"rue-v THE; 315 4 LEL'RARY Midnigan State 29' 0 Universfiy This is to certify that the dissertation entitled Towards a Microscopic Energy Density Functional for Nuclei presented by Biruk B. Gebremariam has been accepted towards fulfillment of the requirements for the Ph.D degree in Physics and Astronomy z‘mréjEPP/OTESSOF'S Signature 2 0/0 Date MSU is an Affirmative Action/Equal Opportunity Employer a.-I-1-I‘O-U----‘¢-O-I---I-I-Q-l-I-I-I-I-o-l-l-n-o--.— .U-.-.-I-I-'-'-I-I----I-l-l- -_-> — PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 5/08 KzlProleocaPresIClRCIDateDue.indd m7 UNA R f) TOWARDS A MICROSCOPIC ENERGY DENSITY FUNCTIONAL FOR NUCLEI By Biruk B. Gebremariam A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Physics and Astronomy 2010 WAR , . . . . o o . r . O .. .. . .. . . . p -.. \ ' I” rutu“b L . I. . , l . l _ t. u. IA. ,‘ ‘4" ‘I 1" A- A. ' \ ' n I, v. ‘1" "‘ '- J . .- I . My. 8.7- n t.‘ A'b - d . ‘ ‘A, ‘ 'J I. u ‘ I 1". . \_ .‘ ”- a ' , U , a. _‘ r" l I ABSTRACT TOWARDS A MICROSCOPIC ENERGY DENSITY FUNCTIONAL FOR NUCLEI By Biruk B. Gebremariam In spite of their tremendous success, the limitations of current nuclear energy density functionals (EDFS), all parameterized empirically in the form of the local Skyrme, the nonlocal Gogny or relativistic functionals, have become apparent in the past sev- eral years. In order to address these deficiencies, a current objective of low-energy nuclear theory is to build non-empirical nuclear EDFs from underlying two-, three- and possibly four-nucleon interactions and many-body perturbation theory (MBPT). In this work, the first step towards that goal is taken by calculating the HF contri- bution from the chiral EFT two- and three-nucleon interaction at N 2LO. The density matrix expansion (DME) of Negele and Vautherin is a convenient method to map the highly non-local Hartree—Fock expression into the form of a quasi-local Skyrme— like functional with density dependent couplings. Reformulating the DME in terms of phase space averaging (PSA) techniques, we show that the resulting DME, PSA- DME, is more general and has a significantly better accuracy for spin-unsaturated systems than the original DME of Negele and Vautherin. This is achieved without compromising the accuracy of PSA-DME for spin—saturated ones. Imposing the as- sumption of time-reversal invariance, we apply PSA-DME to the HF energy from the chiral EFT two- and three-nucleon interaction (at NQLO) and calculate the couplings of the emerging EDF analytically using a combination of analytical and symbolic approaches. Subsequently, we perform preliminary analysis of these couplings and show that their density dependence is driven by the long-range (pion—exchange) part of the interaction. Finally, we discuss the UN EDF semi-phenomenological approach that is attempting to utilize the results of this work. DEDICATION In loving memory of my aunt, Sebchiya Gebremariam, whose love and support I still miss. To my wife, Eden T. E103, whose love I cherish. To my dearest friend, Misganaw A. Gashaw, “guadea!” our time-like worldlines should cross soon! iii ACKNOWLEDGMENT I am indebted to my professors, Thomas Duguet and Scott Bogner, who provided me with all sorts of support in my research endeavor. I was lucky to tap into a much wider knowledge base than the typical graduate student. Besides helping me connect the major dots, the fact that they had the patience to let me explore problems in my own way has been a wonderful experience. I am deeply grateful for all of that. I would also like to thank the members of my guidance committee: Mark Dykman, Piotr Piecuch and Brad Sherrill for their helpful suggestions, comments and encouragements. Next, I would like to show my gratitude to my fellow graduate students, office- mates and friends. I should specially mention Jacob Clifford, Morewell Gaseller, Jeremy Armstrong, Ivan Brida, Angelo Signoracci, Liyuan J la and Rhiannon Meharc- hand. Thanks for making me feel at home. Be it two of my three passions, science and soccer, we always had something to discuss. As to satisfying my third passion, literature and painting, Daniel Berhanemeskel, you played the main role there and I owe you a big thank you. The perfect learning and research environment that the cyclotron and physics and astronomy department provided was one of the main ingredients for my suc- cessful completion of this PhD research project. Starting with the highly helpful administrative personnel and scientists to the availability of top-notch computer soft- wares, it is really a magnificent place to do one’s research. In short, thanks! Talking of top-notch softwares, I should specially mention Wolfram Research’s Mathematica. Here is my appreciation for powering my innovation. I also extend my gratitude to all members of the UNEDF collaboration, and the microscopic EDF group in particular, for letting me be a part of this wonderful collaboration. My unique appreciation goes to my dearest wife Eden T. Elos. Her love, support and pragmatism have been and are very important in pulling me out of the often- romantic world that I create around myself. A friend in need is so goes the saying. iv And I say, Mesge, you are indeed one true friend. I thank you for all the walks that we had in life, be it physically through the streets of Addis or telepathically when I am in East Lansing and you back home. Finally, thank you all my family members in Addis and Denver, especially my mom Belaynesh Seifu, Mulumebet Asfaw and Tekleab Hailu, and friends for all your love, good thoughts and understanding. Hey, mom! I have arrived now. u. I; .4 - I ...'.’ I) fl.) 5. r I .‘1 I TABLE OF CONTENTS List of Tables ................................. List of Figures ................................ Low-energy Nuclear Physics 1. 1 Introduction ................................ 1.2 Conventions and Notations ........................ Nuclear Interactions 2.1 Historical highlights ............................ 2.2 Symmetry Properties of Nuclear Interactions .............. 2.3 Remarks on high-Precision Phenomenological Models ......... 2.4 Chiral EFT Models ............................ 2.4.1 NN part at N2LO ......................... 2.4.2 NNN part at N2LO ........................ The E—term ............................ The D-term ............................ The C-term ............................ Low energy constants and parameters of the N NN interaction at NQLO ......................... The Nuclear Many-Body Problem 3.1 Remark on ab-initio/MBPT-based methods .............. 3.2 Goldstone—Brueckner formalism ..................... 3.2.1 Expansion of the ground-state wave-function and energy Hole-line expansion for non-perturbative potentials ...... Perturbative expansion ..................... 3.2.2 Choice of the one-body potential F ............... Phenomenological Energy Density Functionals 4.1 Phenomenological Nuclear Energy Density Functionals ................................ 4.1.1 Motivation from density functional theory ........... 4.1.2 Single- and multi-reference EDF formulations ......... 4.2 Skyrme energy density functionals ................... 4.2.1 Particle-hole functional ..................... 4.2.2 Particle-particle functional .................... 4.2.3 Self-consistent solution ...................... vi 20 22 22 23 26 26 27 27 29 29 5 Caiistru '1 C. .- -.w. .r- "' P‘l i M. l. 4.2.4 Existing pararneterizations .................... 39 4.2.5 Predictive power of empirical EDFs ............... 39 4.2.6 Outlook .............................. 42 5 Constructing Non-Empirical Energy Density Functionals 45 5.1 Constructing Non-Empirical Energy Density Functional ........ 45 5.1.1 Philosophy, Goals and Limitations ............... 46 5.2 The Density Matrix Expansion (DME) ................. 52 5.2.1 Basics of the DME ........................ 53 5.2.2 Existing variants of the DME .................. 55 5.3 PSA-DME ................................. 56 5.3.1 Motivation for a PSA reformulation of the DME ........ 56 5.3.2 Momentum phase-space of finite Fermi systems ........ 58 5.3.3 PSA-DME for the scalar part of the OBDM of time-reversal invariant systems ......................... 62 5.3.4 PSA-DME for the vector part of the OBDM in time-reversal invariant systems ......................... 67 5.3.5 k}, and isospin invariance of the resulting EDF ......... 71 5.3.6 Extension to non-time-reversal invariant systems ........ 72 Constraints on the 7r—functions ................. 76 5.3.7 Remarks on the DME of the local densities ........... 80 5.3.8 Remarks on the DME of the anomalous densities ....... 86 5.4 Accuracy of DME ............................. 91 5.4.1 Inputs to non-self-consistent tests ................ 92 5.4.2 Fock contribution from VC .................... 93 5.4.3 Fock contribution from VT .................... 99 5.4.4 Fock contribution from V1.3 ................... 106 Basic analysis ........................... 106 Further investigation of the spin-orbit exchange ........ 110 5.4.5 Hartree contribution from VC, VLS and VT ........... 115 5.4.6 Preliminary self-consistent tests ................. 118 6 Non-Empirical Energy Density IMnctional from NN interaction 123 6.1 The HF energy from an N N interaction ................. 123 6.1.1 HF contribution from a central interaction ........... 126 6.1.2 HF contribution from the spin-orbit interaction ........ 128 6.1.3 HF contribution from the tensor interaction .......... 129 6.1.4 Additional contributions to the HF energy ........... 131 6.1.5 The leading-order pairing contribution ............. 133 6.2 Application of the DME to the NN-HF energy ............. 133 6.2.1 Analytical couplings from the chiral EFT NN interaction at N2LOI36 6.2.2 Single-particle fields and equations of motion .......... 138 vii 7 Non-Empirical Energy Density Functional from Chiral EFT NNN Interaction at N2LO 140 7.1 The Hartree-Fock energy from Chiral EFT NNN interaction at N2LO 141 Basic form of the HF energy ................... 142 HF energy from the E-term ................... 147 HF energy from the D-term ................... 147 HF energy from the C-term ................... 148 7.2 DME for the HF energy from chiral EFT N2LO 3N F in time-reversal invariant systems ............................. 149 7.2.1 Generic forms of the 3NF energy expressions .......... 150 Generic-Form-l .......................... 151 Generic-Form-2 .......................... 151 Generic-Form-3 .......................... 152 7.2.2 The DME-coordinate system ................... 152 7.2.3 Generalized PSA-DME ...................... 153 Infinite nuclear matter limit ................... 155 7.2.4 The resulting EDF ........................ 155 Comments on the second-order truncation for spherical systems 158 7.3 Analytical Couplings from the chiral EFT NNN interaction at N2LO for time-reversal invariant systems .................... 160 Comparison of analytical and Monte-Carlo results ....... 162 8 Semi-phenomenological EDF, Future Extensions and Conclusions 168 8.1 The semi-phenomenological approach .................. 168 8.2 Key future extensions ........................... 173 8.3 Conclusion ................................. 176 9 Appendix 178 9.1 Mathematical Formulae ......................... 178 9.1.1 Miscellaneous elementary formulae ............... 178 9.1.2 Clebsh—Gordon, Wigner 3—J and 6—J coefficients ........ 180 9.1.3 A few special functions ...................... 182 Legendre polynomials ....................... 182 Laguerre polynomials ....................... 183 Gamma functions ......................... 184 Spherical harmonics ....................... 184 Bessel functions .......................... 185 9.1.4 Three-dimensional spherical harmonic oscillator eigenfunctions 186 9.1.5 Gegenbaur expansion ....................... 187 9.1.6 Functional derivatives ...................... 188 9.2 The one-body density matrix and densities ............... 188 9.2.1 Properties of single particle states ............... 189 9.2.2 One-body density matrix ..................... 190 9.2.3 Local densities .......................... 191 9.3 9.4 9.5 9.6 9.7 9.8 9.2.4 Properties under time reversal .................. 9.2.5 Extension to anomalous contractions .............. 9.2.6 Relations among the densities .................. Local Gauge transformation of the OBDM and local densities 9.3.1 Local Gauge transformation in many-body physics ...... Conventional formulation ..................... Second quantization formulation ................. 9.3.2 Local Gauge transformation of normal densities ........ 9.3.3 Local Gauge transformation of anomalous densities ...... Densities in spherical systems ...................... 9.4.1 Expression for the normal densities in spherical symmetry Scalar part of the density matrix - matter density ....... Kinetic density .......................... The vector part of the density matrix — Spin density ...... Spin-orbit density ......................... 9.4.2 Expression for the anomalous densities in spherical symmetry pairing density .......................... Pairing kinetic density ...................... Pairing spin-orbit density .................... Details on the density matrix expansion ................ 9.5.1 Husimi distribution and the local anisotropy Pflr”) ...... 9.5.2 Wigner transform of the pq(F1, F2) up to ’12 .......... 9.5.3 Generalized PSA-DME ...................... 9.5.4 Generalized PSA-DME for the scalar part of the OBDM Recovering previous DMEs of the scalar part of the OBDM . . Further approximation with respect to 2? ............ 9.5.5 Generalized PSA-DME for the vector part of the OBDM . . . 9.5.6 Remarks on the generalized PSA-DME ............. 9.5.7 The modified-Taylor series expansion .............. Derivation of EDF from HF energy of local NN interaction ...... 9.6.1 Central contribution ....................... 9.6.2 Spin-orbit contribution ...................... Spin-orbit contribution in time-reversal invariant systems . . 9.6.3 Tensor contribution ........................ 9.6.4 Leading-order pairing contribution ............... 9.6.5 The resulting EDF: EDF-NN-DME ............... 9.6.6 Analytical couplings from chiral EFT N N interaction at N2LO HFB equations from EDF-NN-DME ................... 9.7.1 The mean field from EDF-NN-DME .............. 9.7.2 The Pairing field from EDF-NN-DME ............. Numerical solution of EDF-HF equations in spherical systems . 9.8.1 Full-DME in spherical systems ................. 9.8.2 Exchange-only-DME in spherical systems ........... 9.8.3 Harmonic Oscillator basis expansion method ......... ix 192 193 195 197 198 198 198 200 201 202 203 203 204 205 209 210 211 211 211 212 212 214 217 218 223 224 225 228 228 232 232 234 235 236 237 237 240 243 245 247 248 249 IAN 7-1 9.9 9.10 Matrix elements of the kinetic part ............... 252 Matrix elements of the central potential part .......... 253 Matrix elements of the spin-orbit part ............. 253 9.8.4 Self-consistent iterations and convergence ............ 254 The HF energy of chiral EFT NNN interaction at N2LO ....... 254 9.9.1 Remarks on the symbolic implementation ............ 254 9.9.2 HF energy from the E—term ................... 255 Direct part ............................ 255 Single-exchange part ....................... 255 Double-exchange part ...................... 255 E—term contribution for specific systems ............ 256 9.9.3 HF energr from the D-term ................... 256 Direct part ............................ 256 Single-exchange part ....................... 257 Double-exchange part ...................... 258 D-term contribution for specific systems ............ 261 9.9.4 HF energy from the C-term ................... 262 Direct part ............................ 262 Single-exchange part ....................... 263 Double-exchange part ...................... 265 C-term contribution for specific systems ............ 274 Symbolic derivation of EDF-NNN-DME for time-reversal invariance . 275 9.10.1 Generic DME ansatz ....................... 276 Key points on the DME ansatz ................. 278 Connnents on the DME ansatz ................. 280 9.10.2 The G—tensors and their analytical forms ............ 280 9.10.3 Sample DME simplification ................... 284 Angular integrations for spherical systems ........... 285 9.10.4 Contributions to EDF-NNN-DME ................ 288 Fourth order EDF from the D-term ............... 288 Fourth order EDF from the single-exchange piece of the D-like part of the C-term ................... 289 Fourth order EDF from the double-exchange piece of the D-like part of the C—term ................... 289 Fourth order EDF from the Rl-double-exchange piece of the C-term ......................... 290 Fourth order EDF from the R2-double-exchange piece of the C-term ......................... 291 Fourth order EDF from the R3-double-exchange piece of the C-term ......................... 292 Fourth order EDF from the R4-double—exchange piece of the C-term ......................... 292 9.10.5 Extension to deformed time-reversal invariant systems . . . . 293 X 9.11 Analytical couplings for the EDF from chiral EFT NN N interaction at N2LO ................................... 295 9.11.1 Functional form of the couplings ................. 295 Couplings from Generic-Form-l ................. 295 Couplings from Generic-Form-2 ................. 296 Couplings from Generic-Form-3 ................. 296 9.11.2 Matching generalized PSA-DME against the DME-ansatz . . . 297 9.11.3 Application of Gegenbaur’s addition theorem .......... 298 9.11.4 Analytical and symbolic integration ............... 301 Bibliography ..................................................... 304 xi 1.1 1.2 2.1 2.2 4.1 5.1 5.2 5.3 LIST OF TABLES Acronyms used in this work ........................ Definitions and conventions used in this work. ............. Seven Decades of Struggle: The Theory of Nuclear Forces from Ref. [22]. 10 Parameters for chiral EFT NNN interaction at N2LO, with A: = 700 [MeV]. Note that the values for c3 and c4 are from Ref. [61]. . . . INM properties of Skyrme functionals (from Ref. [81]): saturation den- sity pm, bulk compressibility Koo, isoscalar effective mass (m*/m)s, Thomas-Reiche-Kuhn enhancement factor rev and energy per particle at saturation E /A. ............................ MBPT contributions from NN and NNN interactions up to second- order (Normal contractions) in Hugenholtz representation. ...... The first-order anomalous/ pairing diagrams, otherwise called Bogoli- ubov contributions, from the N N and NN N interactions in Hugenholtz represenation. ............................... The Brink-Boeker force(B1) ....................... Full-DME and Exchange-only—DME for Brink-Boeker interaction and several DMEs ............................... xii 21 39 48 48 119 121 1.1 1.2 1.3 2.1 4.1 LIST OF FIGURES Images in this dissertation are presented in color (Color online) A selection of energy/ length scales in physics. ..... (Color online) Low energy static and and dynamical nuclear properties. (Color online) The chart of nuclide and the domains of applications of the standard nuclear structure method. The black region shows the stable nuclei, the green lines show the traditional magic numbers and the red curve delimits the experimentally known nuclei. From Ref. [81]. Hierarchy of nuclear forces from Chiral Perturbation Theory, classified according to a power counting (Q/AX)”, and restricted to V S 3 for simplicity. Three-body forces appears at next-to—next-to—leading order (N2LO), but some of the associated low-energy constants are already constrained by the two-body domain (black symbols) while others (gray symbols) are to be adjusted on three-body observables. Horn ref. [81]. Illustration of the asymptotic freedom of phenomenological EDF mod- els in the case of two-neutron separation energies. In the major shell where empirical EDFs are adjusted on experimental data, the agree- ment between all relativistic and non-relativistic calculations is clearly seen. In the next major shell where no data exist, discrepancies be- tween these models become more apparent (from J. Dobaczewski et al. [150]). ................................. 3 5 17 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 (Color online) Nuclear matter energy per particle as a function of Fermi momentum kip at the Hartree-Fock level (left) and including second- order (middle) and particle-particle-ladder contributions (right), based on evolved N 3LO N N potentials and 3N F fit to E3 H and mm. Theo— retical uncertainties are estimated by the NN (lines) and N NN (band) cutoff variations (from Bogner et. a1. [28]) ................ (Color online) The S-wave solution of the Bethe-Goldsone equation and the uncorrelated S-wave function ................... The quadrupole anisotropy P§‘(R) of the local neutron momentum dis- tribution in a selected set of semi-magic nuclei. The black, red and blue vertical lines indicate the approximate half-radii (where the density be- comes half of the density at the origin) .................. (Color online) k2, k]; and the isoscalar kp extracted from a converged self-consistent calculation of 214Pb, a neutron rich nucleus. ...... (Color online) pn(r) for 48Cr and 208Pb from a converged self-consistent calculation using Sly4 EDF. ....................... (Color online) The parameters for Cr and Pb isotopic chains obtained after fitting the neutron density, pn(Fl/2), with the 7r—functions as given in Eqs. (5.67)-(5.69) ......................... (Color online) Coherence length 6 (R) for 220, 600a, 6”Ni, 104811, 120371, 212Pb (From Pillet et. al. [187]) ...................... (Color online) [,6,,(R,'r)|2 calculated with HFB-D18 for 104311, 12°Sn, 128311. Scale has been multiplied by a factor of 106 (Horn Pillet et. al. [187]). ................................. (Color online) Comparison of 0541?, r) and CffiDMWR, r) where the latter is computed from the 1r—functions of one of the three DMEs: NV-DME, PSA-DME or PSA-DME—II. Upper panels: two-dimensional integrands. Lower panels: ratios of CngME(R,r) over Cfn(R,r) for fixed values of R. Densities are obtained from a self-consistent EDF calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole part and no pairing. ........................... xiv 47 51 61 72 85 85 89 5.10 5.11 5.12 5.13 5.14 5.15 5.16 5.17 (Color online) Percentage error of EgMEhm] compared to Egan], where the former is computed from: NV-DME, PSA-DME or PSA- DME—II II—functions. Densities are obtained from self-consistent EDF calculations using the SLy4 Skyrme EDF in the particle-hole channel and no pairing. .............................. (Color online) The same as Figure 5.10 but for two different values of the range of the Gaussian interaction ................... (Color online) Comparison of T5,,1(R,r) and T3133 (R, r) where the latter is computed from NV-DME, PSA-DME, PSA-DME—II or from PSA-DME with P2"(R) = 0 which we denote as INM—DME. Upper pan- els: two-dimensional integrands. Lower panels: ratios of TfiAfWR, 7‘) over Tap ,1(R,r) for fixed values of R. Densities are obtained from a n converged self-consistent calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole channel and no pairing ............. (Color online) Percentage error of ETQMEhm] compared to ETF [nn] where the former is either computed from: NV-DME, from PSA-DME or PSA-DME-II. Densities are obtained from self-consistent EDF cal- culations using the SLy4 Skyrme EDF in the particle-hole channel and no pairing. Notice the different vertical scale compared to Fig. 5.10. . (Color online) A few representative nuclei with diffuse T,f,‘,J(R, r) to- gether with absolute Eflnn] for the corresponding isotopic chains. Densities are obtained from a self-consistent EDF calculation using the SLy4 Skyrme functional in the particle-hole part and no pairing. . (Color online) Comparison of L3 5,,( R, 'r) and LS 5;? M E (R, r) where the latter is computed from either NV—DME or PSA-DME. Upper panels: two-dimensional integrands. Lower panels: ratios of LS 5:1 E (R, 7*) over Lan(R,r) for fixed values of R. Densities are obtained from a con- verged self-consistent calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole channel and no pairing ................ (Color online) Percentage error of E5313 [7m] compared to Efshm] where the latter is either computed from N V-DME or from PSA-DME. Densities are obtained from self-consistent EDF calculations using the SLy4 Skyrme EDF in the particle-hole channel and no pairing. Notice the different vertical scale compared to Figs. 5.10 and 5.13 ....... (Color online) Ratio of the DME (Eq.(5.101)) over the exact (Eq. (5.58)) expressions of the toy nonlocal matter density .............. XV 97 98 101 103 105 107 109 112 5.18 (Color online) Gram,(R,F) as a function of 'r for a selected set of (R, A, N) .................................. 5.19 (Color online) Percentage error of Eg’DMEMn] with respect to Eg [nn] for Cr isotopic chain. The upper plots show C3,, and CflDME for NV-DME and the parameterized 7r—function which we call PI-DME. 5.20 (Color online) L85", and LngDME for NV—DME, with densities ob- taiend from a converged self-consistent calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole channel and no pairing ..... 5.21 Comparison of Skyrme HFB and DME-based HFB codes ........ 7.1 (Color online) The percentage error of the truncated Gegenbaur expan- sion with respect to Monte-Carlo based calculation of the contribution to E/ A in INM. Upper plots show the actual values for the calculation based on the truncated Gegenbaur expansion. ............. 8.1 (Color online) Cd" and C1“ couplings from chiral EFT NN interaction at N2LO ................................... 8.2 (Color online) C({J couplings from the chiral N N interaction at N2LO with error bands from naturalness requirement. ............ 8.3 (Color online) C1“ couplings from the chiral N N interaction at N2LO with error bands from naturalness requirement. ............ 8.4 The saturation curves, W(p, I) of INM using the phenomenological SLY4 functional and semi-phenomenological DME-based functionals. Here, NZLO includes the contribution from both NN and NNN inter- actions (From Ref. [164]). ........................ 8.5 The same as Fig. 8.4 but for PNM (From Ref. [164]). ......... 9.1 (Color online) R5(k, 11:2, 333,9) for a set of angles. ........... xvi 114 117 118 119 166 170 171 172 174 Chat 10th Chapter 1 Low-energy Nuclear Physics 1 .1 Introduction Problems in physics are characterized by different energy or length scales as depicted in Fig. 1.1. Low-energy nuclear physics lies well below the energy scale for quantum chromodynamics (QCD), AQCD m 106V, and aims at describing nuclear phenomena that occur in the energy scale of a few tens of MeV, as characterized by the typical Fermi energy, 8;. Even though QCD establishes that nucleons, viz, protons and neutrons have a complex structure in terms of quarks and gluons, low-energy nuclear physics never attempts to resolve their structure as justified by SF/AQCD << 1. Its ultimate goal is the proper description of ground- and excited-state properties of nuclei and nuclear matter in terms of the interaction between and among the relevant low-energy degrees of freedom: protons and neutrons. Fig. 1.2 presents the diversity of nuclear properties one is looking after in the realm of low-energy nuclear physics. There are several size-dependent and size-independent factors that complexify the coherent solution of the nuclear-many body problem. For infinite nuclear mat- ter (INM) and finite-nuclei, the existence of the so—called Coester [[1]-[7]] and Tjon 1 Hi ------- . , ----------- 885 Nuclear 3 Atomic 5 Cosmology VEV ' .' ____________ g | QCD Neutrinos . ....... _. : Astronomy : l ' . l Geophysrcs ; 6111‘ Z Nuclear : QHE t“; ;__: i f-mflauon P311011 t‘] [ : "-CD-s-m -:: i : Solar Z 5 to on - ;: 2 ; I i P m . const :: - i system .32 g -' 5 : 1 , 10 : 1 . I 1m 5 1 5 pc 10 ,6 m l E i ll :: g l 1018 GeV 1] lGeV : I1T3v1K l 101“)“ ‘ : :g : Mug; I“ 'J [ll-lime i on i gaming quarks + giht : i Earth. 5 length . leptons ll CMB: '- -- - -_ . Galaxy [ Ry berg ; Classrcal Mechmncs : Sm lhaory Particle Condensed Matter AdS/CFT- RHIC Figure 1.1: (Color online) A selection of energy/ length scales in physics. lines [8] respectively, point to the fact that the nuclear—many body problem cannot be solved successfully without allowing for many-body forces. In addition, in a sys— tem of interacting nucleons, there exist both single—particle and collective excitations, such as sound waves in nuclear matter and rotational/vibrational modes in finite- nuclei, at about the same energy scale. At the same time, most nuclei (i.e. nuclei with masses typically between 40 and 350) are intermediate between few-body and statistical systems. This renders ab—initio techniques impractical due to computa- tional complexity especially for systematic studies which involve hundreds of nuclei. It also prevents the application of statistical approaches due to the smallness of the number of constituents. Furthermore, the need to describe structure and reaction interfaces (fission, fusion, nucleon emission at the drip-line...), the existence of a large isospin asymmetry, and the essential role of superfluidity adds to the complexity of the problem. Due to these factors, a coherent understanding and description of nuclear phe- 2 «l . a“-- L—A [Semen Spectroscopy reflective modes 'Reaction properties RPA, QRPA, GCM Fusion, transfer, elastic fienvy elements Iontic behaviours Fission, Fusion, SHE Drip lines, halos 'Astrophysics r-process, NS, SN Figure 1.2: (Color online) Low energy static and and dynamical nuclear properties. nomena remains elusive, in spite of several decades of theoretical and experimental - investigations. Still, in the last decade, theoretical nuclear physics has seen significant progress from several fronts. Some of the main ones that are relevant to this work are the construction of nuclear interactions within the frame of chiral effective field the— ory (EFT) [9][[10]-[12]], the application of renormalization group techniques to soften two- and many-nucleon interactions [[13]-[15]], the use of ab-initio approaches to cal- culate the properties of increasingly heavier mass nuclei [23], and phenomenological energy density functional (EDF) approaches to computationally intensive calculations thanks to advances in computing power and numerical algorithms. Although high-precision phenomenological two- (NN) and three-nucleon (NNN) interactions have existed for some time [[16]-[21]] and have been successfully used in nuclear structure and reaction calculations [23], they are inconvenient from both theoretical and practical points of view. These interactions lack a controlled expan- sion scheme that would provide a meaningful estimate of theoretical error bars, and there is no clear relation between their NN and NNN parts. Additionally, these phe- nomenological interactions lack a connection to the underlying low-energy QCD. As 3 - ~ 1 F. , -... r . D u '-~ . ,, J . _ l 7' ' ' ‘ . ____. - \ a ’“ , h . - s ,— P‘ V. ‘ Av <" , ~ .. I , a. v u I 1, ' . I - .—. » - . . ....|‘. ‘ a» L . u . . C v- . . ‘ V ‘ , - _.. u H‘ '- 'V I “ DOA“-- I! F‘; fl‘ V w t..-A. _"' r . I s ._ N. ‘. y p ‘\ I ~4 1"4 ’ u ' .r. . A. u" . ,.. ‘ a result, the role of chiral symmetry breaking of QCD which plays a crucial role in determining the long-range part of nuclear interactions is not consistently treated in such potential models [22]. From the viewpoint of nuclear structure calculations, phenomenological interac- tions contain a strong short—range repulsive core, thereby making the nuclear many- body problem highly non-perturbative. In general, the latter statement also holds for chiral EFT interactions which are built with a rather high intrinsic resolution scale, A, as this significantly couples low and high momenta [27]. Historically and in the context of infinite nuclear matter (INM) calculation, this necessitates an infinite re-summation of ladder diagrams, i.e. compute the Brueckner G-matrix, to obtain a meaningful starting point for more advanced calculations based on the hole-line expansion [[74],[30]]. The hole-line expansion method, usually at the lowest order, has been applied to closed-shell medium to heavy nuclei but with little success [38]. In the case of light nuclei with A S 12, state of the art Green’s function Monte- Carlo (GFMC) and no-core shell model (N CSM) calculations can be performed with impressive results [23]. However, their large computational cost makes them inap- plicable for beyond A > 12 region. Most recently, CC (coupled-cluster), IT-NCSM (importance-truncated no—core shell model) and IT-CI (importance-truncated con- figuration interaction) have been used to extend the applicability of ab—initio meth- ods [24, 25]. Eventually, nuclear interactions necessarily depend on the resolution scale [27]. The realization of low-momentum interactions characterized by a low momentum cut—off, A, through renormalization group techniques results in the elimination of the non-perturbative aspects, viz, short-range repulsion and tensor forces of conventional nuclear interactions [27]. The analysis of Weinberg eigenvalues and the calculations of INM equation of state as well as the calculation of a selected set of finite nuclei confirm the perturbativeness of low-momentum interactions. As a matter of fact, INM 4 — I " . H .— ’ v -- - . A _ . . . A.- _.u. .4 , f‘. .. n1 ' e n- -‘ : "‘ V ' ." a . v .1 ‘ J .I n q- v< v A: “A sh . “v .5‘ IA-‘ . -. ' a... In o , ¢ \ . - . A‘- .- I . u. n - T‘ ._ . O.- . C s‘.’ ‘ k I. _ -. w. .', N. I A». ..'