11institutetext: Technische Universität Darmstadt, Department of Physics, 64289 Darmstadt, Germany 22institutetext: ExtreMe Matter Institute EMMI, GSI Helmholtzzentrum für Schwerionenforschung GmbH, 64291 Darmstadt, Germany 33institutetext: IRFU, CEA, Université Paris-Saclay, 91191 Gif-sur-Yvette, France 44institutetext: KU Leuven, Department of Physics and Astronomy, Instituut voor Kern- en Stralingsfysica, 3001 Leuven, Belgium 55institutetext: CEA, DAM, DIF, 91297 Arpajon, France 66institutetext: Université Paris-Saclay, CEA, Laboratoire Matière en Conditions Extrêmes, 91680 Bruyères-le-Châtel, France 77institutetext: CEA, DES, IRESNE, DER, SPRC, 13108 Saint-Paul-lès-Durance, France 88institutetext: Helmholtz Forschungsakademie Hessen für FAIR, GSI Helmholtzzentrum, 64289 Darmstadt, Germany

Ab initio description of monopole resonances in light- and medium-mass nuclei

IV. Angular momentum projection and rotation-vibration coupling
A. Porro\thanksrefad:tud,ad:emmi,ad:saclay    T. Duguet\thanksrefad:saclay,ad:kul    J.-P. Ebran\thanksrefad:dam,ad:dam_u    M. Frosini\thanksrefad:cadarache    R. Roth\thanksrefad:tud,ad:darm2    V. Somà\thanksrefad:saclay
(Received: July 1, 2024 / Revised version: date)
Abstract

Giant Resonances are, with nuclear rotations, the most evident expression of collectivity in finite nuclei. These two categories of excitations, however, are traditionally described within different formal schemes, such that vibrational and rotational degrees of freedom are separately treated and coupling effects between those are often neglected. The present work puts forward an approach aiming at a consitent treatment of vibrations and rotations. Specifically, this paper is the last in a series of four dedicated to the investigation of the giant monopole resonance in doubly open-shell nuclei via the ab initio Projected Generator Coordinate Method (PGCM). The present focus is on the treatment and impact of angular momentum restoration within such calculations. The PGCM being based on the use of deformed mean-field states, the angular-momentum restoration is performed when solving the secular equation to extract vibrational excitations. In this context, it is shown that performing the angular momentum restoration only after solving the secular equation contaminates the monopole response with an unphysical coupling to the rotational motion, as was also shown recently for (quasi-particle) random phase approximation calculations based on a deformed reference state. Eventually, the present work based on the PGCM confirms that an a priori angular momentum restoration is necessary to handle consistently both collective motions at the same time. This further pleads in favor of implementing the full-fledged projected (quasi-particle) random phase approximation in the future.

1 Introduction

Giant Resonances (GRs) Berman and Fultz (1975); Harakeh and Woude (2001); Garg and Colò (2018); Colò (2022) are collective nuclear excitations that can best be pictured in terms of oscillations of the nuclear surface in an effective liquid-drop model. GRs are, in this respect, one of the clearest manifestations of collective motion in finite nuclei. The other most evident collective behaviour in nuclei is provided by rotational excitations Elliott (1958). These features have traditionally been described through empirical models adopting various resolutions, e.g. the macroscopic Bohr-Mottelson collective models Bohr and Mottelson (1998) and their microscopic counterparts Rowe (1985); Libert et al. (2016), the Interacting Boson Model Arima and Iachello (1976, 1978); Iachello and Arima (1987), the Fermion Dynamical Symmetry Model Wu et al. (1987), etc. In a microscopic framework, a unified and rigorous treatment of both rotations and vibrations is not amenable to simple solutions of the Schrödinger equation. For this reason, the two are often addressed separately. This is justified by the fact that nuclear rotations and vibrations typically pertain to different energy regimes. Indeed, rotations are low-energy excitations, with ground-state rotational bands typically spanning few MeV’s above the nuclear ground state. Instead, GRs typically appear above 10 MeV excitation energy with large differences depending on the system and on the multipolarity of interest.

When addressing doubly open-shell nuclei, both realms are typically approached starting from mean-field reference states breaking rotational symmetry, e.g. deformed Hartree-Fock-Bogoliubov (HFB) vacua. Rotations are then extracted by performing an angular momentum projection (AMP) acting either on one deformed HFB vacuum Sheikh and Ring (2000) or on a linear superposition of them, thus defining the projected generator coordinate method (PGCM) Bender et al. (2003); Robledo et al. (2018). Lately, rotational spectra have also been generated starting from deformed coupled cluster (CC) calculations with explicit AMP Hagen et al. (2022); Sun et al. (2024); Hu et al. (2024). As for GRs, the (quasi-particle) random phase approximation ((Q)RPA) is the usual method of choice, where nuclear vibrations are treated as harmonic fluctuations around the deformed HFB minimum Péru and Goutte (2008); Yoshida (2009); Avogadro and Nakatsukasa (2011); Beaujeault-Taudière et al. (2023), without any AMP.

In a recent work Porro et al. (2024a), an AMP in RPA calculations was attempted a posteriori, i.e. after solving the RPA secular equation based on the deformed reference state. The corresponding monopole response was shown to be contaminated with an unphysical coupling to the rotational motion and an empirical method was designed to subtract it. On a conceptually similar ground, when U(1)𝑈1U(1)italic_U ( 1 ) symmetry is broken, QRPA pair transfer probabilities have been demonstrated to overestimate the exact results within the exactly solvable Richardson model Richardson and Sherman (1964). The most significant discrepancies occur near the transition from the normal to the superfluid phase Gambacurta and Lacroix (2012). On the other hand, the present series of four papers Porro et al. (2024b, c, d) addressing the giant monopole resonance (GMR) from an ab initio standpoint has demonstrated the suitability of the symmetry-conserving PGCM to address GRs. In this context, the goal of the present paper, the fourth of the series, is to investigate whether performing the AMP a posteriori on top of the GCM solutions induces the same shortcomings as those observed within the RPA framework and thus confirm that the a priori AMP is mandatory to properly handling the coupling between rotational and vibrational motions.

The present paper, denoted as Paper IV, is organized as follows111The first three papers of the series are denoted as Paper I Porro et al. (2024b), Paper II Porro et al. (2024c) and Paper III Porro et al. (2024d), respectively.. First, the different strategies to perform a symmetry restoration within the (P)GCM are formally introduced in Sec. 2 before discussing the impact on the strength function in Sec. 3. In Sec. 4, the empirical method used to remove the spurious coupling due to the a posteriori AMP is detailed. Numerical results are eventually presented in Sec. 5 whereas conclusions are drawn in Sec. 6 .

