levels, so that at that pressure the DOS become zero (ϵ2s = ϵ2p). Crossing occurs for confinement radii R0 ∼ 3.4 a.u., 2.4 a.u., and 2.1 a.u. for Li, Be+ and B2+ atoms, respectively. For confinement radii less than the crossing point, the 2p energy levels have lower values than those found for the 2s energy level, and there is photon emission induced by the pressure cavity. We should note here that in a sudden approximation perturbation, for a shrinking of the cavity from radius R0 to R0 + ΔR0, the probability of finding the system in the 2p state is zero owing to symmetry arguments (orthogonal states). Thus, there is a higher probability for the system to remain in the same s symmetry state and then proceed to the 2p state by photon emission. Consequently, for a cavity radius lower than the critical crossing point, f2s2p2s becomes negative owing to photon emission, and some other transitions must increase its DOS value to satisfy the TRK sum rule. In Fig. 5(a), we also show the core results for the f1s2p1s transition, and we can see that the DOS increases as R0 is reduced. f1s2p1s shows an abrupt change near R0 ∼ 5 a.u., 3.5 a.u., and 3 a.u. for the Li, Be+, and B2+ atoms, respectively. For lower values of R0, the DOS transition increases, reaching values near 1 as consequence of confinement, thus becoming a dominant intensity line.

    Dipole oscillator strength for the is → 2p and is → 3p electronic transitions as a function of the impenetrable spherical cavity size R0 for Li, Be+, and B2+ atoms for i = 1 (core) and i = 2 (valence) electrons. The curves without symbols are for the is → 2p transition of the valence i = 2 electron, while the curves with symbols are for the i = 1 core electron.

    Figure 5.Dipole oscillator strength for the is → 2p and is → 3p electronic transitions as a function of the impenetrable spherical cavity size R0 for Li, Be+, and B2+ atoms for i = 1 (core) and i = 2 (valence) electrons. The curves without symbols are for the is → 2p transition of the valence i = 2 electron, while the curves with symbols are for the i = 1 core electron.

    In Fig. 5(b), we show the 1s → 3p and 2s → 3p DOS for Li, Be+, and B2+ atoms as functions of the cavity radius. As can be seen, for cavities with radius lower than the critical crossing point, f2s3p2s becomes larger than unity, although the TRK sum rule is satisfied for all cavity radii. Thus, 2s → 3p becomes the strongest transition, so there is a change in luminosity in the atom as the pressure increases, but in this case due to photon emission induced by the change in pressure, similar to piezoluminescence.40

    E. Static polarizability

    In Fig. 6, we show the static dipole polarizabilities αs2s and αs1s for the valence and core states for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity, as a function of the confinement radius R0. The crossing points are highlighted by vertical lines. Note that owing to the small contribution of the core electrons, the total atomic polarizability is dominated by the valence contribution for all pressures. For comparison, Fig. 6(a) also shows the unconfined Li and Be+ results as reported by Schwerdtfeger and Nagle11 and Tang et al.32 at R0 = 30 a.u., and it can be seen that there is good agreement with our results. For R0, we obtain the dipole polarizabilities for unconfined lithium-like atoms as αs2s,Li=171.188 a.u., αs2s,Be+=27.3836 a.u., and αs2s,B2+=8.99440 a.u., which exhibit a difference of ∼4% with respect to HF results.11 From Fig. 6, we observe that as the confinement radius decreases, so does the polarizability, until the sp crossing point is reached. In Fig. 6(a) for the Li atom, for a cavity with radius R0 = 4.4 a.u., the polarizability decreases to 54.2926 a.u., which is about 30% of the free value. As R0 is reduced, the 2s and 2p energy levels become positive, and the αs2s polarizability increases, diverging at R ∼ 3.4 a.u., which is at the critical crossing point of the 2s–2p energy levels. For lower values of R0, αs2s becomes negative owing to the transition to photon emission. In the case of the Be+ atom, at R0 = 6.5 a.u., we observe a value of αs2s=26.2083 a.u., and then the polarizability decreases for lower values of the confinement radius until R0 ∼ 3.4 a.u., where a minimum value of αs2s=16.3509 a.u. is found. Then, for values of R0 < 3.4 a.u., the polarizability increases rapidly, diverging at R0 ∼ 2.4 a.u., and it then becomes negative for lower values of R0. A similar situation occurs for the Be2+ atom, but with a minimum value of 6.772 95 a.u. at R0 ∼ 2.8 a.u. and a divergence at the crossing point R0 ∼ 2.1 a.u. In Fig. 6(b), we show the results for the core contribution αs1s, where the effect of the impenetrable cavity starts for R0 < 3 a.u. A decrease in αs1s is observed for lower values of R0, where the energy levels become positive. As noted already, the total static dipole polarizability αs=2α1s+αs2s is dominated by the contribution of the valence electron, αs2s.