. u, 'U I 6‘ shows saturation already at the HF level, while the empirical saturation properties are reproduced satisfactorily at second-order in MBPT [28]. For finite nuclei, the energies and radii of a select set of nuclei seem to be remarkably converged at second order with good systematics and relatively small corrections coming from particle-hole states in the RPA [29]. Still, the application of these ab-initio methods for medium to heavy mass nuclei involves considerable numerical complexity. In addition, the 2 accuracy of these methods is not on par with the current tool of choice for calculating ground- and excited-state properties of medium to heavy mass nuclei, namely, energy density functional (EDF) methods [26]. Fig. 1.3 shows the domains of application of the stande nuclear structure methods. ~120 L100 -60N I ' I 0 ' so 100 N150 ' 260 l 2750 Figure 1.3: (Color online) The chart of nuclide and the domains of applications of the standard nuclear structure method. The black region shows the stable nuclei, the green lines show the traditional magic numbers and the red curve delimits the experimentally known nuclei. Ffom Ref. [81]. Currently, EDFs are completely phenomenological by construction. Modern pa— 5 rameterizations of these empirical EDFs such as the Skyrme, Gogny and their rela- tivistic counterparts provide a fair description of bulk properties and certain spectro- scopic features of known nuclei [26]. However, such empirical EDFs lack predictive power away from the valley of stability or known data. In addition, the objective of having spectroscopic quality EDFs does not seem to be attainable with current energy functionals [26]. Consequently, an intense ongoing effort is dedicated to empirically fitting EDFs possessing more complex analytical forms and/or enriched density de- pendent couplings [[31]-[36]]. Along with such phenomenological approach, the quest for predictive EDFs can be complemented by constraining the analytical form of the functional and the value of the couplings from MBPT and the underlying low-momentum two— and three- nucleon (N N and NNN) interactions. The present work is a step towards that goal. In CHAP. 1, we present a brief discussion of nuclear interaction models with special emphasis on chiral EFT. CHAP. 2 introduces the nuclear many-body problem and the diagrammatic approaches that rely on summing a selected set of diagrams. This is followed by CHAP. 3 where we deal with the formalism and performance of phe- nomenological EDFs. CHAP. 4 lays out the philosophy, goals and limitations of our approach for constructing a non-empirical EDF. After introducing the density ma- trix expansion (DME) as the mathematical technique to make an explicit connection between MBPT and quasi-local EDFs, we describe a new formulation of the DME based on phase space averaging (PSA). In addition, non-self-consistent and prelim- inary self—consistent performance tests of this newly formulated DME are given. In the subsequent chapter, CHAP. 5, we give details of the derivation of non-empirical EDF from a generic N N interaction, at the lowest order in MBPT (Hartree-Fock) and the application of the result to the chiral EFT NN interaction at N2LO. CHAP. 6 dis- cusses the contribution to the non-empirical EDF from chiral EFT N NN interactions at N 2LO at the HF level. In addition, an on-going effort to build a universal energy 6 v‘ . cu : [v \' . . '» - ' i. (- e a. - ‘ 9 .’ - . .A i 5.. ”D— ‘I 1.. “Jan. T’r‘ P — LU“ ‘u—y. — a, l O .l i I . \'\~‘. .l \t ' . g. .. r.T M‘l A ya cr- A‘J .. i.. u, - lv-- . .1". .U‘ Ly- 4;“. density functional (UN EDF) that incorporates the results of this work, as well as pos- sible extensions and conclusions are discussed in the last chapter, CHAP. 7. Finally, all relevant definitions, formulae and derivations, both analytical and symbolic, are presented in a set of detailed appendices. 1.2 Conventions and Notations The acronyms, notations and definitions used throughout the thesis are listed below. Table 1.1: Acronyms used in this work. OBDM One-body density matrix EDF Energy density functional DFT Density functional theory NN interaction NNN interaction two-nucleon interaction three-nucleon interaction HF Hartree-Fock HFB Hartree-Fock—Bogoliubov INM Symmetric and unpolarized infinite nuclear matter PN M Unpolarized pure neutron matter EFT Effective field theory RG Renormalization group DME Density matrix expansion MBPT Many-body perturbation theory RPA Random phase approximation .. 'TQ I I Table 1.2: Definitions and conventions used in this work. Qt \ ‘H (19; P9 1.7 PT. U ST V: .. j‘ ~ ‘2. af/s/ /p/9 [IR/57W; [ST NS" (11 [772' [ST p/é' a2 [7r - [ST p/s“ a3 [7% p/5 i p/é' i r 7rp/ 1' my 1(a) Mi?) Denotes cross product Pauli vectors: (0:, 0y. 0;), (TI, Ty, 7'2) Unit vector along vector A, or operator A in case A is an operator The differential solid angle with respect to A th Exchanges the spin coordinates of the ih and ] particles. It IS given by P"-— - 1/2(1 + 51:52) Exchanges the iso-spin coordinates of the i” and 3"” particles. It IS given byPT . —— -1/2(1 + T1 T2) Exchanges the spatial coordinates of the it“ and 3"" particles The particle exchange operator given by Pij = PIT}. Pg P; The spin singlet (i = 0) and triplet (2'. = 1) projectors. These are given by flag/1 = 1/2 (1 IF 61 ~62) The isospin singlet (i = 0) and triplet (i = 1) projectors. These are given by 1170/1 = 1/2 (1 q: f] - 73) NN interaction vertex of type I where I can be C-central LS-spin orbit or T-tensor and ST can take the values 10,01, 11,00 where the first 1/0 refers to spin and the second 1/0 refers to isospin triplet / singlet 2"" 7r—function associated with local densities such as 10(6) NU?) / ill—ib- i‘h 7r—function associated with non-local densities 100:1, F2)/§(F13F2)/15(F17 772) /'§(F19 F2) alSTlvrf/grf’gké) 2 4w f dr VITStr) inf/gum arses/57.5mm) s — M We) ivrf/S‘vrf/ii alST[7rp/37r 0/8“ fd +14st f“ (”‘1 Chapter 2 Nuclear Interactions 2.1 Historical highlights The theory of nuclear forces started in the 19303 when Yukawa introduced the idea that the nuclear strong force is carried by a particle with a mass approximately 200 times that of an electron [37]. Table 2.1 summarizes the major developments of the past seven decades in the attempt to derive NN interactions from first principles. With the conception of effective field theory (EFT)[9], it has become clear that pion-based theories of the fifties, this time with an explicit connection with low- energy quantum chromodynamics (QCD) [[10]-[12]], should be revived. In the last decade, EFT has been applied successfully to the consistent derivation of NN, NNN and many-nucleon interactions at various orders in the low-momentum expansion scale, Q/AX, where Q is the energy scale of the low-energy physics and AX ~ 1 GeV refers to the chiral symmetry breaking scale. Details relevant to the present work are given in section 2.4. In parallel with these efforts to derive nucleon-nucleon and many—nucleon interactions starting from field-theoretic approaches, various high- precision phenomenological N N and N NN interactions have been parameterized [[16]- [21]]. These efforts have been guided by requiring the interactions to satisfy a number 9 am Table 2.1: Seven Decades of Struggle: The Theory of Nuclear Forces from Ref. [22]. 1935 Yukawa: Meson Theory 1950’s The “Pion Theories” One-Pion Exchange: o.k. Multi—Pion Exchange: disaster 1960’s Many pions E multi—pion resonances: 0', p, w, The One-Boson—Exchange Model: success 1970’s Refined meson exchange models, including sophisticated 271' exchange contributions (Stony Brook, Paris, Bonn) 1980’s Nuclear physicists discover QCD Quark Cluster Models 1990’s and beyond Nuclear physicists discover EFT Weinberg, van Kolck Back to Pion Theory! But, constrained by Chiral Symmetry Breaking: success of symmetries. In the following, we discuss the symmetries that are used to constrain the form of NN interactions. 2.2 Symmetry Properties of Nuclear Interactions While the derivation of the strong NN and many-nucleon interactions is an ongoing effort, there are a number of symmetries that a given nucleon-nucleon interaction should satisfy. Since one can denote the most general nucleon-nucleon interaction by its matrix element between two—body states, we use v(1,2) E( -oI I I?! I 7101017202 Q2, ifii'F10141F202Q2) E v(F1k101q1, F2E202QQ), (2.1) 10 to discuss the action of the various symmetries. In Eq.(2.1), the dependence on the momentum of the interacting particles is to allow for nonlocality of the interaction. The following are the basic symmetry properties that a given NN interaction needs to satisfy [38]. Hermiticity. Invariance under an exchange of coordinates v(1,2) = v(2,1), (2.2) e Translational invariance v(1,2) 2 11(7", 1:1 alql, [:2 02(12), (2.3) o Galilean invariance v(1.2) = WE. alql, am). (2.4) Invariance under space reflection ’U(FE, 0'1ql, 02(12) = ’U(—F — E, alql, 02q2), (2.5) Time reversal invariance '00—’13, 01611, am) = 21(7“ — 13, -—01q1, —02Q2). (2.6) Rotational invariance in coordinate space implies that the interaction is a scalar. Additionally, ”(V-‘71:, 01111, 0292) = ”(‘77 — E, 02022. 01(12): (2-7) 11 which is due to Eqs. (2.2) and (2.5). Hence, terms in the interaction which are linear in a,- and q,- depend only on 0 = (01 + 02)/2 and q = ((11 + (bl/2. o Rotational invariance in isospin space which is an approximate symmetry broken by the coulomb interaction and other isospin—breaking effects. If assumed to hold, then ”(F/:3: 0141» 02(12) = 1’0(FE102101) + ”01(ng 02: 01)7'1 '72. (2-8) Even after correcting for electromagnetic effects, there is a strong experimental evi- dence that the nucleon-nucleon interaction breaks charge symmetry [39] and charge independence [[40], [41]]. The experimental evidence comes from the difference in the scattering lengths of pp, nn and pn systems. These values read app 2 —17.3 :1: 0.4fm, am = —18.8 :1: 0.5fm and em = —23.74 :1: 0.02fm. In general, nucleon-nucleon in- teractions can be classified into four classes according allowed isospin operators [42], i.e. 0 Class I forces have only dependencies on [11, (T1 - r2)], and do not break either charge symmetry or independence, 0 Class II forces maintain charge symmetry but are charge-independence-breaking (CIB). They are characterized by the isotensor T12 defined by analogy to the usual tensor 5'12 given in Eq. (2.10), and vanish for T2 = i1 (mm or pp) systems, 0 Class III forces are both charge-symmetry-breaking (CSB) and CIB, but remain invariant under the exchange of the two nucleons, and are thus proportional to (r21 + r22). They do not cause isospin mixing since T2 commutes with T2, and vanish for T 2 = 0 (up) systems, 0 Class IV forces are both CSB and CIB, and are antisymmetric under the ex- 12 1. . . ~i . .— "1 n . 1P . 1‘ .... .-. . ’- 0'; b b‘ , .. . ... a' ' .- . .‘ s. _ . v.-, ‘ . ‘ ‘ e .. ,. . a1 .. 3"; g k - “ p.‘ ‘5 . . . .‘1 1.. . change of the two nucleons, which causes isospin mixing. They are proportional to (r21 — 732) or (r1 (8) r2)z , and vanish for T2 = i1 systems. The most general class-I two-body potential invariant under the fundamental sym- metries recalled above can be decoupled into [43] f 1 ) 1 (01 '02) 211.2) = 2 «2pm 51” p (E - 3') 6212 K 312.1? A (7'1° T2) where the various operators are the so—called central ll, tensor 512,7, spin-orbit (ES), quadratic spin-orbit Q12, and 512,5 components. The operators 812,50, 312.123 and Q12 are given by 3 512,; = fi(0’1’f’)(0’2'fl—0’1'0’2, (2.9) 3 _. s 812,; = 13(0'1 ' k)(0’2 - k) — 01' 02, (2.10) 1 -. .. .. .. Q12 = §[(01 ~L)(02 - L) + (oz-L)(01-L)], (2.11) where all operators in (2.2) have radial prefactors, 2.7,,(r), that can be constrained from microscopy or experimental data. 13 -- ‘ l . ~ ‘~ -. 1 iv 0- 'u' \ 1‘ v v- . r . s..- o... «v 06 i ...‘....I. .. .. n . J . ...L. n. n 0* . r - n f . “.- .. .. o. v . e- - u u. I.' ' ‘ I i r ,. ' ' o- . . n .W ’ - e M_ 'F‘l , . I. ‘ I1. s... Ast' i; ‘ 4 "5 1 .. “Me n '- 3 s.’ .1 -- ‘ .. '9. . . .1 ~ ‘4 '\ 7. r. 1.. 2.3 Remarks on high-Precision Phenomenological Models The construction of phenomenological models for nucleon-nucleon interactions pro- ceeds by parameterizing the radial prefactors vp(r). It is well known that the long- range (r > l/mfl) part of the interaction is given by one-pion exchange, thereby e—mwr 7‘ fixing the radial form factor to the usual Yukawa form, . The phenomenolog- ical models that have been parameterized in the last two decades [[16]-[21]] are said to be high-precision as they are able to fit low-energy (S 350MeV) nucleon-nucleon scattering data with a chi square per degree of freedom, x2/Ndata, close to one. Ad- ditionally, all currently available high-precision phenomenological models are charge dependent (CIB and CSB) and use about 40-50 parameters. The main difference among the various phenomenological models lies in the way they attempt to capture the intermediate- and short-range parts of the interaction. The need to include many-body forces has been suggested by discrepancies be- tween low-energy properties computed with two-body forces only and experimental data, such as differential nucleon-deuteron cross-sections [[44]—[46]], triton and other light nuclei binding energies [47], and the violation of the Koltun sum rule [48]. For instance, the binding energies of 3H versus 4H6 computed with all available NN models align on a so-called Tjon line that excludes the experimental point [8]. This is seen as a necessity to use consistent N N N forces to sneak away from this Tjon line. Likewise, the Coester line on which lies the saturation point of INM computed with NN forces only [[1]-[7]], is another indication that N NN forces are essential to reproduce bulk properties of nuclear matter[[49], [50]]. Phenomenological N NN potentials are available [[51]—[54]], based on mesons ex- changes plus empirical short-range components. Using the same philosophy as phe- nomenological NN forces, they are adjusted on binding energies and scattering observ- 14 ables of three— (and four—) body systems such as proton / nucleon-deuteron diffusion data [[55]-[57]]. In the following section, it will be seen that chiral EFT, N NN inter- actions appear naturally which is one of the main advantages of the EFT approach. 2.4 Chiral EFT Models Potentials based on chiral EFT [9] exploit the separation of scales between the chiral symmetry-breaking scale, Ax z 1 GeV, and typical momenta of low-energy processes at play in the nuclear structure context, Q, usually about m7r z 140 MeV [[10]-[12]]. In that respect, few-nucleon processes can be treated using only nucleons and pions as degrees of freedom, the 7r—N interaction being governed by the spontaneously broken chiral symmetry of QCD. All other heavy mesons and nucleon resonances are integrated out of the theory, and their effects are contained inside scale-dependent couplings. The effective Lagrangian only depends, in this approximation, on a finite number of low-energy constants (LECs), and can be classified using a systematic expansion based on a power counting in terms of (Q/Ax)”, where V is called the chiral order. At a given accuracy (Q/AX)”, only a finite number of terms in the Lagrangian are needed in the low-momentum regime. The leading order interaction corresponding to V = 0 is denoted by LO. There is no contribution for V = 1, and following terms V > 1 are called (next-to-)"‘l leading- orders (N"“1LO). This framework includes effects beyond the NN force, since three, four-... body interactions appear naturally in the perturbative expansion, and the hierarchy vNN >> vNNN >> UNNNN is a direct consequence of the power counting, as shown in Fig. 2.1. At this point, chiral interactions exist up to N3LO [[58],]59]], where most of the NN and one-pion, two-pion and three-pion (OPE/2PE/3PE) diagrams have been computed using various approaches [58, 59]. Improvements of such approaches may 15 consist in (i) increasing the chiral order V of the perturbative expansion, although [power counting implies that higher contributions will be substantially smaller, as al- ready observed in the case of OPE/2PE [62], (ii) the introduction of four-nucleon forces arising naturally at N3LO [63], (iii) treating extra degrees of freedom explic- itly, such as nucleon A excitations that play a role in three—body forces [[66],]68]] and isospin breaking NN forces [68], or (iv) refining the short range phenomenological cutoff schemes. Finally, since chiral perturbation theory is a low-momentum expan- sion, its predictions are by essence only valid for momenta Q << AX. Several families of chiral forces are defined depending on the values of the intrinsic high-momentum cutoff up to which they are defined, whose values typically range between 450 and 750 MeVs. This makes chiral potentials significantly softer than phenomenological hard-core interactions. In general, chiral EFT potentials have the general structure VEFT : Vnn + VLAA) a (2.12) where Vmr are due to n pion-exchanges and Va(A) refers to the contact parts which depend on the high-momentum cutoff scale, A. In chapter 6 and 7, we calculate the HF energy from chiral EFT N N and NNN interactions at N2LO, with emphasis on the contribution from the finite-range parts of the interaction VM. Hence, we now describe the chiral EFT interaction at N2LO in some detail. 2.4.1 NN part at N2LO At N2LO in the low-momentum expansion Q, the pion—exchange (finite-range) part of the NN interaction can be written as VIN : V15?) + V1?) + [4(3) 11’ 9 16 ’0 h‘.’ NN diagrams NNN diagrams LO (V = 0) -- (V = 1) NLO (V = 2) ::: 2:: [Cl -- -- [D] - NQLO (u = 3) :2 [El Nucleon line . £10) ,I'A = 0 term (no field derivatives) ------- 1r fine I £11)le = 1 term (one field derivatives) 9 £12) A = 2 term (two field derivatives) Figure 2.1: Hierarchy of nuclear forces from Chiral Perturbation Theory, classified according to a power counting (Q/AX)”, and restricted to V S 3 for simplicity. Three-body forces appears at next-to—next-to-leading order (NZLO), but some of the associated low-energy constants are already constrained by the two-body domain (black symbols) while others (gray symbols) are to be adjusted on three—body observables. From ref. [81]. 17 us_' 'i t ‘ .1 .. v: [ ‘H_ 1' -l.' _ 1L. 1 , . A. 'f V’ ,A-a ‘ \ r‘v’ A . .. . \\ u. 7 It ' . . ... . c -I . '. 3] 4|, 5 v... = va‘i’wx’. (2.13) Here the superscripts denote the corresponding chiral order and the ellipses refer to the Q4-and higher order terms which are not considered in the present work. As can be seen, contributions due to the exchange of three-and more pions are further suppressed. In [13) ® [(7') ® [77) space, the finite-range (pion—exchange) part of the chiral N N interaction through N2LO takes the form1 .0 (E1, “VII-51132) = ([VC(CI) + Ti 742 VVCWH + [ VS(Q) + 7:1 ~r21Vs(q)]61 '52 where (f = If" - it: is the momentum transfer, with the relative momenta being it: = E1 — E2 and 13’ = If; — E251? = (E1 + 1:2)/2 and K' = (131’ + lip/2 are center of mass momenta of incoming and outgoing interacting particles respectively. The requirement of Galilean invariance is enforced by 6U? — If ’). In passing, we remark that the the contact part of the interaction contains terms that depend on 15' = (13’ +13) / 2 and/ or (f. The subscripts C, S, T, LS label the form factors of central, spin- spin, tensor and spin-orbit components of the interaction. The form factors are scalar functions of the momentum transfer q and are such that (i) only WT gets contribution from one pion-exchange (ii) VC, ,WC, , VT, WT, VLS, WLS get contribution from two- pion exchange. Actual expressions and details on the contact parts of the interaction are given in Ref. [59]. lThe finite-range NN spin-orbit piece is actually zero-range up to N'ZLO. 18 2.4.2 NNN part at N2LO From a general standpoint, three-body forces can be. characterized by _._._. - _._._. 1 - f _. .. (klkzkalvzwlki ikzi > _ Q—F 1+E2+k3 —E 17’ ~. :31V'(k1k72k3lki"-2'”;i)1 (2.15) where Q is the volume used in the box—normalization of the momentum basis states, 61-51 +52 +133: Ef—Eé—Eé is the Kronecker delta and 170-51 62%| “1’ #2’ .3) is a matrix element in momentum space and an operator in spin-isospin space whose dependence on spin and isospin degrees of freedom is not displayed. The NNN x—EF T interaction first appears at N2LO where it is composed of three components [69] (i) the E—term (ii) the D-term and (iii) the C-term. The E—term The E-term, which is a three—nucleon contact interaction, is the simplest part of the x—EFT 3NF at N2LO. Its expression reads VE(IE'1E2E3IIZ;E[)= Emma-mas), (2.16) where E E %. (2.17) The D-term The D-term involves one-pion exchange plus contact interaction. Its analytical reads . ~~~ ~-~ CD 014720252 02-43036"; VDkkk It, [kl E g—A AX( T-T +———‘T.T. (”3'1“) 4—f.f£A (13+m?r 1 2 Q§+mi 2 3 03'5101'51 qf+m§r 19 where q,- = ki— —k,~’. The C-term The C-term of the interaction involves two-pion exchange. Its analytic form reads 2 -. _. . -*-'-' ~-'-* . 01°4102'42 ma 3 V, k '- k't'” E q—A F... 0‘ C( 1 2 3! l 2 3) (2f7r (q¥+nlg)(q§+m72r) 1237-172 02 6203 {if} 05 a [3 F 7' r ((12 +mv2rl((13 +m12r) 231 2 3 03 4301 C11 08 Ta 7’3 F. , 2.19 ((13 +mi)(q1 +mi) 3127.37) ( ) with 01mg, F013 : C113 [._4 f3 ijk — c, _, _, a _, + 2f—Zqi-qj] + %6 (3777 0;, (q,- x qj). (2.20) Low energy constants and parameters of the NNN interaction at N2LO Values of the various coupling constants appearing in Eqs.(2.16)-(2.20) can be found in Table 2.2. There are several ways to extract fix the CD and CE low-energy constants, one of which is adjusting these constants such that the binding energies of 3H and “He from ab—initio calculations with NN and N NN interactions match experimental values. The above statements also hold for c1. On the other hand, there is still some controversy over which set of values is “right” for c3 and c4 with extractions from 7r—N scattering and NN begin different with large uncertainties. Resolving these differences is important as many quantities are sensitive to the values of c3 and c4 [60]. 20 Table 2.2: Parameters for chiral EFT NNN interaction at N2LO, with AI = 700 [MeV]. Note that the values for c3 and c4 are from Ref. [61]. gA 1.29 f,r [MeV] 92.400 m7r [MeV] 138.040 C] [ GeV—1] -0.760 c3 [ GeV-1] -4.780 04 [ GeV‘l] 3.960 CD -2.062 E -0.625 21 Chapter 3 The Nuclear Many-Body Problem 3.1 Remark on ab-initio/MBPT-based methods Ab—initio methods for the nuclear many-body problem such as no-core shell model (NCSM) solve the A—body problem in a given model space while quantum Monte— Carlo methods such as Green’s function Monte-Carlo (GFMC) rely on stochastic integration of the many-body Schrodinger equation [23]. Currently, they are able to incorporate both NN and N NN interactions. However such methods show exponen- tial scaling with A, thus limiting their applicability to only A S 12 [23] due to their computation costs. In this regard, CC (coupled-cluster), IT-NCSM (importance- truncated no—core shell model) and IT-CI (importance-truncated configuration inter- action) should be mentioned as ab-initio methods that solve the A-body problem approximately in the given model space. They have lower computational complexity and thus extend the applicability of ab-initio methods to heavier nuclei [24, 25]. In contrast, MBPT-based methods rely on partial finite/infinite—order summation of MBPT diagrams according to some organizing principle. Infinite—order summation may be necessitated by the non-perturbativeness of the starting interaction, and may not be necessary if one starts from perturbative low-momentum interactions [27]. The 22 non-perturbative behavior of conventional phenomenological interaction models can be traced to [27] i.e. (i) the hard-core repulsion that makes nucleons scatter up to very high energies and requires large basis sets, (ii) the tensor force coming from OPE which is singular at short distances, and (iii) the presence or virtual (di—neutron) or bound (deuteron) states. On the other hand, vacuum nuclear interactions are strongly renormalized in the nuclear medium. This suggests that expressing the many-body energy in terms of an unperturbed Slater determinant coupled to an effective in-medium interac- tion that already includes many-body correlations might be possible. That is, the minimal set of in-medium correlations that have to be included to reach a reasonable description of the system, i.e. infinite nuclear matter or finite nuclei, need to be in- corporated in the definition of the in-medium interaction. This can be achieved for simple systems in the context of Goldstone—Brueckner theory [70]. 3.2 Goldstone—Brueckner formalism As long as pairing is not explicitly included, the Hamiltonian H = t+ U can be decomposed in terms of a one-body hamiltonian ho that has Slater determinants [,~) as eigenstates, and a perturbation hl, i.e. H = h0+h1, (3.1) ha 2 t+r=Zt,+Zr,,-ala,=Zené’ga, (3.2) i ij n h] = ’U—F. (3.3) The quantities 6,, are the eigenenergies of ho corresponding to single-particle states app, whereas 5,- will denote many-body eigenenergies of ho associated with unperturbed 23 Slater determinants, i.e. A A 5i = Z ftp [(1%) = Héj’pIO) . (3.4) p=1 P=1 According to Gell—Mann-Low’s adiabatic theorem [71], the true ground state [60) of H can be obtained from the adiabatic evolution of the ground state of ho from t 2 —00 to t = 0 by gradually turning on the residual interaction [72], i.e. |90)=1im._>0( U.(0,—oo)|<1>0) ) <o|U.(0, —oo)|0) (36) where the adiabatic evolution operator U€(t, to) from t to to is defined in the interaction representation starting from the Hamiltonian in the Shrodinger representation H (e, t— to) as .. 'h t U(t,t0) E exp 2’2“] U€(t,t0) exp [—z,_:) ] (3.6) [- . t U€(t,t0) _=_ exp —1/ TH(€,7'). . (3.7) . h t0 ‘ From an expansion of U6 in powers of the residual interaction and integrations over time in Eq. (3.5), a series expansion of the ground state [80) is obtained [73], i.e. 50— lo leo>=;( 1, h.)"|<1>o>u...., (3.8) where the sum runs only over linked diagrams, i.e. where [@0) does not appear as an intermediate state. The latter is enforced at the level of (3.5) where the denomi- nator fixes the normalization of [90) by eliminating disconnected vacuum-to—vacuum 24 diagrams [79]. Likewise, a similar expansion of the ground-state energy E0 reads 1 11 E0 = 50 + Z<¢0lh1 (c h hl) [(1)0>connectede (39) __ ,0 \— n 0 where the sums now only runs over connected diagrams. However, if the expansions of Eqs. (3.8) and (3.9) are truncated at a given order, non-converging results arise if the vacuum interaction contains a non—perturbative hard core. On the other hand, it is possible to extract a series of ladder diagrams where a succession of interactions '0 scatters nucleons into particle states. This series can be replaced by a reaction matrix G which resums those Brueckner’s particle-particle ladders and can be represented by the self-consistent Bethe-Goldstone equation [[75]- [78]] C(w) = v + v C(w), (3.10) w - ho where w is the starting energy that corresponds to the in—medium energy of the nucleons at the location where G is inserted, whereas the Pauli operator Q excludes occupied states, i.e. those below the Fermi level 61? associated with the unperturbed vacuum |0), that is Q: 2 lpp’)(pp’l- (3-11) €p,€p/>CF The replacement of the initial interaction by the re-summed G-matrix modifies the short-range part of the in-medium two-body wave function, such that it is strongly suppressed over a distance of the order of the range of the repulsive core, that is the healing distance, or wound [70]. .- Ill .1 . , an a . “ .’ \ ' “ ... n J I .‘, .— 1 'I -. . . m, r‘ ' . .. A . . ,, I a-.. Y»- ' i. II a“ ' w ‘9' u. .. l-. l: ._r-' \ d.‘ r17 3.2.1 Expansion of the ground-state wave-function and en- er gy The general idea consists in regrouping, if necessary, clusters of diagrams under G in such a way that a converging series is obtained, i.e. a truncation at a given order provides a result of a given precision [74]. Once the G—matrix has been computed, it replaces all instances of v in diagrams, excluding those where successive G-matrices are connected by a two-particle intermediate state, that is no particle-particle ladder connecting two G-matrices must be written. Hole-line expansion for non-perturbative potentials While the G-matrix regularizes the hard-core repulsion, an expansion in terms of G for the ground-state and singleparticle energies remains non-perturbative, in such a way that the proper expansion parameter is the number of hole lines [80]. At lowest order in the hole-line expansion, the ground-state energy E is given by the Bruckener-Hartree—Fock (BHF) approximation. The BHF approximation consists of a self-consistent solution of the equations ,, .. .. .. 1 -- 1 E0~(z][G(w)[2]) — (zylvlzj)+§ng>CFw_€m_€n+z-n x (mnIG(w)|z'j), (3-12) where e,- are the on-shell single-particle energies that are obtained by a functional derivative of the ground-state energy and the two-body matrix elements of G (w) and v are anti-symmetrized. Thus, the lowest order in hole lines for E0 (two hole lines) leads not only to a term with one line in the self energy but also to a rearrangement term containing two hole lines and coming from the functional derivative of the particle- particle ladder propagator. 26 Pmurl | 'Yjw . IL’. »a ‘-i Perturbative expansion If the starting interaction is in fact perturbative, as it will be the case for low- momentum interactions, in—medium correlations can be treated through converging perturbative series in powers of v for E0 and 61‘. Indeed, the ladder series from (3.10) becomes perturbative, such that it can be truncated at a given order in intermediate ladders. For instance, the ladder series for vlow—k is almost converged at second order in MBPT [81]. For the relevant diagrams that appear at second order from NN and N NN interactions, refer to Table 5.1 in section 5.1.1. 3.2.2 Choice of the one-body potential P The proper choice of the unperturbed hamiltonian ho is crucial to have a rapidly convergent series [82]. Several choices for the one-body field F are possible, among which (i) a phenomenological expression that is fixed a priori, (ii) the Hartree-Fock approximation where 6,, are eigenenergies of the Schrodinger equation associated with the vacuum force, or (iii) a more involved approach necessary for non-perturbative potentials, e.g. where the one-body field 1" is constructed at lowest order in the on— shell G-matrix or includes rearrangement terms (extended Brueckner-Hartree-Fock calculations) [[83]- [86]] . Note that the truncation orders can be different in the series for the energy E0 and the self-energy 6,, e.g. E0 can be computed at second order while single-particle energies are derived from a more simple (Woods-Saxon...) potential or only at first order in 1). Still, adding more orders in the expansion of the single-particle energies adds extra diagrams in the series for E0 such that it converges faster. Finally we remark that description of pairing within a diagrammatic framework is possible by defining anomalous propagators and allowing for anomalous contractions in addition 27 to the normal contractions. Refer to [87] for details. 28 Che Ph Fm 4.1 Chapter 4 Phenomenological Energy Density Functionals 4.1 Phenomenological Nuclear Energy Density Functionals The nuclear energy density functional (EDF) approach, due to its computational tractability, is the many—body method of choice to study medium- and heavy-mass nuclei in a systematic manner [26]. The central element of EDF approach is the energy density functional. Currently available realizations of the EDF approach, all empirically constructed, vary in the way they parameterize this energy density func- tional [26]. These include the quasi-local Skyrme, the nonlocal Gogny and relativistic models. 4.1.1 Motivation from density functional theory Historically, nuclear EDF based approaches were motivated by starting from effec- tive interactions in the particle-hole and particle-particle channels and solving the 29 self-consistent mean-field equations [26]. Recently, the focus has shifted towards considering the energy density functional approach as motivated from effective field theory where the various densities of the system are the basic low—energy degrees of freedom [35]. In parallel, the development of density functional theory (DF T) [[88]-[91]] in quan- tum chemistry and condensed matter physics seems to have given nuclear energy den- sity functional approaches a starting theoretical basis. DFT has been applied success- fully to the structure of quantum many-body electronic systems (atoms, molecules, solids...). The comparatively small computational cost of the approach makes DFT the only feasible solution for systems with large number of electrons [92]. Instead of the many-body wave-function, DFT takes the fermion density as the “fundamental” variable. The two building blocks of DFT are 0 The Hohenberg-Kohn theorem [95], which states the existence of a functional F [p] such that the ground-state energy of a system of N particles in a one-body external potential u(r) can be written as Em = F[p1+ / drumpm. (4.1) where F [p] only depends on the Hamiltonian of the interacting system, thus is independent of the external potential u(f'). The ground- state density p00") and energy E0 = Eu[p0] are then obtained by minimizing Eu[p] with respect to a variation of the density p0") under the constraints that p is positive and f df'p(f) = N. It should be noted that this existence theorem does not imply that all the information about the ground state is contained in the electron density p('F) [93]. 0 Due to its practical difficulties, DF T is not implemented as a pure functional 3O . P. T lv 1 E". M of the density, a la Thomas-Fermi theory [94]. Rather, one makes use of the Kohn-Sham implementation [97], which asserts that for any interacting system, there exists a unique local single-particle potential uK3(7"') such that the ground- state density of the interacting system equals the ground-state density of the auxiliary non-interacting system in the external potential U.K3(T_'), that is 2 , (4.2) pm = pram = 2 am expressed using the lowest N single-particle orbitals, $.01"), which are solutions of the one—body Kohn-Sham equation v2 [7— + mm] am = mm. (4.3) m where e,- are the Kohn—Sham eigenvalues. In the Kohn—Sham scheme, F is split into PM = Tlpl + U [pl + 5.1-rip]. (44) where (i) T [p] is the universal (kinetic) energy functional of the non-interacting sys- tem, (ii) U [0] is the Hartree functional depending on the two-body interaction po- tential V(|7"',~ — Fjl), and (iii) Exc[p] is the so-called exchange-correlation functional, including the Fock term and all remaining many-body correlations. When Eu [p] is ne- glected, the Kohn—Sham equations reduce to the standard self-consistent Hartree ones. Additionally, the Kohn-Sham potential is given through the condition that ground- state energies of the interacting and non-interacting problem (U [p] = E51:(:[p] = 0) are met for the same density p(F), i.e. 6 Ere [pl 670'? (4.5) ukslfl 5 “(Fl + 31 While the Kohn—Sham potential is local / multiplicative, the exchange-correlation func- tional might be highly non-local. The main difficulty for DF T practitioners lies in the fact that no prescription is given to construct F [p], i.e. the universal exchange- correlation part Exc[p]. Several levels of realization exist to construct Exp] = / mum, (4.6) and they correspond to adding more complex dependencies in the functional Eu [p]. The standard classification separates, from the most simple to the most involved level of description [89]: e The local density approximation (LDA), where Eu [p] only depends on the local density, p(7"') and is matched onto the energy per unit volume of the correspond- ing infinite homogenous system, 0 The generalized gradient approximation (GGA), where additional specific de- pendencies on the gradient V7p(f') are added to Exc[p], e The meta-GGA, which introduces as an additional degree of freedom the kinetic energy density of occupied Kohn—Sham orbitals 2 as = Zlvcbm (4.7) e The hyper-GOA, which takes also into account dependencies of Em[p] on single— particle energies e,- and occupations p,- , o The generalized random phase approximation (RPA) which involves unoccupied Kohn-Sham orbitals, and can be seen as the ultimate goal in terms of global accuracy. 