2 Formalism

The (P)GCM formalism was introduced in detail in Paper I Porro et al. (2024b). The essential elements needed to discuss the effects of the symmetry breaking and restoration in strength functions are only briefly recalled below.

2.1 The generator coordinate method

The original GCM formulation is introduced first. The GCM wave-function ansatz Hill and Wheeler (1953); Griffin and Wheeler (1957) is a general superposition of so-called generating functions reading as

|Ψνqfν(q)|Φ(q),ketsubscriptΨ𝜈subscript𝑞subscript𝑓𝜈𝑞ketΦ𝑞\ket{\Psi_{\nu}}\equiv\sum_{q}f_{\nu}(q)\ket{\Phi(q)}\,,| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ ≡ ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ , (1)

where q𝑞qitalic_q denotes a set of variables referred to as the generator coordinates. The index ν𝜈\nuitalic_ν refers to a principal quantum number while fν(q)subscript𝑓𝜈𝑞f_{\nu}(q)italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) is a weight function to be determined222The mixing coefficients {fν(q),q[q0,q1]}subscript𝑓𝜈𝑞𝑞subscript𝑞0subscript𝑞1\{f_{\nu}(q),\,q\in[q_{0},q_{1}]\}{ italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) , italic_q ∈ [ italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] } are defined such that |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ is normalized.. The ensemble {|Φ(q),q[q0,q1]}ketΦ𝑞𝑞subscript𝑞0subscript𝑞1\{|\Phi(q),\,q\in[q_{0},q_{1}]\rangle\}{ | roman_Φ ( italic_q ) , italic_q ∈ [ italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] ⟩ } denotes a set of non-orthogonal Bogoliubov states typically obtained as solutions of constrained HFB calculations requiring that the solution satisfies

Φ(q)|Q|Φ(q)quantum-operator-productΦ𝑞𝑄Φ𝑞\displaystyle\langle\Phi(q)|Q|\Phi(q)\rangle⟨ roman_Φ ( italic_q ) | italic_Q | roman_Φ ( italic_q ) ⟩ =q,absent𝑞\displaystyle=q\,,= italic_q , (2)

where Q𝑄Qitalic_Q denotes a set of operators defining the collective coordinates.

The unknown coefficients fν(q)subscript𝑓𝜈𝑞f_{\nu}(q)italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) are determined variationally based on Ritz’ variational principle, namely by minimising the energy associated with |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩

δΨν|H|ΨνΨν|Ψν=0.𝛿quantum-operator-productsubscriptΨ𝜈𝐻subscriptΨ𝜈inner-productsubscriptΨ𝜈subscriptΨ𝜈0\delta\frac{\braket{\Psi_{\nu}}{H}{\Psi_{\nu}}}{\braket{\Psi_{\nu}}{\Psi_{\nu}% }}=0.italic_δ divide start_ARG ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | start_ARG italic_H end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ end_ARG start_ARG ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ end_ARG = 0 . (3)

The variation with respect to the weights fν(q)superscriptsubscript𝑓𝜈𝑞f_{\nu}^{*}(q)italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_q ) eventually leads to a generalised eigenvalue problem known as the Hill-Wheeler-Griffin (HWG) secular equation Griffin and Wheeler (1957) reading as

q[(p,q)Eν𝒩(p,q)]fν(q)=0,subscript𝑞delimited-[]𝑝𝑞subscript𝐸𝜈𝒩𝑝𝑞subscript𝑓𝜈𝑞0\sum_{q}\Big{[}\mathcal{H}(p,q)-E_{\nu}\mathcal{N}(p,q)\Big{]}f_{\nu}(q)=0\,,∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT [ caligraphic_H ( italic_p , italic_q ) - italic_E start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT caligraphic_N ( italic_p , italic_q ) ] italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) = 0 , (4)

where the so-called Hamiltonian and norm kernels are respectively defined as

(p,q)𝑝𝑞\displaystyle\mathcal{H}(p,q)caligraphic_H ( italic_p , italic_q ) Φ(p)|H|Φ(q),absentquantum-operator-productΦ𝑝𝐻Φ𝑞\displaystyle\equiv\braket{\Phi(p)}{H}{\Phi(q)}\,,≡ ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_H end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ , (5a)
𝒩(p,q)𝒩𝑝𝑞\displaystyle\mathcal{N}(p,q)caligraphic_N ( italic_p , italic_q ) Φ(p)|Φ(q).absentinner-productΦ𝑝Φ𝑞\displaystyle\equiv\braket{\Phi(p)}{\Phi(q)}\,.≡ ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ . (5b)

2.2 The projected generator coordinate method

Constrained HFB solutions typically break symmetries of the Hamiltonian, so that restoring those symmetries is mandatory to obtain approximations to exact eigentates carrying good symmetry quantum numbers. The symmetry restoration is achieved within the PGCM by adding a projection operator onto good symmetry quantum numbers to the symmetry-breaking GCM state. While presently focusing on rotational symmetry associated with angular-momentum conservation, the approach is general and consists of modifying Eq. (1) according to333The mixing coefficients {fνσ(q),q[q0,q1]}subscriptsuperscript𝑓𝜎𝜈𝑞𝑞subscript𝑞0subscript𝑞1\{f^{\sigma}_{\nu}(q),\,q\in[q_{0},q_{1}]\}{ italic_f start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) , italic_q ∈ [ italic_q start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] } are defined such that |ΨνσketsubscriptsuperscriptΨ𝜎𝜈\ket{\Psi^{\sigma}_{\nu}}| start_ARG roman_Ψ start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ is normalized.

|Ψνσ=qfνσ(q)Pσ|Φ(q),ketsuperscriptsubscriptΨ𝜈𝜎subscript𝑞superscriptsubscript𝑓𝜈𝜎𝑞superscript𝑃𝜎ketΦ𝑞\ket{\Psi_{\nu}^{\sigma}}=\sum_{q}\,f_{\nu}^{\sigma}(q)P^{\sigma}\ket{\Phi(q)}\,,| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ = ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q ) italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ , (6)

where Pσsuperscript𝑃𝜎P^{\sigma}italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT is the projection operator associated with the (a) symmetry (sub)group 𝒢𝒢\mathcal{G}caligraphic_G of the Hamiltonian. The projection operator selects the components of each |Φ(q)ketΦ𝑞\ket{\Phi(q)}| start_ARG roman_Φ ( italic_q ) end_ARG ⟩ carrying the good symmetry quantum numbers σ(JMΠNZ)𝜎𝐽𝑀Π𝑁𝑍\sigma\equiv(JM\Pi NZ)italic_σ ≡ ( italic_J italic_M roman_Π italic_N italic_Z ), i.e. the PGCM ansatz in Eq. (6) has good total angular momentum J𝐽Jitalic_J and angular momentum projection M𝑀Mitalic_M, parity Π=±1Πplus-or-minus1\Pi=\pm 1roman_Π = ± 1 as well as neutron N𝑁Nitalic_N and proton Z𝑍Zitalic_Z numbers. The projector Pσsuperscript𝑃𝜎P^{\sigma}italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT can be generically written as