    Static dipole polarizabilities αs2s (a) and αs1s (b) as functions of cavity size R0 for Li, Be+, and B2+ atoms. The solid triangle (▴) and the solid circle (•) at R0 = 30 a.u. are the HF results of Schwerdtfeger and Nagle11 and Tang et al.,32 respectively.

    Figure 6.Static dipole polarizabilities αs2s (a) and αs1s (b) as functions of cavity size R0 for Li, Be+, and B2+ atoms. The solid triangle (▴) and the solid circle (•) at R0 = 30 a.u. are the HF results of Schwerdtfeger and Nagle11 and Tang et al.,32 respectively.

    F. Mean excitation energy

    In Fig. 7, we show the mean excitation energies I01s, I02s, and I0, for Li, Be+, and B2+ atoms confined by an impenetrable spherical cavity as functions of the confinement radius R0. We can see that at R0 = 30 a.u., the results for the free mean excitation energies are in good agreement with previous HF results from Oddershede and Sabin,27 Kamakura,41 and Dehmer et al.42 From Fig. 7(a), we can see that as R0 decreases, I02s increases, showing an abrupt change near R0 ∼ 10 a.u., 5 a.u., and 4 a.u. for the Li, Be+, and B2+ atoms, respectively, with I02s,Li=3.87967 eV, I02s,Be+=14.73933 eV, and I02s,Be2+=28.5561 eV. For the Li valence electron, we observe an increase of ∼40% with respect to the free mean excitation energy at R0 = 4.4 a.u. For Be+, we find an increase of ∼11%, and for B2+ an increase of ∼9% for the same confinement. Figure 7(a) also shows I01s for Li, Be+, and B2+ atoms, and we observe an increase in the mean excitation energy as R0 is reduced. In Fig. 7(b), we show results for the total mean excitation energy I0. We find I0 = 33.708 90 a.u., in good agreement with the value of 34.004 13 a.u. obtained by Oddershede and Sabin,27 Eq. (17), and the value of 34 a.u. reported by Dehmer et al.42 for Li atoms. Note that as a consequence of 2s–2p energy level crossing, the photon emission produces a negative energy transfer, so the logarithmic contribution is undetermined, as defined by Eqs. (12) and (13). This is observed in the I02s contribution and in the total I0 mean excitation energy for R0 less than the critical cavity radius at the sp crossing energy levels. Thus, a different approach may be required to determine it, such as that proposed by Smith et al.43

    (a) Mean excitation energies I01s (curves with symbols) and I02s (curves without symbols) as functions of cavity size R0 for Li, Be+, and B2+ atoms. For comparison, we also show at R0 = 30 a.u. the values of the free atoms obtained by Oddershede and Sabin27 (○), Kamakura41 (▵), and Dehmer et al.42 (□). (b) Total mean excitation energy I0.

    Figure 7.(a) Mean excitation energies I01s (curves with symbols) and I02s (curves without symbols) as functions of cavity size R0 for Li, Be+, and B2+ atoms. For comparison, we also show at R0 = 30 a.u. the values of the free atoms obtained by Oddershede and Sabin27 (○), Kamakura41 (▵), and Dehmer et al.42 (□). (b) Total mean excitation energy I0.

    G. Slater’s X-α contribution

    One advantage of Slater’s X-α approach is that we can estimate the electron exchange contribution to the energy for a confined quantum system through a single parameter. In Fig. 8, we show the behavior of Slater’s X-α parameter αX as a function of the cavity confinement radius R0. We find that the largest contribution occurs for the Li atom, followed by the Be+ ion and then the B2+ ion for low-pressure cavities. The contribution increases as R0 decreases, reaching a maximum, and it then decreases as the cavity becomes small. For the ions, the 2s electrons are tighter and the electron exchange parameter is lower for large spherical cavities. However, this behavior is inverted as the cavity increases the pressure. For cavities whose radius is smaller than the critical radius, the αX parameter is largest for B2+, followed by Be+, and then Li. So, electron exchange is important as long as the valence electrons remain bounded.