32 However, in spite of several recent developments, a rigorous connection between nuclear EDF and DF T approaches is yet to be found [[98]—[103]]. The key aspect of this problem is the fact that unlike the systems that are studied in condensed matter physics and quantum chemistry (bound by external potentials), the nuclear many- body problem involves a self-bound system. In contrast to the standard Hohenberg- Kohn theorem which is symmetry-conserving, the nuclear Kohn-Sham potential im— plementation of EDF approaches breaks symmetries of the Hamiltonian such as trans- lational and rotational symmetries. Even though projection techniques can be used to restore these symmetries, understanding its implications for DF T requires further theoretical development. Additionally, the presence of both spin and isospin degrees of freedom and the importance of pairing correlations need to be considered in nuclear EDF approaches. For a related formulation of pairing within the DFT framework, refer to Ref. [104] although the formulation corresponds to a system coupled to a particle reservoir. 4.1.2 Single- and multi-reference EDF formulations As mentioned in section 1.1, the fact that nuclei are self-bound fermionic systems with both collective modes and individual excitations existing on the same energy scale make the nuclear many-body problem a complex one. In order to handle this problem, nuclear EDFs incorporate the assumption that these correlations can be divided into two different classes that can be incorporated in two different steps (i) short-range in-medium correlations which are recovered at the level of single-reference energy density functional (SR-EDF) calculations and commonly referred to as mean- field calculations (ii) long-range correlations that originate from collective modes and symmetry restoration. These are handled by multi-reference energy density functional (MR-EDF) calculations. In SR—EDF calculations, the EDF is a functional of the normal, p,,-. and anoma- 33 H...— _ ..__.-.—-.... n—n a.“ cut Hi lous, Hij, parts of the OBDM defined in appendix 9.2.2 and 9.2.5. In general, the energy density in SR—EDF is given by [105] gsnlq’ol E ESR[pij7f$ijafi;j 1 PP 1 m; : tiJ’ pji + '2‘ Z vijklpik plj + 1 Z vijle’kik Klj ijkl ijkl PPP * 1 pm: -4: Z vijklnm pli pmj pnk + Z Z vijklmn pli h‘jk h3mn 1 ijklmn ijklmn + QIH (4.8) where 2) denotes the effective interaction in the respective channel. Traditionally, SR- EDF calculations have been referred to as self-consistent mean-field theory where one starts from an effective two- and three—body interaction and calculates the Hartree- Fock (HF) or Hartree-Fock-Bogoliubov (HFB) energy density. However, SR—EDF calculations are distinctly different from mean-field calculations in that specific prop- erties of the interaction vertices, e.g. 1135;, = v32, are not satisfied [105]. SR—EDF calculations can reproduce static collective correlations such as pairing and deformation through the symmetry breaking of the auxiliary state |0) with respect to which the OBDM is defined. This does not hold for collective modes and dynamical correlations, which require Multi-Reference (MR) calculations. Motivating from Hamiltonian-based generator coordinate method (GCM) calculations [105], MR- EDF is formulated as g _ 230,16le foISAml‘I’o) ‘I’il ((1)0lq)1> MR = * 20.16MB f0 f1 (‘I’ol‘l’d (4.9) where EMR[0,1] is the MR—EDF and the weight functions f0, f1 are determined by symmetry consideration and/or diagonalization. If one follows the Hamiltonian formalism, the most natural guidance for the construction of €MR[0,1] is pro- vided by the generalized Wick theorem (GWT) [105] which asserts that one obtains 34 in ft? .' EMR[(), (1)1] by replacing the SR density matrices by transition densities [105]. N ev- ertheless, the application of this prescription to currently available EDFs encounters several pathologies which have been traced to the occurrence of non-integer pow- ers of the density matrix in the functional. One proposed solution [105] is the re- parameterization of EDFs in terms of only integer-powers of the density matrix. 4.2 Skyrme energy density functionals In the Skyrme-EDF model [26, 106], the energy density functional 8 is given as the sum of kinetic, particle-hole, particle-particle (pairing), Coulomb and center-of-mass correction terms, i.e. £[p7 5: 5*] = 8km. [p] + splilp] + £pp[pa I‘E, Hf] + 6Com. [P] + 800.11). [,0] - (410) 8 is quasi-local and is expressed as the single integral in coordinate space of a local energy density. The expressions for Skin, ECO“), and Scam, can be found in the lit- erature [26]. They are also discussed in section 6.1.4 in relation to the application of the density matrix expansion [[107],[170]] to the HF energ;r from a generic NN interaction. 4.2.1 Particle-hole functional The particle—hole part of the Skyrme-EDF resembles meta-GGA functionals in a DFT context as it uses explicit dependencies on several local densities and currents, includ- ing spin-orbit densities. This is crucial for the proper treatment of finite nuclei. The functional is the most general bilinear combination of all local densities, built from the density matrix up to second order derivatives, in such a way that 8 remains invariant under the transformations associated with all symmetries of the nuclear 35 Hamiltonian, i.e. parity, time-reversal, rotation, translation, gauge and isospin trans- formations [126]. The functional reads gphlpl ZESkyrmelPl = Z de-rApp pq pq + ApAp qupq + APT (quq _]"(1.jq) +A33§q-§Q+A8m§q-A§q+Apv" (pqv- jq-l-jq'fi X s”) + 41‘7sz - 5"’)(V - 4'1) + A""( Z JZVJ:,..— seq - T‘qq) uu +21“. [(ZJg,)(Z Jg,,)+ ZJZVJL’H— 25"! - F0] 41 qséq’ + Bssgq - 5”, + BSAss'q - A5", + BPVJ (9‘16 - j”, +jq ° 6 x §q’) +BV5VS(V-s )(v-g ’)+B“(ZJZ,,J§,C—"q Tq’) +B”[(ZJ3,,)(Z Jg,,)+ 2J3 J3], —2§" F‘I',] (4.11) where the coupling constants AX /BX refer to the interaction between particles with identical/different isospins, respectively. The densities that occur in Eq. (4.11) are given in appendix 9.2.3. The coupling constants AX/BX may further depend on densities that do not involve spatial derivatives. Historically, Eq. (4.11) was derived starting from the HF expectation value of a Skyrme interaction [108] which contains zero-range terms plus gradient corrections to encode finite-range effects, and is a sum of central, spin-orbit and tensor terms. i.e. vSkyrme(fia F) = vcent. (1313,?“ ) + ULs(I—f, F) "l" 'vtens.(fi1 F) (4‘12) 36 _. A 1 . vcem(R,F) : t0 (1 + :roPa) 6(7") + 6 t3 (1 + 3:3Pa) 1 . 2. , - 2. . 2. :. +§t1(1+41P.>[k%+ok’-6k +p7(7_") 6(4) (4.13) 01,30?) = “42) ((27.1 + 02:2) '11:, X 60%)]? (4.14) —o 1 -+ I a ”'1 -o —o "' _, vtens.(r) : E te { [3 (01 k )( 2 k ) _ (01 02) km] 60‘) +4., [3 (a. -k’) 50?) (52 .115) — (a, .52) 1}."- (sm i5] . (4.15) In this context of viewing 8?), [p] as the HF energy from a zero-range Skyrme force, the timeeven and time-odd terms of the coupling constants of the Skyrme energy functional are related through the underlying parameters of the Skyrme inter- action [106]. However, in the general EDF formulation, the time-even and time—odd couplings are independent of each other, aside from relations dictated by local gauge invariance. Even though this most general second-order particle—hole functional has been known for quite some time, traditional studies concentrated only on those terms which were deemed most important. Recently, the impact of all couplings is being analyzed in various studies [[26], [159], [158]]. 4.2.2 Particle-particle functional Neutron-neutron and proton—proton pairing acts mostly in the spin-singlet channel 5' = 0 of the nuclear interaction, as shown by the properties of the bare NN force [117]. At the same time, it occurs mainly in the 3 wave, that is a local pairing functional. This is usually used to justify the expression of the particle-particle functional 8”, as 841mm = /(1FA’3"Z|(5”|2. (4.16) q 37 where usually ~ " t” (”ll App E — 1 — . 4.17 4 i 77 psat ( ) The latter expression derives from a density-dependent delta interaction (DDDI) [127, 128,129,130,131] vf’fi(F,f—f) E v”(ff)6(f’) E {0 (1:21)”) 1_n(_pop(fl) ] 6(F). (4.18) It is bilinear in the pair density (3", defined in Eq. (9.99), whereas the strength {0 is taken to be the same for neutron-neutron and proton-proton pairing. £W[p, ,5, 5*] enforces pairing correlations only in the T = 1 channel, as proton-neutron pairing is usually neglected. The introduction of T = 0 pairing requires a more involved formalism, since pairing correlations can now couple between superblocks of different signature in the HFB equations [118, 119]. Two parameters 77 and a control the spatial dependence of the coupling constant through the overall isoscalar density-dependent coupling. A zero value of 77 corresponds to a pairing strength that is uniform over the nuclear volume (“volume pairing”) while 1) = 1 corresponds to pairing strength which is stronger in the vicinity of the nuclear surface (“surface pairing”). A value 17 = 1/2 corresponds to an intermediate situation (“mixed-type pairing”). Values 0 < 1 correspond to stronger pairing correlations at low density. 4.2.3 Self-consistent solution After the construction of the densities pi,- and #3,,- from an auxiliary [), the variation of the EDF (Skyrme-EDF) with respect to these densities results in Hartree-Fock Bogoliubov (HFB) equations. Refer to appendix 9.7 for a brief discussions of these equations. One solves these equations self-consistently. For detailed discussion on this, refer to Ref. [81]. 38 p38, [fm'3] Koo [MeV] (m"‘/m)8 It.) E /A [MeV] Ref. SLy4 0.160 229.9 0.70 0.25 -15.97 [135, 145] 3111 0.145 355.4 0.76 0.53 -15.85 [146] m‘l 0.162 230.0 1.00 0.25 —16.07 [144] pg, 0.145 230.0 0.70 0.25 —15.69 [144] p3,, 0.160 230.0 0.70 0.25 —15.99 [144] p3,, 0.175 230.0 0.70 0.25 —16.22 [144] T6 0.161 235.6 1.00 0.00 —15.93 [147] SKa 0.155 263.1 0.61 0.94 —15.99 [148] T21-T26 0.161 230.0 0.70 0.25 —16.00 [142] Table 4.1: IN M properties of Skyrme functionals (from Ref. [81]): saturation density p53,, bulk compressibility Koo, isoscalar effective mass (m*/m)s, Thomas-Reiche—Kuhn enhancement factor K,” and energy per particle at saturation E/A. 4.2.4 Existing parameterizations About 150 parameterizations of the Skyrme EDF have been defined so far and ad- justed for various purposes (see [120] and references therein for the most common parameterizations). Sample parameterizations and associated properties of INM are shown in Table 4.1. These functionals differ in what quantities were emphasized dur- ing the fits. For instance, T6 has an isoscalar effective nucleon mass (m*/m)3 = 1, providing a denser single-particle spectrum, while SKa has a different isoscalar effec- tive mass, but also a different density dependence (density—dependent term with an exponent of 7 = 1 / 3 instead of 'y = 1/6). T21 to T26 incorporate tensor terms that differ by their neutron-neutron couplings [142]. 4.2.5 Predictive power of empirical EDFs The discussions in the previous several sections were for the Skyrme EDF. Even though we have not discussed Gogny and relativistic [26] realizations of the EDF, the key points of this section regarding the predictive power of currently available EDFs holds for all three implementations. This is due to the fact that these EDFs generally 39 provide comparable predictions, in spite of some variations for particular observables [[26], [106]]. The application of phenomenological EDFs for a broad range of nuclear structure problems has been a success story in the past few decades [26]. Recently, the growth of available computational power has allowed large-scale projects, such as deformed calculations of ground-state properties over the nuclear chart. Systematic calcula- tions of ground-state properties, as well as some collective excitations, for all known and theoretically predicted nuclei, are now available. Mass residuals over about two thousand known nuclei obtained at the SR-EDF level are of the order of one MeV, which is an accuracy sufficient for a direct comparison with experimental data [[121] -[123]]. Such calculations also provide a reasonably good description of static proper- ties beyond the ground-state energy, e.g. shell structure, pairing gaps, charge radii, individual excitations or deformation. Likewise, MR—EDF calculations have already met a lot of success, in particular regarding the description of dynamical correlation energies, vibrational/ rotational excitations and super-deformed bands or shape transitions [[124], [125]]. Among other challenging areas of interest, extensive studies have for instance been dedi- cated to [106] (i) (asymmetric) fission properties of heavy elements, (ii) the forma- tion of superheavy nuclei, (iii) the application of dynamical approaches based on the time-dependent HF / HF B formalism to describe nuclear fission / fusion, and (iv) collec- tive motions through the self-consistent (quasiparticle) random phase approximation ((QlRPAl- However, many challenges are still ahead in order to (i) further increase the overall precision of EDF-based methods, e.g. decrease mass residuals, (ii) describe excited states with spectroscopic accuracy (of the order of 300 keVs), as it is achieved for sd—shell nuclei using the Shell Model [[132]-[134]], (iii) control spin and ferromagnetic instabilities and (iv) improve the predictive power of EDFs in the unknown region 40 TW “-0 ti ' 1 V I ' l o-neu n separa on energies Mass Formulae 25 20 - $ - .-' 0 s .2, <5 V3 5 15 5' ' data 't ‘9'! ' 0 - "'46 P cm s _-‘ data do not ex15t , 10 1 l . | g 1 l 1 l 1 I 1 i 4_ [4L l 50 60 70 30 80 90 100 110 120 130 Neutron Number Neutron Number Figure 4.1: Illustration of the asymptotic freedom of phenomenological EDF models in the case of two-neutron separation energies. In the major shell where empirical EDFS are adjusted on experimental data, the agreement between all relativistic and non—relativistic calculations is clearly seen. In the next major shell where no data exist, discrepancies between these models become more apparent (from J. Dobaczewski et al. [150]). 41 of the nuclear chart. Indeed, while all empirical models constrained by experimental data mostly agree with each other within the major shell they are adjusted in, extrap- olations towards the nucleon drip-line do not agree with each other. This divergence in the next major shell is seen for most standard observables such as the two-nucleon separation energy or the pairing gap and is exemplified by Fig. 4.1. Furthermore, empirical EDF models give rise to spurious effects. For instance, the particle-hole effective vertex extracted from typical empirical functionals is rarely fully antisymmetric (e.g. fractional density-dependencies). This leads to a series of difliculties in SR— (self-interaction and self-pairing) and MR- (poles and spurious steps) EDF calculations. Some of these issues have been identified and practical cures have been proposed [105]. However further developments are required in order to develop a fully satisfactory theory. 4.2.6 Outlook Various groups are pursuing different strategies to overcome the deficiencies of phe- nomenological EDFs and make them of spectroscopic quality. In this context, spectro- scopic quality refers to the ability to describe and predict not only the bulk properties such as mass and radii but also low-energy spectroscopy and collective states of nu- clear systems far below the MeV accuracy. On the one hand is the effort to empirically improve the analytical form and couplings of the EDFs [[31]-[36]]. This includes 0 The construction of EDFs containing beyond second order derivatives [35]. Re- cently, the authors of Ref. [35] undertook the construction of nuclear EDF with up to sixth order in gradients. It is possible to reduce the large number of cou- plings significantly by the successive application of symmetry constraints such as Galilean (gauge) invariance. Further reduction can be accomplished due if one requires time-reversal and spherical symmetries. Furthermore, the number 42 of couplings also depends 011 whether one incorporates density dependencies 011 all or some of the couplings. Approaches that rely on the pseudo—potential perspective, start by selectively enriching various parts of the effective interaction. There have been several suggestions to augment the traditional Skyrme interaction given in Eq. (4.12), e.g. adding a spin-density dependent term [33] 1 _. . 1 _. V = 660 +scaPasors>6<0 + gtz‘i'u+43‘Pa{§1(R)P-9t)6(0. (4.19) where the exponents 73 and ’73: are even integers in order for the EDF to remain time-even. The contribution of these terms vanish in even-even nuclei. These additions seem to remove spin and ferromagnetic instabilities [33] from conven- tional EDFs, an improvement that must be seen in light of the fact that the spin-isospin components of nuclear EDFs are less understood / constrained than their scalar/isoscalar counterparts. Systematic fitting of the nuclear EDF. This does not necessarily imply improv- ing the form of the functional. Rather, it focuses on the application of advanced algorithms to explore the manifold of permissible parameterizations with the use of a large set of experimental data as a reference [106]. Traditionally, practition- ers have taken the easier route of only using ground state properties of magic and semi-magic nuclei to constrain the couplings. The availability of data on nuclei far from the valley of stability have provided more stringent constraints on the couplings, with special emphasis on the isovector properties [106] that are less understood. The experimental data identified for this purpose include (i) bulk properties such as binding energy and charge radii (ii) spin-orbit split- ting in nuclei for which accurate data exists such as 40Ca, 48Ca, 90Zr or 132811 43 in addition to 16O and "me which are usually employed (iii) neutron radii (iv) odd-even staggering of binding energies in medium to heavy nuclei (v) isotopic shifts, deformations, excitation properties and (vi) nuclear matter saturation properties and the equation of state of pure neutron matter. While no definite proof exists that one can not obtain significant improvement by following this method, recent results [109] indicate that the form of both the functional and couplings might be too limiting to obtain predictive EDFs. A complementary approach is one that relies less on fitting empirical functionals to known data, but rather attempts to constrain the analytical form of the functional and that values of its couplings from many-body perturbation theory (MBPT), based on realistic two- and three-nucleon (NN and NNN) interactions [[110]-[154]]. This is the path followed in this work, which is similar in spirit to OEP (orbital-dependent energy potential or ab—initio DFT ) [115, 116]. The main techniques, results, possible future extensions and outlooks are presented in the next several chapters. 44 Cl Chapter 5 Constructing Non-Empirical Energy Density Functionals 5.1 Constructing Non-Empirical Energy Density Functional It is commonly asserted that the nuclear many-body problem is intrinsically non- perturbative [38]. The strong short-range repulsion, the strong tensor force from iter- ated pion-exchange, and the presence of nearly bound states in the S-wave constitute the main reasons as to why the nuclear many-body problem is non—perturbative [27]. However, this argument relies on the assumption that the nuclear many-body problem is driven by an absolute, unique Hamiltonian, without making explicit reference to the intrinsic energy or resolution scale that underlies the modeling of such a Hamiltonian. However, recent studies have shown that the above statements need qualifica- tion as the nuclear Hamiltonian depends on the energy resolution scale [27]. In this context, an important recent development is the construction of low-momentum in- teractions starting from chiral effective field theory (EFT) interactions and using 45 renormalization group (RG) methods. Even though these methods can be applied to any interaction that originally couples low and high momentum states, chiral EFT interactions are preferable starting points because of the consistency that character- izes their many body-forces forces and operators as well as because of the possibility to systematically improve their precision by going to higher chiral orders. Refer to section 2.4 for details. The use of low-momentum interactions simplifies the nuclear many-body problem as it eliminates, or at least weakens, the main origins of non-perturbativeness [27]. In particular, the consistent three-nucleon interactions become perturbative as one lowers the intrinsic momentum scale of the two-nucleon piece [28]. Calculations of in- finite nuclear matter using MBPT in terms of low-momentum two— and three-nucleon interactions show convergence, at least in the particle-particle channel. As Fig. 5.1 shows, including the second-order contribution from the two— and three-nucleon inter- actions, one obtains reasonable saturation properties of infinite nuclear matter, with weak dependence on the resolution scale [28]. Moreover, the freedom to vary the order of the input EFT interactions and the cutoff via RG provide a powerful tool to assess theoretical errors arising from truncations in the Hamiltonian and the chosen many-body approximations. All these features point to the fact that it may be possible to construct non- empirical energy density functionals. Indeed, Hartree-Fock becomes reasonable, if not quantitative, starting point [28], which suggests that the theoretical developments and phenomenological successes of EDF methods for Coulomb systems may be applicable to the nuclear case for low-momentum interactions. 5.1.1 Philosophy, Goals and Limitations Calculations in INM [28] and the binding energies and radii of finite nuclei [29] show that at least second-order contributions from MBPT have to be incorporated to eb- 46 r s ‘. e I I 9‘... .\ u — \’ 10.2 C fill'II sly u‘. .. . -.d-I-\>.Ufi .110... I 51u1r161'..;>"£F'I‘I'I'I1"Tf'-1'l .,,.. HA=L8fm 5" v NN from N3LO (500 MeV) H A =20 66' g ' 5 low k H A g 2.2 an" ._. 3NF a: to E and r ._. A = 2.3 fm " 5 ." 3H 4He '1 § ' 2.0 < Am: <25 fm / é-w -: 9 Empirical : 5 -15 C] saturation - ._ 1 . 1 Hartree-Fock 1’0"" 2nd order 8 pp ladders ‘ -20 1 1 l 1 l 1 l 1 1 l 1 l 1 l 1 l 1 1 l 1 l 1 l 1 l 1 l 4 0.8 1.0 1.2 1.4 1.6 0.8 1.0 1.2 1.4 1.6 0.8 1.0 1.2 1.4 1.6 -1 -l '1 RF [fm 1 1‘1: [fm l 19; [fm 1 Figure 5.1: (Color online) Nuclear matter energy per particle as a function of Fermi momentum hp at the Hartree-Fock level (left) and including second-order (middle) and particle-particle-ladder contributions (right), based on evolved N 3LO NN potentials and 3N F fit to E3}! and mm. Theoretical uncertainties are estimated by the NN (lines) and NNN (band) cutoff variations (from Bogner et. al. [28]). tain quantitative success. Likewise, first-order treatment of pairing correlations using low-momentum two-nucleon interaction show good agreement with experimental re- sults [112]. On the side of the interaction (chiral EFT interactions in this case), one needs to go up to N3LO in the chiral expansion in order to describe elastic scattering phase shifts in the two-nucleon sector with x2 / data close to one [12]. In addition, these interactions still contain significant coupling of low and high momentum modes which necessitates their consistent evolution to loW-momentum to make HF a rea- sonable starting point and obtain a convergent MBPT. Hence, a microscopic / non- empirical calculation of the nuclear many-body problem should incorporate at least the contribution of the diagrams shown in table 5.1 for the normal and table 5.2 for the anomalous / pairing contributions, starting from low—momentum interactions. Though the perturbativeness of the nuclear many-body problem when using low- momentum interactions is quite comforting, MBPT is still numerically too expensive for a systematic calculation of hundreds of heavy open-shell nuclei. Additionally, the accuracy of currently favored approaches such as empirical EDFs cannot be met, at this point, with completely non-empirical MBPT calculations. Hence, a method 47 Table 5.1: MBPT contributions from NN and N NN interactions up to second-order (Normal contractions) in Hugenholtz representation. MBPT Order NN-interaction NNN-interaction First Order in MBPT OO Second order in MBPT Table 5.2: The first-order anomalous/ pairing diagrams, otherwise called Bogoliubov contributions, from the NN and NNN interactions in Hugenholtz represenation. <30 from NN: from NNN: 48 is sought to map MBPT contributions to numerically tractable forms, such as local EDFs, with the aim of refitting some parts of the functional in a controlled and theoretically motivated way. In this work, we do not attempt to derive a completely non-empirical EDF. Rather, we have a more pragmatic goal of enriching and improving current phenomenological Skyrme EDFs by identifying and incorporating novel density dependencies arising from missing pion physics. We further restrict the work in that only the first-order (HF) contributions from the un-evolved chiral EFT NN + NNN interactions at N2LO have been calculated. Subsequently, we apply the DME to the resulting nonlocal energy functional to obtain a quasi-local Skyrme—like EDF. In practical implemen- tations, this is to be followed by refit of the couplings, which has the added benefit that the whole scheme can be implemented in existing codes with minimal modifi- cation. Refer to section 8.1 for more details. With the goals and limitations of the work in perspective, the justifications to concentrate only on the HF contribution from non-evolved chiral interactions and subsequent application of the DME are as follows: 0 First, it is well known that RG evolution of interactions to low-momentum modifies only their short distance structure [15]. The input chiral interaction has both contact and finite-range pion exchange parts, as given by Eq. (2.12). The RC evolution modifies mostly Vd(A). However, the energy contribution from Vd(A), at least at the HF level, is of the same form as conventional Skyrme EDFs. Thus, refit of the Skyrme parameters should compensate for the RG evolution of this part of the interaction. As we are primarily interested in identifying the dominant density dependencies arising from finite-range physics, it is justifiable to apply the DME to the energy contribution from 17,. 0 Second, inclusion of second-order contributions necessitates the development of 49 non-trivial extensions of the DME technique, as those expressions involve non- localities both in space and in time [166], while the currently available DMEs can only treat nonlocalities in space [[170],[107]]. This can be illustrated by contrasting the contributions to the energy from the HF and second-order di- agrams. Discarding all spin and isospin coordinates for the sake of simplicity and considering only NN interaction, VNN , £HF O( del (IT-‘1 VNNUFI —F2[)p(F1,F2)p(F2,7—'.l), (5.1) 62"" cc Z/dfidfidf‘sdfl[453(F1)¢3(F2)VNN(|F1-F2|)¢)(F1)¢5(F2) 0376 >< 4:05) 42%) WWI-Fa — el) 40(5) 4446) X Pan 9643 (1 - PM) (1 — p56) 60 + 65 — e, -— 66 , (5.2) where pm, is the density matrix, defined in Eq. (9.70), in the canonical single- particle basis of the reference HF reference state and 60, is the energy of the single-particle level. While the HF contribution, U”, can be expressed as a functional of p(F1, F2) only, the same cannot be said about the second-order con- tribution, 82M, or any beyond-HF contribution. This is due to the occurrence of energy-denominators. A satisfactory extension of the DME that can properly handle beyond-HF contributions and in particular the energy-denominators is yet to be invented [166]. Third, it is well known that the dominant contributions to bulk nuclear proper- ties are of Brueckner-HartreeFock (BHF) type [38]. Operationally, this amounts to replacing the vacuum interactions in the HF expression by a Brueckner G- matrix, which is discussed in section 3.2. But, the G—matrix “heals” to the bare interaction at long distances. This is usually demonstrated by studying the behavior of the S—wave in—medium pair wave-function (at zero center of mass 50 momentum) of the Bethe—Goldstone equation in a repulsive hard-core spherical potential [38] __ sin(kr) sin(k7‘c) 90372) ‘4’“) — ‘7..— ‘ a. g(rc,7‘c)’ ’" Z (5'3) $0") = 01 T < To) (5.4) where To is the radius of the hard—core, k is the relative momentum of the two-particles and _ i 1 wdk,sin(k’r)sin(k’r) — 27r2 rr’ kF k2—k’2 ’ 9(7) 7") (5.5) Figure 5.2 shows the solution of Bethe-Goldstone S—wave solution for rela- tive momentum, k = k; / 2, and the uncorrelated two-body wavefunction, 630(7') = sin(kr)/(kr). Simple analysis shows that g(r,7") decreases rapidly ¢(X) 1.0_ - Uncorrelated 0.8_ 0.6: . 1 , - -* - Correlated(BG) 0.4: 0.2: _ \ ‘ l ‘ g X =kFI‘ ; 6 8 10 —0.2~ ———— Figure 5.2: (Color online) The S-wave solution of the Bethe-Goldsone equation and the uncorrelated S-wave function. 51 with increasing 1“, with a distance scale of l/kp. One defines the healing dis- tance, rh, which refers to the distance beyond which the the wave-function effectively attains the unperturbed value. This is given by the approximate re- lation kprh % 1.9 , more or less independent of the relative and center of mass momenta [38]. Then, one can use Eq. (3.10) to show that G—matrix heals to the bare interaction in the same manner. Hence, applying the DME to the finite-range part of the interaction, viz, 17,, at the HF level will capture the same contributions to the density-dependent couplings as given by the finite— range part of the G-matrix in a more sophisticated BHF calculation. In this way, the dominant density-dependence that arises from the finite-range of the interactions is accounted for. Finally, the algebra required to obtain even the starting point for the DME (viz, DME on the HF energy from chiral EFT NN + NNN interactions at N2LO) is so tremendous that most of the work can be done only using some form of automation [[156], [161]]. This is especially the case if one wants to have the complete form of the functional without any restricting assumptions regarding time-reversal invariance and /or spherical symmetry. This work is just the first step in the long-term project of building non—empirical 11(3lear EDF. There are several pos31ble extensmns that can be made 1n the future. R efer to section 8.2 for a related discussion. 5 ‘ 2 The Density Matrix Expansion (DME) T hQ DME was originally proposed by Negele and Vautherin [170] to derive an ef- fe - Qtlve nuclear Hamiltonian. In the first paragraph of their paper [170], Negele and \r autherin note that the purpose of the density matrix expansion is to relate the compu- 52 tationally simple efiective interactions of 6 and Skyrme forces to the computationally cumbersome theory derived directly from the nucleon-nucleon force. In deriving an effective nuclear Hamiltonian, Negele and Vautherin avoided fol- lowing the moment based expansion which were considered in earlier works [167] in which one considers expansions of the fourier transform of a short-range interaction. Their rational for doing that was the fact that the long range part of the nuclear G-matrix heals to the bare one-pion-exchange-potential (OPEP), which causes con- vergence problems for moment based expansions. Hence, they invented an expansion, the density matrix expansion, that exactly includes the long-range OPEP tail for the nuclear density matrix [170]. 5.2.1 Basics of the DME The central idea of the DME is to factorize a local or nonlocal density obtained from the one-body density matrix (OBDM) by expanding it into a finite sum of terms that are separable, usually, in the relative and center of mass coordinates, (F, it). There are a few exceptions to the (7", If) choice as the DME-coordinates. These exceptions are mostly relevant to the the application of the DME to the HF energy from the chiral EFT NNN interaction at NQLO. Refer to section 7.2 for details. Adopting (F, If) as our DME-coordinates and the notation introduced in Ref. [168], one writes the general DME formulae lmax p,()=',,r~'2) z 2117(4)) 79,(R'), (5.6) [:0 mmax emf.) x Z 11:1,,(kr) 0,,(12) , (5.7) m=0 "max ache) e 2 112a.) ”..(R’). (5.8) 11:0 53 olllilX 3W2) e 211-SW) 141?), (5.9) 0:0 “max —. 9.05/2) 2 Zflmwm. (5.10) 1120 where k is a momentum scale to be determined that sets the scale for the decay in the direction of the relative coordinate 7", II,f (k r) are the so—called n—functions that remain to be specified, and {731(6), 9114(6)} E {04(R))Tq(1f), Jaw/(é): WG-5), qu(§),«§'q(§), FAR), “fill, (5‘11) ~ {7346). 90(6)} 6 {71(0), qu—i), Jam/(R): 6154(15): AMI-i), «311(6). 6(5), 6(a)), (5.12) refer to the local anomalous densities, while gqm/Q) and Hum?) are from the set of local normal or anomalous densities. The DME emphasizes separability of the expansion in the relevant expansion- coordinates above the approximation of nonlocality. That is, even for local densities that depend on a single coordinate and hence with no nonlocality, one can talk about an expansion in terms of the DME-coordinates as stated by Eq. (5.10). In a sense, one is approximating the nonlocality in one of the DME-coordinates. For exam- p18, pq(1"'1) E pq(E + F/ 2) can be expanded in terms of quantities that depend 011 If and 1" separately. In practice, however, the emphasis on separability above the approximation of nonlocality is of limited use as most DME approaches rely on an— alytical techniques that fail to work when there is a long-range of nonlocality in the 54 ml‘rt‘ \. v . ‘, lbs w r .. ,0 v- T‘ 1 fl . ‘ l‘ Willi...‘ , ,, ..-i, . 1., . l “-031 \, 1 ..r: ml. .13) .1. . .’ pr -. ' u r - n ,r; , .'Ls lULL‘utLuu V P'ry‘ e. W {1:16 519. "P 7 .~ '} 3.. 9.1%.”; l 7 lo, ' 5'91”), _:r\ I if“ ”i . ' tA“ .‘\\ T‘.