Pσ=dφgσ(φ)R(φ),superscript𝑃𝜎d𝜑superscript𝑔𝜎𝜑𝑅𝜑P^{\sigma}=\int\text{d}\varphi\,g^{\sigma}(\varphi)R(\varphi)\,,italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT = ∫ d italic_φ italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) italic_R ( italic_φ ) , (7)

where gσ(φ)superscript𝑔𝜎𝜑g^{\sigma}(\varphi)italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) represents the irreducible representations of 𝒢𝒢\mathcal{G}caligraphic_G whereas R(φ)𝑅𝜑R(\varphi)italic_R ( italic_φ ) denotes the unitary symmetry transformation operator changing the orientation of a state by an angle φ𝜑\varphiitalic_φ. The PGCM ansatz can thus be expanded as

|ΨνσketsuperscriptsubscriptΨ𝜈𝜎\displaystyle\ket{\Psi_{\nu}^{\sigma}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ =qfνσ(q)dφgσ(φ)R(φ)|Φ(q)absentsubscript𝑞superscriptsubscript𝑓𝜈𝜎𝑞d𝜑superscript𝑔𝜎𝜑𝑅𝜑ketΦ𝑞\displaystyle=\sum_{q}\,f_{\nu}^{\sigma}(q)\int\text{d}\varphi\,g^{\sigma}(% \varphi)R(\varphi)\ket{\Phi(q)}= ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q ) ∫ d italic_φ italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) italic_R ( italic_φ ) | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
qfνσ(q)dφgσ(φ)|Φ(q,φ),absentsubscript𝑞superscriptsubscript𝑓𝜈𝜎𝑞d𝜑superscript𝑔𝜎𝜑ketΦ𝑞𝜑\displaystyle\equiv\sum_{q}\,f_{\nu}^{\sigma}(q)\int\text{d}\varphi\,g^{\sigma% }(\varphi)\ket{\Phi(q,\varphi)}\,,≡ ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q ) ∫ d italic_φ italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) | start_ARG roman_Φ ( italic_q , italic_φ ) end_ARG ⟩ , (8)

where the φ𝜑\varphiitalic_φ-rotated Bogoliubov state |Φ(q,φ)R(φ)|Φ(q)ketΦ𝑞𝜑𝑅𝜑ketΦ𝑞\ket{\Phi(q,\varphi)}\equiv~{}R(\varphi)\ket{\Phi(q)}| start_ARG roman_Φ ( italic_q , italic_φ ) end_ARG ⟩ ≡ italic_R ( italic_φ ) | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ has been introduced. Applying the variational procedure based on the PGCM ansatz (Eq. (2.2)) leads now to a set of σ𝜎\sigmaitalic_σ-dependent HWG equations

q[σ(p,q)Eνσ𝒩σ(p,q)]fνσ(q)=0,subscript𝑞delimited-[]superscript𝜎𝑝𝑞subscriptsuperscript𝐸𝜎𝜈superscript𝒩𝜎𝑝𝑞superscriptsubscript𝑓𝜈𝜎𝑞0\sum_{q}\,\Big{[}\mathcal{H}^{\sigma}(p,q)-E^{\sigma}_{\nu}\mathcal{N}^{\sigma% }(p,q)\Big{]}f_{\nu}^{\sigma}(q)=0\,,∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT [ caligraphic_H start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) - italic_E start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT caligraphic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) ] italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q ) = 0 , (9)

where the so-called symmetry-restored Hamiltonian and norm kernels are defined as

σ(p,q)superscript𝜎𝑝𝑞\displaystyle\mathcal{H}^{\sigma}(p,q)caligraphic_H start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) Φ(p)|HPσ|Φ(q)absentquantum-operator-productΦ𝑝𝐻superscript𝑃𝜎Φ𝑞\displaystyle\equiv\braket{\Phi(p)}{HP^{\sigma}}{\Phi(q)}≡ ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_H italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
=dφgσ(φ)Φ(p)|H|Φ(q,φ),absentd𝜑superscript𝑔𝜎𝜑quantum-operator-productΦ𝑝𝐻Φ𝑞𝜑\displaystyle=\int\text{d}\varphi\,g^{\sigma}(\varphi)\braket{\Phi(p)}{H}{\Phi% (q,\varphi)}\,,= ∫ d italic_φ italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_H end_ARG | start_ARG roman_Φ ( italic_q , italic_φ ) end_ARG ⟩ , (10a)
𝒩σ(p,q)superscript𝒩𝜎𝑝𝑞\displaystyle\mathcal{N}^{\sigma}(p,q)caligraphic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) Φ(p)|Pσ|Φ(q)absentquantum-operator-productΦ𝑝superscript𝑃𝜎Φ𝑞\displaystyle\equiv\braket{\Phi(p)}{P^{\sigma}}{\Phi(q)}≡ ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
=dφgσ(φ)Φ(p)|Φ(q,φ).absentd𝜑superscript𝑔𝜎𝜑inner-productΦ𝑝Φ𝑞𝜑\displaystyle=\int\text{d}\varphi\,g^{\sigma}(\varphi)\braket{\Phi(p)}{\Phi(q,% \varphi)}\,.= ∫ d italic_φ italic_g start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_φ ) ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG roman_Φ ( italic_q , italic_φ ) end_ARG ⟩ . (10b)

2.3 PAV-GCM

In the PGCM described above the secular equation (Eq. (9)) is solved in presence of the symmetry projection, i.e. the variational minimization is restricted to each irreducible representation of the symmetry group. For this reason, this scheme is presently denoted as the variation after projection GCM (VAP-GCM).

In between the GCM and the VAP-GCM, a scheme can be considered in which the symmetry projection is performed only after the GCM solution based on a symmetry breaking ansatz has been obtained. Such an intermediate approach is naturally denoted as the projection after variation GCM (PAV-GCM) scheme.