    Slater’s X-α parameter αX as a function of cavity size R0 for Li, Be+, and B2+ atoms.

    Figure 8.Slater’s X-α parameter αX as a function of cavity size R0 for Li, Be+, and B2+ atoms.

    In Tables II and III, for reference purposes, we show the 1s, 2s, and 2p energy levels, the total HF energy, the first allowed DOS transition 2s → 2p, the dipole polarizability, the mean excitation energy, and Slater’s αX parameter [Eq. (9)] for selected values of the confinement radius R0 for Li, Be+, and B2+ atoms,

    R0ϵ01sϵ02sϵ02pEHFf1s2p1sf2s2p2sf2s3p2sαs1sαs2sI01sI02sαX [Eq. (9)]
    0.71.320 7731.446 016.934 431.131 00.990 43−0.606 481.571 510.005 350.000 16378.6090.351 35
    0.750.536 4926.840 514.525 625.115 20.989 92−0.605 381.571 750.006 740.000 48337.2660.357 48
    0.790.033 4923.793 812.929 221.173 30.989 19−0.604 201.571 540.008 010.000 87309.6990.362 48
    0.8−0.078 0623.105 912.568 320.288 00.988 97−0.603 871.571 430.008 350.000 99303.4490.363 73
    1.0−1.524 9413.559 07.5403 38.240 710.980 14−0.593 311.564 100.016 800.006 45215.3030.389 80
    (8.513 92)
    2.0−2.749 712.002 961.295 42−5.192 260.799 73−0.412 731.370 960.077 550.726 01111.3580.517 36
    (−5.084 19)
    3.0−2.791 160.415 510.343 33−6.820 140.543 18−0.112 391.046 190.100 1621.263 4103.9570.577 90
    4.0−2.792 330.018 790.067 24−7.217 590.385 240.136 820.779 410.102 3558.957 8103.62522.106 90.588 02
    (−7.185 99)
    5.0−2.792 36−0.112 58−0.040 06−7.348 980.310 790.305 190.595 470.102 4559.181 9103.61611.353 70.585 90
    (−7.323 00)
    6.0−2.792 36−0.163 46−0.088 74−7.399 870.279 080.418 020.465 070.102 4576.600 6103.6167.478 420.583 19
    (−7.376 79)
    8.0−2.792 36−0.194 16−0.125 16−7.430 560.262 820.550 790.287 460.102 45118.705103.6164.702 040.580 69
    (−7.409 85)
    10−2.792 36−0.199 96−0.135 15−7.436 360.261 290.614 550.172 210.102 45150.076103.6163.879 670.580 13
    (−7.416 58)
    −2.792 32−0.201 16−0.138 81−7.437 490.261 190.651 270.020 210.102 45171.188103.6133.567 840.580 02

    Table 2. Similar to Table I, but for a Li atom confined by an impenetrable spherical cavity for several selected values of the cavity size R0. The values in parentheses are the theoretical results from Sañu-Ginarte et al.29