‘ r II: E ' ~ Lu“ (Jy- "1‘ A 1‘ ‘, r- ”n w» . f“"“-‘A“ , A;.F‘.. P ,‘I L.» 11111.1: wlv Nib—1'... I'll ‘-_, 1 .M~I "‘1' (1‘ v in (h H b (If; .1 { i;_v. ' expansion—coordinate. This work concentrates mainly on the expansion of the nonlocal scalar and vector components of the normal part of the density matrix, viz, pq(F1,F2) and §q(F1,F2). The extension of the approach to non time-reversal invariant systems is important for constraining the nuclear EDF for those systems. This is discussed in section 5.3.6. The apparent need for the DME of the local densities (pq(F1/2) and L(F1/2)) that appear in the exact HF energy of time-reversal invariant systems, justifications for why one should avoid expanding these densities and related technical problems and their possible solutions are discussed in section 5.3.7. The expansion of the nonlocal anomalous densities, especially ,5(7'~'1, F2), has drawn some interest due to the need to enrich the pairing part of the nuclear EDF. Nevertheless, unlike the nonlocal normal densities, there are some conceptual and technical difficulties to be overcome. These are discussed in section 5.3.8. We gauge and compare the accuracy of the various DME approaches using non self-consistent measures. Finally, we augment this with preliminary self-consistent tests. These tests are discussed in section 5.4. 5.2.2 Existing variants of the DME The main problem to be solved in constructing a viable DME technique is the deter- mination and optimization of the various 7r—functions and the identification of which local densities occur in the expansion of the given density. The currently available DME techniques [[170]-[173]] approach this problem in two distinct ways. On the one hand are those methods that resum infinite order “Taylor-series” expansion terms in a clever way, while on the other are those that start with an inspired ansatz and paramterize and optimize the the 7r—functions phenomenologically. In the first group, we have the original DME of Negele and Vautherin and its variants [170, 171, 173], while in the second group we have those that are mostly based on gaussian approxi- mations of the scalar part of the OBDM [172]. In the phenomenological optimization 55 g: war it}. 1 h ad'ilii' rpruxiizmtv lllilr: mm: 1135’: that ar' an . 41¢; XXL.» Illt 5 a ‘-"rif'.-~_i‘.1i . TFI. ‘I ‘ Y‘:_ . rusL DAY‘\“~~“\ ’r-r-‘n 1-; ‘ . I‘ly “““lekit 11‘ 'IY\Y\- <1 ; i a) X, 0" F'\ r. v.3: PbA‘I 3'" ’~ '1‘“. 1 lf‘pdill‘u“ C—‘r, -‘ ..:!’ll;_\"g ll. ,. UV”: 1",: ‘d In; “~‘l 1. of the 7r—functions, the parameters are optimized to recover various properties of the OBDM such as the correct local semiclassical kinetic energy density and integrated projector identity of the OBDM (see Eq. (5.50)). In addition, there is yet another classification based on whether the techniques approximate the full quantal or semi-classical approximations of the density matrix. While most of the existing DME techniques approximate the full quantal OBDM, the ones that are based on Wigner-Kirkwood expansion of the single-particle propagator fall into the second/semi—classical category [173]. Further differences appear with regards to the choice made to fix the momentum scale k. In fact, the DME of Ref. [171] is a variant of the original one proposed by Negele and Vautherin (NV-DME) [170] that improves the accuracy of the expansion obtained at first order (nmax = 0) by optimizing the momentum scale k. In appendix 9.5.3, we recover the original DME of Negele and Vautherin using the PSA-DME discussed in the next section and the generalized PSA-DME, while appendix 9.5.2 contains the key points of the semi-classical Wigner-Kirkwood based expansion of the density matrix. 5.3 PSA-DME 5.3.1 Motivation for a PSA reformulation of the DME One of the main shortcomings of all existing DME formulations is that they are mostly focused on the scalar part of the OBDM. For instance, Negele and Vautherin acknowledge in their seminal paper that they were not able to design an approximation of the vector part of the OBDM on the same level, and thus with the same accuracy, as the one they obtained for the scalar part. This is an essential problem in view of constraining the nuclear EDF non-empirically. Indeed, the vector part of the OBDM is non zero in spin-unsaturated nuclei, i.e. in almost all nuclei. Moreover, 56 .1- lr .?ixii‘I‘J'P « of the OED? a the? Dixiil 2f the 0813 119:: aerate airmen Hence. r 37%;(1‘5 I0!’ :2: (1erng ilfl'Ta’l‘x’lI' . “-4 Al "'. till: L1 i-H‘r- ' J H‘J, j-‘T' .1 . .. «a. 1‘ l ,'h '- "i4 \‘ all available DME techniques hold only for time-reversal invariant systems, with no apparent extension to non time-reversal invariant systems. These problems convinced us to formulate a DME approach that has the following qualities: (i) the accuracy for the scalar part of the OBDM should be comparable to, if not better than, the existing DME techniques. It should be mentioned that the percentage error of existing DME techniques for the scalar part of the OBDM is quite small for various measures, which should be enough to capture the correct density dependence of the couplings in the resulting EDF. (ii) The DME of the vector part of the OBDM should have a comparable accuracy to that of the scalar part. Except for the DME of Negele and Vautherin [170] which performs badly for the vector part of the OBDMI, the other techniques either do not refer to the vector part at all or their accuracies are not gauged properly. (iii) It should readily be extended to non time-reversal invariant systems. Hence, we formulated a new DME technique which we call PSA-DME where PSA stands for phase space averaging. Note that the PSA formulation of the DME is not completely new. In fact, Negele and Vautherin start using the “local energy approximation” technique of Ref. [174] and mention the possibility of phase space averaging in infinite nuclear matter. For the actual derivation, they revert to a formal Bessel-function plane-wave expansion. From a formal point of view, the PSA approach developed below differs from that mentioned in Ref. [170] and is applied consistently to both the scalar and the vector parts of the OBDM. For instance, in spite of the weak angular dependence of the scalar part of the OBDM [176], the inconsistency in the order of application of the anglaaveraging and series expansion that exists in Ref. [170] is not an issue in the present case. Still, it is shown in appendix 9.5.3 that our PSA—DME approach can be used to recover the original DME. In the following, some of the key properties of the momentum phase space of 1Refer to section 5.4 for actual percentage errors of the. various DMEs. 57 a . f vnwf-JOL" -rr.oapana.'-I . A '1‘ . f..“1 \,‘ -‘ UI ‘LF‘ Ill! 1 lite Fermi systems are identified with the aim of incorporating these features into re 7r—functions with the PSA-DME approach. We implement two different strategies . incorporate these phase space features: analytical derivation and phenomenological )timization. .3.2 Momentum phase-space of finite Fermi systems he momentum phase-space distribution of quantum systems can be studied via a ultitude of quantum phase-space distribution functions [169]. Studies using the 'igner distribution in Ref. [177] and the Husimi distribution in Ref. [178] Show that .e local single-particle momentum distribution displays a diffuse and anisotropic :rmi surface at the (spatial) surface of the finite system. These are peculiar features the momentum phase-space distribution that are not present for homogeneous stems. The Wigner distribution function [175] is often used to approximate the phase ace distribution of nuclei. It has been studied both analytically and phenomeno- gically for various models applicable to nuclei (see Refs. [176], [179] , [180]). The odels include pure harmonic oscillator with sharp and smeared occupations, har- onic oscillator with orbital occupation from DDHF and meanfield calculations with Woods—Saxon potential. The analytical calculations of the various models give the same general form for e Wigner distribution function. For the case of magic nuclei and in the absence spin-orbit interaction, the distribution function fq(§,p‘) in a harmonic oscillator rtential depends solely on the dimensionless parameter 5 [180] . ‘ , I 1.: ;.~ mill la .-u.r.'r .1 ~ fl&‘[‘r_'. Iii-ll; r ‘.l 1) L‘\a. ‘Hl . '. -‘ - llbfflll‘flll '6. fl . ~ V 1 1 r5 .tflal“ l 4 r . LAA .1“ ?\ [r-rv~u«;=.21;u:.«i ‘- Y The at. ' 1 , . m dlil>-.' ””1 _ . d“ I: 'f‘ ‘1). “All [ll-J! §[. ‘ 9 r i'wv. I A -i “111‘ “4'1 and is given by £41117) )KL2( (28m, , (5.14) where hw = 41A‘1/3 is the oscillator size parameter, K is the principal quantum number and L‘}; is the associated Laguerre polynomial, given in appendix 9.1.1, and rig is the occupation probability. In Ref. [177], the authors parameterize the Wigner distribution using the Fermi distribution function. All these studies indicate a diffuse fermi-surface for the local momentum distribution with the diffuseness being much pronounced around the nuclear surface. The above model calculations are able to capture the diffuseness, but they do not show anisotropy/ deformation of the local fermi surface. In Refs. [178] and [177], the authors solve for the single particle wave functions in spherical Woods-Saxon potential with no spin-orbit interaction and show that the local fermi surface is anisotropic. This has no counterpart in the phase space distribution of infinite-fermi systems (IN M). The anisotropy of the local single particle momentum phase space distribu- tion can be quantified with the lowest order deformation of a spherical phase space distribution, viz, quadrupolar deformation. In Ref. [178], the local quadrupolar de- formation of the momentum Fermi surface (for a given isospin) is given by2 fdfi[3(é'r 'PT" - I72]Hq(F,I5) f 6151721141317) 2 Ha- - mom)? p3.- — 1 + 0<(kz~ro>2 ). (5.15) PM a = lab where Hq(f',15) is the Husimi distribution, To is a length scale used in the Husimi distribution and k}. is a short-hand notation for the local Fermi momentum kflff) 2As the anisotropy is usually not large. it is not necessary (at least in this work) to go to higher multipolas to quantify the deformation. v . 1 . . -..r._-"_ {[\ ,f‘un'fil ‘ l d ‘ V t I “"WI‘[VIO~ , , | \ ALE...” ‘A‘I '._ ‘0» t . 1|." , 1"" 31‘“ IAll‘n.“ defined in a local density approximation through _, 2 _, 1/3 km) E kg. = [37r pq(R)] . (5.16) In subsequent formulae, the If dependence of kHl—f) is mostly not shown explicitly for notational simplicity, except in formulae/ places where we have to remind its If dependence. Equation 5.15 is computed in the basis 90,-(Fq) that diagonalizes pij, i.e. the basis from which the Slater determinant |) is built. Details on the Husimi distribution and simplified expression of P30“) in spherical symmetry suitable for semi-magic nuclei is provided in the appendix 9.5.1. Fig. 5.3 shows the quadrupole anisotropy of the local neutron momentum distri- bution calculated for a selection of semi-magic nuclei. Single-particle wave-functions are obtained from a Skyrme—EDF calculation performed with the BSLHFB code [181] using the SLy4 parametrization of the Skyrme EDF. The pairing terms in the EDF were switched off. Fig. 5.3 also displays the local neutron Fermi momentum (Eq. 5.16) in order to locate the position of the nuclear surface. In spite of pronounced shell fluc- tuations, the result corroborates the conclusions drawn in Ref. [178]; P511?) becomes negative just inside the surface, denoting an oblate momentum Fermi surface while, outside this region, the local momentum Fermi surface becomes strongly prolate. In both cases, we have taken an axis normal to the nuclear surface as the reference axis. The next two sections show how we make use of these properties of the phase-space distribution of finite Fermi systems to design our PSA-DME of both the scalar and vector parts of the OBDM. 60 ~ . II-i — at“: R [fm] Figure 5.3: The quadrupole anisotropy P2"(R.) of the local neutron momentum distribution in a selected set of semi-magic nuclei. The black, red and blue vertical lines indicate the approximate half-radii (where the density becomes half of the density at the origin). 61 '. rm, (Mic. ‘ . O K K I ...-x . I O i “ll... [9. i ... , A kg)?" ‘ 5.3.3 PSA-DME for the scalar part of the OBDM of time- reversal invariant systems In PSA-DME, there are three key steps that are used to determine the local den- sities that occur in the expansion of the given nonlocal density and optimize the 7r—functions. These are: (i) Identifying of the nonlocality operator as an exponential derivative operator acting on the OBDM. (ii) Performing a Taylor series expansion of the operator about some momentum scale 1:. This is the point at which a momentum scale is introduced in the DME, though the actual form of If; is not fixed yet. (iii) Averaging the momentum scale over the local momentum distribution of the system of interest. Applying the first step, viz, extraction of the exponential nonlocality operator of the scalar part of the OBDM, one writes ) = Z (p:(7"'20q) W(fiaq) pi: i0 MI‘M .. F .. Pq(R+§’R_ _ 6 —6 = e’"(_lT2) Z @5200) 9947:1011) P211 id (5.17) _. _. T1=1'2=R. In the next step, one extracts a phase factor eff": in order to perform a Taylor series expansion of the non-locality about the momentum scale 1:. Hence, —o fl-o ._. --o" 7116—6 _i-. *4 —v pq(R+%,R—-72:) =e‘r'k e ( 9 k) Z< Zenaaqwxaam , (518) 2'0 -o r1=F2=§ Where we truncated the expansion at second order. In principle, nothing prevents to one from including higher order terms. This is especially true in light of recent empha- 62 \] r" 1 J . r. . k- . ‘ 1 - ’ l Hun sis on the inclusion of beyond-second—order gradient terms in Skyrme like EDFs [35]. In that case, one needs to define additional local densities in addition to the ones given in section 9.2.3. Noting that the derivation is restricted to time-reversal invariant systems, the next step consists of angle averaging over the orientation of 7", which is a reasonable step as the scalar part of the OBDM has negligible dependence on the orientation of F [176]. The final step involves averaging the dependence on the momentum scale I? over a model phase space that characterizes the system under study. As mentioned in section 5.3.1, we make two different choices. First, we perform the PSA with the phase space of the locally-equivalent pure isospin infinite matter. Denoting the function to to be averaged as 9,103), this operation amounts to setting the local phase space distribution, fq(§, If), as ME) = e 2 H8+gns(k‘;«r)[;,-qu¢i(7‘101(1)10?i 270102 - - v —6 .- =eir.ker.(_12_2_zk) 2 $3730”) 210102 x (O'QIElUl) 9915('F101(I)P?1 F1=F2=R -- _, <7 _ v7 .- 26”“{1-1—7‘ [( 1 2 2—2k)} 2 (p1(T20'2(I) 10102 X <02l5l01>93i(FIUIQ) 10:11 a (5-30) 71=F2=R where only the first order term in the expansion of the non-locality operator is kept. The zero-order term in the above expansion provides the local spin density §q(§) Which is zero for the time-reversal invariant systems. In fact, for time—reversal invari- ant systems, the cartesian spin-current pseudotensor density Jq,pu(R') and its gradients are the only standard local densities at hand to express the DME of the vector part of the OBDM. Consequently, we could not express the higher-order (beyond first-order) terms in the above expansion in a closed form in terms of the cartesian spin-current Pseuclotensor density and its gradients. Nevertheless, section 5.4 shows that PSA- DA’IE attains a high accuracy even at this level of approximation. Still, there is a Possibility of studying higher-order terms in the context of the generalized Skyrme 67 EDF discussed in Ref. [35]. Here also we carry out the two strategies of incorporating the phase space infor- mation: analytically and phenomenologically (parameterically). We start with the analytical procedure which was also discussed in Ref. [170], though with no reference to phase space of finite systems. Sticking with the first term, the authors in Ref. [170] argued that averaging over the orientation of E and setting k = k}? should be sufficient to provide a reasonable account of the vector part of the exact OBDM. This gives _, "' F -' F , g 7' -‘ sq,” (R + 5, R — 5) 2 z H1(k%r) ETpJqW(R), (5.31) where HiU‘i‘fT) = jo(k}'»(1§)'r) - (532) If instead one applies the same procedure as for the scalar part of the OBDM, i.e. one performs the PSA over the locally-equivalent pure-isospin infinite matter phase-space, as given in Eq. (5.19), one rather obtains 3151:1452) _ Hflkg‘r) : kq (R)T‘ F (5.33) However, and as mentioned in section 5.3.2, the local momentum phase-space distribution of finite nuclei has a markedly different behavior than that of INM around the nuclear (spatial) surface. Given that the vector part of the density matrix peaks around the nuclear surface, it seems more appropriate to perform the PSA over a diffuse and anisotropic phase space. Given that we do not have a parameter free Way of introducing the diffuseness and the primary quantity to be averaged, emf, couples the orientation of 1" and If, we limit ourselves to invoking the anisotropy of the phase space. As a. completely analytical approach, we perform the PSA over a 68 deformed Fermi sphere that incorporates the information contained in the function P§(R) discussed in section 5.3.2. We do this by averaging over a spheroidal local momentum distribution given by MRI—5) = 902’ - k2) (5.34) where k2 k2 2 k; = r; 3. + 5’. + '2... (5.35) 0(R)2 “(3)2 C(R)2 with a(R) and C(12) being position dependent quantities that relate to P30?) The specific relations and various details of the derivation are given in appendix 9.5.3. The final result differs from that in Ref. [170] only in the analytical form of Hi. The result reads ~ ~q a Hf(k}',r) = 3919(5)” , (5.36) 52am where ~ -' 1/3 _. kg. 2 (fl) 1am). (5.37) 2—1‘25’(R) The PSA over the locally-equivalent neutron or proton infinite matter modifies the analytical form of 11? compared to NV-DME, i.e. compare Eqs. (5.32) against ( (5.36) and (5.39)). In addition, and contrary to the scalar part of the OBDM for which it is unimportant, taking into account the deformation of the local momentum distribution of the finite system leads to a modification of the relevant momentum scale 12%. In view of isolating the significance of such an effect, while preserving the benefit of using PSA, one can set Pfllf) = 0 in Eq. 5.37. In the second strategy, we incorporate the phase space information by parameter- 69 1~~v ‘ 7’ .iL , . (al.. A , ‘. f 1171‘”. I‘m -‘ ‘ u, ~r‘" "T .. 1.1 Llf ] izing the anisotropy of the Fermi surface. I11 leading order, one talks about quadrupo- lar deformation of the Fermi Surface. Thus, the PSA is performed in a phase space distribution with fq(1'2',i5) = (9(5- 4:3.) (1+ a(3Cos2(6) — 1)) , (5.38) where a is a quadrupolar deformation parameter to be optimized, 6 is the angle between R and if and Si is the SinIntegral function. Even though the deformation actually couples R and If, we approximate this as a coupling between F and I: to actually obtain the final form of the 7r—function. Thus, Hi“ (R, 7') reads _ 3 3.10627) (1 5' q H1(kp7‘l— (1524‘) +(k%T)3 (—18 Si(k‘,’,r) — 61:24" c0s(k‘}+r) + 24 sin(k%r)) . (5.39) To parameterize the deformation parameter a, we take hint from Wigner-Kirkwood expansion of the scalar part of one-body density matrix. As explained in appendix 9.5.2, the h2 Wigner-Kirkwood expansion of the Wigner transform of the scalar part of one- body density matrix reads - 52 , 52 1 .. pwmam = as 425) — ,g—TgAvqé (A— ht) + 54—m[(vm)2 + 5(p-vrvq] x 6”(A — 53,) + 005‘), (5.40) Where ha, 2 H, = $172; + Vq(R) is the single particle Hamiltonian and /\ is the Chemical potential. The origin of the deformation at this order is the (13' - V)2Vq(R) term. It is well known that close to the nuclear surface, the self-consistent potential that acts on the nucleons, Vq(R), and the density, pq(R), have similar profiles. Both are usually approximated by the Woods-Saxon shape, of course with opposite signs. Hence, we make the series of approximation [Fla/(KR) % AVq(R) o< qu(R), with 70 AW ... hp 1 ll.“ .Al-blll r l 1 Err]. . V‘ a - ' .v ..2. . l . V ‘ ‘ ‘1 A ... our final parametrization being a. = m qu(R) + b (5.41) where m and b are constants to be fit. If one sets a. = 0, the l—IflkZm) given in Eq. (5.39) is recovered. In Sec. 5.4, we discuss and compare the accuracy obtained using all of the preceding DME variants for the vector part of the OBDM. 5.3.5 It} and isospin invariance of the resulting EDF Dealing separately with the neutron or proton OBDM in a finite nucleus, it is natural to perform the corresponding PSA over the phase space of neutrons or protons of the system. However, this provides 7r—functions with an explicit isospin dependence. Even though this does not have any implication at this point, it does when we apply the DIVIE to the HF energy of two- and three-nucleon interactions as discussed in the next two chapters. This is because the EDF that results from the application of the DIVI E breaks isospin invariance (but not its isospin symmetry). As mentioned in section 2.2, there are isospin-breaking parts of the nuclear interaction, still the fact that We get an EDF that breaks isospin invariance even when we start from one that has that, symmetry might not be a welcome feature. A simple prescription to recover Symmetry of isospin invariance is to replace all k} with the isoscalar kp which is defined through 1 3 352 / VV —0 —o —o here 10(R) E pn(R) + pp(R). Fig. 5.4 shows k2, k}? and the isoscalar kp extracted fr 0m a. converged self-consistent calculation of 214Pb. This is a neutron rich nucleus an d thus, the difference between the three momentum scales should be maximized. C j. _. .. .. onsldering the small difference between kHR), k%( R) and the isoscalar kF(R) that 71 ..V- '0 l'r' 1.4‘ 1.2 1.0- :5: 0.8- E' . “i 0.65 l 0.4- ) 0.2~ 0°50 ' 2L0 ' 4L0 ' 6.0 ' 8L0 ' 10.0 R[fm] Figure 5.4: (Color online) k2, k}; and the isoscalar kF extracted from a converged self-consistent calculation of 214Pb, a neutron rich nucleus. we see in Fig. 5.4, the prescription of replacing k} with [CF in the 7r—functions might be a satisfactory method to recover isospin invariance in the resulting EDF. 5-3-6 Extension to non-time-reversal invariant systems The PSA-DME approach has enabled us to obtain both analytical and parametrized forms of the various 7r—functions that occur in the expansion of the scalar and vector parts of the OBDM in time-reversal invariant systems. However, from a formal point of View, PSA-DME uses the assumption of time-reversal invariance only to turn off the tifile—odd densities. Thus, one can envision direct extension of PSA-DME to non- time rexrersal invariant systems, where the time-odd densities such as 34R) in the Case of 10.10"}, F2) and §q(R), TAR), PAR) in the case of §(f'1,772) start playing a vital r Ole. Nevertheless, it will be clear from the subsequent discussions that the appearance 72 3 l' h".‘ r-r~'v 1 in“ of time-odd densities gives rise to various constraints that the 7r—functions have to satisfy in order for the EDF (that results from the application of the DME to the starting HF energy) to respect certain global and local symmetries. This requires a systematic study of the problem. We tackle this by formulating a generic DME, which we call the modified-Taylor series, and using it to perform a formal study of the issues related to the extension of the DME to non-time reversal invariant systems. The complete development of a DME for non-time reversal invariant systems with specific analytical/parameterical 7r—functions is outside the domain of this work. The modified-Taylor series approach was introduced in Ref. [168] and expanded in this work. It consists of replacing the numerical coefficients in the Taylor series expansion of the density being expanded with 7r—functions which are yet to be deter- mined. These 7r—functions can depend on one or several variables. Illustrating the eXpansion with the nonlocal scalar density and the nonlocal vector density, we have -' -o "’ F --0 -¢ pq(7"1, 7'2) 2110(9) pq (R) + 111(9) 5 (V1 — V2) pq (r1, T2) .. F1=F2=R 1 F 5 2 F 5 2 + 5112(9) [(5 V1) + (5 V2) [pq(7‘1 , 7'2) (5.43) i=fi=fi F 7" -» _, _, — I13(9) (— V1) (2 V2) Pq(T1, 7‘2) 5 F :17 =1? 1 2 and 35(R :1: 12:12 4: g) = ng(r2):,(f2') :t 113(9); (6. — 62);,(51, a) - 71:52:] 1 8 F _. 2 F 2 - - - +‘H2(Q) - ' V1 + - V2 3q(7“1."2) 2 2 2 _- 5 T1—T2— g F -* F '0 _. _. _. _H3(Q) _ ' V1 — ' V2 Sq(r1 a 7‘2) (5.44) 2 2 F1=172=if Wh ere the 7r—functions are to be found analytically or optimized phenomenologically. 73 fl represents the variable 011 which the 71‘ functions depend. Requiring Q to be scalar, dimensionless and depend on 1‘ implies that Q = rk. In case where Q is assumed to depend on F, then 9 = h(F,lZ) for some scalar function h. Here, we assume the 7r-functions to be independent of the orientation of F. The choice of having four 7r—functions instead of five in Eqs. (5.43) and (5.44) is motivated by the need to get a symmetric expansion in R + g and R — g. Even though a definite approach with which to constrain Q and the n—functions is be discussed, it should be mentioned that the modified-Taylor series approach is formally applicable to all local/nonlocal and/ or normal/ anomalous densities in both time-reversal and non time-reversal invariant systems. In line with this, the modified— Taylor series expansion for all the densities defined in section 9.2.3 and 9.2.5 is given in appendix 9.5.7. In the construction of this expansion for the various densities, we have not made any reference to a constructive way of fixing the basic expansion variable, 9, and the 7r—functions. This is where explicit connection is realized between the modified-Taylor series and PSA-DME (and/ or other DME variants) discussed in the previous few sections. In other words, the modified-Taylor series can be seen as a template which can be adapted to various DMEs which in turn can be considered as as approaches to fix 9 and the 7r—functions. However, there is a technical problem in that, one cannot, at this point, fully express the modified-Taylor series expansion in terms of standard local densities. Hence, we need one more layer of assumptions to realize the explicit connection. Both problems are solved at once by realizing that the basic quantity that one is re- ally interested in approximating is the energy density (or energy) at the Hartree-Fock- Bogoliubov level instead of the local densities. One starts from the exact expression for the energy, and approximates it by replacing the exact densities (local / nonlocal) With their counterparts as given by the modified-Taylor series expansion. Requiring th e reslilting expression to be a local EDF which fulfills various local and global sym— 74 n' ’Y".'. T..- ‘VV‘V- , . [f ‘ -‘-» «1. .,,. at .A- . v. metries results in several constraints relating the different 7r—functions. The required local or global symmetries can be rotational invariance, parity, particle number, time— reversal invariance, isospin invariance and local gauge invariance or its traditional counterpart, Galilean invariance. Naturally, the validity of some of the symmetries depends on the starting interaction. In order to make the procedures clear, these steps are applied to typical terms from the central and tensor exchange parts of the Hartree-Fock energy (of two nucleon interaction). Since we are interested only in the form of the expression, numerical coefficients and spin-isospin labels of the interaction are dropped. As discussed in section 9.6.1, a typical term from the central part of the interaction takes the form (@[VC|) = Z / at]? d? M.('r)pq(F1,F2)pq(F2,F1), (5.45) «1 Where [(12) is the Slater determinant HF wave function. Next, we apply the modified Taylor series expansion of the densities, use our assumption that the 7r—functions are independent of the orientation of F and perform angle integration f er, thereby obtaining <|Vo|) = 47.2 dedrVU") [(Hgm» p§(R)+ 012(0)) j.(R')-iq(127) 7.2 _3_ +—122Hfi(9 )HP(Q) Pq(R) (A9002) W270 )) — % 113(0) 115(0) p.02) 5.02)] . (546) w . . . . . here We have truncated terms contammg beyond-second order derlvatlves. Applymg th . . . . e Same set of steps to a typlcal term from the tensor 1nteract1on, ([VT|), we obtaln (|VT|) = ZdedFVT(r).§,(F1,F2)-.§‘q(ffz,F1), q 75 = 4nZ/(11’2'drv7.(r) [(ng(§2))2§q(1§).§,(1§) q .2 y _. 4 + :3- (H‘ilfllf Z qulequwrlR) 2 _, _. _. .. +{5 3(1) WEAR) MAR) -2 (1(3)) 7‘2 ~ —- -' -' -' _ 6’ H30") H30) .§},(R) - mm] , (5-47) with similar procedures being applicable to the remaining terms in the HF B energy. In order to make the explicit connection between the modified-Taylor series ex- pansion and PSA-DME (and / or other DME variants), we make use of the PSA-DME (and / or other DME variants) to approximate the Fock energy of time-reversal invari- ant systems and set it equal to the corresponding expression obtained from for the modified-Taylor series expansion. In this way, the 7r—functions and their arguments Q in Eqs. (5.46) and (5.47) that multiply the time—even densities can be fixed. Still, the TF—functions that multiply the time-odd, local and anomalous densities are not yet determined. This is where the relations between the 7r—functions through symmetry and other constraints come in to the picture. Constraints on the n—functions Requiring the 7r—functions to be independent of the orientation of 7" and the need to Obtain gauge (Galilean) invariant bilinear combinations of densities in the resulting EDF iIIlpose strong constrains on the 7r—functions. The gauge transformation of the one‘body local densities is discussed in appendix 9.3. These constraints are obtained by applying the modified Taylor series expansion to the various local and nonlocal densities that occur in the HF B energy of a finite-range two-body interaction and reqlflring the resulting EDF to be gauge invariant. Dropping the arguments of the 7r\ funCtions for ease of notation, the resulting constraints read 76 ”wa— L 'v M)? (i) H3015 + 115;): 2[n’;]2 and 115012;” + 113“) = 2[Hf]2 —o --pJ"_T.§‘. P.T_.§‘j (11) 7’0 7rl — T6771 and 7rl 71'0 — 7r07r1 0011811? = 11‘: H5 (iv) 3; 2 I13 where ”U is either 5,151,772) or 5:1(F1J'2) (v) 7r—functions of the pairing densities have to be real-valued functions. One of the most important qualities of the original DME of Negele and Vautherin is its exact treatment of unpolarized, symmetric INM at the Hartree-Fock level. Con- straining the 7r—function DME to reproduce INM limit of the direct and exchange parts of the energy density separately, we obtain the following two constraints on the 7r—functions («(02 = 1 (5.48) p p p j1(rk) 2 (NC)2 10 (1192 - 1 Where k must reduce to k} = [37rzpq]3 when one goes to INM. Thus either one has to use a fixed 19 = k}, or the parameter k should be such that it evolves to 1 k “+ k2. = [3w2pq] 3 as one goes to INM. F11I‘ther constraint is obtained using the idempotency of the density matrix to €XpreSS the particle number. I.e. from pg 2 p3, the particle number can be expressed as N. = Trpq=/d§df'lp(1?fl|2 (5.50) Thus a a constraint on the 7r—functions can be obtained by inserting the DME expres- 77 sion for pq(fi, F) 2 4 a N, = f deF Hampqua+m€+ —2'"—4H5(n>(qu-2TQ(R)) r2 p ~ 2 —EH3(9>Tq(R> = 47r [dédr [H3(Q)2pg(§) — 7.2 12 2 — % 113(9) Ham) pqu‘i) mfé) , (5.51) + 113(9) H§(Q)pq(§) (A qu-é) _ 2771”?» which is an integral constraint that can be utilized a posteriori to calculate some parameters. In the original DME and its variants, it can be shown that the num- ber constraint is satisfied exactly [202], while the PSA-DME breaks this constraint slightly. Finally, constraints on the 7r—function come from the large and small limits of r. The W—functions should go to zero for large 7' and for small 7', the modified Taylor series has to reduce to ordinary Taylor series expansion. In addition, we require the 7r ~functions to be such that the gradient, 6;, and the gradient squared, [57;]2, of the densities are reproduced exactly at 7" = 0. The resulting constraints are 110(0) = 111(0) = H2(0) = 113(0) = 1, 6(0) = H'1(0)=0, ng(0) = 0, (5.52) lim H0(r) = lim H1(r) = lim 112(7‘) 2 lim H3(r) = 0, (5.53) W . . here we have dropped the denSIty label on the 1r—functlons to denote that these CO IlStI‘aints hold for the 7r—functions of any density. Additionally, we used only II,- 78 which refers to the it“ 7r—function from nonlocal densities, even though the constraints are valid for 7r, (from local densities). The only exception is there is no 53 in the local case. In the DME of Ref. [172], the authors impose a local constraint = mi?) — 15(5), (5.54) F=0 4 — [mama] which is to be satisfied by the DME of pq(F1, F2). They refer to this constraint as the local imposition of the correct kinetic energy density. One recovers Eq. 5.54 by combining Eqs. (5.52), (9.108) and (9.109). It can easily be shown that 7r—functions satisfying the small 1' limits given in Eq. (5.52) satisfy this constraint. I.e. this particular constraint is a subset of the constraints listed in Eq. (5.52). Concluding, the explicit relationship that we established between the modified- TaYIor series expansion and PSA-DME (and / or other DME variants), together with the various constraints obtained through symmetry and other subsidiary conditions enable us to reduce the number of independent unknown w—functions significantly. Still, the number of unknown 1r—functions is larger than the number of constraint relations. Thus, the complete determination of all the 7r—functions requires further parametrization of some of the 7r—functions. In practice, it may not be possible to Satisfy all the relations among the 7r—functions and at the same time obtain a reasonable accuracy. In that case, some of the less stringent constraints have to be 1‘EIE‘J‘ied. Since this work is confined to the development of non-empirical EDF for timfireversal invariant systems, most, if not all, of these constraints are satisfied by de fault as one can simply choose the 7r—functions of the time-odd densities in such a “’3. y that they satisfy the constraints. 79 --.-..~~-_ t—Auu 5.3.7 Remarks on the DME of the local densities The apparent need for the DME of local densities can be seen from the Hartree con- tribution to the energy originating from the central part of a two-nucleon interaction. Reproducing the expression derived and discussed in section 9.6.1 ()4...) = / daczapqm) 12,/(a). (5.55) where for simplicity, we have dropped the singlet-triplet label of the interaction and numerical coefficients. Thus, if one requires a local EDF, one needs to approximate Eq. (5.55) utilizing a suitable DME expansion for the local densities. Equivalently, one can approximate the energy density, Eden, which is defined as 1 _. .. Eden 5 Z7? /dQ77/)Q(Ir1)pq’('r2)a (556) With f d9; referring to angular integration with respect to the orientation of F. In line with this, Negele and Vautherin, in Ref. [170], approximate the energy density given in Eq. (5.56) as - _. 35 . 