Projecting the GCM states solution of Eqs. (1)-(4), one works with the set of projected states

|Ψ~νσketsubscriptsuperscript~Ψ𝜎𝜈\displaystyle\ket{\tilde{\Psi}^{\sigma}_{\nu}}| start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ NνσPσ|Ψν=Nνσqfν(q)Pσ|Φ(q),absentsuperscriptsubscript𝑁𝜈𝜎superscript𝑃𝜎ketsubscriptΨ𝜈superscriptsubscript𝑁𝜈𝜎subscript𝑞subscript𝑓𝜈𝑞superscript𝑃𝜎ketΦ𝑞\displaystyle\equiv N_{\nu}^{\sigma}P^{\sigma}\ket{\Psi_{\nu}}=N_{\nu}^{\sigma% }\sum_{q}f_{\nu}(q)P^{\sigma}\ket{\Phi(q)}\,,≡ italic_N start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ = italic_N start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT | start_ARG roman_Φ ( italic_q ) end_ARG ⟩ , (11)

where Nνσsuperscriptsubscript𝑁𝜈𝜎N_{\nu}^{\sigma}italic_N start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT is a normalising factor provided by the condition

11\displaystyle 11 =Ψ~νσ|Ψ~νσ,absentinner-productsuperscriptsubscript~Ψ𝜈𝜎superscriptsubscript~Ψ𝜈𝜎\displaystyle=\braket{\tilde{\Psi}_{\nu}^{\sigma}}{\tilde{\Psi}_{\nu}^{\sigma}% }\,,= ⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG | start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ , (12)

such that

(Nνσ)2=pqfν(p)fν(q)𝒩σ(p,q).superscriptsuperscriptsubscript𝑁𝜈𝜎2subscript𝑝𝑞superscriptsubscript𝑓𝜈𝑝subscript𝑓𝜈𝑞superscript𝒩𝜎𝑝𝑞(N_{\nu}^{\sigma})^{-2}=\sum_{pq}f_{\nu}^{*}(p)f_{\nu}(q)\mathcal{N}^{\sigma}(% p,q)\,.( italic_N start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT = ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_p ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) caligraphic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) . (13)

2.4 Discussion

{NiceTabular}

c:c Symmetry breaking
GCM (Q)RPA
      Large-amplitude superposition       of deformed HF(B) states       Harmonic fluctuations       around a deformed HF(B) state
Status: available Status: available
PAV GCM PAV (Q)RPA
      Angular-momentum projection       of deformed GCM states       Angular-momentum projection       of deformed (Q)RPA states
Status: developed in this work Status: developed in Refs. Erler (2012); Porro (2023); Porro et al. (2024a).
PGCM P(Q)RPA
      Proper treatment of rotation-vibration       coupling within the GCM       Proper treatment of rotation-vibration       coupling within the (Q)RPA
Status: available Status: formalism available Federschmidt and Ring (1985); Tsuchimochi and Van Voorhis (2015)
Symmetry conserving

Table 1: Schematic representation of different projection levels based on symmetry-breaking GCM and (Q)RPA.

The PAV strategy assumes that rotational and vibrational degrees of freedom are strictly decoupled, i.e. intrinsically deformed solutions are first obtained assuming that rotational degrees of freedom are frozen before adding the rotational motion to each vibrational state thus obtained.

From a physical standpoint, such a decoupling presuppose that vibrations and rotations relate to very different time scales, such that they can be addressed separately in a Born-Oppenheimer-like approximation. Specifically, rotations are assumed to be infinitely slower than nuclear vibrations, which is a direct consequence of the rotation being ideally associated with a zero-energy (Goldstone) mode.

However, nuclear rotations happen in fact at finite frequencies, such that they cannot be decoupled a priori from vibrational modes Bohr and Mottelson (1998). Thus, the variational/diagonalisation process at play to determine physical states should be performed in a Hilbert subspace simultaneously accounting for vibrational and rotational degrees of freedom. This is achieved in the VAP scheme that is the method of choice to consistently treat the coupling effects between nuclear vibrations and rotations. A schematic summary of the different levels of symmetry breaking and restoration in the GCM and the (Q)RPA is displayed in Table 2.4.

3 Transition strength

Generically, the ground-state strength function associated with an arbitrary excitation operator O𝑂Oitalic_O reads as

SF(ω)νσ|Θ0σ0|O|Θνσ|2δ(EνσE0σ0ω),subscript𝑆𝐹𝜔subscript𝜈superscript𝜎superscriptquantum-operator-productsubscriptsuperscriptΘsubscript𝜎00𝑂subscriptsuperscriptΘsuperscript𝜎𝜈2𝛿subscriptsuperscript𝐸superscript𝜎𝜈subscriptsuperscript𝐸subscript𝜎00𝜔S_{F}(\omega)\equiv\sum_{\nu\sigma^{\prime}}|\braket{\Theta^{\sigma_{0}}_{0}}{% O}{\Theta^{\sigma^{\prime}}_{\nu}}|^{2}\,\delta(E^{\sigma^{\prime}}_{\nu}-E^{% \sigma_{0}}_{0}-\omega)\,,italic_S start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ω ) ≡ ∑ start_POSTSUBSCRIPT italic_ν italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUBSCRIPT | ⟨ start_ARG roman_Θ start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG italic_O end_ARG | start_ARG roman_Θ start_POSTSUPERSCRIPT italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_δ ( italic_E start_POSTSUPERSCRIPT italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT - italic_E start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_ω ) , (14)

where |ΘνσketsubscriptsuperscriptΘ𝜎𝜈|\Theta^{\sigma}_{\nu}\rangle| roman_Θ start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ⟩ (Eνσsubscriptsuperscript𝐸𝜎𝜈E^{\sigma}_{\nu}italic_E start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT) denotes an eigenstate (eigenenergy) of the nuclear Hamiltonian.

The present work focuses on the monopole response and on the K=0𝐾0K=0italic_K = 0 component of the quadrupole response, respectively associated with the excitation operators

O𝑂\displaystyle Oitalic_O r2i=1Ari2.absentsuperscript𝑟2superscriptsubscript𝑖1Asuperscriptsubscript𝑟𝑖2\displaystyle\equiv r^{2}\equiv\sum_{i=1}^{\text{A}}r_{i}^{2}\,.≡ italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT A end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (15)

and

O𝑂\displaystyle Oitalic_O Q20i=1Ari2Y20(ϑi,ϕi),absentsubscript𝑄20superscriptsubscript𝑖1𝐴superscriptsubscript𝑟𝑖2subscript𝑌20subscriptitalic-ϑ𝑖subscriptitalic-ϕ𝑖\displaystyle\equiv Q_{20}\equiv\sum_{i=1}^{A}r_{i}^{2}\,Y_{20}(\vartheta_{i},% \phi_{i})\,,≡ italic_Q start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT ≡ ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_A end_POSTSUPERSCRIPT italic_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Y start_POSTSUBSCRIPT 20 end_POSTSUBSCRIPT ( italic_ϑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (16)

where (ϑi,ϕi)subscriptitalic-ϑ𝑖subscriptitalic-ϕ𝑖(\vartheta_{i},\phi_{i})( italic_ϑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) denote spherical angular coordinates. Below, the ingredients entering Eq. (14) in GCM, PAV-GCM and VAP-GCM calculations are specified.