    R0ϵ01sϵ02sϵ02pEHFf2s2p1sf2s2p2sf2s3p2sαs1sαs2sI01sI02sαX [Eq. (9)]
    Be+
    0.52.137 1260.173 131.471 160.325 60.990 70−0.608 331.572 960.001 41−0.000 21737.3080.346 04
    0.550.123 4848.159 925.269 944.587 60.990 17−0.607 041.573 590.001 94−0.000 14628.2270.353 88
    0.557−0.107 0846.742 924.537 942.747 00.990 03−0.606 801.573 600.002 02−0.000 16615.2580.354 98
    0.57−0.507 7444.255 423.252 539.525 30.989 73−0.606 311.573 560.002 18−0.000 21592.4330.357 06
    0.6−1.312 1539.150 520.612 932.954 10.988 82−0.604 971.573 170.002 58−0.000 36545.3470.361 89
    0.8−4.145 5919.084 910.201 97.861 460.973 72−0.587 661.559 610.006 20−0.003 52356.7270.395 24
    1.0−5.095 4710.385 15.645 46−2.410 340.940 73−0.553 611.525 120.011 02−0.016 18274.0380.428 96
    2.0−5.663 140.637 480.342 34−12.970 40.609 97−0.194 041.134 740.026 48−2.156 21198.7830.530 27
    3.0−5.667 09−0.414 12−0.332 76−14.025 30.395 470.118 460.787 490.027 4818.111 5197.52341.12820.534 69
    4.0−5.667 10−0.614 43−0.488 48−14.225 60.325 820.279 470.581 180.027 4918.076 4197.51520.37020.529 12
    5.0−5.667 09−0.659 95−0.531 32−14.271 20.308 630.357 810.441 370.027 4922.364 1197.51514.73930.526 99
    6.0−5.667 09−0.670 23−0.543 24−14.281 40.305 470.393 280.336 290.027 4925.356 9197.51512.85790.526 44
    10−5.667 08−0.672 89−0.547 30−14.284 00.304 960.413 090.147 450.027 4927.374 4197.51412.05470.526 30
    −5.667 11−0.672 82−0.547 31−14.283 80.304 970.413 150.117 100.027 4927.383 6197.50512.05290.526 32
    B2+
    0.42.057 8691.865 347.233 090.81370.990 79−0.609 091.574 080.000 570.01157.600.345 26
    0.4250.026 8679.606 140.953 174.73720.990 50−0.608 351.574 570.000 700.01045.470.350 04
    0.43−0.328 0677.419 939.832 971.88630.990 41−0.608 161.574 620.000 730.01025.370.351 00
    0.6−6.736 8833.751 617.428 116.43870.978 38−0.593 421.565 240.002 07−0.002 80614.9800.384 90
    0.8−8.761 8515.098 47.7981 9−5.712 240.938 23−0.551 231.523 060.004 37−0.006 89435.7360.425 00
    1.0−9.319 567.291 263.712 27−14.396 30.869 37−0.480 511.447 140.006 72−0.030 34365.2450.460 22
    2.0−9.541 72−0.670 23−0.699 29−22.656 20.477 74−0.018 720.938 650.010 21−22.104 5321.5150.510 64
    3.0−9.541 95−1.309 56−1.135 42−23.295 70.352 200.204 470.646 290.010 246.908 57321.39440.52560.501 88
    4.0−9.541 94−1.387 56−1.200 28−23.373 70.331 750.277 280.469 210.010 248.205 13321.39428.55610.499 53
    5.0−9.541 93−1.396 37−1.209 33−23.382 50.329 760.295 810.343 880.010 248.831 21321.39326.18550.499 23
    6.0−9.541 92−1.397 17−1.210 37−23.383 30.329 640.299 160.264 510.010 248.970 87321.39225.81350.499 20
    10−9.541 89−1.397 23−1.210 49−23.383 30.329 630.299 660.186 630.010 248.993 81321.39025.76540.499 20
    −9.541 58−1.397 22−1.210 49−23.382 60.329 650.299 630.184 990.010 248.994 40321.36825.76550.499 22

    Table 3. Similar to Table II, but for the Be+ and B2+ atoms.

    IV. CONCLUSIONS

    We have studied lithium-like atoms confined by an impenetrable spherical cavity of radius R0. We find good to excellent agreement when comparing orbital and total energies, as well as when determining dipole transitions, static polarizability, and mean excitation energies for the unconfined systems. For the lithium atom, we find excellent agreement for confined ground state energies in comparison with available theoretical results.

    We confirm that, as a consequence of the confinement, the system orbital and total energies increase as the pressure increases owing to a reduction in cavity size. However, the first allowed dipole transition, 2s → 2p, decreases, while 2s → 3p increases. Consequently, as the pressure increases, the intensity of light emitted by the atom in the cavity is shifted. However, there is a crossing point (critical pressure) at which the 2s and 2p energy levels are inverted; consequently, the DOS for that transition becomes zero at that critical pressure. For higher pressures, the DOS become negative owing to photon emission. In addition, the 2s → 3p DOS reach values larger than unity for high pressures, and the 2s → 2p DOS becomes negative. Thus, we can confirm that the static dipole polarizability is reduced as the pressure increases, as the electrons become highly localized within the cavity and less prone to be polarized, and diverges at the point of transition from photon absorption to photon emission. We also find that the mean excitation energy, which measures the ability of the atom to absorb energy due to excitations, increases as the pressure is increased, with implications for material damage under extreme conditions. As a result of the existence of the crossing point, the valence and total mean excitation energy become undetermined owing to a logarithmic indeterminacy, and thus a different approach may be required.