1 - - 1 .. - goes ,5 pa ”9.1/(R)+Wnlrkr‘)[qu/(RlqulRl‘l'zpqulAPqKR) F 1" -0 -o —o 5mm) - vmm] , (5.57) by applying the expansion technique they devised for the nonlocal density, p(7"'1, 7‘32), to the product of the local densities, pq(f'1) pq,(1'"2). However, subsequent numerical teStS U183],[184]] indicated that the expansion of local densities is at the root of most of the error propagation and enhancement in self—consistent tests of the DME. This iS discussed in detail in section 5.4.6. Even though the DME makes no direct reference to the range of nonlocality, as 80 mentioned in section 5.2.1, the fact that the range of nonlocality with respect to F is very large for the local density, pq(F1/2) (where 771/2 means the argument can be 7‘} or F2), is mentioned to be the main reason why the DME does not work as accurately as it does for the nonlocal density, pq(F1,F2). In Ref. Bhaduri78, using a one-dimensional harmonic oscillator model with partial occupation of the single- particle states, the authors show that the nonlocality with respect to F of the local nucleon density, pq(F1/2), varies on the scale of the whole system, while the scale of the nonlocal nucleon density, pq(f'1, F2), is set by the local Fermi momentum k5. Even in the surface of nuclei, one can see that £071,152) falls off much faster than pq(7"’1)pq(f'2) in the relative coordinate, 1", by considering a one-dimensional surface With an exponential decay [183]. Le. by modeling the local density with a schematic eXponential decay function. Hence, the fact that both the local density and energy density involved in Eq. (5.57) have a large nonlocality scale with respect to F make the DME approximation inherently inaccurate, at least in one-dimensional problems. Nonetheless, in problems with dimensions greater than one, the simple charac- terization of the failure of the the DME of pq(F1/2) based on the scale of nonlocality needs refinement. This becomes obvious when one considers closed-form analytical 9X1)? ESSions for pq(f'1, F2) and pq(1"'1/2) in various model systems. In Ref. [208], a closed form eXpression for pq(1'"1, F2) is given for the case of an isotropic harmonically trapped ideal Fermi gas in any dimension. The more relevant expression is the one given in RBf- l185] for a three-dimensional oscillator with a smeared occupancy _, _, _. 1 1 + t pq(r1,7'2) = quR) exp [“1027‘21—_—t] a (5-58) pq(T1/2) = W(1 — t2) 3/2 eXp [—Q2r¥/2T_—t:| (5.59) Where t = 8—6 h”, ,3 is the inverse temperature, a2 = mw/h with the energy 8 = (N + 3/2) hw. Thus, one can argue that both the local and nonlocal nucleon densities are 81 y . Ullznl‘ (Jr‘ 4 V 1 1 . r ,. .‘ .1. it‘ ( I 1‘)" governed by comparable scales, relegating the supposed large scale of the nonlocality in F as an incomplete or limited explanation for the failure of the DME of pq(F1/2). The missing piece of the explanation can be identified once the DlVIE—coordinates are replaced in Eq.(5.59), viz, F1 /2 = 1:1" :1: 1 / 2 F. This makes the difference between the local and the nonlocal densities to be transparent. The nonlocal density falls- off exponentially (Gaussian fall-off) with respect to r independent of the orientation of F, which is also the case for pq(F1, F2) extracted from a converged self-consistent calculation of nuclei as shown in section 5.4. In contrast, the local density shows maximally different behaviors depending on the orientation of F. In short, sitting at a particular location in the nucleus, R, one can go to the surface or deep into the interior of the nucleus with the same 1" but different directions of F. Thus, the Significant dependence of pq(F1/2) on the orientation of F is partly responsible for the failure of the DME of pq(F1/2) as DMEs invariably average over the orientation of F. In Ref. [183], it is argued that the DME of the Hartree contribution can be avoided by treating it exactly, especially as the exact Hartree treatment does not result in 8. Significant increase in numerical complexity. In a related work, Ref. [184], the authors show that treating the Hartree contribution exactly removes most of the errors in the self-consistent numerical test of the DME. This and related issues are discusSed in section 5.4.6. Our numerical tests include both expanded and exact Hartree treatments. In the expanded case, the 7r—functions of pq(F1/2) are fixed by equ&ting Eq. (5.57) with the expression obtained from replacing pq(F1/2) in Eq. (5.56) with 1 F _. Dam/2) e «am pm) i «1(9) Viz/W?) + 57510) (— Var/1.15), (5.60) Ml‘h 82 . 1 .. y, . I l . n I 1 y . E ....u w u) .14 l, cl 1 . .) .nd. and truncating at second-order in the gradient , 2 7' ... _. ... a... s [75(9)] p.112»), [(1%) - —[ {(9)} qum) vp,)(R) +1—27T 7mm)[pq(R) )q(A/) I )+ pq /(R R)qu(R) , (5.61) which results in the 7r—functions 105 . 105 . 775(9) = mhh‘ki) , (5.64) Where (2 = rk‘}. The parameterized version of the 7r—functions for pq(F1/2) is inspired by the an- alytical form of pq(F1/2) in the three-dimensional harmonic oscillator with a smeared OCCuDancy [185] as given in Eq. (5.59). First, we fix the oscillator frequency w and the Oscillator length b according to the Blomqvist-Molinar formula [38] ha) = 45A‘1/3—25A‘2/3, (5.65) 197.33 b = —, 5.66 V940fiw ( ) where A is the mass number of the nucleon under consideration. The parameterized Tr"functions are given by p 0‘2 2 a4 4 —r2/b2 50(9) = (10+ —b— + 3 e , (5.67) WW?) = e‘TQ/bz, (5.68) 2 2 713(9) = err/b , (5.59) (5.70) 83 where the gradient corrections are damped with a gaussian of range b. From the short range limit of the 7r—~functions as given in Eqs. (5.52), the leading parameter a0 = 1. The rest of the parameters, viz, {(12, (14} are fixed by fitting the exact Eden as given by Eq. (5.56) with Eq. (5.61), with densities extracted from a converged self-consistent calculation of a selection of nuclear chains. A direct justification for the form of the 7r—functions given in Eqs. (5.67)-(5.69) comes from the fact that pq(a:) of spherical nuclei can be fit to a very good accuracy with the ansatz 11:4 III" _r2 2 p.99 = qu(0)an5;e (b , (5.71) n=l where pq(0) is the value of the central density and :1: stands for 1‘1, 7'2, 7' or R. The fact that p(0) is used instead of an additional free-parameter is due to there being local densities that play a similar role in the DME. Fig. 5.5 shows the neutron density Obtained from a converged self—consistent calculations of 48Cr and 208Pb and their Corresponding fit curves Our extensive tests show that the pq(F1/2) length scale b given in Eq. (5.66) remains uniformly valid and, perhaps not surprisingly, one can pr 0duce the same high-quality fits to almost all nuclei. However, the parameters show Strong Shell fluctuation, as can be seen from Fig. 5.6 which shows the parameters for CI‘ and Pb isotopic chains. The above discussions consider only the local nucleon density, pq(F1/2). The strong fluctuation of the other time-even local density, .Z,(F1/2), with respect to F due to its strong dependence on the shell structure of the particular nucleus under study, did “0‘ permit a systematic analytical study. Hence, for the numerical tests carried out in Section 5.4.6, a simple Taylor series approximation is used to fix the W—fllflCthnS of j1/2 7Tb = 7r? = 713?: 1, (5.72) Where the arguments must be extended in non time-reversal invariant systems as the 84 0.101 ‘ _ \ 3.0008. '5: . fl 4.30.06- : . G 3.10.041 5 g 1 0.021 0'000 ' 2 ' Z: ' 6‘50'2'4'6'8'1‘0 R [fm] R [fm] Figure 5.5: (Color online) pn(r) for 4807" and 208Pb from a converged self-consistent Calculation using Sly4 EDF. Parameters ‘ VVVO’O’O\ rro’ \ 2°5l —B— :12 u . \ 3 - \,/ \ 1.5] —_A- a‘ "+' as 0.54 *0" 4 1 +++_++_+_.,.+.+-+- ++-+-o--+-‘-\+-o--+-+-+-+-+-+-+-+-+'+'* l 'OOS“ ‘ J) \o—o—o—O-o-o-oor' H ‘ 1 ~ AH. . .1.5‘ , ° Whhh‘fl/ 2'0 - 2‘8 T316 ' 4'4 - 5'2 - 9.6 T 194 71127 120 - 118 I N N Figure 5.6: (Color online) The parameters for Cr and Pb isotopic chains obtained after fitting the neutron density, pn(F1/2), with the 7r—functions as given in Eqs. (5.67)-(5.69). 85 time-odd local densities jq(F1/2), .§’q(F1/2) , Tq(F1/2) and Rq(-F1/2) do not vanish. In this context, the discussion in section 5.3.6 is relevant. Finally, it should be mentioned that these expansions of the local densities can be avoided once the Hartree contribution to the HF energy from two-nucleon interaction is treated exactly. The situation is different and more complex for the HF contribution from the chiral EFT three-nucleon interaction at N2LO. However, with a particular choice of DME-coordinates, we avoid the expansion of local densities altogether. This is discussed in section 7.2. 5.3.8 Remarks on the DME of the anomalous densities Currently, there are several simple effective interactions that are being used in the pairing channel to perform HFB and related calculations. From a practical point 0f view, the simplicity of the effective interactions is necessitated by the numerical COmplexity that one would have to overcome in order to perform a 3D (deformed) HF B calculation for deformed nuclei. Still, the accuracy of current pairing part of C"Trent functionals calls for further improvements and constraints on the form and CDuplings of the functional [26]. Similar to what is being implemented in the case of particle-hole part, two com- plementary approaches, viz, phenomenological parameterizations and non-empirical construction of the functional are being undertaken [112]. The non-empirical ap— proach tries to address the role of the the bare NN + NNN interaction and their finite“ranges by successive addition of MBPT contributions in the pairing channel. Recently, the first step towards non-empirical pairing functional for nuclei has been taken in Ref. [112] where the first systematic calculation of pairing gaps in semi-magic mldei is carried out. By fixing the normal self-energr contributions with conventional Skyrme functionals, using low-momentum NN interaction and accounting only for the contribution of 130 partial wave to the pairing gaps, the results show that it is indeed the leading order (Bogoliubov diagram) that contributes the bulk of the pairing gaps 86 in finite nuclei. However, including N NN interaction in the treatment degrades the agreement between theoretical and experimental results leaving plenty of room for coupling to collective fluctuations. One caveat of these works is the fact that they are limited to spherical nuclei, due to the aforementioned numerical complexity to perform 3D HFB calculations. The nonlocal contribution to the total energy from the pairing part can be seen from the form of the leading order (Bogoliubov diagram). Reproducing the expression given in Eq. (9.272) for the spin-singlet, isospin-triplet channel from the central part of the interaction, (tn/31(5),... or Z / dfi ng V810) |,3,(7~1,7=‘2)|2, (5.73) q Where one notes that the nonlocal pairing density has the same role as the role of Pq(F1, F2) in the particle-hole part of the functional. From a formal point of view, the modified-Taylor series DME can be directly aPDIied to the anomalous densities, thereby approximating the exact leading-order nonlocal pairing functional with a local one. For instance, the nonlocal pairing density can be expanded as ~ -o 77 __, 7? ~ ~ —0 r F " -' ~ —o —o pQ(R It 5, R 3F E) % Hg(Q)pq(R) :l: Hf(fl)§ ' (V1 — V2) q(T1,T2) ., Tl=d2zR + ‘2‘ 112(9) 5 ' V1 + 5 V2 pq(T1, 7‘2) _, 1=®=R - 4 —o F 4 ~ --0 _' —H§(9) (5 ° V1) (5 ' V2) Pam) 7‘2) 1 (5-74) F1=72=R Where the 1r—functions are yet to be specified. However, both the technical and conceptual problems that need to be settled in 01‘der to obtain accurate 7r-functions for the anomalous densities seem to be signifi- cantly harder than for the normal densities. To start with, the size of the nonlocality 87 with regards to F of the nonlocal pairing density, 5,,(R :l: g , R IF ’5), in finite nuclei is still under discussion [[186],[187]]. It is commonly characterized by the coherence length, 6, of nucleonic Cooper pairs. Arguments based on infinite nuclear matter and the local density approximation (LDA) seem to suggest that f as given by Pippard’s relation [188] R2 A F __ 5.75 7rm*A’ (a ) f: is of the size of the nucleus. In Eq. (5.75), m" is the effective mass and A is the pairing gap. If the supposed large spatial extension of the nucleonic Cooper pairs in finite nuclei were to held without any modification, a DME approach which does not rely on any assumed short-range nonlocality needs to be invented. In fact, the stronger versions of these arguments stipulate that the existence of a small parameter ro/fi, Where r0 is the interaction range, implies that pairing in nuclei should be insensitive 130 the details of the nuclear interaction. This is supposed to justify phenomenological Paralneterizations using a local functional with no derivative corrections [189]. Practical calculations in finite nuclei paint a moderately different and favorable picture- Indeed, results from the recent non-empirical pairing calculations [112] sug- gest t311%?“ some details of the interaction may be important. Furthermore, the intuitive arguments that one builds starting from INM regarding such quantities as the size 0f the nucleon cooper pairs require precise qualification. In Ref. [187], the authors StUdY the neutron correlation length, 5,,(R), having defined it as fdr 1"4[)5,,(I_f,F)|2 fdrr2[5n(R,1:)[2’ MR) = (5.76) Where the subscript n denotes that the quantities are extracted for neutrons. The Strong position (density) dependence of the correlation length can be seen from Fig. 5.7. In addition, the authors extract [fin(R,7‘) I2 which is shown in Fig. 5.8. 88 14 F. 2115:“? 5.7: (Color online) Coherence length 5 (R) for 220, 600a, 6”Ni, 104572, 12oSn, 5 (From Pillet et. 51. [187]). 89 It can be seen that the nonlocality in F is in general in the range of 2 — 3 fm. In contrast, the correlation length is much larger except close to surface where it attains values in the range of 2 —— 3 f m. Nonetheless, the basic quantity that is approximated by the DME is [ 5,,(R, 1") [2, which reduces the significance of the large values of the coherence length. In Ref. [190], the authors conduct a related study of a slab of in- finite nuclear matter, confirming the smallness of the local correlation length at the surface and its largeness inside the slab. 8 A6 5.4 [I . 2 , . 0 . , 0246 024.3000“H r(fm) r(fm) mm Figure 5.8: (Color online) [5,,(R, 7‘)[2 calculated with HFB-D18 for 10487), 120811, 12sign. Scale has been multiplied by a factor of 106 (From Pillet et. al. [187]). Hence, these realistic studies of pairing in finite nuclei point to the possibility of developing a DME for the anomalous densities. In Ref. [191], a leading-order semi—classical expansion of the anomalous density based on the Thomas-Fermi ap- Pl‘oximation is given as 545.5) s Chases/245'). (5.77) Where jo denotes the spherical bessel function of order zero and 0 stands for constants Characterizing the pairing field strength and the local Fermi momentum. In Ref. [187], it is Shown that even at this level of approximation, there is a qualitative agreement 90 between the exact and the corresponding DME approximation. In Ref. [38], the Wigner-Kirkwood h—expansion of the Bloch propagator of a superfluid system is derived (up to 112). Leaving aside any further approximation that might be required, performing inverse laplace transform of the h—expanded Bloch propagator should recover the gradient corrections to the leading-order expression given in Eq. (5.77). In relation to the DME of anomalous densities, further works along these lines include working out the analytical expressions for the inverse laplace transform, recovering the expressions for the n—functions and extensive systematic accuracy tests. 5.4 Accuracy of DME The accuracy of a particular DME can be tested by comparing the exact density with its DME—approximation. But, our main objective is approximating the HF energy density and energy from two— and three-nucleon interactions. As discussed in section 6.1 for the two-nucleon case and section 7.1 for the three nucleon case, the HF energy expression involves a bilinear or trilinear combination of densities extracted from the OBDM. The numerical tests we conduct in this work concentrates on how well the DME approximates various contributions to the two—nucleon HF energy, in both non self-consistent and self-consistent HF calculations3. Prior to going in to the details, the following remarks are in order: (i) We consider only time-reversal invariant systems, (ii) Primarily, we concentrate on the DME of pq(F1, F2) and §q(F1,F2). In the two-nucleon HF energy, these nonlocal densities oc- cur in the Fock/exchange contribution from the central, tensor and spin-orbit pieces, While the picture is both different and significantly more complex in the case of three- mlCleon interaction (refer to section 7.2). The fact that we concentrate mainly on the Fock contributions is because the DME is inherently inaccurate for the Hartree con- 3 . . . _ In this section, we make repeated references to the HF energy from the generic ave-nucleon Intel‘action derived in the next chapter. 91 tribution from N N interactions [[l83],[l84]]. This is indeed confirmed in section 5.4.6, where we advocate treating the Hartree terms exactly. (iii) In a related note, the accuracy of the DME in reproducing the various contributions to the HF energy from three-nucleon interaction (chiral EFT three-nucleon interaction at N2LO) has not been tested. However, we use analytical approaches that ensure that the approxima- tions used in these calculations are exactly the same as the ones we use for the HF energy from N N interactions. (iv) Besides the two PSA-DMES (the analytical and parameterized), we include the original DME of Ref. [170] for the accuracy test, as all the other DMEs [ [171]- [173]] concentrate only on pq(Fl, F2) and give comparable accuracy. The three DMEs that we test in the following several sections are labeled as PSA- DME, PSA-DME—II and NV—DME. For PSA-DME, the 7r—functions are given in Eqs. (5.22), (5.23) and (5.36), PSA-DME—II (the parameterized version) is the one with w—functions as given in Eqs. (5.27), (5.28) and (5.39) and NV—DME refers to the original DME of Negele and Vautherin with w—functions as given in Eqs. (5.22) (the same as PSA), (5.29) and (5.32). 5.4.1 Inputs to non-self-consistent tests The generic form of the central, spin-orbit and tensor interactions in the different Spin-isospin channels are discussed in section 9.6.1, 9.6.2 and 9.6.3. The radial form factors in the present calculations are either a gaussian or a renormalized Yukawa according to Ref. [192]). Specifically we use ()0 e-'2/“2 , va(7) = (5.78) 329[(ra""7r"'erfc("—]71 — TA) — (T -> —T)] a 92 independently of the spin/isospin-singlet/triplet channel, (S, T), and with co = 50 MeV, a = 1.5 fm, m, = 0.7 fm’l. The momentum cut-off A is set equal to 2.1 fm‘1 while erfc is the complementary error function which is defined as e7'fc(;1:) = % foo (It (42. (5.79) It must be stressed that none of these interactions are realistic two-nucleon interac- tions, but rather schematic representatives. Still, they are reasonable form factors as the objective of this section is to gauge the accuracy of the DME variants against a reasonable reference point that is not itself meant to provide useful or realistic results. Finally, note that neutron density matrices and local densities used in the following sections have been obtained, for all semi-magic nuclei of interest, through spherical self-consistent EDF calculations employing the SLy4 EDF parameterizations with no pairing. 5.4.2 Fock contribution from VC In time—reversal invariant systems, the expression of the Fock contribution to the energ from the central part of the two-nucleon interaction contains a bilinear product of non-local matter densities as well as a bilinear product of non-local spin densities. Since the latter also appears as part of the tensor contribution to the Fock energy, we postpone the discussion regarding the spin-density product to section 5.4.3. Before comparing the Fock energy to its DME counterpart, we first conduct a m Ore stringent test 011 the energy density in which the integration over the angle of I“ has already been performed, i.e. we compare the integrand _. 1 _. _' _. ... C:n(R1T) E E[(iQrPralrlyr'z)0,.(7‘2J‘1) , (5.80) 93 to its DME counterpart —o 1'2 Ohm“? 1'2" 2 HP 2 1? 12 up I." n” ii 1A R nn ( J“) — 0( r") Pnl )Pn( ) + 0( Fr) 2( F’)Pn( ) Pn( ) 3 4 .. 3 _. - 7.71(1?) + 3A}1‘2p12(R)) 7 (5'81) where the latter depends on which variant of the DME has been adopted. We de- note such integrands as energy densities throughout this section. Strictly speaking, it is necessary to multiply them by the interaction to obtain the dimension of an energy density. Still, we postpone the folding with the interaction to the second mea- sure introduced below. In the definition of CngMWR, r), we have truncated terms with beyond-second order gradients, in line with current phenomenological imple- mentations of Skyrme EDF. In addition, a consistent account of such fourth-order derivatives in the EDF would require to go also to fourth order in the DME itself, which is not addressed in this work. This is an important point that underlines our philosophy that the primary purpose of the DME method is not to reproduce the fine details of the OBDM, but rather to reproduce as best as possible the energy density and the total energy at a given order in the expansion. The latter two are precisely what are gauged in this work, whereas no tests dedicated to the reproduction of the OBDM by itself are performed. The parameters of PSA-DME—II 7r—functions of the nonlocal matter density, as. defined by Eq. (5.27), read a = —0.4130, )3 = 1.2430. (5.82) We obtained these values fitting Eq. (5.80) with Eq. (5.81) using densities extracted fiOIn a converged self-consistent calculations of a selected set of nuclei. 94 NV-DME PSA-DME PSA-DME-II Exact 9 E6 L‘. I"3 0 M 7“) mi Mi 0 3 6 3 6 9 9 0 3 R [fm] R [fm] R [fm] OUQO — NV-DME — — PSA-DME ' " ‘ PSA-DME-II R = 3.0[fm] \ R = 6.0[fm] \ / ' o i i 3 (1 i 2 3 r [fm] r [fm] Figure 5.9: (Color online) Comparison of 05,,(R, r) and thDM E (R, r) where the latter' lS computed from the 7r- —functions of one of the three DMEs: NV~DME, PSA-DME or PSA-DME—II. Upper panels. two-dimensional integrands. Lower Pa-nels: ratios of CF51) M E (R 7*) over 05,,(R, 7‘) for fixed values of R. Densities are 0h tained from a self-consistent EDF calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole part and no pairing. 95 Figure 5.9 shows " that all the three DMEs provide comparably good profile- reproduction of the integrand C5,(R.7‘) within the typical range of nuclear inter- actions (7‘ ~ 2 fm). Beyond such a non locality, the quality of the reproduction deteriorates significantly, with that of PSA-DME deteriorating slightly faster. In ad- dition, one sees from the lower panels of Fig. 5.9 that the quality of the reproduction decreases as one goes to the nuclear surface, i.e. for R Z 4 fm. This is slightly improved by taking into account the diffusivity of the local momentum distribution when designing the PSA-DME for the scalar part of the OBDM, as shown by the better accuracy of PSA-DME-II. In general, PSA-DME—II stays much more close to one than the other two DMEs, with slight overestimation in the range of nuclear inter- actions. Note also that, although the plots are provided for two sample nuclei, more systematic tests have been performed over several semi-magic isotonic and isotopic chains that support such conclusions. Coming to the energy itself, i.e. to the integrated product of the interaction 110(7‘) with the central energy density, we compare Eghm] : 47r/dR (17‘ 7‘2 11(7(r) GEAR, T) , (5.83) Eg’mwhm] = 4Tr/dR (17‘7‘2 ’UC(T‘) CHF,’PME(R, 7'). (5.84) Figure 5.10 shows the relative error obtained from the three DME variants com- pared to the exact Fock contribution for both the Gaussian and the renormalized- Yukawa radial form factors and for three semi-magic isotopic chains. Let us start with Fig. 5.11 that shows that the dependence of the accuracy on the range of the (Gaussian) interaction used is significant, i.e. about a factor of 1.5 hatween a = 1.0 fm and a = 1.5 fm. As can be expected from the two—dimensional 4 . . . . . . . . .~ « Note that for semi-maglc spherical nuclei used 1n this work, the energy densities C,’,,,(R. r) and F. - , _. CrinD M E (R, 1‘) only depend on the magmtude of R. 96 Gaussian + NV-DME - - — - Yukawa + NV-DME ------------ Gaussian + PSA-DME -—---— Yukawa + PSA-DME ------- Gaussian + PSA-DME-II -—-- Yukawa + PSA-DME-II 8.0- In: e 6.0 L: a.) o O 3: 4 0 Cr [ N1 E z‘\ "x 8 I, \\ \\~ I.- 2 0‘ I ‘\ ’/’~\\\ ] ‘\\ ’ a: \\/’ ~ \‘v” 0.0 .1 .. .............. 18 2'6 3'4 4'2 5'0 5'8 24 3'2 4'0 48 N N I- o In In (a 35 a 4.0+ Sn * Pb H 5 ’,/”‘\‘ 2 2.01 ‘~~~--- ,x— - -_/ \~——- 5: a,” .............. 0.0 ----- \~~‘p' """ -_.-T 48 5'6 6'4 7'2 8'0 8'8 94 102 1i0 lfii 126 134 N N Figure 5.10: (Color online) Percentage error of EgM E [7171] compared to Eg[7m], Where the former is computed from: NV—DME, PSA-DME or PSA-DME—II Y\‘functions. Densities are obtained from self-consistent EDF calculations using the SLY 4 Skyrme EDF in the particle—hole channel and no pairing. ' 97 a = 1.0[fm] + NV-DME — - — -a = 1.5[fm] + NV-DME ------------ a = 1.0[fm] + PSA-DME --‘--- a = 1.5[fm] + PSA-DME ------- a = 1.0[fm] + PSA-DME-II - ---a = 1.5[fm] + PSA-DME-II 6.01 h 8 4.01 l- a Q 4 if a 2.01 Cr 0 G) In 0 m ”"’\\ 'a—. 0.0 /l’ ......... “A\\ ’IA””__---______::~- I” ............ :.::,_~’{:— ---------- 18"'2'6fi1'3'4" '4'2"'5'o"'5's' N Figure 5.11: (Color online) The same as Figure 5.10 but for two different values of the range of the Gaussian interaction. 98 ;,», 6r \ . 1 ‘iav D ~41 ' 1 I29 6 1 V ‘0. -"T‘ if] k '1 a ‘v Y Anlfll: v7.“ .10, o. :J—a. 'II' 316-- arm 1 density profiles in Fig. 5.9, the accuracy decreases as the range of interaction increases. which holds for all available DME techniques [[170]-[173]]. This stresses that the local quasi-separability of the OBDM with respect to 7" and if underlining the DME, which is exact in IN M, deteriorates with increasing non-locality 'r in finite nuclei. As long as the hypothesis of quasi-separability is well realized within the range of the interaction, the DME can be quantitatively successful. On average, the error obtained with PSA-DME and NV—DME are similar as can be seen in Fig. 5.10, i.e. about 6—8% for the three isotopic chains and for both for the Gaussian and the renormalized-Yukawa interactions. While PSA-DME—II gives a better accuracy with a percentage error between —0.5 and 2. Similar improvement over that of Ref. [170] is reported in Refs.[[171],[172]]. PSA-DME—II, while it is much better than PSA-DME and NV-DME, it shows a gradual drift of the sign of the error, from overestimation to underestimation of the exact value, as the interaction range increases. 5.4.3 Fock contribution from VT We now turn to the Fock contribution coming from the tensor part of the NN inter- action. Such a contribution involves bilinear products of non—local spin densities As a matter of fact, two terms with different analytical structures emerge such that the exchange tensor energy-density reads ram-‘2'») a T.f;,.<é,r>+T.§.,2<fi,r>. (5.85) . _. 1 T,f,,,1(Rr) : 117? er§u(f1,fi2)-§,,(f'2,7'~‘1), (5.86) "' 1 rrru ... —. T512(R9r) E a; er Z :3 8,14,01,73) . W X Sun/(7:727 F1) , (5.87) 99 . . . u 1 .K c I .4 .‘v, ’- ~A..L’~. 2.. ‘ 1 {4" "r4 ._.. . 1 4 where T,f;,J(R, I) also appear in the central contribution to the Fock energy. The two DME counterparts, which eventually depend on which variants of the DME is being adopted, read .2 ——[I1f(k;tv >121 ....(V R)Jn,,...(fi). 11 11:1: INF i:(J..,(,..12).]...(1‘2’)+J..,,.,.(I"2')J..W(R‘) TF,DA!E(1'§, 7') nn,l 7'2 F. 11E “’ Tn? (RJ') 71 , +Jn,pu(R)J nun (2)111) and reduce for spherical systems to ”I nn,1 TF ”Mai, r) —% [11f (12:2.r)]2j;,(1‘2’) - HR), (5.88) “I o TF‘DA'!E(1§3 7‘) 11712 (5.89) One recovers a pattern which is seen when deriving the empirical Skyrme EDF from an auxiliary Skyrme effective interaction. That is, the central part of the inter- action only produces the so-called symmetric bilinear tensor terms proportional to Jn,p,,(R) Jn,,w(R) while Tfigh I E (R, r) that contains asymmetric bilinear tensor terms proportional to Jn,,w(R) Jn,.,,,(R) solely comes from the tensor interaction [196]. This can be easily traced back to the spin-space coupling that characterizes the tensor operator. Since the numerical tests are presently carried out for spherical systems, we are only concerned withT m, F1(R, 7') andT TF’DME(R, 7'). For spin-unsaturated nu- clei, Tan1(R,r) IS highly localized around the nuclear surface as seen in Fig. 5.12 for 208Pb. The same figure shows the progressive and significant improvement that the PSA approach brings to the DME of the vector part of the OBDM. This is realized in both PSA-DME and PSA-DME—II. The optimal parameters that we obtained for 100 0 7.. IL ..I. 0. q 1.. l 1 l hlllllllll I 9 40 xi. 0 9 6 3 Av .v-al... ... . A- ~.v3¥..¢~\w-\CA~ ......fi .. ... ... ...... . ...-b- .- -=h— .- u -\ nut “L1 ..4 at. ....H 51 .. INM-DME PSA-DME 9 ‘6 3 NV-DME 0 ------------ INM-DME 3 6 9 0 3 6 9 """" ' R [rm] R [rm] PSA DME _ _ _ — PSA-DME-II ‘g 1.5 m R = 4.0[fm] 2.0 - I \ E/Exact I-1 H e in M P . u- D P e r [fm] 1' [fm] Figure 5.12: (Color online) Comparison of Tnle(R, r) and T33 E (R, 7") where the latter is computed from NV—DME, PSA-DME, PSA-DME—II or from PSA-DME with P;(R) = 0 which we denote as INM-DME. Upper panels: two-dimensional integrands. Lower panels: ratios of T333 E (R, 1*) over Tfl‘fiR, r) for fixed values of R. Densities are obtained from a converged self-consistent calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole channel and no pairing. 101 PSA-DME—II of £1071, 1:72), as defined by Eq. (5.39), read m = 2.543, b = —0.0799. (5.90) Within the typical range of nuclear-interactions, N V—DME falls off much faster than both PSA-DMES. Less importantly, NV—DME also introduces artificial and pro- nounced structures in a region that corresponds to the tail of the interaction. Both of these drawbacks are rectified progressively by the PSA-DME approach. While most of the improvement is already brought by the spherical PSA (P2(R) = 0, which is the same for both PSA-DME and PSA—DME—II), an even better accuracy is obtained by incorporating the quadrupolar deformation P2(R) of the local momentum Fermi distribution. The overestimation of T5,,1(R, r) at very small 7' seen for all DMEs in the lower panels of Fig. 5.12 corresponds to a region where the integrand is small and where its weight is further reduced in the integrated energy by the r2 phase-space factor. Coming to the energy itself, i.e. to the integrated product of the interaction 'vT(r) with the tensor energy density, we compare Eflnn] = 47r/dR dr 1‘2 vT(r) T,f;,(R, r), (5.91) Eg’DMEhm] = 47r/dR dr 7‘2 117(7) 1:3,ME(R, r) , (5.92) which for spherical nuclei reduce to the contribution from T5“ and T5,?” Figure 5.13 shows the relative error of the three DMEs compared to the exact Fock contribution, for both the Gaussian and the renormalized-Yukawa radial form factors and for three semi-magic isotopic chains. For both types of interaction, the percentage error of N V- DME easily reaches 40%. This is in contrast to PSA-DME and PSA-DME—II whose percentage errors are typically within 21:10% for most parts of the three isotopic 102 Gaussian + NV-DME - - —- Yukawa + NV-DME ------------ Gaussian + PSA-DME ------ Yukawa + PSA-DME ------- Gaussian + PSA-DME-II — - -- Yukawa + PSA-DME-II 60.0- I- 2 40.0 I- 0 q, J M 3 1: 20.0J 0 O l- 0 a. l 0.0- 18 “ 60.01 G l- 8 0 4000 no 3 :1 8 20.0( In 0 n- 0.01 48 56 64 72 80 88 94 102 110 118 126 134 N N Figure 5.13: (Color online) Percentage error of EgMEhm] compared to E; [7m] where the former is either computed from: NV-DME, from PSA-DME or PSA-DME—II. Densities are obtained from self-consistent EDF calculations using the SLy4 Skyrme EDF in the particle-hole channel and no pairing. Notice the different vertical scale compared to Fig. 5.10. 103 chains. This can be traced to the fact that, while both NV-DME and PSA-DMEs overestimate the reference quantity for small 7“ (typically less than 1 fm), NV-DME decreases much faster with 1', thereby overcompensating for its initial overestimation. In contrast, PSA-DMES stays close to the exact value for a much larger range of 7‘ values There exist short sequences of isotopes for which the percentage error shows a considerable increase. The fact that all three DMES display such a feature suggests that the problem is independent of the specific form of the 11:7 function used. To iden- tify the source of the problem, Fig. 5.14 shows T "11(R, r) for three nuclei displaying a sudden loss of accuracy. One notices that T F (R, 1') extends over larger intervals nn,1 in R and 1" than for 20st (see Fig. 5.12). This corresponds to the fact that the selected nuclei are nearly spin-saturated and generate very small Eff: [nn] in absolute value, as seen from the lower panels of Fig. 5.14. As a result, the relative inaccuracy of any DME becomes large and the percentage error increases suddenly. Of course, the resulting error in the total EDF remains very small as the corresponding tensor contribution is anyway negligible, i.e. the local spin-orbit density J;(R) is close to zero in nearly spin-saturated nuclei. Therefore the sudden losses of relative accuracy are not as worrying as Fig. 5.13 initially suggests. In conclusion, the use of PSA techniques has allowed us to bring the DME appli- cable to the bilinear product of non-local spin densities on the same level of accuracy as for terms depending on the scalar part of the OBDM. One could certainly work even harder to bring down the overall DME percentage error. This could be achieved by (i) allowing for additional parameters in the H—functions to be optimized on a set of reference calculations. However, our extensive optimization of the two-parameters of Hf have convinced us that one cannot remove the sudden loss of relative accuracy discussed above for spin-saturated nuclei. As already stated, this is not a problem in the end as the corresponding contribution to the energy is negligible anyway. (ii) One 104 Exact - 64Cr Exact - 114Sn 20 28 36 44 52 60 Sn 40_0 . — Gaussian £5. , """ Yukawa ‘ V20.0 l 50 58 66 74 82 96 104 112 120 128 N N Figure 5.14: (Color online) A few representative nuclei with diffuse T5,,1(R, 7‘) together with absolute E; [rm] for the corresponding isotopic chains. Densities are obtained from a self-consistent EDF calculation using the SLy4 Skyrme functional in the particle-hole part and no pairing. 105 can go to higher orders in the DME, consistently for both the scalar and the vector parts of the OBDM. This should however be done within the frame of the generalized Skyrme EDF proposed in Ref. [35]. 