3.1 GCM

The unprojected transition amplitude between the GCM ground state and a GCM excited state reads as

Ψ0|O|Ψνquantum-operator-productsubscriptΨ0𝑂subscriptΨ𝜈\displaystyle\braket{\Psi_{0}}{O}{\Psi_{\nu}}⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG italic_O end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ =pqf0(p)fν(q)Φ(p)|O|Φ(q)absentsubscript𝑝𝑞superscriptsubscript𝑓0𝑝subscript𝑓𝜈𝑞quantum-operator-productΦ𝑝𝑂Φ𝑞\displaystyle=\sum_{pq}f_{0}^{*}(p)f_{\nu}(q)\braket{\Phi(p)}{O}{\Phi(q)}= ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_p ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_O end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
pqf0(p)O(p,q)fν(q).absentsubscript𝑝𝑞superscriptsubscript𝑓0𝑝𝑂𝑝𝑞subscript𝑓𝜈𝑞\displaystyle\equiv\sum_{pq}f_{0}^{*}(p)O(p,q)f_{\nu}(q)\,.≡ ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_p ) italic_O ( italic_p , italic_q ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) . (17)

Using more compact notations, the transition matrix element between any two GCM states can be written as

Ψμ|O|Ψνquantum-operator-productsubscriptΨ𝜇𝑂subscriptΨ𝜈\displaystyle\braket{\Psi_{\mu}}{O}{\Psi_{\nu}}⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_μ end_POSTSUBSCRIPT end_ARG | start_ARG italic_O end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ pqfμpOpqfqν(fOf)μν,absentsubscript𝑝𝑞subscriptsuperscript𝑓𝜇𝑝subscript𝑂𝑝𝑞subscript𝑓𝑞𝜈subscriptsuperscriptfOf𝜇𝜈\displaystyle\equiv\sum_{pq}f^{*}_{\mu p}O_{pq}f_{q\nu}\equiv(\textbf{f}^{% \dagger}\cdot\textbf{O}\cdot\textbf{f})_{\mu\nu}\,,≡ ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUPERSCRIPT ∗ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_μ italic_p end_POSTSUBSCRIPT italic_O start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_q italic_ν end_POSTSUBSCRIPT ≡ ( f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ⋅ O ⋅ f ) start_POSTSUBSCRIPT italic_μ italic_ν end_POSTSUBSCRIPT , (18)

where the indices of the matrix O associated with the operator kernel O(p,q)𝑂𝑝𝑞O(p,q)italic_O ( italic_p , italic_q ) run over the generator coordinates whereas the linear coefficient matrix f indices run on the generator coordinates for the lines and on the GCM states for the columns. Naturally, the energies entering Eq. (14) are the GCM energies delivered by Eq. (4).

3.2 VAP-GCM

In PGCM calculations, the transition amplitude reads as

Ψ0σ0|O|Ψνσquantum-operator-productsuperscriptsubscriptΨ0subscript𝜎0𝑂superscriptsubscriptΨ𝜈𝜎\displaystyle\braket{\Psi_{0}^{\sigma_{0}}}{O}{\Psi_{\nu}^{\sigma}}⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG | start_ARG italic_O end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ =pqf0σ0(p)fνσ(q)Φ(p)|Pσ0OPσ|Φ(q)absentsubscript𝑝𝑞superscriptsubscript𝑓0subscript𝜎0𝑝superscriptsubscript𝑓𝜈𝜎𝑞quantum-operator-productΦ𝑝superscript𝑃subscript𝜎0𝑂superscript𝑃𝜎Φ𝑞\displaystyle=\sum_{pq}f_{0}^{\sigma_{0}*}(p)f_{\nu}^{\sigma}(q)\braket{\Phi(p% )}{P^{\sigma_{0}\dagger}OP^{\sigma}}{\Phi(q)}= ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_p ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q ) ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_P start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT italic_O italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
pqf0σ0(p)Oσ0σ(p,q)fνσ(q)absentsubscript𝑝𝑞superscriptsubscript𝑓0subscript𝜎0𝑝superscript𝑂subscript𝜎0𝜎𝑝𝑞superscriptsubscript𝑓𝜈𝜎𝑞\displaystyle\equiv\sum_{pq}f_{0}^{\sigma_{0}*}(p)O^{\sigma_{0}\sigma}(p,q)f_{% \nu}^{\sigma}(q)≡ ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∗ end_POSTSUPERSCRIPT ( italic_p ) italic_O start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_p , italic_q ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ( italic_q )
pqf0pσ0Opqσ0σfqνσabsentsubscript𝑝𝑞superscriptsubscript𝑓0𝑝subscript𝜎0superscriptsubscript𝑂𝑝𝑞subscript𝜎0𝜎superscriptsubscript𝑓𝑞𝜈𝜎\displaystyle\equiv\sum_{pq}f_{0p}^{\sigma_{0}*}O_{pq}^{\sigma_{0}\sigma}f_{q% \nu}^{\sigma}≡ ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 italic_p end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∗ end_POSTSUPERSCRIPT italic_O start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_σ end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_q italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT
(fσ0Oσ0σfσ)0ν,absentsubscriptsuperscriptfsubscript𝜎0superscriptOsubscript𝜎0𝜎superscriptf𝜎0𝜈\displaystyle\equiv(\textbf{f}^{\sigma_{0}\dagger}\cdot\textbf{O}^{\sigma_{0}% \sigma}\cdot\textbf{f}^{\sigma})_{0\nu}\,,≡ ( f start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT ⋅ O start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_σ end_POSTSUPERSCRIPT ⋅ f start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT ) start_POSTSUBSCRIPT 0 italic_ν end_POSTSUBSCRIPT , (19)

where the matrix OσσsuperscriptO𝜎superscript𝜎\textbf{O}^{\sigma\sigma^{\prime}}O start_POSTSUPERSCRIPT italic_σ italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT associated with the projected kernel Oσσ(p,q)superscript𝑂𝜎superscript𝜎𝑝𝑞O^{\sigma\sigma^{\prime}}(p,q)italic_O start_POSTSUPERSCRIPT italic_σ italic_σ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT ( italic_p , italic_q ) carries the symmetry quantum numbers of both the bra and ket. Indeed, and contrary to the Hamiltonian, the operator O𝑂Oitalic_O is not necessarily a scalar under symmetry transformations. Naturally, the energies entering Eq. (14) are the PGCM energies delivered by Eq. (9).