    Our work shows the reliability of Slater’s X-α approach in the context of HF theory to study confined N-electron quantum systems. This approach has the advantage that it can be extended to larger systems to provide excitation spectra in different confinement environments, thus shedding light on the behavior of N-body quantum systems under extreme conditions.

    References

    [1] A. Bijl, J. De Boer, A. Michels. Remarks concerning molecural interaction and their influence on the polarisability. Physica, 4, 981-994(1937).

    [2] S. A. Cruz, J. Sabin, E. Brandas. Advances in Quantum Chemistry(2009).

    [3] S. A. Cruz, J. Sabin, E. Brandas. Advances in Quantum Chemistry(2009).

    [4] K. D. Sen. Electronic Structure of Quantum Confined Atoms and Molecules(2014).

    [5] W. Jaskólski. Confined many-electron systems. Phys. Rep., 271, 1-66(1996).

    [6] E. Ley-Koo. Recent progress in confined atoms and molecules: Superintegrability and symmetry breakings. Rev. Mex. Fis., 64, 326-363(2018).

    [7] G. McGinn. Atomic and molecular calculations with the pseudopotential method. VII One-valence-electron photoionization cross sections. J. Chem. Phys., 53, 3635-3640(1970).

    [8] J. N. Bardsley. Pseudopotential calculations of alkali interactions. Chem. Phys. Lett., 7, 517-520(1970).

    [9] P. G. Burke, W. D. Robb. The r-matrix theory of atomic processes. Adv. At. Mol. Phys., 11, 143-214(1976).

    [10] M. J. Seaton, H. E. Saraph, G. Peach. Atomic data for opacity calculations. IX. The lithium isoelectronic sequence. J. Phys. B: At., Mol. Opt. Phys., 21, 3669(1988).

    [11] P. Schwerdtfeger. The pseudopotential approximation in electronic structure theory. ChemPhysChem, 12, 3143-3155(2011).

    [12] C. Y. Lin, Y. K. Ho. Photoionization of atoms encapsulated by cages using the power-exponential potential. J. Phys. B: At., Mol. Opt. Phys., 45, 145001(2012).

    [13] E. Buendía, A. Sarsa, F. J. Gálvez. Study of confined many electron atoms by means of the POEP method. J. Phys. B: At., Mol. Opt. Phys., 47, 185002(2014).

    [14] J. W. Cooper. Photoionization from outer atomic subshells. A model study. Phys. Rev., 128, 681-693(1962).

    [15] E. V. Ludeña. SCF Hartree-Fock calculations of ground state wavefunctions of compressed atoms. J. Chem. Phys., 69, 1770-1775(1978).

    [16] J. C. Slater. A simplification of the Hartree-Fock method. Phys. Rev., 81, 385-390(1951).

    [17] A. Zangwill. A half century of density functional theory. Phys. Today, 68, 34-39(2015).

    [18] A. Szabo, N. S. Ostlund. Modern Quantum Chemistry: Introduction to Advanced Electronic Structure Theory(1996).

    [19] M. S. Pindzola. Excitation and charge transfer in proton-lithium collisions at 5–15 kev. Phys. Rev. A, 60, 3764-3768(1999).

    [20] M. Higuchi, M. Miyasita, K. Higuchi. An alternative scheme for calculating the unrestricted Hartree–Fock equation: Application to the boron and neon atoms. Physica B, 407, 2758-2762(2012).

    [21] S. A. Cruz, R. Cabrera-Trujillo. Confinement approach to pressure effects on the dipole and the generalized oscillator strength of atomic hydrogen. Phys. Rev. A, 87, 012502(2013).

    [22] S. Koonin, D. C. Meredith. Computational Physics: Fortran Version(1998).