5.4.4 Fock contribution from VLS Basic analysis We now turn to the spin-orbit contribution to the F ock energy. Unlike for the central and tensor forces, such a contribution involves both the scalar and the vector parts of the OBDM. In this case, we first compare the spin-orbit energy density 2' Lsf,(1'2',r) = Z1; d9.§,.(fi,52).rxmaxi-8,51), (5.93) to its DME counterpart 4 1 - Z . - - LSS.D“’E = gflflkiidd Z 6“"‘3anuavg(nspn<8)). 11.u.;‘3=:r which eventually depends on which variants of the DME is being adopted and that reduces for spherical systems to , -‘ 1 ~ -0 -v -¢ _. LSS.D"’E(R,1~) = 6 113% ~24) r2Jn(R) - V11 (”Sumpam) ° (5.94) Note that we have truncated terms with more than two gradients in LSfif’fiR, r). The numerical tests shown in the present section actually use PSA-DME only (no PSA-DME—II) with the quadrupolar deformation parameter set to zero (PQ"(R) = 0). we still label the results as PSA-DME. It will be seen that these simplifying choices have no bearing on the discussion at hand. Figure 5.15 shows that PSA-DME significantly overestimates (in absolute values) 106 NV-DME PSA-DME J11 '. 9 ‘ 9 E“ 6 1.13. 5 3 0 . ~ . ~ 0 0 3 6 9 0 R[fm] 2.0] 1.5“ —NV-DME ——PSA-DIVIE DME/ Exact rd ’9 P u- 0.0 . . . 3.0 A y 2.5 1 I \ t 92.0 I \ DME/Exa r-I 1-1 c u: l I I l \ \ \\ B / / I l I l \ \ \ 0 5, R = 6.0[fm j R = 6.5[fm] 0.0 - . . . . . 0 l 2 3 0 1 2 3 1' [fm] 1' [fm] Figure 5.15: (Color online) Comparison of LS:,,(R, r) and LS:;,DME(R, 'r) where the latter is computed from either NV-DME or PSA-DME. Upper panels: _, tWo-dimensional integrands. Lower panels: ratios of LSEKWR, r) over LS:,,(R, r) f0r fixed values of R. Densities are obtained from a converged self—consistent Calculation of 208Pb with the SLy4 Skyrme EDF in the particle—hole channel and no Pairing. 107 the maximum peak of LS:,,(R,7') at the nuclear surface. In addition, oscillations at larger 1', i.e. in the tail of the two-nucleon interaction, are not captured by PSA— DME. In contrast, NV-DME reproduces relatively well the density profile Lan(R, r), in particular as for the main peak at the nuclear surface. This suggests that the significant improvement for PSA-DME over NV—DME as to reproducing the tensor energy density does not carry over to the spin-orbit energy density. The previous assertions are supported by tests carried over several isotonic and isotopic chains. Looking for possible improvements, we tested that including truncated higher-order terms associated with the action of 618' on (1 /4Apn — Tn + 3/5k}2pn), when going from Eq. 5.93 to 5.94, does not improve the accuracy of PSA-DME. Coming to the energx itself, i.e. to the integrated product of the interaction 111,5(7') with the spin-orbit energy density, we compare EELS-[7171] = 47r/derr2 vL3(7')Lan(R,1‘) , (5.95) 1111 Efglqnn] = 47r/derr217L3(r)r2LSDME(R,r) . (5.96) Figure 5.16 shows the percentage error obtained for three isotopic chains. In agree- ment with the analysis done for the spin-orbit energy density, the percentage error of PSA-DME is impractically large and negative, in the range of -15% to -50% for the two schematic interactions used. In contrast, NV-DME provides a much better accuracy with percentage errors within :1: 10% for most studied isotopes. Last but not least, one notes that the spikes in the percentage errors already discussed in section 5.4.3 arise for the same isotopes and relate to the vanishing non-local spin density in near spin-saturated nuclei. 108 Gaussian + NV-DME — - — - Yukawa + NV-DME ------------ Gaussian + INM-DME ~----- Yukawa + INM-DME ._. 25.0- 2 ,4- r. s 0 0 I ‘\‘ I \'7—‘-——h.. >\ If; at) / \\ ' \\\ 7 E -25.0- ~. \J Ct ........ ' I """"" Niki.“ 2 I .\_ .4 ............. \_\ ........... , \ _____ o I \. --.,~ ......... \.\ 7 \ , °" -50 0 ' . h: ’ f . \. . '''''' I . ' 18 26 34 42 so 58 24 32 40 48 N N 1... 25.0' 1 "s 2 /’ \-\ g _\_ ’ _-- -~- O 0'0 ... \/ x—Q‘ 393,111 ‘r\ Pb u \'-, _. \ 5 -25.0- ........... -’ YES“ l I?" ~ ---------------- U “I | '. x \ ....... In x ................. -~. _ 0 -—-~ »' ‘ ~. ------- A. -50 0 . . ._._'._._._'. - . t__._.'_._.- . ° 48 56 64 72 80 88 94 102 110 118 126 134 N N Figure 5.16: (Color online) Percentage error of EESMEUm] compared to Efs[nn] where the latter is either computed from NV—DME or from PSA-DME. Densities are obtained from self—consistent EDF calculations using the SLy4 Skyrme EDF in the particle-hole channel and no pairing. Notice the different vertical scale compared to Figs. 5.10 and 5.13. 109 Further investigation of the spin-orbit exchange The results of the previous section show that N V—DME is better suited than PSA- DME to reproduce the spin-orbit contribution to the Fock energy. This can be c011- founding in light of the better accuracy obtained using PSA-DME to reproduce the tensor contribution to the Fock energy. We can infer from Fig. 5.12 that NV—DME underestimates the main peak of the nonlocal spin density while the latter is well cap- tured by PSA-DME. It is thus puzzling to find the opposite for the Fock spin-orbit energy density. In the following we employ a toy model of the OBDM of finite nuclei to show that this is due to a fortuitous cancelation of errors. Having already a handle on the non-local spin density §q(F1, F2), we focus on the term it multiplies in the spin-orbit energy density, i.e. 7" x 62pq(771,772), which we first approximate by 7" x 61'? pq(7"'1, F2) thanks to the weak dependence of the non-local matter density on the orientation of 7" [176]. Hence, and focusing arbitrarily on neutrons, we want to compare the two quantities GE 2 612014117?) 3 (5'97) V8 (118(kpr)pn(§)) , (5-98) GDME = where the latter is independent of whether NV—DME or PSA-DME is used. To do so, we employ the toy model we discussed in section where the expressions for the local and nonlocal neutron densities are given by Eqs. (5.58) and (5.59). From these equations, one easily obtains 6w}? + g, 8‘ — g) = exp[—1/4a2r21———i:] [mm] (5.99) " " 405—1/21a2 2____ 110 The corresponding PSA-DME reads .. ,1. .. F j Mr 7'2 1+t . 2 n .. ,1; such that, given the definition of WAR), one can easily obtain V711 [1W ‘2‘7‘)p(§)] = 10( ’2‘7')6§pn(R) (5.102) and show that " GDME( 9, F) . [ 2 21+t] Gnu-0 R,r E _, = "17‘ ex 1 4a 7‘ — . ‘( 9 GE(R,f) 30(F) P / 1—t In order to study Gratio quantitatively, we fix the inverse oscillator length, (1, using the Blomqvist and Molinari formula, i.e. a = (0.90/11/3 + 0.70). In subsequent discussions, we take reasonable combinations of A and N although we show that the conclusions of the present section are independent of the actual value of A. Before analyzing the behavior of Gmtio(R,1—‘), it is worth noticing that the toy nonlocal matter density is exactly separable in relative and center-of-mass coordi- nates. Such a separability being one inherent, usually only approximate, aspect of the DME, we expect the latter to work well in the present case [185]. Computing the same ratio as in Gran-(,(R, 7‘) without the gradient operators, we do indeed obtain the good performance of the DME as is visible in Fig. 5.17. Note in particular that the ratio is independent of the value of R. Such a result proves that the toy model provides a situation comparable to the one studied in Sec. 5.4.2, i.e. the DME of the scalar part of the density matrix performs well. Such a performance sets the stage in View of qualifying the results obtained below for Gran-(,(R, 7"). In order to identify the short distance behavior of Gran-AR, 7"), we perform a Taylor 111 'A=120 N=80 .\ j 'A=200 N=120 A = 120 N = 60 g 4.0 ‘ R = 0.0 [fm] a ----- R = 2.5 [fm] \ --—-R=4.9[fn1] E 2.01 G \ 0.0 - - - A = 200 N = 100 l g 4.0! R = 0.0 [fm] g3 ----- R = 2.9 [fm] \ —-—— R=5.8 [fm] E” 2.01 a 0.0 a - \ 0.0 1.0 2.0 0.0 r [fm] k \ 110 2.0 r [fm] Figure 5.17: (Color online) Ratio of the DME (Eq.(5.101)) over the exact (Eq.(5.58)) expressions of the toy nonlocal matter density. 112 series expansion in 7' (5.103) _. kn? ,‘2 1 t Gratio(Refl % 1 + (_ [l 1:2)72 6 4(1—t) Looking close to the surface of the nucleus, one can neglect k}? / 6 in comparison with the second term of Eq. (5.103). Defining Germ,.(R,f) E Gmti0(R,f') — 1, one obtains 4 .2 G.,.....(R,F) a» (0.326 — 0—03) 7'2. (5.104) 711/3 Equation 5.104 is valid around the nuclear surface. Inside the nucleus, one cannot neglect the first term 162:2 / 6 of Eq. (5.103). This is irrelevant as the spin-orbit energy density is concentrated around the nuclear surface. Figure 5.18 bears our expectation i.e. overestimation of GE by GDME around the nuclear surface for a wide range of R, A and N values. It can also be seen that there is a gradual and systematic shift from slight underestimation to overestimation as one moves from inside the nucleus to the nuclear surface. Keeping the results shown in Fig. 5.17 as a reference, we conclude that the applica- tion of the gradient operator on the scalar part of the density matrix deteriorates the quality of the DME that overestimates the exact results, in particular as one goes to the surface of the nucleus where the exchange spin-orbit energy density is maximum. Combined with the good approximation of the vector part of the density matrix, such a semi-quantitative analysis explains the overall overestimation (in absolute value) of the exchange spin-orbit energy provided by PSA-DME (see Fig. 5.16). Contrarily, the underestimation of the vector part of the density matrix by N V—DME provides a fortuitous, but rather accurate, cancelation of errors such that the nonlocal spin-orbit energy density is much better reproduced overall (see Fig. 5.16). Even though we can be satisfied with such a situation in the short term future and advocate the use of the NV-DME variant for the spin-orbit contribution to the Fock energy, it would be 113 A = 120 N = 60 R=0.0[fm] A=120 N=80 22.0 ----- R = 2.6 [fm] ’ ‘a —-—R = 5.1[fm] ‘/' ‘/,, Ul- _____/’ "‘/ 0.0 - A = 200 N = 100 R = 0.0 [fm] A = 200 N = 120 32.0 ----- R = 3.0 [fm] / I a? —-—R=6.1 [fm] ,/’ ‘/’ L51- _‘/.-% ’_/—/ 0.0 0.0 1.0 2.0 0.0 1.0 2.0 r [fm] 1' [fm] Figure 5.18: (Color online) Gmm(R, F) as a function of r for a selected set of (1%, A, N). 114 more satisfying 011 the long run to design a suitable DME for the gradient of the scalar part of the density matrix that can be combined with the improved PSA-DME for the vector part.This loss of accuracy for the spin-orbit part does not have any impact on the application of the DME to the HF energy from chiral EFT NN + N NN interaction at N2LO. This is due to the fact that the N N spin-orbit interaction that we have at N2LO is zero-range/contact, thus does not require the application of the DME. In the NNN case, the problem does not seem to be relevant. Refer to section 7.2 for details. 5.4.5 Hartree contribution from VC, VLS and VT The numerical results given in section 5.4.6 confirm the View that the DME should be applied only to the Fock part of the HF energy. However, for the sake of completeness, we gauge the accuracy of the DME when applied to Hartree contributions. As shown in the next chapter, the Hartree contribution from the tensor part of the two-nucleon interaction vanishes for time—reversal invariant systems. For central and spin-orbit parts, the exact integrands for the profile comparison are Craig: T) E 41—17;]er pn(Fl) pn(7-“2) v (5.105) _. 1 _, LSEH(R, 7') E :1"; fdflr P1471) F ' 41,1072) , (5.106) while the corresponding DME expressions read 0.52”“ .3 [.g(6)]2p,(é)pq,(a_’_[.((m]25p,(é).epq,(5) 2 -O —O + {inseam pq(R)Ap,/(R) + pawn/2(8)] . (5107) 1.3.7.0“ z gri’mfiwvmfi). (5.108) where the 7r—functions in Eq. (5.107) are fixed in two ways: (i) The first set (NV- DME) are given in Eqs. (5.62)-(5.64). (ii) The second set consists of the parameterized functions given in Eqs. (5.67)-(5.69) and whose optimized parameters are: a2 = 0.850 and (1.1 = 0.3000. Eq. (5.108) results after fixing the 7r—functions according to Eq. (5.72) (simple Taylor series expansion). The corresponding integrated energy contributions are given as Eg[nn] = 47r dedrr2 210(T)C,g,(R,r), (5.109) Eg’DME[nn] = 47r dedrr2 120(1‘) CflDME(R,r), (5.110) Efshm] = 47r dedrr2210(r) Lan(R,r), (5.111) EggDMEMn] = 47r dedrr2 110(7')LS£I,’,DME(R,T), (5.112) Fig. 5.19 shows that for large 1' values, Eq. (5.107) does not reproduce the correct profile of Eq. (5.81) for both sets of 7r—functions. The two DMEs (sets of 7r—functions) have opposite effect in that region. In contrast, one can achieve an accurate repro- duction of the integrated contribution, Eq. (5.109), with Eq. (5.110), when the range of the interaction is short. Furthermore, the plots contrast the accuracy of the two sets of 7r—functions, with the parameterized version performing significantly better as the range of the interaction increases, though with a decreasing overall accuracy. Even though this decrease in accuracy is a general trend for all DME approximated quantities, the deterioration is more significant in this case than the Fock contribu- tions. Perhaps this is due to the wrong prediction of the exact profile (Eq. (5.81)) with Eq. (5.81). Fig. 5.20 compares the profile of the exact Hartree contribution from the spin-orbit part of the interaction to its DME approximation. One can see the DME fails to properly capture the profile, with a very large error resulting in the integrated contributions. 116 Percentage error 10.0 ‘ ————————— -— " a = 1.0 [fm] + NV-DME -20.0‘ 1 .17:- ’ '1 ----------- ------------ a = 1.5 [fm] + NV-DME 0.0 , ------------------- Yukawa + NV-DME -10.0* Cr ———- a=[l.0 fm]+PI-DME ..... — a = [1.5 fm] + PI-DME — — — — Yukawa + PI-DME ..— ——- ___________ .................. -—----——-— ----- 18" 2'6 3'4" ' '4'2' 5'0 58 N Figure 5.19: (Color online) Percentage error of Eg’DME[nn] with respect to Eghm] for Cr isotopic chain. The upper plots show 03,, and CflDME for NV—DME and the parameterized 7r—function which we call PI-DME. 117 - Exact NV-DME 10 ' 10 I [fill] R [fin] R [fin] Figure 5.20: (Color online) L35... and L5,?“ for NV-DME, with densities obtaiend from a converged self-consistent calculation of 208Pb with the SLy4 Skyrme EDF in the particle-hole channel and no pairing. 5.4.6 Preliminary self-consistent tests In practical terms, one of the important benefits of the DME approximations is the fact that existing Skyrme HFB codes require minimal modifications to be used with EDFs obtained from the DME. Fig. 5.21 contrasts how a code is implemented for Skyrme HFB against one for the DME based functional. As can be seen the main change is in replacing the eventually constant Skyrme couplings, with density— dependent couplings obtained from the DME. We carried out a limited set of self-consistent test to gauge the accuracy of the DME, in both full- and exchange-only-DME. In full—DME, both the Hartree and Fock contributions are approximated with the DME, while in exchange-only—DME, only the Fock contributions are approximated with the DME while the Hartree ones are treated exactly. In addition to confining the test to time-reversal invariant and spherically symmetric systems, there are several simplifying choices that we made. 118 Roam Potentials Kohn—Sham Potentifls A A Skyrme DME energy HFB energy HFB t0, t1: tV Alp], BUM Orbital: lid Occupation t’e Orbital. and Occupation I's (a) Skyrme-HFB code. (b) DME-based HFB code. Figure 5.21: Comparison of Skyrme HFB and DME—based HFB codes. These are e We used the Brink-Boeker [193] force which has only a central component with gaussian form factors. The actual form and parameter values are 4 2 2 Ve(r) = Z a,- e—(T12/“al +b,-e‘(’"12/“b> , (5.113) i=1 with pa = 0.7fm and ab = 1.4fm. The magnitude and range parameters in the four spin-isospin channels read Table 5.3: The Brink-Boeker force(B1) 113113 11311; 11:11“; 113115 a,- 389.5 389.5 801.6 801.6 b,- -140.6 -140.6 -3.82 -3.82 Even though one usually adds a zero-range spin-orbit part to this interaction, this is not done in this work as our target is to compare the DME approxima- tions, and a zero-range interaction is treated exactly in the DME. A complete self-consistent test of DME approximations should make use of modern N N in- teractions that have central, spin-orbit and tensor components. In this regard, 119 building a local chiral interaction that. has a low-momentum cutoff, and thus gives sensible HF results, will be useful. 0 we calculate only the total energy, and its components such as kinetic, Hartree and Fock contributions of closed-shell nuclei: 16O, 40Ca, “Ca and 90Zr. The fact that we do nOt resolve the single-particle spectra (such as spin-orbit split- ting) prevents us from assessing the DME of 81,071, 8'2) for spin-unsaturated 48Ca and 90Zr. Specifically the impact of the significant improvement in the DME of 82,61,772) brought by our PSA-DME is not yet gauged with self-consistent calculations. Additionally, we do not calculate the corresponding exact HF re- sults. Rather, we use results from Ref. [184] when we need to refer to exact HF results. Derivations related to the self-consistent numerical test are given in Appendix 9.8. For faster convergence of the calculations, we implemented both Broyden’s [194] and imaginary-time methods. Table 5.4 lists the results of the self-consistent calculations. At this point, we remind the reader that there are four densities that appear in the exact HF energy of time-reversal invariant systems: pq(f'1, F2), §q(F1,F2), pq(F1/2) and .2071”) (which appears only if the given NN interaction has a finite-range spin-orbit part). The second column show how the 1r—functions of these densities are fixed. Obviously, pq(1"'1/2) and LG} /2) are not expanded in exchange-only-DME calculations. We start with the nonlocal densities. First, we have the labels N V-full and NV- exc-only. The full and exchange-only labels should be self-explanatory. NV refers to fixing the 7r—functions of pq(F1,F2) and §q(F1,F2) according to the original DME of Negele and Vautherin (Ref. [170]). PSA-II-exc-only use the parameterized versions of the 7r—functions of pq(f'1, F2) and 5.1071372). The parameters that we used are the ones that we optimized for the non self-consistent test, while for pq(F1/2), the parameters and 7r—functions are as discussed in the previous section ( 5.4.5). For the 7r—functions 120 —o of Jq(F1/2), we use Taylor series as was done in Ref. [170]. Table 5.4: Full-DME and Exchange-only-DIVIE for Brink-Beeker interaction and several DMEs EM kin Dir Exch lBO NV —full -6.204 13.948 - - NV-exe—only -5.600 13.474 18.839 -37.914 PSA-II—full —7.932 15.417 - - PSA-II—exc—only ~5.635 13.338 18.513 ~37.487 40Ca NV-full -8.526 16.822 - - NV-exc—only -7.516 15.793 22.567 -45.878 PSA-II-full -10.359 17.575 - - PSA-II—exc-only -7.539 15.583 22.075 -45.198 48Ca NV-full -7.447 16.678 - — NV-exc—only -6.625 15.762 21.334 -43.803 PSA-II-full -9.304 17.57 - - PSA-II—exc-only -6.646 15.529 20.884 -43.062 902r NV-full -9.339 18.778 - - NV-exc—only -8.388 17.320 24.322 -50.041 PSA-II—full -11.543 19.038 - - PSA-II-exc-only -8.389 17.040 23. 700 -49.140 A complete self-consistent test should include the exact self-consistent HF calcula- tion and the calculation of several other quantities such as matter, proton and neutron radii, proton and neutron densities. For the single-particle energies, the balance of Hartree to Fock contributions need to be assessed [184]. Still, our preliminary test is consistent with the main conclusion of Ref. [184], viz, the full—DME gives an excess 121 of 1 MeV per particle binding energy compared to exchange-only-DME, irrespective of the performed DME. The exact HF calculations given in Ref. [184] show that the error in the exchange-only-DME (compared to the exact HF of Ref. [184]) are much smaller than that of the full-DME. Consequently, one can obtain a significant reduc- tion of the error in DME approximations by treating the Hartree contribution exactly as exact treatment of the Hartree contribution does not add to the numerical com- plexity of the problem. Comparing the two DMES, it can be seen that the difference between the exchange-only versions of NV-DME and PSA-DME—II is marginal, while for the case of the the full-DME version, there is a significant difference with NV- DME being much closer to the exact HF results reported in Ref. [184]. This must be due to the strong parameter dependence and self-consistent error enhancement as the non—self~consistent percentage error from PSA-DME—II (which we called PI-DME) of the Hartree contribution, given in section 5.4.5 is less than that of NV—DME. This requires further investigation. 122 Chapter 6 N on—Empirical Energy Density Functional from NN interaction In this chapter, we calculate the HFB energy from a generic two-nucleon interaction that contains central, spin-orbit and tensor components. Furthermore, we apply the DME to the HFB energy to obtain a local EDF. Analytical couplings of the particle- hole (HF) part of the resulting EDF are calculated using the finite-range part of the chiral EFT two-nucleon interaction at NZLO. Following the usual convention, we represent momentum transfers with q. To avoid ambiguity, the isospin coordinates of the particles are labeled with T. 6.1 The HF energy from an NN interaction Starting with a two-body interaction, the Hartree-Fock energy is given, in an arbitrary basis, by l . . * ENN = E: Z (iTJT’|Vlkr”lT"’) (F101T1;F202T2|dF1dF2 = 11 , (6.3) ”162 717-2 and the definition of the density matrix in h") 69 I6) 8) I?) space as given in Eqs. (9.70), the HF energy can also be written as 4 1 . NN— — ‘1“ 1 132"...“ I; —- 2 22::‘/{:I:[ ctr,(73¢71717,g727§|l/’ ITJCK§T3T4CT4TQ) or 1:1 X p(F303T3, F101T1)P(F404T4, 71.20272) 1. 4 ......~132....(1)_...(2)_... : ETI'ITTQ H dr,(r1r2|V lT3T'4)/) (7'3, 7'1)p (7'4, 7‘2), (6.4) i=1 where V18” = V(1 — P12), with P12 being the particle exchange operator defined in Table 1.2. A matrix notation is used in the second equation and the traces, denoted by Tr, denote summation over the spin and isospin indices of “particle 1” and “particle 2”. The quantity MOO-3,71) is defined in Eq. (7.12). As discussed in section 7.1 in relation to a similar calculation for the chiral three-nucleon interaction at NQLO, this notation makes the direct Mathematica implementation of the equations transparent. Refer to that section for details on this notation 1. In this chapter, we eventually qualify all results for the finite-range (pion-exchange) 1The calculations for the NN case are relatively much simpler than for the NNN case. Conse- quently, the complete HF + DME calculation was carried out manually in this case. 124 part of the chiral EFT two-nucleon interaction at NZLO, which is discussed in sec- tion 2.4.1. As the HF energy from the contact part is already in a quasi-local (Skyrme— like) form, it does not require the application of the DME. Thus, we do not discuss it any further. The actual expression for the contribution to the EDF from the HF energy of the contact part can be found in Ref. [153]. The application of the DME to the HF energy requires expressing the HF energy in terms of the scalar/vector-isoscalar/isovector components of the OBDM in IF) 8) IE) (8 I?) space. This is due to the fact that the DME, as formulated in section 5.2, is most intuitively expressed in that space. For its formulation in momentum space, consult Ref. [154]. Hence, we need to perform inverse-Fourier transformation of the chiral interaction given in Eq.(2.14). This results in (R‘r' (v: if?) = ([VC(T) + T1 .T21/VC(T)] + [ v5.6) + T, .T21I11'5(T)]5, ~02 + [W(T) + T1 .T,WT(T)] s, 8.51, - 5,. + % [v,,,.(T) + T1 -T21;V,,,(T)] (s, + 5,) . (A. a 53)) 6(F- 1'”) 6(R — 1?!) , (6.5) where R’, R denote the center of mass coordinates, f", r are the relative coordi- nates and 6,. and FR refer to gradients with respect to 7" and R respectively. The {Vc(r), Wc(r), ...} denote the inverse-Fourier transform of the respective form factors given in momentum space [153] d" ._._. W") = / (2:)36“""%(q) forz'=C.s.T, (6.6) i d” .~~_,. _, , = E (273)36” (q-F)V,-(q) for1=LS. (6.7) One should note that chiral interactions come with a regulator that cuts off high- 125 momentum components. It should be noted that we have not included regulators in Eqs. (2.14), (2.16), (2.18) and (2.19). The commonly used regulators result in a non-local interaction since (TH-cg |V| E172) which can also be written as (11" WIT) is replaced with f(k’/A)(E’|V|i5)f(k/A) for some momentum scale A, where f(k/A) —> 0 for k >> A and f(k/A) z 1 for 1: < A. In contrast, Eq. (6.5) shows that the I7“) <8) I5) <8 I?) space representation of the interaction given in Eq. (2.14), viz, without the regulator, is diagonal and depends on the gradient \"7 with respect to 1". The spin-orbit part in Eq. (2.14) is actually the only term that depends on 6,3. This dependence on V7,}. is usually remarked by referring to the interaction as quasi-local. In order to obtain a local interaction, one could use a regulator that suppresses large momentum transfers instead of large relative momenta. In any case, we neglect the regulator since we work at the HF level which samples only the low-momentum spectrum of the single-particle Hilbert space. This argument will remain valid as long as the local Fermi momentum kp << A. 6.1.1 HF contribution from a central interaction A local two—nucleon central interaction can be split into four different spin-isospin channels as V's-j : l/cijh‘) (SO-:1 — F3) 6(F2 - F4) H02. HT]. , (6.8) where i, j€{0, 1} denote the singlet and triplet channels, I/Cij(r) is the form factor and Hat. and HT]. are the spin and isospin projection operators defined in Table 1.2. For the chiral interaction given in Eq. (6.5), the form factors of the central interaction in the different channels is given by vg“(r) = v.(T) -— 3ll”(.(’l‘) + V..(r) — 3w,(T) , (6.9) V8 (r) W) + W) — 3v..(r) — 3W.(r), (6.10) 126 v3. (T) = no) + l"l"7,.(1') + mm + W..(r) , (6.11) V80(r) = W(r) — 31'1“},(1‘) — 3V3(r) + 91V,(r) , (6.12) where Vc(r), Wc(r), are the coordinate space form factors given by Eq. (6.6)- (6.7). Starting from Eq. (6.4), the HF energy in the four different channels can be derived by replacing V with the corresponding interaction given in Eq. (6.8). The details can be found in appendix 9.6.1. For the spin-triplet and isospin-singlet channel, one has I . 1 _, _, 3 _, _, 3 _, _, _, _, E35110] = g;/ dr, dT2VciO(T)[§/)T(7‘1)Pr(7“2) + §pT(r1,r2)pT(r2,r1) 1,, _, _, _, 1 _, _, _, _, _, _, +534“) ' 5r(7'2) + §ST(7‘1T7“2) ' 36(7‘2TTI)] T (6-13) While for the spin-singlet, isospin-triplet-channel NN _ l —: -: 01 _. T -. - ~ ~ EC l01l — 82/ d71d72 Vc (T)[PT(T1)PT(7‘2) + Pr(T1TT2)Pr(7‘2T7‘1) 1 _, _, 1 _, _, 1 _, __, _, _, + g2] (1T1 (1712 V810") [5PT(T1)Pr(T2) + —p,(r1,r2)pT(7‘2,T1) The HF energy for the triplet-triplet and singlet-singlet channels read 1 .. - _. - _. - - .. EgNllll = gZ/drldmvcill'r)[3Pr(rllpr(7‘2) — 3PT(7‘1=7‘2)PT(T2T7‘1) —O +§T(7'l) ' §T(F2) _ §T(F1T'F2) ' §T(F2’Fl)] + 6;] d’fidfé Vél(7')[gPT(F1)Pr(F2) — ng(F1.7‘72)pT(F2T7-‘i) 1_, _, _, _, 1_, _, _, _, _, _, +5.46)- .6.) —§(>()] (6.15) and The corresponding expressions for the finite-range part of the chiral NN interaction at N 2L0 can be found by utilizing Eqs. (6.9)—(6.12) in the place of the generic form factors V” (1‘). In time reversal invariant systems, the symmetry properties of the one- body density matrix discussed in section 9.2.4 can be used to simplify the expressions. In particular, those terms that depend on the local spin density vanish as §q(7"') = 0 in this case. Finally, it should be mentioned that the channel by channel expressions for the central interaction agree with the unpolarized and symmetric infinite nuclear matter limit of the same expression given in [170]. 6.1.2 HF contribution from the spin-orbit interaction A given quasi-local two-nucleon spin-orbit interaction can be split into its spin-isospin- singlet-triplet channels as . 2.. 2' i. _. _, _. _‘ _. _. _. V139 = -—§VLJS(T) 6(r1 — r3)6(r1 — r3) V - (01 + 02) Ilai HTj' (6.17) The spin-orbit orbit interaction vanishes in the spin-singlet channels lxsmgm) = 1/\/2(|T1)—|11)) as S2 IXsingzet> = 0, with two-body spin operator 5'. = (31+62)/2. For the chiral interaction given in Eq.(6.5), the spin-orbit form factors read Vigil") = 2V1.50“) - 6WLS(7‘) T (6.18) V2.66) = 2163(7) + 2WLS(T). (6.19) 128 The HF contribution from the spin-orbit interaction in the spin-triplet and isospin- singlet channel is given by .0 1 -o -o r -o _. —o —o —o -9 - -o E53110] = :1- Z/ (17‘, (Ir-2 V2g(r)r- I: T(r1)p;(r2) + 8,.(1‘1) X ]T(r2) i —o —o —o -—o a -o d —o —o +1 2/ (17', (173 V1330) [.s,('r1,r2) . r x Vgpf(r2,r1) +pT(F1,T_“2)FX 62 ' TTf-(F2,F1):l , (6.20) while in the spin—triplet and isospin-triplet channel, the result reads —o 1 _. _. _. ~ .. _. _. _. . .. Elllelll = ‘2‘ Z/ drl (”2 V1366)” [J,(r1)p,(r2) + 3r(7‘1) X Jr(7“2)] an. -o 1 4 ... _’ ... _, _, ‘3 _. + Z Z] dTTdTTvgarV' [Jr(r1)/2T(r2)+ sm) X “ml i —p -O "’ d q -. -' ._. -. — 21:, /d7‘1dT2VL‘§-(r)[BdrnrzlrxV20T("2~7‘1) + pT(F1,F2) 7? X fig ' .Tf-(Fg, F1) ] . (6.21) The actual derivation is given in appendix 9.6.2. 6.1.3 HF contribution from the tensor interaction A local two-nucleon tensor interaction in the four spin-isospin channels is given by V1? 2 V71;j(r)6(f'1 — F3) 6(7-3 — 705121-10’. HT]. , (6.22) 129 where 512 is the tensor operator given in Eq. (2.10). The tensor interaction acts only in the spin-triplet channels. This becomes obvious once 512 is written as $12 = 6 (§ - f)2 — 2 S2, where the total spin operator, 5;, has zero expectation value in the spin-singlet state. Hence, concentrating on spin-triplet channels, the chiral NN interaction at N2LO given in Eq.(6.5) has the following components V7100) = éVTU') — W’TU) , (6.23) v;‘(T-) = -;—V,(T) + éWfi-r) . (6.24) The HF contribution to the energy in the spin-triplet and isospin-singlet channel is given by _‘ _. 3T 1 TV _, ... —-o —o -' —’ ENNllol: 211;: Z / drl (1T2 VTI‘O(r) I: 7,-2 8T.#(r1)81-’,V(T2) — STU‘I) . 371(7‘2) 1' pu 3r 1‘ _, _, _. _, _. _. _, _. _. _. —71'/§—uSTTH(TlaT2) 8?,U(r23r1) — 87(7‘1’7'2) . 71(7‘2’7‘1) ’ while for the spin-triplet and isospin-triplet channel 37' 73, _, _. _. _. _. _. E¥”{111=—2 Z Z / dr. dr2v1(w~)[——,—“ s.,.(r.)s...(r2) — 3.6.) - 3.62) 1[31' r _, _. _. _. _. _. .. ——:2 u S’Tp(7‘1 T2)3Tu(1‘2,r1) + s,(r1,r2) . 37(r2,r1)] 37' 71, _, _. _. _. _. _. +—41:WZ/dT1dT2L/TII(T) [Tp23‘r.p(r1)31",u(7‘2) — ST( 1) ' .‘7-(7‘2) 3r 7‘ _, _, _. _. _. _. _. _ —%_VsT-H(T17T2) Sit/(r'brl) + 37(7‘127'2) ° ‘ ;(7‘2,7‘1)]. (6.26) Once more, the relevant expressions for finite-range part tensor part of the chiral NN interaction at N2LO can be obtained by making use of Eqs.(6.23)-(6.24). For time- 130 reversal invariant systems, those terms that depend 011 the local spin density, £107), vanish. This recovers the expression derived in Ref.[170]. Refer to appendix 9.6.3 for details. 6.1.4 Additional contributions to the HF energy In addition to the contributions to the HF energy that come from the starting N N interaction, there are several additional terms that are due to the kinetic energy, the center of mass correction and the coulomb interaction. The simplest is the uncorre- lated kinetic energy associated with the reference product (HF) state 2 Em = h— : / dT’T,(T*). (6.27) Since (HF) meanfield solutions are localized in space, translational invariance of the actual nuclear hamiltonian is broken. Consequently, one needs to correct for the center of mass motion, which can be done by defining an intrinsic Hamiltonian [38]. In addition to the the starting Hamiltonian, the intrinsic Hamiltonian contains a correction term ECM which reads . _ (q’chzwl‘I’) _ (‘I’I(Zkfikl2|‘1’> ECM — 2Am _ T 2Am (6'28) where POM = 2k 13'), is the sum of single-particle momentum operators and A is the number of nucleons and the |) is the reference independent particle or quasi-particle 131 state. Expressing EC)” in terms of densities, one obtains E _ h2 2 / dF (r) h? Zr, // Jada/1' (.3 ‘3 ('3!) C“ — 2Am 7’ 2Am (T r 9" JT’ ' ,_. ~I +2Am 2; f/drdr p,( r, I‘IT)T /(r,r) (6.29) + mm 2, f / df'df" c616") g,(T=*,F'). Several comments are at play concerning such a correction to the the HF energy (i) the first term is a one—body center of mass correction with an overall effect of rescaling the kinetic energy term, Ekin (ii) the second term is really local and is zero if the single-particle states have a good parity, which we assume to be the case (iii) the non-local third term is the so-called two-body center of mass correction and is often omitted; if single-particle states have a good parity, the two coupled densities are labeled with opposite parities (iv) the non-local fourth term contributes to the pairing energy appears if one considers a reference independent quasi-particle state. It has never been considered in practical calculations [26]. If the single particle states have a good parity, the two coupled densities are labelled with opposite parities. In its full generality, such a term generates neutron-proton pairing. These nonlocal (the third and fourth) terms will be neglected in our application of the DME. The last correction arises from the Coulomb repulsion among the protons. It has both direct and exchange parts. The nonlocal exchange contribution is usually approximated with the slater approximation [155]. This is due to the fact that for the long-range Coulomb interaction, the simple slater approximation seems to perform at least as good as the DME techniques discussed in the previous section. Hence, we write the contribution from the Couloumb interaction as the HF energy reads so,“ _= —eW//d1"’pfifl_pp£ll)—(223(;)1/3/de:/3(77) . (6.30) 132 6.1.5 The leading-order pairing contribution The leading-order pairing contribution is obtained by calculating the expectation value of the interaction in a Bogoliubov quasi-particle vacuum. At this point, we enforce several restrictions: (i) we neglect proton—neutron pairing, hence no isospin- singlet contribution. This is justified in most cases as the Fermi energies of protons and neutrons are quite different for most nuclei [38]. (ii) only central interaction is considered, as it is the 1So channel that exhausts most of the pairing contribution in nuclei, i.e. the contribution of other partial waves are negligible [117]. Leaving the details to appendix 9.6.4, the spin-singlet isospin—triplet contribution reads 1 _. _. ~ _. .. (¢|V31|)m,-r = :1- : / drld'rg V310)|p7('r1,r2)|2 , (6.31) while for the spin-triplet and isospin-singlet channel, we have 1 _, _, :3, 4 _, ~ _, _, ((blvéll (I) >pair = Z Z / drl (1T2 V8710) 87(T1, 7'2) ' .97-(7'1, T2) . (6.32) The Coulomb interaction has an important effect on proton pairing gaps [110]. Specif- ically, its repulsive nature reduces proton pairing gaps (anti-pairing effect). To cal- culate these contributions, one simply replaces the form factors, V6110“) and V810), with the corresponding Coulomb interaction form factor. 