3.3 PAV-GCM

Given the PAV-GCM states introduced in Eq. (11), the corresponding transition amplitude reads as

Ψ~0σ0|O|Ψ~νσquantum-operator-productsubscriptsuperscript~Ψsubscript𝜎00𝑂subscriptsuperscript~Ψ𝜎𝜈\displaystyle\braket{\tilde{\Psi}^{\sigma_{0}}_{0}}{O}{\tilde{\Psi}^{\sigma}_{% \nu}}⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG italic_O end_ARG | start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ =N0σ0Nνσpqf0(p)fν(q)Φ(p)|Pσ0OPσ|Φ(q)absentsubscriptsuperscript𝑁subscript𝜎00subscriptsuperscript𝑁𝜎𝜈subscript𝑝𝑞subscript𝑓0𝑝subscript𝑓𝜈𝑞quantum-operator-productΦ𝑝superscript𝑃subscript𝜎0𝑂superscript𝑃𝜎Φ𝑞\displaystyle=N^{\sigma_{0}}_{0}N^{\sigma}_{\nu}\sum_{pq}f_{0}(p)f_{\nu}(q)% \braket{\Phi(p)}{P^{\sigma_{0}\dagger}OP^{\sigma}}{\Phi(q)}= italic_N start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ∑ start_POSTSUBSCRIPT italic_p italic_q end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_p ) italic_f start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( italic_q ) ⟨ start_ARG roman_Φ ( italic_p ) end_ARG | start_ARG italic_P start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT † end_POSTSUPERSCRIPT italic_O italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG | start_ARG roman_Φ ( italic_q ) end_ARG ⟩
=N0σ0Nνσ(fOσ0σf)0ν.absentsubscriptsuperscript𝑁subscript𝜎00subscriptsuperscript𝑁𝜎𝜈subscriptsuperscriptfsuperscriptOsubscript𝜎0𝜎f0𝜈\displaystyle=N^{\sigma_{0}}_{0}N^{\sigma}_{\nu}(\textbf{f}^{\dagger}\cdot% \textbf{O}^{\sigma_{0}\sigma}\cdot\textbf{f})_{0\nu}\,.= italic_N start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ⋅ O start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_σ end_POSTSUPERSCRIPT ⋅ f ) start_POSTSUBSCRIPT 0 italic_ν end_POSTSUBSCRIPT . (20)

Up to a normalising factor, the PAV-GCM transition amplitude combines the projected kernels introduced in Eq. (19) with the mixing coefficients of the two involved GCM states.

In this scheme, the energies entering Eq. (14) are the PAV-GCM energies delivered by

E~νσsubscriptsuperscript~𝐸𝜎𝜈\displaystyle\tilde{E}^{\sigma}_{\nu}over~ start_ARG italic_E end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT Ψ~νσ|H|Ψ~νσΨ~νσ|Ψ~νσ=|Nνσ|2(fHσσf)νν,absentquantum-operator-productsubscriptsuperscript~Ψ𝜎𝜈𝐻subscriptsuperscript~Ψ𝜎𝜈inner-productsubscriptsuperscript~Ψ𝜎𝜈subscriptsuperscript~Ψ𝜎𝜈superscriptsubscriptsuperscript𝑁𝜎𝜈2subscriptsuperscriptfsuperscriptH𝜎𝜎f𝜈𝜈\displaystyle\equiv\frac{\braket{\tilde{\Psi}^{\sigma}_{\nu}}{H}{\tilde{\Psi}^% {\sigma}_{\nu}}}{\braket{\tilde{\Psi}^{\sigma}_{\nu}}{\tilde{\Psi}^{\sigma}_{% \nu}}}=|N^{\sigma}_{\nu}|^{2}(\textbf{f}^{\dagger}\cdot\textbf{H}^{\sigma% \sigma}\cdot\textbf{f})_{\nu\nu}\,,≡ divide start_ARG ⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | start_ARG italic_H end_ARG | start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ end_ARG start_ARG ⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ end_ARG = | italic_N start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( f start_POSTSUPERSCRIPT † end_POSTSUPERSCRIPT ⋅ H start_POSTSUPERSCRIPT italic_σ italic_σ end_POSTSUPERSCRIPT ⋅ f ) start_POSTSUBSCRIPT italic_ν italic_ν end_POSTSUBSCRIPT , (21)

where in fact a single projector is sufficient to compute the kernel given that the Hamiltonian is a scalar under rotation.

4 Spurious coupling to rotational motion

In this section the concept of spurious coupling to the rotational motion in theories breaking angular-momentum conservation is introduced. For the reasons mentioned in Sec. 2.2, vibrational GCM excitations obtained without AMP may be non-orthogonal to rotational states. This feature is considered spurious given that the neglect of the rotational degrees of freedom in the HWG equation precisely assumes that the intrinsic GCM states are fully decoupled from them. Based on this consideration, a method is now designed to subtract a posteriori the spurious coupling between excited GCM states and a pure rotational motion of the corresponding ground state.

4.1 Subtracted GCM

Given the overlap between an arbitrary GCM excited state |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ and the PAV-GCM ground state

aνsubscript𝑎𝜈\displaystyle a_{\nu}italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT Ψ~0σ0|Ψν,absentinner-productsubscriptsuperscript~Ψsubscript𝜎00subscriptΨ𝜈\displaystyle\equiv\braket{\tilde{\Psi}^{\sigma_{0}}_{0}}{\Psi_{\nu}}\,,≡ ⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ , (22)

the spurious coupling can be subtracted by redefining the excited state444The procedure can be applied to any many-body method accessing symmetry-breaking ground and excited states. as Faessler and Nojarov (1988)

|Ψ˘νNν˘[|Ψνaν|Ψ~0σ0].ketsubscript˘Ψ𝜈subscript𝑁˘𝜈delimited-[]ketsubscriptΨ𝜈subscript𝑎𝜈ketsubscriptsuperscript~Ψsubscript𝜎00\ket{\breve{\Psi}_{\nu}}\equiv N_{\breve{\nu}}\big{[}\ket{\Psi_{\nu}}-a_{\nu}% \ket{\tilde{\Psi}^{\sigma_{0}}_{0}}\big{]}\,.| start_ARG over˘ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ ≡ italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT [ | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ - italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ⟩ ] . (23)

It is immediate to check that the orthogonalisation condition

Ψ~0σ0|Ψ˘ν=0inner-productsubscriptsuperscript~Ψsubscript𝜎00subscript˘Ψ𝜈0\braket{\tilde{\Psi}^{\sigma_{0}}_{0}}{\breve{\Psi}_{\nu}}=0\,⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG over˘ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ = 0 (24)

is satisfied and that the normalisation constant Nν˘subscript𝑁˘𝜈N_{\breve{\nu}}italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT is given by

(Nν˘)2=1|aν|2.superscriptsubscript𝑁˘𝜈21superscriptsubscript𝑎𝜈2(N_{\breve{\nu}})^{-2}=1-|a_{\nu}|^{2}\,.( italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT = 1 - | italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (25)

Replacing |ΨνketsubscriptΨ𝜈\ket{\Psi_{\nu}}| start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ by |Ψ˘νketsubscript˘Ψ𝜈\ket{\breve{\Psi}_{\nu}}| start_ARG over˘ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ in Eq. (17), the subtracted GCM (sub-GCM) transition strength can be computed.