    [23] M. Inokuti. Inelastic collisions of fast charged particles with atoms and molecules—The Bethe theory revisited. Rev. Mod. Phys., 43, 297-347(1971).

    [24] H. Friedrich. Theoretical Atomic Physics(2006).

    [25] R. Jackiw, H. A. Bethe. Intermediate Quantum Mechanics(1997).

    [26] H. Bethe. Zur theorie des durchgangs schneller korpuskularstrahlen durch materie. Ann. Phys., 397, 325-400(1930).

    [27] J. Oddershede, J. R. Sabin. Orbital and whole-atom proton stopping power and shell corrections for atoms with z < 36. At. Data Nucl. Data Tables, 31, 275-297(1984).

    [28] C. Froese-Fischer. A general multi-configuration Hartree-Fock program. Comput. Phys. Commun., 14, 145-153(1978).

    [29] R. Betancourt-Riera, L. Ferrer-Galindo, L. A. Ferrer-Moreno, R. Riera, A. D. Sañu-Ginarte, E. M. Guillén-Romero. New approach to obtain the analytical expression of the energy functional in free or confined atoms. Results Phys., 13, 102261(2019).

    [30] A. W. Weiss. The calculation of atomic oscillator strengths: The lithium atom revisited. Can. J. Chem., 70, 456-463(1992).

    [31] J. K. Nagle, P. Schwerdtfeger. 2018 Table of static dipole polarizabilities of the neutral elements in the periodic table. Mol. Phys., 117, 1200-1225(2019).

    [32] Z.-C. Yan, J. Mitroy, T.-Y. Shi, L.-Y. Tang. Dynamic dipole polarizabilities of the Li atom and the Be+ ion. Phys. Rev. A, 81, 042521(2010).

    [33] A. W. Weiss. Wave functions and oscillator strengths for the lithium isoelectronic sequence. Astrophys. J., 138, 1262(1963).

    [34] P. A. Lakshmi, V. K. Dolmatov, J. P. Connerade. The filling of shells in compressed atoms. J. Phys. B: At., Mol. Opt. Phys., 33, 251(2000).

    [35] J. P. Connerade, R. Semaoune. Atomic compressibility and reversible insertion of atoms into solids. J. Phys. B: At., Mol. Opt. Phys., 33, 3467-3484(2000).

    [36] R. Cammi, R. Hoffmann, M. Rahm, N. W. Ashcroft. Squeezing all elements in the periodic table: Electron configuration and electronegativity of the atoms under compression. J. Am. Chem. Soc., 141, 10253-10271(2019).

    [37] A. Banerjee, C. Le Sech. A variational approach to the dirichlet boundary condition: Application to confined H, He and Li. J. Phys. B: At., Mol. Opt. Phys., 44, 105003(2011).

    [38] A. Sarsa, C. Le Sech. Variational Monte Carlo method with Dirichlet boundary conditions: Application to the study of confined systems by impenetrable surfaces with different symmetries. J. Chem. Theory Comput., 7, 2786-2794(2011).

    [39] P. Anantha Lakshmi, J. P. Connerade, P. Kengkan, R. Semaoune. Scaling laws for atomic compressibility. J. Phys. B: At., Mol. Opt. Phys., 33, L847-L854(2000).

    [40] N. A. Atari. Piezoluminescence phenomenon. Phys. Lett. A, 90, 93-96(1982).

    [41] H. Tsuchida, H. Ogawa, S. Kamakura, M. Inokuti, N. Sakamoto. Mean excitation energies for the stopping power of atoms and molecules evaluated from oscillator-strength spectra. J. Appl. Phys., 100, 064905(2006).

    [42] M. Inokuti, R. P. Saxon, J. L. Dehmer. Systematics of moments of dipole oscillator-strength distributions for atoms of the first and second row. Phys. Rev. A, 12, 102-121(1975).

    [43] M. Inokuti, E. Shiles, D. Y. Smith, W. Karstens. Mean excitation energy for the stopping power of light elements. Nucl. Instrum. Methods Phys. Res., Sect. B, 250, 1-5(2006).

    C. Martínez-Flores, R. Cabrera-Trujillo. High pressure effects on the excitation spectra and dipole properties of Li, Be+, and B2+ atoms under confinement[J]. Matter and Radiation at Extremes, 2020, 5(2): 24401
    Download Citation