6.2 Application of the DME to the NN-HF energy In this section, we apply the DME to the HF B energy derived in the previous section to obtain a local EDF. In section 5.4.6, we have verified that the DME of the Hartree contribution is the main source of the descripancy between the DME-approximation and exact HF, thereby advocating the exact treatment of the Hartree contribution. 133 Still, for the sake of completeness, we apply the DME to all contributions of the HFB energy: Hartree, Fock and Bogoliubov. As the HF B energy is derived for a generic two-nucleon interaction, we perform the derivation of a local EDF using the modified-Taylor series detailed in appendix 9.5.7. As explained there, all available DMEs, including PSA-DME developed in this work, can be mapped in to this formal expansion. Since, the starting point is the strict Hartree, Fock and Bogoliubov contributions (diagrams), the energy functional is intrinsically a bilinear functional of p and K, i.e. 5[p,n,n*] = 5;; + 552 + 5;; + 55.2, + 5”” (6.33) coul ’ where the right hand side corresponds to the uncorrelated kinetic energy, the particle- hole (HF), the particle-particle (Bogoliubov/ B), the center of mass and Coulomb corrections. The p and n exponents denote genuine, original dependence on the density matrices. To recap the steps for the application of the DME, first we replace the densities in the HFB energy with their formal expansion given in appendix 9.5.7. This is followed by the simplification of the expression using the angle independence of the 7r— functions and the relations among the 7r—functions discussed in section 5.3.6. After neglecting terms with beyond second-order gradients, the particle-hole part of the EDF takes exactly the same form as given by Eq. (4.11), where in this case, the A/ B couplings are functionals of the 7r—functions and the starting interaction. Through the DME, finite-range contributions of the starting interactions are en- coded into density-dependencies of the EDF couplings. For instance, App 2 3 fair?"2 [V31(r) ((7r3(kp7'))2 + (H8(kpr))2) + 3Vél(r) ((7rf,’(kpr))2 — ( Ham)?” , (6.34) 134 where we used the isoscalar kp as the DME length scale and suppressed the I? de- pendence of kp for brevity. It should be noted that due to the density dependence of the couplings, the usual integration by parts that is used in traditional Skyrme EDFs to reduce the number of independent terms can not be applied here. For instance, in conventional Skyrme EDFs, it is possible to convert the pAp term of the EDF into Vpr , thereby reducing the number of terms. Generally speaking, this is not possible in the previous case. The couplings depend on the central, spin-orbit and tensor parts of the interaction as follows Central ——+ {App, A”, APT , APAP, AVPVP, AJJ , ASAS, AVsoVs, Bpp’ B” , BPT , BPAP, BVPVP ’ BJJ 3 BsAs ’ BVsoVs } Spin-Orbit -—> {ApVJ, AVPJ, BPVJ, BVPJ} Tensor —> {AJJ, A” , A3133, AVSVS, AVSOVS ? BJJ , BJJ , BsAs , BVsVs , BVsoVs } . The complete expression of the couplings is given in appendix 9.6.5. In the particle—particle (pairing) channel, the application of the DME to the pair- ing contribution results in a functional that is more complex than the usual phe- nomenological forms, given in Eq. (4.16). It reads epp[p,n,n*] a 5; / diz' [Ac war + prg) (Am) — we) +Ac m1?) (AMI?) — 4mm) +Ajj Z j:,pu(1-?:) jT-#V(§):| 3 [JV (6.35) 135 where the A” terms originate from spin-triplet pairing while the rest originate from spin-singlet pairing. The actual expressions for all couplings is given in appendix 9.6.5. The EDF that results from the correction terms, namely, uncorrelated kinetic energy, center of mass and Coulomb corrections are simply given by 1‘12 —. -» 6': = 'fiZ/(IRTT(R)’ h2T .. -+ 555, = —2Am Z/dRT.(R). 535111 = / dI—f {CPP pPpP + CpApppApp + CVpr 6:01) ' 6p], 3 2 3 1/3 " 43 " —Ze (g) [dRpp/ (R), (6.36) where as noted in section 6.1.4, we have neglected the third and fourth terms of Eq. (6.29) while its second term vanishes due to the assumption of good parity for the single-particle states. For the Coulomb correction, we have applied the DME to the direct piece, while leaving intact the Slater approximation for the exchange part. The application of the DME to the Hartree contribution is given just for the'sake of completeness. In fact, even the exchange contribution from the Coulomb interaction can not be treated accurately due to the long-rangedness of the Coulomb interaction. One can expect the DME of the direct part to be much worse. Still, we give the values for the C couplings in appendix 9.6.5. 6.2.1 Analytical couplings from the chiral EFT NN interac- tion at NQLO There are three steps necessary to obtain the analytical calculation of the couplings of the local EDF derived in the previous section. First, we have to restrict the discussion to time—reversal invariant systems as the analytical forms of the 7r—functions for time- 136 odd densities are not completely determined yet. This is discussed in section 5.3.6 in detail. The next two steps involve (i) fixing the interaction which in this case is the finite-range part of the two-nucleon chiral EFT interaction at NQLO. The respective use of the three-nucleon interaction is the subject of the next chapter. (ii) Specifying the 7r—functions. This can be PSA-DME, or any of the other available DMEs. In fact, using different DMEs results in different couplings, which is mentioned in section 8.2 as a way to estimate the error of the DME couplings. In our case, we calculate the couplings for PSA-DME. A similar calculation can easily be done for the original DME of Ref. [170]. The derivation is discussed in appendix 9.6.6. As the final expressions are too lengthy, we discuss only the skeleton expressions of the couplings. For the more on the couplings, consult section 8.1 for a relevant discussion. The lengthy analytic expressions for the DME couplings tend to obscure their underlying structural sim- plicity. Therefore, it is more illuminating to display the couplings in “skeleton form”. Each coupling Ct“) is given by the sum of the LO (n = 0), NLO (n = 1), and N2LO (n = 2) contributions 2 cm) = 20m 2'6 {pt/mm), . . .}. (6.37) n=0 where the dimensionless variable u E kp/m,r and t = {0, 1} is the isoscalar/isovector index. The fact that we express the couplings in terms of isoscalar/isovector no- tation instead of proton-neutron is for conformance with the notation used in the derivations related to the three-nucleon interaction. Refer to [156] as to why the isoscalar/isovector notation is more convenient in that case, and Eqs. (9.72)-(9.75) for the simple algebra relating isoscalar/isovector notation with that of proton-neutron 137 notation. Now, each coupling can be written as Cgf,),(u) = Z cry-Rn, t, u)fj(n, u) (6.38) i where (lg-001,13, u) are rational polynomials in u and .73-(11, u) are functions which may exhibit non-analytic behavior in u due to the finite-range of the NN interaction. In the skeleton expressions listed below, we use a more compact notation where the dependence of the a’s on u, t, and n is not explicitly shown: 0 L0 couplings C(i) = as) + a?) log(1 + 4u2) + a?) arctan(2u) (6.39) o NLO couplings : 2 C”) = of,” + a?) log(1+ 2u2 + 2uv1+ 23)] + agi)\/1 + u2 log(1 + 2n2 + 2u\/1 + u2) (6.40) o NZLO couplings C“) = as) + all” log(1 + U2) + agi) arctan(u) (6.41) 6.2.2 Single-particle fields and equations of motion In appendix 9.7, we give the derivation of the single-particle fields and HF B equa- tions of motion that result from the variation of the Skyrme-like EDF given in 138 Eqs. (4.11),(6.35) and (6.36). Additionally, we give similar derivation for the case where the DME is applied only to the exchange part of the HF energy. As discussed in 5.4.6, all numerical tests are carried out for the case of spherical symmetry. Hence, we give the most simplified single-particle fields and equations of motion that result when spherical symmetry is imposed. Furthermore, the numerical methods used to solve the self-consistent spherical HF equations are also discussed. All this can be found in appendix 9.8. 139 Chapter 7 Non-Empirical Energy Density Functional from Chiral EFT NNN Interaction at N2LO In this chapter, we calculate the HF energy from the chiral EFT NNN interaction at NZLO. Next, PSA-DME, formulated and discussed in chapter 4, is generalized in such a way that it becomes applicable to the NZLO chiral EFT NNN HF energy. This is followed by the application of this generalized PSA-DME to obtain a local EDF and the analytical calculation of the couplings. Additionally, we make several references to the actual symbolic implementation of the calculation. Again, following the usual convention, we represent momentum transfers with q. To avoid ambiguity, the isospin coordinates of the particles are labelled with r. 140 7.1 The Hartree-Fock energy from Chiral EFT N NN interaction at N2LO The consistent application of an MBPT calculation starting from a chiral EFT in- teraction, at a given order, requires utilizing all the components of the interaction: two- and many-nucleon interactions. As discussed in section 2.4.2, the leading three- nucleon interaction appears at N2LO in the chiral expansion and it has three main pieces: the three—nucleon contact which is referred to as the E—term, the one-pion ex- change plus contact (D-term) and the two-pion exchange which is called the C-term. In this section, we calculate the HF energy from these pieces of the NNN interaction and apply the DME to obtain a quasi-local EDF. Unlike the NN case, the algebra required to arrive at our final target, namely, a quasi-local EDF, is so complicated that one can simply rule out a manual derivation. This is due to the tremendous size of the algebra required in both layers of the problem. Firstly, we have to derive the exact HF energy in terms of the scalar / vector- isoscalar/isovector parts of the OBDM. This has to be followed by the application of the DME to obtain the final quasi-local EDF. However, the whole problem displays several features that make it amenable to symbolic automation [151]: (i) it involves many similar and repetitive algebraic steps (ii) most of it does not involve numerical computation, and (iii) the part of it that seems to require numerical computation, such as multidimensional integrals, can be performed using a combination of analytic expansion and symbolic integration. In the following section, the HF energy from chiral EFT NNN interaction at N2LO is expressed in a form that makes the symbolic implementation transparent. The complete symbolic derivation is discussed in Ref. [156]. 141 Basic form of the HF energy A three-nucleon interaction can in general be decomposed as a sum of three terms V3N E V12 + V23 + V13, (7-1) where 17,5 is symmetric in nucleon i and j. Specifically, for the chiral EFT three- nucleon interaction at N2LO, Vi]- depends on momentum transfers (j;- and fl} and, in general, on the spin-isospin coordinates of the three nucleons. Refer to section 2.4.2 for details. Starting with the HF energy from a three-nucleon interaction . 1 . . ~ . . (V3513?) 2 —6— :(ZJICIWAI(1+ P13P12 + P23P12)(1— P12)ll]k> , (7.2) ijk a few basic algebraic manipulations are in order to express the HF energy in terms of only one of the three 17,-]- operators, e.g. I723, as . 1 . . ~ . . (‘65?) = 5 Z fulfill/230 — 2P13 — P23 + 2P12P23)|’1Jk)a (7-3) ijk where sz denotes the exchange operator (of particles 1 and m) defined in Table 1.2 whereas i, j andk denote occupied HF single-particle states. Note that for ease of notation, we are using the single-particle basis that diagonalizes the one-body density matrix of the HF Slater determinant. One can identify three groups of terms in Eq. (7.3): direct, single-exchange and 1 double-exchange terms . The direct term corresponds to the expectation value of V23, the single-exchange term to the expectation value of I723(—2 P13 — P23) and the 1This should not be confused with one- and two-pion exchanges contribution to the three-nucleon interaction. 142 double-exchange term to that of 2 127231912ng 7(ir 1 .. " .. my!) a ‘Q‘ZQJklléalUkl. (7.4) ijk. 0353’") 2 5 Emmet—21913mutate). (7.5) ijk main 2 Z. (7.6) ijk As the derivation of the Skyrme-like quasi-local EDF from the exact HF energy requires the application of the DME, we need to express the HF energy in the IF) (8) IE) ® IF) single-particle basis. Hence, we perform inverse-Fourier transformation of the interaction. This transformation leaves the spin-isospin dependencies untouched. Furthermore, just as in the case of the NN interaction, the fact that the calculation is confined to the HF contribution enables us to neglect the regulator so long as hp << A, the momentum cutoff scale. The absence of the regulator makes the interaction local in coordinate space and simplifies the form of the interaction in IF) 8) lo) 69 Ir) space. Confining the discussion to the spatial dependence, we have -OI -OI -OI -OI (F1772F3IV23I'7‘1T27‘9 = (5071 - 7‘1)5(F2 — 772') 5(773 r 7‘3) X V2303 — F1, F3 - F1), (7-7) where 1 (2r)6 v23(r2 — F1. F3 — r.) s [dadaeiiwirfi’eta-(i341)vixen). (7.8) At this point, we do not actually perform the integrals over the momentum co- ordinates in Eq (7.8), except for the E—term of the interaction which is a trivial three-nucleon contact interaction, thereby yielding simple delta functions as shown in Eq. (9.347). Rather, Eq. (7.8) is used as it is, resulting in fifteen-dimensional integrals 143 in Eqs. (7.13)-(7.15). As discussed in section 7.2, the application of the DME prior to the actual multi-dirnensional integrations is crucial. The next target is to rewrite Eqs. (7.4)-(7.6) in a form transparent for Mathemat- ica implementation. We illustrate the steps required to achieve that with Eq. (7.6), for which we have MW") 2 :(ijkll/ggPlgP23lz'jk) ijk 3 3 = Z Z 2 f Hng, Hdfil (ijlc|Fllol'r1'F/202'r2'FI3oérg) ijk m=1 n=1 0f..03 T1,..T3 .. I I a I I —» I I " 1- or -+ .. _. X (rllolr1 ”2027'2 rl3o3r3ll/23Pf’2 P23 lrlolrl rgogrg r3o3r3) ><(71.1017’1775027'2 71.3037'3lpffzpziilijkla (7-9) where we used completeness relations in the three Hilbert space 3 E E f H dr, |r101r1 rgogrg r3o3r3) (rlolrl T202T2 r303r3| = 11 , (7.10) 01 ..03 1'1 ..T3 i=1 and ("I E (in (In. We split the particle-exchange operator such that the coordinate part acts on the wave-functions while the spin-isospin piece is taken care of along with the interaction. Let X:- represent (Fioiri) such that the one—body density matrix reads as 9(Xjan) —=— QffiUjTjstUka) E Z ‘pfffkak Tk)‘19i(7?j 03' 7'1), (7-11) I where the sums is over occupied single-particle HF states. Making use of this, we define another quantity, which we call the auxiliary density matrix, as 07X)“, Xk) E etfim’n’, From), (7.12) 144 where ie {1, 2, 3}. Basically, the spin—isospin coordinates of this quantity are those of the i‘" particle. Applying the steps demonstrated in Eq. (7.9) and using Eqs. (7.12), (7.7)-(7.8), one can express the direct, single-exchange and double-exchange parts of the three-nucleon interaction HF energy as ‘ ( ir 1 -o —0 -¢ -' a ' _. (VS/['1 l = 2 Tr1Tl‘2rpf3 [/ (17'1d7'2d7‘3 91(X1) 920(2) 930(3) X V2303 - F1, F3 — F1)] , (7.13) (VII/PIX) = - TriTI‘2Tr3 [/ 61771617726177“; 91(X3.)?1) 92032) 03031, X3) XV2:5(F2 — 77197-33 — F1) :57] 1 _. _. _. _. _. —. . _. —. — 5 rI‘I‘1TI‘2rIlr3 [/ drl(17‘2dT3 91(X1) 92(X39 X2) 93(X21X3) XV23(F2 — F1, F3 — 71.013537] , (7.14) (Mffvmxl = ITITF2TT3/df'id7—‘tzdf3 [01(X2,X1)02(23,X2)Q3(X1,X'3) XV2;3(F2 — F1, F3 — F1) Pég-PIOJ] , (7.15) where gi(Xj) E Qi(X-.j, X3) and Tr,- refers to tracing over spin and isospin coordinates of the ith particle. The key to understand the form of these equations is the splitting of the particle exchange operator, performed in Eqs. (7.9), that results in the spin-isospin coordinates of each particle to be grouped in a single auxiliary density matrix. These are the basic equations that are implemented directly in Mathematica. In Ref. [156], it is shown that the implementation of these equations is transparent, viz, directly interpretable to the language that Mathematica understands. This would not have been the case without the trick used to group spin-isospin coordinates of each particle in a single auxiliary density matrix, Eq. (7.12). The following sections state the contributions to the HF energy of time-reversal invariant systems from the E-, D— and C-terms of the chiral EFT NNN interaction 145 at N 2L0. The complete expressions where the assumption of time-reversal invariance is relaxed are given in appendix 9.9. Even for time—reversal invariant systems, some of the expressions are too long. In those cases, the expressions are relegated to the same appendix, where we also give the corresponding results for INM and PN M (pure neutron matter). Prior to delving in to the details of the expressions, the following observations can be made regarding the HF energy: (i) Each term in the energy expression should contain three local/nonlocal densities. (ii) As discussed in appendix 9.2.4, the various local and non-local densities that result from the one-body density matrix have specific timereversal properties. Hence, considering that energy is a time-reversal invariant quantity, there can be no term that contains one/ three time-odd densities. Note that at the level of exact HF, the only time-odd density that we have is the local spin density, s'q(F). Nonetheless, the application of the DME extends this set to include all the time-odd densities that are discussed in appendix 9.2.4. (iii) The fact that the starting interaction is isospin invariant makes the energy isospin invariant as well. Therefore, there can be no term in the energy expression that contains one/ three isovector densities. (iv) For each part of the interaction, there are the direct, single- and double-exchange contributions as given by Eqs. (7.13)-(7.15). Finally, we remark that the HF expressions and the resulting quasi-local EDF from the chiral EFT NN N interaction at NZLO are given in terms of isoscalar-isovector no- tation instead of proton-neutron notation. In Ref. [156] where we discuss the Math- ematica implementation, it is shown that the isovector-isoscalar notation is better suited to the implementation. Finally, keeping in mind that the NNN chiral EFT interaction at N2LO does not have isospin invariance breaking terms, the isospin in- variance of the energy expressions of both exact HF and quasi-local EDF become transparent in isoscalar-isovector notation. 146 HF energy from the E-term The actual operator structure and analytical form of the E-term of the chiral EFT NNN interaction at N2LO are given in Eq. (2.16). The HF energy that results from it for time-reversal invariant systems is given as = —,%E/dr(p3s€(ae>sY(p.+ (V ) 3N 3N 3N + (vQPLFJnXJNM) (7.39) ... (’8 3 _. 2 d7‘1 CCJNM P001) a (7-40) 163 where P8 vii (’8 vii CCJNM E C(fDlrJgV'M + CC'DQIJNM + CRZIJNM' (741) We have separated the three different contributions to the coupling in Eq. (7.41). As discussed in section 7.1, the C-term of the chiral EFT three-nucleon interaction at N2LO has what we call the D-like and R pieces. In the coupling shown in Eq. (7.41), the first two terms of the couplings are from the single- and double-exchange parts of the D-like piece, while the last one is from the double-exchange part of the R—piece. These are given by 7% 54 9A CCDIXJNIU -10247I'6 2f 6 7 513171 q21q31 01777.3, C3 _. _. 2 2 —4 + 2—(12 ' (13 ((12 + "13.)(73 + m3.) f3 f1? j1(krli"3 — 52]) j1(’€F|53 — f2|) kFlf3 — 552] (9154553 — 552] 2 ) / (fl—:2 £133 (152 (lg—:3 €262.52 81-6333 (7.42) 3 2 p 81 9A ~~~~rr 11:? CC(i)2x,1NM = 1024176 (2f ) f (1332 C1333 (1‘12 d(13 6 q2"’2 6 (,3. ‘3 1r [3 ‘7 x 56171 (121931 ( 4017713r + 203 .. s) — —(12 ' (13 ((1% + "”3013 + "13.) f3 f3 x J'IUCFJI2) j1(kF$3) j1(kF|53 — 52]) (9171132 kF$3 kFlfs — 1332] . (7.43) p3 _ 243 57.4 2C4 _. —. —. -. 7.7.5 127.5 CR'ZXJNM = —10247T6 2ft 712— (11172 dedq2 dq3 8 2 26 3 3 31 ’71 B2 “'2 Q2 Q3 Q2 (13 68171116732721! ((1% + m3)(t1§+ m3.) leko332) j1(kF333) j1(kFlf3 '" 5‘2” (@1132 ($271133 (CPI-”173 — 552] (7.44) As can be seen from Eqs. (7.42), (7.43) and (7.44), all of them require the appli- cation of Gegenbaur’s addition theorem. This is due to the occurrence of j1(kp|f2 — f3|)/(kp]fg — 553]) in these terms that are not separable in 552 and E3. The numer- 164 ical test discussed in appendix 9.11.3 shows that we can truncate the Gegenbaur expansion of this term at fifth order as j1(/.~,.~(3'3 — .721) /144 r(3/2) 5 3 , , .. _. z + — ' 1 k :1? 1.? ~17. kFll‘s—l‘zzl 7T #7133132 gm 2)J’+l( I 2)]“+1( f 3) x CZ/2(cos(6)) , (7.45) where 6 is the angle between :52 and $3 This is followed by analytical integration of the couplings. Furthermore, one notes that Eq. (7.42) requires the application of two Gegenbaur expansions while Eq. (7.43) and (7.44) require the application of only one Gegenbaur expansion. To assess the accuracy of the couplings/EDF in Eq. (7.39) when calculated with the truncated Gegenbaur expansion, we compare the result with the case when the couplings are calculated with the essentially-exact Monte-Carlo integration (without Gegenbaur expansion). In Fig. 7.1, we show the percentage error of the Gegenbaur-based calculation with respect to the Monte—Carlo ones, for the contribution of the three terms of Eq. (7.39) to the energy per particle of INM as a function kp. For each of the three terms, we have two curves where the insets show the actual contribution to the energy per particle when the couplings (Eqs. (7.42)- (7.44)) are calculated analytically with the truncated Gegenbaur addition theorem, at fifth order, and the lower curves (main curves) represent the percentage errors. The constants of the chiral EFT three-nucleon interaction that are used in this particular calculation are specified in Table 2.2, with he = 197.327 [MeV fm]. The results show that, the percentage error resulting from truncating the Gegen- baur expansion is less than 0.5% over the range of physically interesting kp values. In fact, for the double-exchange from the D—like and R pieces, the percentage error shows a strong fluctuation between 0 and 0.5%. For the single-exchange from the D- like term, the percentage error shows a steady increase from 0 to about 0.3%, which 165 $100.0- 5 0.0--— a 3 0.01“. > . > . \ 3 . s g 80.0- g -5.0- \ g .501 \ I—a ) I—l ‘ g 60.0- =-10.0< \ 2.100 30 ' 3’0 ‘ \ 0 ‘ \ :3 40.0: 5-15.0] \ gin-15.0: a. 20.0- 2-20.03 $9-200] \ . . x S 0.0- S - . . - . N . a . . . o 0.75 1.25 1.75 O 0.75 1.25 1.75 °‘ 0.75 1.25 1.75 .1 .1 -1 k1,. [fm ] anm ] kF[fm ] P UI ‘— “—— Percentage error -0.5 0.75 ' 1.25 ' T75 k1,. [fm'I] Figure 7.1: (Color online) The percentage error of the truncated Gegenbaur expansion with respect to Monte-Carlo based calculation of the contribution to E/ A in INM. Upper plots Show the actual values for the calculation based on the truncated Gegenbaur expansion. 166 is not surprising as we needed to apply Gegenbaur’s addition theorem twice in that case. Considering the unavoidable numerical errors/flucuations in the Monte-Carlo calculation and the smallness of the percentage errors, obtained we can conclude that the truncation of Gegenbaur’s addition theorem at fifth order provides a practically exact truncation. At this point, one should realize yet another reason for the need to automate the whole calculation. I.e. the application of Gegenbaur’s addition theo- rem replaces each term in the couplings with about five terms when truncated at fifth order. For instance, in the integration of Eq. (7.42), the single Monte—Carlo integra- tion is replaced with about 25 integrals due to the double—application of Gegenbaur’s theorem. Hence, even though it enables us to obtain completely analytical couplings, Gegenbaur’s addition theorem comes with a tricky overhead: about two orders of magnitude increase in the number of integrals to be calculated. Finally, we remark that the conclusion of this section, viz, the truncated Gegenbaur’s addition theorem enables us to calculate the couplings in a practically exact manner, holds for all other couplings as the truncated Gegenbaur expansion is the only “approximation” that goas into the calculation of the couplings. In the next chapter, besides the possible future extensions and conclusions, we perform preliminary analysis of the couplings and the ongoing semi-phenomenological approach that is attempting to make use of this work. 167 Chapter 8 Semi-phenomenological EDF, Future Extensions and Conclusions 8.1 The semi-phenomenological approach Based on the arguments discussed in section 5.1.1, we advocate a semi-phenomenological approach in which the phenomenological Skyrme functional is to be augmented with the DME-functional. Here, DME-functional refers to the EDF that we obtained from the application of the DME to the Fock energy contributions of finite-range N N and the complete HF of finite-range NNN chiral EFT interactions at N2LO. In this scheme, the Hartree contributions from the NN part are to be treated exactly. Finally, the phenomenological Skyrme parameters are to be re—fit to IN M and finite nuclei properties, leaving those couplings/terms that Originate from the DME intact. Actually, the so—called phenomenological Skyrme parameters can also considered to have originated from the contact part of the chiral EFT N N and N NN interactions. The generic structure of EFT interactions given in Eq. (2.12), VEFT = Vw+Vct(A)-. shows a clean separation between long- and short-distance physics. Consequently, each DME coupling at the HF level can be decomposed as the sum of a density- 168 independent, A-dependent piece, which are subsequently re—fit, coming from the con- tact terms of the EFT NN and NNN (E-term) interaction (VCt(A)) and a density- dependent, A-independent piece coming from the finite—range pion exchanges cf“? = cf1‘2(A;v..)+Cfl‘2(1‘2’;v..), (8.1) 631“?“3 = cf1‘2‘3(A;v..)+cfl‘2‘3(fi;v.), (8.2) where (1 C2 and <19 (3 are bilinear and trilinear combinations of densities that occur in the EDF [[153], [160]]. In this sense, the re—fit parameters can be viewed as containing the effects of the HF contribution from the contact interactions, Vet, plus higher order effects that would arise in a more sophisticated Brueckner—HartreeFock or 2nd-order MBPT cal- culations. In this regard, through the loose connection of the refit Skyrme parameters to the EFT contact terms, the EFT concept of naturalness might provide useful the- oretical constraints for the fitting procedure [163]. The following several plots show sample 6:1 <2. As can be seen from fig. 8.1, the novel density-dependence is controlled by the long-range parts of the NN interaction. Therefore, it is not surprising to see that the density profile of the couplings shown in the figures is driven by the LO term (one-pion exchange) since the NLO and N2LO interactions are of shorter two-pion exchange range. Even though the couplings in fig. 8.1 seem to satisfy the hierarchy requirement, viz, LO > NLO > NQLO, it is not guaranteed that this will always hold. This is because we are including only HF contributions to the couplings, with our focus being primarily on finite-range pieces. The fact that the hierarchy might not be maintained should not seem to be a big problem as HF amounts to comparing the LO, N LO, NQLO potentials (which are not observables) and therefore are they not required to obey any hierarchy. Fig. 8.2 and 8.3 show the CO” and C1” couplings with +/— error bands as de- 169 400 CH [MeV-fms] 0. c> [MeV-fms] '3 I"; H -150 C -200 Illllrffi l I I I l Tllllll — LO ‘ T=0 ——- NLO a _ NNLO . lllllllllllllllllll 0 OJ. (115 (12 p [fm'3] (105 Figure 8.1: (Color online) GP and Cl“ couplings from chiral EFT N N interaction at N2LO. 170 500 I I I I I I I II I I I I I I I II 1 A = 500 MeV SFR — SLy4_Tse1f 400 ——- SLy5_T SkM* w"; 300 IT $31.21 «.7. _ --- TZA % 200 — E _ i: o 100 — 0 ':'.'-L—'.'—.' é-'_-'-_'—;'—_':';—'_'—_'—i __ - I I I I I I] II I I I I I l I II 0.01 0.1 -3 P [fm ] Figure 8.2: (Color online) 65’ couplings from the chiral NN interaction at N2LO with error bands from naturalness requirement. 171 50 TiT TIIIWII I I IIIIIII I r ................................ 7 0 ,_ -50 _______ “a “—1 > -100 a) 2-150 2 _, — SLy4_T self 0 -200 --- SLy5_T _ _ SkM* _ 250 -- T22 - — _ -—-- T44 - _ASFR—SOOMeV ___ TZA _ _ I l I I l I I II I I I I I I III I 300 0.01 0.1 -3 P [fm ] Figure 8.3: (Color online) (31” couplings from the chiral NN interaction at N2LO with error bands from naturalness requirement. 172 termined from the naturalness requirement [163], compared with the corresponding phenomenological values. The error bands cover all phenomenological values in the density region of interest. Thus, the main conclusion that can be made at this stage is that the DME couplings are close to phenomenology, but with a novel density- dependence, as long as one allows for natural-sized contact terms. For a detailed discussion, refer to Ref. [153], while for the corresponding discussion on 6:19:31 refer to Ref. [160]. The first calculations following the semi-phenomenological approach advocated in this section are underway [164]. Figs. 8.4 and 8.5 shows one of the exploratory “results” regarding the saturation curve, W(p, I), of INM and PNM. Here, I = (pn — pp) / p. The parameters of the DME—based functional used for the saturation curves are not optimized, rather they are simple educated guesses. Preliminary indication from this study is the Skyrme functional that is augmented with the DME functional is more flexible in that it relaxes some of the interdependencies that one observes in phenomenological functionals [164]. 8.2 Key future extensions In this section, we revise the main directions in which this work can be extended in the future. These are 0 Extensive self-consistent test of the PSA-DME. As discussed in section 5.4, our tests can be judged to be extensive only for non self-consistent ones. Along with the self-consistent tact, the invention of a local N N and N NN chiral interaction that is soft enough to be used for these tests is important. 0 Studying the impact of the different DMEs on the couplings of the resulting EDF. Note that the non-self consistent tests that we performed in this work are using schematic interactions and it will be beneficial to extend this and perform 173 W(p) .. ’ " SLY4 50 _ : SLY4—”.0 F SLY4+LO+NLO w ’— ~ SLY4+LO+NLO+N2LO ; L- 30 >— )- f p 20 ~ I0 7 I F 1 L 1 l n i I I 1 I 1 4 1 0.1 0.2 0.3 F' - IO *- f Figure 8.4: The saturation curves, W(p, I ) of INM using the phenomenological SLY4 functional and semi-phenomenological DME-based functionals. Here, N2LO includes the contribution from both NN and NNN interactions (From Ref. [164]). W(p, 1:1) - - - SLY4 100 : SLY4+LO ~ SLY4+LO+NL0 80 : SLY4+LO+NLO+N2LO )- w i— L 40 ... b O )- 20 r J L J l L J l A A A A 1 A I l I l l l l l l L I 1 L I A 1 0.] 0.2 0.3 0.4 0.5 0.6 Figure 8.5: The same as Fig. 8.4 but for PNM (From Ref. [164]). 174 p extensive comparison of the actual couplings that result from the application of different DMEs. This should provide a better estimate of the associated DME errors /uncert ainities. o Non-self—consistent and self-consistent tests of generalized PSA-DME. This is important to gauge the accuracy of the DME approximation that we made to the HF energy from the chiral EFT NNN interaction at NQLO. Even though we have shown in appendix 9.5.3 that the approximations that we used to obtain the generalized PSA-DME are equivalent to the ones used for PSA-DME, the existence of more than one non-locality coordinate may change the relative accuracy of generalized PSA-DME with respect to PSA-DME. Furthermore, the effect of DME-coordinate optimization parameter a in Eq. (7.24) should be investigated. 0 Calculation of Bogoliubov contribution from NN + NN N , extension of PSA- DME or its variants for pairing densities. Furthermore, the required renormal- ization should be designed along with the DME. 0 From the interaction side, the extension should include the contributions from the N3LO component of chiral EFT interactions. It should be noted that, even limiting the calculation at the HF level, the four-nucleon interaction which first appears at this order will make the extensions more complex. In principle, one needs to incorporate the contribution from the four-nucleon interaction. Nevertheless, current estimations of its effect on nuclei, at least in light-nuclei, suggest that it can be ignored safely. For instance, estimates in 4He show that the additional binding energy it provides is of the order of a. few hundred keV [165]. 0 Extension of the DME scheme to approximate higher-order contributions as 175 discussed in section 5.1.1. 0 Analysis of self-interaction and self-pairing issues that arise in the context of the DME [105]. 8.3 Conclusion This work is a part of a long-term project to develop nuclear EDFs starting from many—body perturbation theory and the underlying two- and three—nucleon interac- tions [[110]-[154]]. This is necessitated by the fact that empirical EDFs lack solid mi- croscopic foundations and often result in uncontrolled (i.e., parameterization-dependent) predictions away from known data. We used the DME as a tool to explicitly build microscopic physics associated with long-range pion-exchange interactions into existing Skyrme functionals in the form of novel density dependencies. An important component of this endeavor is the improved PSA-DME and its NNN counterpart, the generalized PSA-DME, which are crucial if we want to provide microscopic guidance to the description of spin-unsaturated nuclei. The rich spin/isovector dependence of pion-exchange interactions gives us hope that their inclusion via the DME will give valuable microscopic constraints on the isovector properties of the EDF. Moreover, it is comforting that these constraints are coming from the best-understood part of nuclear interactions. The EDF obtained as a result of the present work contains only HF physics such that further correlations must be added to produce any reasonable description of nuclei. In the short term, such an addition is being implemented empirically by adding the DME couplings to empirical Skyrme functionals and performing a refit of the Skyrme constants to data [164]. While this is a purely empirical procedure, it is motivated by the well-known observation that a Brueckner G-matrix differs from the starting vacuum NN interaction only at short distances. Therefore, one can interpret 176 the refit to data as approximating the short-distance part of the G-matrix with a zero- range expansion through second order in gradients. Eventually though, it is the goal of the UN EDF (universal energy density functional) project to design a generalized DME that is suited to higher orders in perturbation theory and move closer and closer to complete microscopy. 