4.2 Subtracted PAV-GCM

The subtraction method is extended to the PAV-GCM via the introduction of

|Ψ~˘νσNν˘σPσ|Ψ˘ν,ketsuperscriptsubscript˘~Ψ𝜈𝜎superscriptsubscript𝑁˘𝜈𝜎superscript𝑃𝜎ketsubscript˘Ψ𝜈\ket{\breve{\tilde{\Psi}}_{\nu}^{\sigma}}\equiv N_{\breve{\nu}}^{\sigma}P^{% \sigma}\ket{\breve{\Psi}_{\nu}}\,,| start_ARG over˘ start_ARG over~ start_ARG roman_Ψ end_ARG end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ ≡ italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT italic_P start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT | start_ARG over˘ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ , (26)

where the coefficients aνsubscript𝑎𝜈a_{\nu}italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT are now defined to fulfill the orthogonalisation condition

Ψ~0σ0|Ψ~˘νσ=0.inner-productsubscriptsuperscript~Ψsubscript𝜎00superscriptsubscript˘~Ψ𝜈𝜎0\braket{\tilde{\Psi}^{\sigma_{0}}_{0}}{\breve{\tilde{\Psi}}_{\nu}^{\sigma}}=0\,.⟨ start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG | start_ARG over˘ start_ARG over~ start_ARG roman_Ψ end_ARG end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ = 0 . (27)

In fact, |Ψ~˘νσketsuperscriptsubscript˘~Ψ𝜈𝜎\ket{\breve{\tilde{\Psi}}_{\nu}^{\sigma}}| start_ARG over˘ start_ARG over~ start_ARG roman_Ψ end_ARG end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ differ from the PAV-GCM state |Ψ~νσketsuperscriptsubscript~Ψ𝜈𝜎\ket{\tilde{\Psi}_{\nu}^{\sigma}}| start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT end_ARG ⟩ only if it carries the same symmetry quantum numbers as the ground state, i.e. σ=σ0𝜎subscript𝜎0\sigma=\sigma_{0}italic_σ = italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, which in the case of present interest corresponds to J=0𝐽0J=0italic_J = 0 states. Correspondingly, the coefficient aνsubscript𝑎𝜈a_{\nu}italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT satisfies Eq. (22) and the normalising factor Nν˘σ0superscriptsubscript𝑁˘𝜈subscript𝜎0N_{\breve{\nu}}^{\sigma_{0}}italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT reads

(Nν˘σ0)2superscriptsuperscriptsubscript𝑁˘𝜈subscript𝜎02\displaystyle(N_{\breve{\nu}}^{\sigma_{0}})^{-2}( italic_N start_POSTSUBSCRIPT over˘ start_ARG italic_ν end_ARG end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT =Ψν|Pσ0|Ψν|aν|2=(Nνσ0)2|aν|2.absentquantum-operator-productsubscriptΨ𝜈superscript𝑃subscript𝜎0subscriptΨ𝜈superscriptsubscript𝑎𝜈2superscriptsuperscriptsubscript𝑁𝜈subscript𝜎02superscriptsubscript𝑎𝜈2\displaystyle=\braket{\Psi_{\nu}}{P^{\sigma_{0}}}{\Psi_{\nu}}-|a_{\nu}|^{2}=(N% _{\nu}^{\sigma_{0}})^{-2}-|a_{\nu}|^{2}\,.= ⟨ start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG | start_ARG italic_P start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG | start_ARG roman_Ψ start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ - | italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = ( italic_N start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT - | italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT . (28)

Replacing |Ψ~νσketsubscriptsuperscript~Ψ𝜎𝜈\ket{\tilde{\Psi}^{\sigma}_{\nu}}| start_ARG over~ start_ARG roman_Ψ end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ by |Ψ~˘νσketsubscriptsuperscript˘~Ψ𝜎𝜈\ket{\breve{\tilde{\Psi}}^{\sigma}_{\nu}}| start_ARG over˘ start_ARG over~ start_ARG roman_Ψ end_ARG end_ARG start_POSTSUPERSCRIPT italic_σ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT end_ARG ⟩ in Eq. (20), the subtracted PAV-GCM (sub-PAV-GCM) transition strength can be computed.

5 Applications

The above considerations are illustrated below in the case of 28Si that acts as a typical example. The general conclusions reached below have been checked to be valid in all the nuclei studied in Papers I, II and III.

5.1 Numerical setting

Calculations whose results are presented below are realised employing a spherical harmonic oscillator basis characterized by ω=12Planck-constant-over-2-pi𝜔12\hbar\omega=12roman_ℏ italic_ω = 12 MeV and emax=10subscript𝑒max10\;e_{\!{}_{\;\text{max}}}=10italic_e start_POSTSUBSCRIPT start_FLOATSUBSCRIPT max end_FLOATSUBSCRIPT end_POSTSUBSCRIPT = 10 and the chiral effective field theory (χ𝜒\chiitalic_χEFT) Hamiltonian of Ref. Hüther et al. (2020) built at next-to-next-to-next-to-leading order (N3LO). Two-dimensional (P)GCM calculations are performed in axial symmetry using (r,β2)𝑟subscript𝛽2(r,\beta_{2})( italic_r , italic_β start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) as generator coordinates. All details of the calculations can be found in Paper II Porro et al. (2024c). Specifically, here particle-number projection is always included such that GCM calculations only omit the AMP, which is focus of the present work.

5.2 Levels of calculation

The different levels of calculations discussed in the present work are

  • GCM: no AMP.

  • PAV-GCM: AMP performed a posteriori on GCM states.

  • VAP-GCM: full PGCM with AMP a priori.

  • sub-GCM: GCM with a posteriori subtraction of the rotational coupling.

  • sub-PAV-GCM: AMP performed a posteriori on sub-GCM states.

5.3 Results

Refer to caption
Figure 1: Monopole (top) and quadrupole (bottom) responses for GCM and VAP-GCM calculations in 28Si.

The monopole (quadrupole) GCM and VAP-GCM responses in 28Si are compared in the upper (lower) panel of Fig. 1. The GCM monopole response is fragmented among four peaks in the interval [11,23]1123[11,23][ 11 , 23 ] MeV, the two dominant ones being located around 18181818 MeV. These four peaks are fully correlated with those appearing in the quadrupole response, even though their relative weights are different in the two cases. This correlation is the fingerprint of the coupling between both modes due to the intrinsic (oblate) deformation of 28Si. This topic was discussed at length in Paper II Porro et al. (2024c).

The inclusion of the AMP in VAP-GCM calculations impacts the responses in two ways. First, the excitation energy of the four dominant peaks are shifted up by 1.41.41.41.4, 1.01.01.01.0, 1.51.51.51.5 and 0.70.70.70.7 MeV, respectively, in the J=0𝐽0J=0italic_J = 0 (monopole) channel and by 2.42.42.42.4, 2.12.12.12.1, 2.52.52.52.5 and 1.71.71.71.7 MeV in the J=2𝐽2J=2italic_J = 2 (quadrupole) channel, i.e. the J=0𝐽0J=0italic_J = 0 and J=2𝐽2J=2italic_J = 2 states originating from the same intrinsic state are no longer strictly degenerate. Second, the intensity of the peaks is modified. The quadrupole response is significantly suppressed, i.e. while the strength of the first two peaks is divided by about a factor of three, the third peak has entirely disappeared and the fourth peak has been severely shrunk555The third (fourth) peak visible in the VAP-GCM quadrupole response corresponds to the fourth (fifth) peak in its GCM counterpart.. The intensity of the peaks in the monopole response remains overall unchanged except that the relative weight of the two main contributions near 18181818 MeV is strongly modified. Overall, the a priori inclusion of the AMP impacts the monopole and quadrupole responses non negligibly but it does so in a way that maintains a close connection to their intrinsic GCM counterparts.

Refer to caption
Figure 2: Monopole (top) and quadrupole (bottom) responses for VAP-GCM and PAV-GCM calculations in 28Si. For the monopole response, the overlap between excited GCM states and the PAV-GCM ground state is also displayed, the corresponding scale being shown on the right y-axis.

Next, VAP-GCM and PAV-GCM results are compared in Fig. 2. The quadrupole responses are very similar except for a shift up by about 1111 MeV for the J=2𝐽2J=2italic_J = 2 VAP-GCM excitation energies compared to the PAV-GCM ones. Contrarily, the monopole responses differ very notably. As a matter of fact, the a posteriori AMP impacts the monopole amplitudes much more significantly666Notice the rescaling of the PAV-GCM response by the factor 1/5151/51 / 5. than in the VAP-GCM calculation777The J=0𝐽0J=0italic_J = 0 energies are little affected, i.e. PAV-GCM and VAP-GCM energies typically differ by 0.50.50.50.5 MeV.. This is particularly true beyond the first three peaks where the originally subleading fourth peak is very strongly enhanced, along with many significant peaks appearing at even higher energies where no GCM strength was visible in the first place. The anomalously large impact of the a posteriori AMP compared to the VAP-GCM results makes the validity of the PAV-GCM monopole response dubious.

In order to analyse the content of these results, the overlap aνsubscript𝑎𝜈a_{\nu}italic_a start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT between excited GCM states and the PAV-GCM ground-state is also displayed in Fig. 2. The PAV-GCM monopole strength happens to be anomalously large888The overlap reaches 0.050.050.050.05 in 28Si and is up to three times larger in other studied nuclei such as 46Ti or 24Mg. In the studied cases, the coupling is larger for low-energy (high-energy) states in prolate (oblate) nuclei. for GCM excited states that are strongly coupled to the PAV-GCM ground state, even though no associated strength was originally present in the GCM response.

Refer to caption
Figure 3: Monopole (top) and quadrupole (bottom) responses for GCM and sub-GCM calculations in 28Si.
Refer to caption
Figure 4: Monopole response for VAP-GCM and sub-PAV-GCM calculations in 28Si.

As discussed in Sec. 4, the spurious coupling to the rotational motion can be subtracted a posteriori from GCM and PAV-GCM results. Figure 3 demonstrates that, even though intrinsically-deformed GCM states carry such a spurious coupling, it has no impact on the associated monopole strength function. In particular, while the full AMP accomplished via VAP-GCM eventually improve over GCM results the latter still deliver a meaningful approximation of the former999As discussed above in connection with Fig. 1, in 28Si the GCM monopole strength actually provides a quantitatively satisfactory approximation to the VAP-GCM one.. Contrarily, the subtraction of the spurious coupling strongly corrects the PAV-GCM strength function as seen in Fig. 4 such that the sub-PAV-GCM monopole response becomes consistent with the VAP-GCM one. In fact, the sub-PAV-GCM monopole response remains very close to the original GCM one, i.e. once the spurious component is removed, the effect of the AMP is underestimated when performed a posteriori.

6 Conclusions

The impact of angular momentum projection (AMP) on the monopole and quadrupole responses of doubly open-shell nuclei has been investigated within the frame of the (projected) generator coordinate method ((P)GCM) based on intrinsically deformed mean-field states. More specifically, the objective was to investigate whether the AMP can be safely performed a posteriori, i.e. solving the secular equation for intrinsic states within the GCM and projecting the solutions on good angular momentum only afterwards. To do so, results were confronted with results from full PGCM calculations where the AMP is performed a priori, i.e. where the secular equation is solved directly for good-symmetry states.

Using 28Si as typical example, the angular momentum projection was shown to have a non-negligible impact on both the monopole and quadrupole responses in full PGCM calculations. First, the position of the dominant peaks are shifted up by about 1111 MeV (2222 MeV) in the J=0𝐽0J=0italic_J = 0 (J=2𝐽2J=2italic_J = 2) channel. Second, while only the relative weight of certain monopole transitions are modified, quadrupole transitions are strongly suppressed.

Next, the a posteriori angular-momentum restoration was shown to contaminate the monopole response with an unphysical coupling to the rotational motion, a result that is fully consistent with the one recently observed in (quasi-particle) random phase approximation calculations based on a deformed reference state Porro et al. (2024a). Eventually, the present work based on the PGCM confirms that an a priori angular momentum restoration is necessary to handle consistently rotational and vibrational collective motions at the same time, which further pleads in favor of implementing the full-fledged projected (quasi-particle) random phase approximation in the future.

Acknowledgements

Calculations were performed by using HPC resources from GENCI-TGCC (Contract No. A0130513012). A.P. was supported by the CEA NUMERICS program, which has received funding from the European Union’s Horizon 2020 research and innovation program under the Marie Skłodowska-Curie grant agreement No 800945. A.P. and R.R. are supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) – Projektnummer 279384907 – SFB 1245. R.R. acknowledges support though the BMBF Verbundprojekt 05P2021 (ErUM-FSP T07, Contract No. 05P21RDFNB). M.F. is supported by the CEA-SINET project.

Data Availability Statement

This manuscript has no associated data or the data will not be deposited.

References