177 Chapter 9 Appendix 9.1 Mathematical Formulae In this section, we list the Mathematical definitions, relations and formulae that have been used in the rest of the work. Only the relevant mathematical relations and formulae are listed, and for a more extensive list, refer to classic references such as [[152],[151]]. 9.1.1 Miscellaneous elementary formulae In various parts of this work, we use the following general linear coordinate transfor- mation. Starting with two coordinates (f1, :52), we define a new coordinate system (2?, 2?) as f=f1—f2 X=(1—a)f1+af2, (9.1) where the unspecified parameter a is a real number satisfying a 6 [0, 1] . The corre- sponding gradient operators are given by 6,1, = a6 —(1—a)€71 1'1 2? 178 VX __— V_.1'1 + 61‘21 (93) with a = 1/ 2 recovering the usual center of mass and relative coordinates. In the derivation of local densities, detailed in appendix 9.4 and other parts of the work, the following elementary results are important. ... _., ., A 2 VV (7' '7') 17le = :2“, (9.4) ,. A, 7'; 6,37" —Ti7'j .. (9;,(7' r) = 2-7( 7‘3 )k (95) ii -—1- d0~(7.A”)(7 B') — 12.4.1? (96) 47f r 7‘ 7' — 3 . The manual derivation of the HF energy from a generic two-nucleon interaction involves a modest amount of spin-isospin algebra. First, the Pauli matrices are given They satisfy the commutation and anticommutation relations 0'in - O'jO'i = i2€ijk0k7 (9.8) 030,- + 03-0,- which can be used to prove (A-3)(§-5) .—. (X-B)I+ia-(A'x§), cos(a.) + i(7’i - ('7') sin(a,) , 179 (9.10) (9.11) for any two vectors A, B and A = (1A. Additional, elementary relations are 5..., = 0 , (9.12) 2051'. a 050,, = 221' for all A’. (9.13) 9.1.2 Clebsh—Gordon, Wigner 3—J and 6-J coefficients Representing Clebsch-Gordon, Wigner 3-J and 6—J coefficients by . . . J1 J2 J J1 J2 J (Jm I J17711J2m2> , , , (9.14) m1 mg m 771.1 771.2 m respectively, the following is a list of the relations that are important for different parts of this work. 1 2" 1 2317721500772)? = 2] :1 , (9.15) for all m and m, such that m = 7m + a. . . . - - . J1 J2 J (J1miJ2m2lJml = (“llmfllfi2 v 2] + 1 (9-16) 7711 7712 m E ( 1),, J1 J5 J6 J4 J2 J6 J4 J5 J6 m4m5m6 m1 m5 m6 —m.1 772.2 7716 7714 -m5 m6 J1 J2 J3 J1 J2 J3 = , (9.17) j4 js jfi m 1 "2.2 m;; where a = 3'.) + 7m + J}, + ms, + 3'6 + m6 . 180 J1J2J ,---J1JJ2 = (_1)11+J2+J (9.18) 7711 mg 771 ml m 7712 L l 1 l L 1 1 Z(2L+1)f(l,L) = ——(—1)’ l(l+1). (9.19) L 1 1 z 0 0 0 V5 where —l 2111—13 if L = 1+ 1; f(l,L) = —(l+1) fl iszl—l; . 0 otherwise. I l 1 : (-1)'-mz m’ 5... _m, (9.20) mi m; 0 \/l(l+ 1)(2l+ 1) l’ 1 l l 1 (l—m1)(l+m1+1) = —1 "ml 6 9.21 m: m; 1 ( ) \/ 21(1 + 1)(21 + 1) "’l""‘z"‘ ( ) l l 1 _ :1)-ll+m +1 (1+ ml)( )(l ml + 1) m .4” [+1 (9.22) mi ml, _1 21 (1+1) (21+ 1) z~ 1 1 1 J x/E \/l(l+1)(2l+1) 181 9.1.3 A few special functions Legendre polynomials Starting with the associated Legendre differential equation, for integer l and m, 51— [(1—1132)P,"‘($)] + [l(l+1) — m2 JPNJT) = 0, (9.24) d2? 1—2:2 where a: e R (the set of real numbers), P,"‘(a:) is the associated Legendre polynomial. The associated Legendre polynomials satisfy —m m (l — m)! m Pt (17) = (-1) WP: (ll (9-25) For m = 0, the differential equation given in Eq. (9.24) can be reduced to (ll—1:2)i2—dI-Z—(2£2 — 22:52}? + l(l+1)H(;I:) = 0, (9.26) where H(:L') is Legendre polynomial of order I. The first few Legendre polynomials are Petr) = 1. (9.27) p101.) = 1;, (9.28) 102(2) = $13924). (9.29) The Legendre polynomials are orthogonal over the. range (—1, 1) and satisfy 2 - = m Omn . (9.30) / da: Putz) Pmo) 182 Additionally, Pz(1) = 1 for any I. The derivative of Legendre polynomials satisfy the following properties (1P 1 l 1 ’(x) = ( + ) , (9.31) :L' I=1 2 "2 — ldP . l [(73) = 2‘P,(;r.) — 131-1(2). (9.32) 71 d1): Laguerre polynomials The Laguerre polynomials are solutions of the Laguerre differential equation (12 Ln(:r) (11),,(93) a: (13:2 + (1— ) d2: Ln(:r) — 0. (9.33) The can also be defined using Rodrigues formula 0 , _ x—a ex dn -—.r n+0 Ln(:L) — n! dx" (6 :1: ), (9.34) where L303) is the associated Laguerre polynomial. The Laguerre polynomials are recovered by setting a = 0 Ln(:1:) 2 L200). (9.35) The first few associated Laguerre polynomials are given by L3(;r) = 1, (9.36) L?(:r) = —x + a + 1, (9.37) ..2 ' 2 . 1 13(2) = 32— — (a+2):1: + (“Jr )5” ). (9.38) 183 Gamma functions The Gamma function, which appears from the extension of the factorial with a down- ward shift of the argument by 1, is given by r(.r) = / dttT-le-fi (9.39) 0 where :1: is a complex number with a positive real part. In this work, we need only the Gamma function for positive integer arguments, which is given by I‘(n) = (n — l)!. (9.40) Spherical harmonics Spherical Harmonics are eigenfunctions of angular momentum operators L2 and L2, and are given by We. 9) = \/(2’4:(1,)ff g)?” 1D("(00899) (9.41) with their orthonormality relation being give by 11' 27r / d6 (131/ma, a) Y,';"(9, a) = (SH/6 . (9.42) 0 m m I 0 There are various relations satisfied by spherical harmonics. and which are of interest to this work. These are Z Kmr(f',)Y;"-Il (7") = 471- 131 (,1! . 1A,) 9 (943) "II 184 where the sum extends over all allowed values of 112.). ram 6 (f9) 131,49] 2 ; ___—— Yz’"(v") + fir) V7Yz’”(7‘% 07' where f (r) is any function dependent only 011 7“. ml . 1 , . V11)”. (1) = -,: Z f(l. L) (um mlLM>Y£’(r). LM where —z 2% ifL=l+1; f(l.L)= —(z+1),/,,’—_1 ifL=l—1; 0 otherwise. K7116, ¢) = ¢<2l+1)(l_ m)! Bm(COSB) eimgb, 47r(l+m)! 03,1109, 0) F (155,1) me. 0) , —m , _ m (l — m)! m P. (x) — (—1) ((+m)! P. (x). Bessel functions Bessel functions are solutions of Bessel’s differential equation 2 . $2.1 Jam) + deaa) _. 2 _ 2 = (11.2 (1.17 + (.1: a )Jc,(a:) 0, (9.44) (9.45) (9.46) (9.47) (9.48) (9.49) where a is a complex number and Ja(:r) is Bessel. function of order a. In most physical problems with spherical symmetry, 0: takes half-integer values, a = n + 1/ 2 . 185 Consequently, one defines the spherical Bessel functions of integer order, n, as . (71 n 1 (I "sin(.r) r Jn(I) E ‘27'In+1/2(l') E (—$) (I E'l') fl? - (9.00) The first few spherical Bessel functions are jotr) = (9.51) 11(1) = 12(1’) = l E ' ECO 13(4) = (13") 5‘“ NE}— —)C—(—°: ), (9.54) ( 105 45 3:) sin( :13) (105 10) cos(1) + 1 — . 331(1’) = 2:3 :r :17 9.1.4 Three-dimensional spherical harmonic oscillator eigen- functions The isotropie three-dimensional harmonic oscillator is described by Schrodinger’s equation h 1 —-2-EA + 5771922 7‘2 (b,,),,,(r, 19,99) = 6n1¢n1,,,(r,6,.¢) (9.56) The wave-function is separable in the radial and polar coordinates as 12.. ... ¢1l_1n,,(7‘,9,¢) : :_(r) Y1 (6,96) a (9'57) where 1 2(n — 1)! ”2 1+1 _ .. 1+1 2 R1, = — .- 1/21; / ., [(7‘) r \/31"(n+l+1/2)3 fl 6 ""1 (r) a: = 137"? 186 7714.) n e...) = hw(2n+l—1/2). (9.58) Note that there are two conventions in use regarding the possible values of integer n: n 2 0 and n 2 1. The latter is used in this work. 9.1.5 Gegenbaur expansion Gegenbaur’s addition theorem of bessel functions of the first kind reads TV ml/y JV(T) _ 2” F(:’) 2 (V + H) Ju+n($) Ju+u(y )CV(COS(0))’ (9'59) p=0 where u > 0 and for all values of :17, y and 6 (the angle between :E and 37). The variable r is given by r = \/.T2 + y2 — 2:1:ycos(6). I‘(V) is the Gamma function and C: refers to Gegenbaur polynomials. The first few Gegenbaur polynomials are given by 05’ = 0, Cf” = 3cos(6), 3/2 _ _E 15 COS2(6) C2 _ 2 + 2 a 03/2 _ _15 003(6) + 35 cos3(0) . 2 2 (9.60) A formula related to Eq. (9.59) is J_,,(7‘): 2”F( (II—___) V W D—I (u + 1)4)1_.,_..( >449) 0.144(6)), (9.61) p=0 and it holds only when Iyeiwl < |;L‘|. Combining Eqs. (9.59) and (9.61), one obtains the relevant expansion for Bessel functions of the second kind and those of the mod- 187 ified Bessel functions. For details, refer to [195]. Gegenbaur’s addition theorem is a key ingredient for the analytical calculation of the EDF couplings obtained from the application of the DME to the HF of chiral EFT three-nucleon interaction at N2LO. Refer to section 9.11 for details. 9.1.6 Functional derivatives A functional maps functions into a number. Analogous to the derivative of functions, one defines the functional derivative of a functional, F [f (13)], with respect to f (:17) as 6F . 0F[f(:l/) + 8591-33)] - F[f(y)l. = lzm€_. 5f(117) E (9.62) The functional derivative satisfies several relations which are analogous to the ones satisfied by the derivatives of functions. For instance, if F and G' are two functional of fix). 6(F G) 6(F) 6(0) 5m) 2 CW + FW’ (963) while for F[f(:l:)] = I: (11' [f(:r)]". 6(F) _ TL ’13 n—l W) — [11)] - (964) 9.2 The one-body density matrix and densities The basic quantity in the EDF approach for nuclei is the OBDM and the various nonlocal and local densities that can be extracted from it. An extensive discussion of these basic quantities is given in this section. 188 ' 9.2.1 Properties of single particle states The wave function of a particle having spin S is a spinor of rank 25', i.e. is composed of 25' + 1 components. The particles constituting the nucleus are protons and neutrons which have spin, S = g. The single particle states are assumed to have a good isospin projection, but mix spin states We use a = i%, q = :l:% to designate the spin, isospin and i, j, k... the remaining quantum numbers of the single-particle states respectively. Thus the single-particle states can be designated as I 2'4) --— XI 2'04). (9.65) In spinor notation, the single particle spinors are given by four real functions 1,91 , ..., 1,9,, , _, 996(FU = +6 (I) 991,i(FQ) + i 992,1(‘F9) (7129) = 994(7‘9) = _‘ 1 = _ fl . (9.66) 1,9,0“ 0 = -5 (I) 993,i(7‘ (I) + 2 9911,1(7‘9) The orthonormality and closure relationships are given as / (1179-9: (.41) 44-974) = 5.6%,, (9.67) 2 (01(F0q) ,9,()="a'q’) = 6(F— F’)6w,6qq, . (9.68) It is important to characterize the single-particle properties under time reversal. The time-reversal operator is given by T = iayKo where K 0 denotes the complex conjuga- tion operator and cry is a pauli spin matrix. Thus the time—reverse of the single-particle states is (T90). (F09) = 20901-5559). (969) where 6 = — a. For an extensive discussion, refer to Ref. [81]. 189 9.2.2 One-body density matrix The one-body density matrix can be written in (Faq) space as pq(Fa.F’0’) = ((PIC (7 a’cq)(roq)l<1>=Z 9021 7" ’0’)q 991900) 937,-, (9-70) where N?) defines the many-body wave function and p2, = (<1) | CI c_,- | (D) defines the density matrix in the basis {Ci/1,9,}. Since the single—particle states have definite isospin quantum number, the density matrix is diagonal in isospin subspace. The density matrix can be separated in its sealar/vector-isoscalar/isovector parts , 1 p844? 0') = 1 [42866) 6..) + 40(66').6..I + (—1)1/2“’(p1(f’,7"’) (5M; + .9'1(F,F').6wr)] , (9.71) where the scalar-isoscalar, scalar-isovector, vector-isoscalar and vector-isovector parts are respectively po(.7'"’) = qu(r0r”)0 )6OJ=ZZ¢:(T’09) 99) ”19)/2). aa’ q 09 ii = EPA-‘17?) (9°72) p.626") = Z pq('F0,F'0’)6 .. =ZA(4) 214994) 46606;); I ij = z 9607‘"), (9.73) 56(11‘”) = Z pq(Fa.F’0’)5’.,/a = Z 2 99217" ”0(1) 991(FUQ) 50/. p3.- ii 00 q 00’ q = Z 6;,(177’), (9.74) §1(F,F') = Z pq(_'0,-f’0')a :2, A( (q) .2]- 99,-( 7‘ ")0q 1,91-(r0q)00 I 0p],- 00’ q aa'q 190 Z A(q) 6;,(71 7"), (9.75) 1- where A(q) = (—1)? q . The extraction of the scalar/vector—isoscalar/isovector parts from Eq.(9.71) can be done easily using properties of the Pauli matrices given in Eqs. (9.12)-(9.13). 9.2.3 Local densities Working in neutron / proton representation and taking derivatives up to second order, the following local densities that can be formed from pq(7"', 7"") and §q(7"', F’) pq(7‘) = ZwJ 1(1(7'9) p.J-. (976) 76(7) = ZVwJ-(Fq) -V|(%|>* 1—<lf>l<1>>* (914,619) (914,699) (9.92) where l’ andl are elements of the single particle configuration space and |) is a quasi—particle vacuum [38]. The operator p“ is hermitean (p‘l = p”) and R. is skew T anti symmetric k = —k. Here kT refers to the trans ose of it , not the time P reverse of Fe. Two important relations hold for )6 and F; . These are fife = ftp“. (9.93) Using the above two relations, it can be easily seen that the generalized matrix R is idempotent i.e. R2 = R. In fact, R is also Hermitean. The pairing tensor can be transformed into the pairing density matrix, )5. This is given in (Faq) space as -'III 6 Foqn‘ aq E 26’ (I) 6.1-; (2,7,, (I) = 25’H(F0’(],'F’5"(1’). (9.94 7 aq’ (I We still assume that the density matrix is diagonal in isospin subspace. In pairing terms, this assumption means that there is no proton-neutron pairing. This implies 193 that the pairing density [3(Faq, F ’ 0’ q’ ) can be written in the form [1,, (Fa, F’a’) (5 / qq' The pairing density matrix can be resolved into its scalar/isoscalar, vector/isovector parts in exactly the same way as the normal density matrix. Thus, appropriate contractions in spin and isospin space yield 2 pq( (F0 "’ a’ )6 =222699.(7’ 69) 996(F69)77" 75.. 00 ,q aq ij M7"? F’) 1 (9.95) q 2 6q(Fa,F’6’)6JJ/A(9) = Z A(9) Z 26 94(F’69)91(F69)K§., 09 if 2 A(9)f)q("1F’) . (996) )6,(Fa,r “’ a’) —2 2 26’ 991(F’6’9) 930690. a 63.7 2 6",(7', 7") , (9.97) q 2 6,60, r'a’) (7.7. 4(4) aa’q 2 Mg) 2 26' 4247071) 44904) 4.7.. ii 00’, q 2 s,(". F’) (9.98) (1 Working in neutron/ proton representations, and going directly to the local densities, the following is the list of the local pairing densities that can be formed by taking derivatives up to second order = 2 2 26 9.- (F69) 99j(FUQ) 773%. (9-99) 2 2 :2: 2a ,(mq) VqJJ-(Faq) R3,, (9.100) = 2 2:: 26’910"9)699j(T0(1) ua 0.0! ij 4.996279) 9j(mq)v99i(7'aq) + V), 991-(Fc’7'q)6- (a’lé’la) ap,(7"aq)) 7117-. (9.106) Jl One notes that most of these anomalous local densities are not used in current empirical parameterizations of nuclear EDFs (see section 4.2.2). In fact, only the local pairing density, fig, is used. As discussed in section 5.3.8, the application of the DME to the anomalous part of the OBDM results in these local densities. Thus, these densities may be useful in future non-empirical construction of the particle- particle/ pairing part of nuclear EDFS. 9.2.6 Relations among the densities One can establish a number of relations among the various local and nonlocal densities defined in the previous section. We start with those relations which are important in the derivation of the DME of the scalar/ vector components of the normal part of the OBDM. These are [(V2 + V0)pq(7", F’)] = V2pq(F) — 2Tq(F) (9.107) sz’ 2 Wm? r') = $172M)— W) -19 m H (9.109) 2 $6M?) +134?) (9.110) [(v2 + v”) Mm F) 4 = stmh“) — 2T,,,.,(F) (9.111) [Vt-91,407,?) F—F’ : [Vim/AF fl] *4 = évpsu,q(1r)+w,w,q(r). (9.112) Most of these relations were initially given in Ref. [197]. Here, we extended the list by deriving additional relations which are found be useful in the derivation of the generalized PSA-DME, discussed in section 9.5.3. We illustrate the derivation of these relations by taking Eq. (9.107) as an example. Starting with the left hand side of Eq. (9.107) [(v2 +v'2)p.,(r,f~)] 9 ~=r [) in space, spin and isospin coordinate space A (:1:1,172,...,1t,,,t) = exp{i§:¢(;rj)} (:c1,;r2,...,:r,,,t), (9.116) j=1 where x,- = (fl, 0,, q.) and (9(13) 2 45(1‘}, t) is an arbitrary, differentiable real function of the position 7" and time t. In general 975(13) are independent of spin and isospin coordinates, and in the static picture, they do not depend on t. Second quantization formulation For generalization of local gauge transformation to other many-body approximations where there is no explicit conservation of particle number, one has to formulate local gauge transformation in second quantization. Since the state vectors in second quanti- zation are elements of abstract Hilbert space or Fock space, local gauge transformation must be performed by an abstract unitary operator [198]. When the transformed- state vector is projected in to the N-particle subspace, the transformed wave-function in the subspace is equal to the projection of the original-state vector times the ap- propriate space and time dependent phase factors. In second quantization the state 198 vector |), which is not necessarily an eigenstate of particle number, is represented in Fock space by the column vector |) 2 col {|0), |1), |"), ...}, (9.117) where |i) refers to i-particle component of the state-vector. A local single-particle operator 95, which is assumed to be diagonal in isospin space, can be represented in second quantization as an operator in Fock space 2 2: / dfl'FalqlcblFU-qua*(F019)a(Fa-2q). (9.118) ”1‘72 ‘1 Defining the unitary operator U as U = exp 2': Z / drIFazq>a*(ralq)a(razq) . (9119) “1‘72 ‘1 the local gauge transformation of the state-vector is given by [199] |)' = U|). (9.120) One can easily verify that this gives back the previous formulation when applied to an N -particle Hartree-Fock wave-function. Once the unitary operator U is defined, local gauge transformation can be carried out by transforming the creation and annihilation operators (1'(Faq) :2 Ua(f"aq)Ul = (”WU") (1(F'0q) al,(f'0q) = Ua*(r~'aq)U* = eWWaMFOq). (9.121) 199 9.3.2 Local Gauge transformation of normal densities The local gauge transformation of the normal part of the OBDM is p; (17101, F202) 2 exp {i(¢9(f‘]) — 99030} pq('F101,1-"202). (9.122) When the various local densities involved in the EDF are calculated from the 10- cally gauge-transformed density matrix Eq.(9.122), one obtains the following relations [200] 10W) = (90’) (9-123) 999 = rq<9+2L<9-vo<9+p.(9(v¢(9)2, (9124) 3529(6) 2 62,907), (9.125) 1.3.99 = j...99+p.<9v..¢(9. (9.126) T4,.99 = T....99 + 2 Z J......99v..999 + 3,..99 (W99). (9.127) J5....99 = J,,,,,.(.-)+s,:,(.=)v,,(.-), (9.128) Fé,u(f) 21701474) +VV¢ (F) Evil-9% 7:) 991a (7") +Zvu¢( flJunF) + v., Zyml( nljm mla 1 x (ln1,§0|jn1)466' (6'|6|6) . In the orthonormal coordinate system defined according to 60:521 Gil ::F—(é..ri“7y) (9.147) A) 1') (9.148) (9.149) and using I 1 <5’l011l5l = «3501145091 (9-150) one can write the nonlocal spin density as 71)an7‘ m’: m M = $2 ,(1 3;” Z 1119')Y1’9=)(21+1) "(1m mlaml’a’ X (_1)2l+2m+130-2 2 J 2 J m, a ——m m, 0’ -m l 1 1 x 2 2 . (9.151) 6 11 0’ Making use of the 3j—6j symbols relation stated in appendix 9.1.2, one obtains 11. .1 . l/nq.'( 7") n ' m, 1 W() = 193me ’;, ’3 Z 11;"! Y1""1~()(—1)“t+ nlj m m, l l l l 1 l l 1 )( x , . (9.152) é % j m, m,’ 11) One can plug in the algebraic values for the 3j and 6j symbols in the above equation and do further simplification. To proceed further than that, it is imperative to choose a coordinate system in which one of the vectors (7‘,r ’) is along the z—axis. Let 7" ml, :0 be the one along the z-axis. This implies that 6’ = 0 and (9’ = 0. Thus, only Ylm contributes. Making use of the relations in Eq.(9.20)-Eq.(9.23) and simplifying, Sq,0(77,7?l) = 0 ....1 . 111 . V1311” V1310") -1, 1.. 1 1 s,,19,r) =—— 223—1)] (21+1) 1' 1 Y, (1)1», (6:019:01 nlj ><3/4+l(l+1)—j(j+1) fil(l+ 1)(21+ 1) 206 V" (1') V" (r) 81.102?) = iZ(—1)21(21+1) "Ii, "'3 Y‘(7‘)Y°*(6’ =0,¢’=0) nlj X3/4+l(l+1)— 1(1+1). ([21(1+1)(21+1) One obtains the prefactor 7'. after properly summing the exponents of —1 which takes the form (—1)“ where a 2 2J' + 3/2 or a : 2J' + 1/2. Writing the components in 1‘, y, z coordinate space, one obtains sq,z(7"',7'"") = 0 (9.153) saw") = [Z (7")"‘J()P111°<0'=o,1'=0> nlj xcos(¢)(3/4+l(l+1—) j(j+1)) 1 l\/(1+ 1)(21+ 1) T, (1 31.1121") =—\/4:D21)"3”)31(cos(1)mo(1'=o,1:0) nlj xsin(cb )(3/4+l(l+ 1)—J (J+1)) (9.154) (9.155) 1 1\/(1+ 1)(21+ 1) ° The above result can be used to show that the nonlocal spin density is in the direction of 7" x 7” . This can also be shown to be true from a different perspective : using the properties of nonlocal spin density under time-reversal and symmetry arguments. In spherical systems, the general form of the nonlocal spin density can be constrained as follows. There are only three vectors available for the construction of any vector physical quantity. These are 7", 7“" and 7" x 7"". Thus §(7",7"’) = sa(r,7",0)7" + sb(7‘,7",0)7""’ + sc(r,7",0)7"x 7’", (9.156) where the sa(r, 7". 0,) 91,3(7 7" ,0) and 36(7‘, 7" ,0) are scalar functions dependent 011 the magnitudes of" 7‘, r’and 0 which IS the angle between 7‘ "and" 7" In spherical systems, 207 the nonlocal spin density should satisfy .,§',,(F 7'") = .§'q(—f', «~F’). (9.157) This condition can be satisfied only by setting all except sc(r, 7") to zero. Thus the nonlocal spin density is proportional to F X 1"". To further constrain the form, let us invoke the property of nonlocal spin density under time-reversal invariance. Under time reversal #015") = —§'*(r,r') = —§q(r',F), (9.158) from which one recovers that §q(f') = 0. Using the above property, one can easily show that the nonlocal spin density has only an imaginary component. Thus (1(7‘37’”) = if’x F’sq(r,r’,6). (9.159) This result has been verified by the derivation in Eq.(9.153) : _ Vigil? anj; (7‘) 0 1311(C036) sq(r,r, 6) — —(/4—17}- 2 Y (O O)—sin0 23+1)(3/4+l(l+1)j(j+1)) l(l+1)\/2l+1' (9.160) In ref. [170] a similar expression is given for the nonlocal spin density with sq(r, r’, 6) being 1 K? 7" V:. r sq(r,r',0) = :t— 2 if]: ) if: )P,'(cos6), (9.161) 21r , n1] where :1: is for j = l :t 1 / 2, P,’ is the derivative of Legendre polynomial P, and 6' is the angle between 7" and 7”. One can show that Eq. (9.161) reduces to Eq. (9.160) by using the relations Eq.(9.46) given in the appendix. Obviously, in time—reversal 208 invariant systems, the local part of the spin density is zero. I.e. 3:,(F) = 0. (9.162) Spin-orbit density Starting with the definition of the local spin-current tensor Jthi/(f‘) : “% (V); ““ VL) Sq,V(F,F,) _, _,,, (9.163) and making use of Eq.(9.159), we can write the local spin-orbit tensor as .. Ta Jq.pu( r) = — rsq(r) Z 6W0 7. (9.164) 0 One can write JQWU‘) as a sum of pseudoscalar, (antisymmetric) vector and (sym- metric) traceless pseudotensor parts _, 1 9 1 z 9 1 , Jq.w( 7.) : 35W JéO)(r) + E Zéwk Jé‘lgh‘) + JéizuU‘), (9.160) k=1: where the three components read ngm = EdeqWW) (9.166) pu 153m) = Zewkttm (9.167) pu 1 J96) = Jam— 3 _. 1 . _. 6,”, JéO)(r) — 5 26“” J5,1),)(r). (9.168) k=:r. Combining the above results, it can easily be shown that both the pseudoscalar and the pseudotensor parts are zero i.e. J (0)07) = 0 and J(2)(f') = 0. Thus one needs to simplify only the vector part. Even though one can perform a series of angular 209 momentum coupling operations to obtain the most simplified form for the vector part of the local spin-current tensor in spherical systems, a simple physically motivated derivation is given in Ref. [201]. In spherical systems j;( F) must be proportional to F. Thus K'1 \_/ ‘3 l‘31 = ([15), (9.169) resulting in ’5 ' "'_-_. J—fi IL- L(F)= r 2(2j+1)[j(j+1)—l(l+1):](V,,,,())2. (9.170) 9.4.2 Expression for the anomalous densities in spherical sym- metry For the anomalous part, starting with the anomalous density matrix as defined in the traditional representation [81] pq (7'0” Fla!) = _ Z 26Unqu[k](T—r 0)anjq[k]* («I6 I) nljm U 2 )(lm,’% a’ljm) 26’—1”—(-—) nljm T x Zle'O" [)(ml— éaljm) 26. (9.171) mla For the following anomalous densities, we follow exactly the same mathematical steps as their normal counterparts. Thus only the results are stated. The two changes are (i) overall sign becomes opposite to that of the normal densities (ii) one of the lower part of the quasiparticle wavefunction is replaced with its upper part. Thus ringing 210 the change in the respective densities, one obtains the following results. pairing density The nonlocal part of the pairing density reads Vq fig“? _ _ :4 23.71. +1 nlj‘(,r 70’) Unlj(r)P P('f° 'f’) , (9172) 7' nlj while for the local part, one simply sets 7‘" = “F in Eq.(9.172) to obtain ‘3‘](7?) : —Z 2.74:1 anj( 12:1] ”(7) . (9.173) The local pairing gradient and laplacian densities are given by _ 2.7 + 1 8 anzgl )UZUU') . _. —Z M 0r r2 7‘ (9.174) 111] ~ 2j+1 02 2 .. "(r)U" () qum = _Z 4” (W + ;)V ’9 r2“? . (9.175) 1in Pairing kinetic density The local pairing kinetic density reads = 222;: [<99 - ——> (___) ll(+1) V9.1 )Uz..( >] (9.176) Pairing spin-orbit density The pairing spin-orbit density is given by fir):— 3 7220141)j([(j+1)-l(l+1)-4 14:1,()U:'.,,<-> (9.177) nlj 211 9.5 Details on the density matrix expansion In this part of the appendix, we derive and discuss the generalized PSA-DME. First, we start with brief discussion of the Husimi distribution and the derivation of the quadrupolar deformation, P30"). This is followed by few remarks and derivations related to the Wigner transform of the pq(F1, F2) up to 52. Subsequently, we derive the generalized PSA-DME, from which we recover all the special cases such as the PSA-DME discussed in section 5.3, the original DME of Ref. [170] and its subsequent generalizations [202]. Finally, we give the formal modified-Taylor series expansion, discussed in section 5.3.6, of all the local densities. 9.5.1 Husimi distribution and the local anisotropy P51 (7") The Husimi distribution is one of the many quantum phase-space distribution func- tions. It possesses the key property of positive definiteness [[203],[169]] and is defined as g .. %5'(F*71)—#2(F‘1) .. 11.6.17) N jam-(rune 0 dn pa. (9.178) where N E 1/(1r3/‘4rg/2) and To is a chosen parameter. In the following, we use the HF‘ single-particle wave functions. The occupation probability of a given spherical shell pqnj’ is one or zero, except for open-shell semi-magic nuclei where the so-called filling approximation provides the valence shell with a partial occupation. Modifying the derivation to include pairing (HFB) can be done through the proper formulation of the Wigner distribution as given in Ref. [204], as the Husimi distribution is a COarse-graining of the Wigner distribution using gaussian phase space factors[169]. To derive Eq. (5.15) for the quadrupolar local anisotropy of the momentum Fermi 212 surface P2”('F') we start from the definition (13' 36;.- 2— *2 H, '1, P.2"(F) —:- f ’ [ (11.721161?) (T ’37, (9.179) and make use of the relations 2'. -1 s , ,_. _. J ... [(1131) e WWI—"1’: (27r)“h"V{-V0(F1’— v1), (9.180) _ 1 F _,...12 312(11)~6r—*' k" ‘2 6 ~ (7‘1 7'1) + O(( Fro) ). (9.181) Through direct application of the above relations, one obtains [@152 1147113) x (279355 ZIF¢1(Fq)|ZP§-’i + 0((k%7'0)2)1 / (190.13)?qu w (270312"Zlcvwrrqn2pz.+0066?) which, plugged into Eq.(9.179), gives PM = [ 37321617) lep..-1]+O((k%ro)2)- Further simplifications can be performed for spherical systems, using single-particle wave-functions essentially the same as the ones given in Eq. (9.140). However, note that we are working in the HF picture. Using the angular momentum relations in section 9.1.2, one obtains 213 Ill ‘2 + ZFUJ )(L—nlj2(r)) pqnfl, 7in where F (l, j) is given in Eq. (9.1.3) and pm”l is the occupation probability of the (71,1, j) shell with q labeling the protons/ neutrons. Plugging these intermediate results into Eq. (9.182) yields the expression of P2(F) as P2(r : 7,15)" 227r +1[(g_3V.?_z_-_;( ))2_z(z2:21)(v,3;( ))] pq,,,-,(9.182) where Tq(F) is the kinetic density as defined in Eq. (9.77). 9.5.2 Wigner transform of the pq(F1, F2) up to it? In the parameterized PSA-DME of 5,,(F1, F2) , we used the Wigner transform of pq(-F1, F2) up to h2 to motivate the form of Eq. (5.41). In this section, we derive Eq. (5.40). We restrict the derivation to the HF approximation. Refer to [204] for a recent work on the formalism of the Wigner distribution in systems with pairing (HFB). One can write the scalar part of the normal part of the OBDM as 0377,?) p( W2 :99. T2q7‘1(I)9()\q—€i)=2LE_1.,\q [_(Ta—l (9.183) where G is the unit step function, Aq is the fermi energy, 6,- is the single-particle energy, ,3 = it and L“1 refers to the inverse Laplace tranform. C5(F1, F2) is the single-particle propagator which, in the HF approximation, reads C”(F F =2 99,- (F)2q to, (qu)e '6‘?" 9(/\q — 6,). (9.184) 214 Since the Wigner transform of the right and left hand sides should be equal, one has 0? 1‘2', M]. (9.185) [Adj-Eel?) : QLELAQ [ d The derivation of the Wigner transform of the density matrix up to h2 can be ob- tained by working out the inverse Laplace transform up to h2 of the single particle propagator. It reads [38] , )2 .. ~ 2 2 —. _. .. Cf(R,p‘) : e—li(%I—H+W(R))(1+251(—AVQ(R)+§3(VV(I(R))2 13 3—m —p(- 6) mm) + 0014)). (9.186) 2 . Defining the single-particle Hamiltonian hq = “ail—MA + V, where Vq IS the self- consistent HF potential, one can use the following relations 7+ioo [3(N hq ) , . — _ q / did 6 ’ ll — 6(/\q — hw), '7 563T]: ""38d(i""'(‘1'"” = 6'eq— hit), for integer n to obtain Lfiixq Be; - ugiwquaf ___ 00F 22:7_Vq(§)), Lgiiq [@752 '3 12“?!” mm); = gmu?) “(Ag—gt gun), Lag [gfi‘iw (vi/Amy] = ‘ giggmmf x 6”()\q — 21); (Km), 215 _ ml .1“ 2+1"(fi)) _. ~ 2 ~ h? 1 ~ 2 , ~ 133—”VI 247726 £7 (1 (p-V) mun] : -24m(p-V) ”(12) 5" A ”2 V I? X (q—‘Qg- q( )) (9.187) Plugging into Eq.(9.185), we obtain the Wigner transform of pq(F1, F2) up to h? a if I pWK.q(Rvm :e(’\q - hilt”) — g—AVQ‘S (“\q — hft') m (9.188) )1? 1 + 2—477— [(VVq)2 + 505‘- V)21/,,]6”(Aq — hip) + 00'1“), where hfv 2 HQ = — V q(R) and the derivatives of the dirac-delta functions m are performed with respect to Aq. Even though it is not the main target of this section, one can calculate the inverse Wigner transform of Eq.(9.188) to obtain the density matrix up to order 152. This effectively gives the extended Thomas Fermi approximation to pq(f"1,f'2). We can call EFT-DME. The important relations one has to use are m a :Bkz. 6 fz—ZITF 5(1: —- kp) and —— —— (9.189) - _h, = W" W) 0A 8Aq0k}’ 2 (12 his where Ag = 7,111:- + V q(R). Since the derivatives act on the dirac-delta functions, one has to perform integration by parts using the relation n n n a [(1136 (A!) —h(fv) F(paT:/\q) : (— 1) 87" F(p,T A (”'11:va (9190) where F (p, 7‘, Aq) is any well behaved function and pp is the value of 1) that satisfies the equation Aq — hfv = 0. Applying these mathematical relations, one essentially 216 obtains EFT-DME of p,,(7"'1,1"72) _, _, Ath33j1(kZ.r) 1 pq(7'la 7'2) = 37f? k? T‘ + 127r2 A”? [Joflifl _ ki'l‘jfikfild] F 1 Vk" 2 . . . . 24”; HF) [30(kgm) — 4kr}.r]1(k‘,’,r) + kg‘2r2]2(kz.r)] F 1 1 " kqfikq F F k0 - k0. qu 2° k9 R4 ___247T2EV F F-; ; —3 Fr]1( F7)+ ,1.~7‘J2( FT) +O( ) (9.191) In Ref. [173], the authors make angle-averaging (with respect to the orientation of F) approximations, followed by the expansion of the Fermi momentum up to ’12 1/3 -1/3 2 1 (W) Ap kq=(37r2 ) +—(37r2 ) [——" -2—" , 9.192 p pq 72 pa pg ,0, ( ) and K6193)? 2 kgAkg to obtain - ~ _ ~ 3310924") T2 . ‘q 3.1047") pETI‘.q(Ra T) _ pQ(R) kg”, + '72qu 30(kFr) 6 kg”, 7'2 (qu)2f - q j1(k%r) _ ET [4J0(LFT) — 9w] . (9.193) 9.5.3 Generalized PSA-DME In section 5.3, we discussed PSA-DME of the scalar and vector components of the normal part of the OBDM of time-reversal invariant systems. In that derivation, the chosen set of DME—coordinates (R, 7") were integral parts of the derivation. In addi- tion, we had a single nonlocality coordinate, namely, 1". Here, we give a generalized formulation of PSA-DME for time-reversal invariant systems where we relax these two constraints. This allows us to recover the DME of the nonlocal densities that occur in the HF energy of both NN and NNN interactions. It is to be noted that the three nonlocal densities that occur in the expression for the HF energy from the 217 chiral EFT NN N interaction at N2LO are of the form 9,071, F1 + 1’2), gq(F1,F1 + 1‘73) and 1 v F2=F3=Fl (9.195) Note that for the PSA-DME developed in section 5.3.3, the next approximations involve angle averaging with respect to the orientation of the relative coordinate and 218 averaging with respect to I: over a Fermi sphere. For details of the logical arguments in favor of performing these approximations, refer to that section. Here also, we apply exactly the same approximations. First, We define a new coordinate system as given in Eq. (9.1) H1 = 52 _ 53, (9.196) X’ = (1—a);f7'2+afg, (9.197) where ac [0, 1]. The essence of this parameter will be clear later in this section. Angle averaging with respect to the orientation of the relative coordinate, 12', en— tails performing },Lfl f (195 where ()5 denotes the orientation of 53'. Let us apply these approximations to Eq. (9.195) term by term. 0 The Leading term gets simplified as 3 -' de i111 ... .. "' —fi / dk / —e* (p.17-29,) = H8(k%x)pq(r1),(9-198) 4W3kp )1.)ng 47‘ F2=F3=F1 where ' kqfll‘) j1(kq|52-5‘73|) 11” «0.. = M I“ r = F * . .1 0(kFT) 3 16%;]: p(7‘1) 3 [£24235 _f3l (7'1) (9 99) o The linear (first-order correction) term has two origins. The first one is from —ik - (532 — £3) and the second is from :32 - V2 + 553 - V3. Hence _W /“ q (“C / 4 6* ["k ' $]pq(r-2,»r3) = L1(k2~$)pq(7‘1), (9.200) 219 where j1(k;’ L‘) , _. .. J1(kflf2 — f3” L Icq.." : — Irq : _ ' kq‘ —‘/"' 9 1( P 1) 3J0(1 1‘)+ 9J 112.1. 3J0( 1"“ 13') + 1:24:32 -J'33| (9.201) For the simplification of the second linear term, first note that according to Eqs. (9.196)-(9.197), the operator in this term, viz, f2 . V72 + 53 - 63 simplifies to f- (a 62 — (1 — (1)63) + X - (F2 + 63). Let us designate V7,, E a9. — (1 — a) 63. (9.202) Thus, ffizgfr A'I/erP iIx‘-1‘(f2 oV2+l3' V3) 2 /dQI: /de€iE'f[i-I°€a +x.(v.+03)] = j0(k.z:))? - ($72 + 63), which implies, 3 / ~/d95 15,—. (a .. .. ) , dk —e" z-V +f-V Ff- 47r3k3r3 Wish-,3; 47f 2 2 3 3 Pq( 2 .3) 1"‘2="3=”1 with .' q, - .9 ~ __ -1 L2(I.:‘;.a.~) = 3’10““) — 391””“1‘2 “D. (9.204) kid: _ lama—m o The second-order correction term is generated by the operator 220 .1 fl. [ii-2 - (62 —— H?) + it}, - (Vi, + 2'13] 2 / 2 , where we are not showing the phase factor explicitly. Using Eqs. (9.196)-(9.197), this operator reduces to [i’ (—2'I:f+ 6,1) + X - (62 + 63)]2/2. The relations /% 5%] ‘19? eif'f’w-Exrfia) 2: 0, (9.205) 19,; d9: ~ _. -» ~ /(__47T W—f (3“ 1(j‘. ° Ix’)(X ' (V2 + V3» 2 0, (9.206) and Eq. (9.5) can be used to simplify the second-order correction as 3 ‘* dad? 'T H. .., F T = _——./ q dk/— "ml/2W1 zk+Va.)+X'(V72+‘73)l2 “<1. My}: , _.F 47r X Pq(F2= 11.3) 7 F2=F3=F1 3 ' qur _. _. _, .. .. _. .. = 2J1I5q; )[X - V1]2pq(r1) + L1(k%13) -V1pq(r1) + L:3(k‘}.ar) pq(’l‘1) F 1:2 ‘ kqa: _, _, +419.” )Aapqem) . (9.207) 2 kpx ? _.. _2 12—1‘3—1‘1 where L kq. _ 3(kg‘x)2 ' kQ. ' ..q 2 8 A we) — 62,... m Fr) — 620w). (9. 0 ) A“ 5 Va. (9.209) The last term involving Aa requires further simplification. Expanding the oper- ator and using relation Eqs. (9.108)- (9.109) and the definition of kinetic density 221 given in Eq. (9.77) Aap0(f:2= F3) ®=fi=fi 2 : (av-2 — (1 — (I)V3> /)q(f:23 F5) (1:2 2 (1 - (1)2 2 — 2a(1— a)Tq(7"'1) 20.2 - 2a + 1 172:173 :51 : — (qu(7_"1) —‘ 27(1(F1) +126 ' 511071)) ’ : qu(F1) _ Tq(Fl) 1 2 (qum) — 2mm) — 2'26 5.971)) (9.210) where we used the fact that 311071) 2 O for time-reversal invariant systems. Collecting all the contributions, the complete generalized PSA-DME for pq(f'1 + :32, F1 + 553) in time-reversal invariant systems takes the form with Pq('F1 + 53:2, F1 + :33) 118 (A321) 2 Hakim) = A1(kz.1') '2 A2(kg‘1') Z +§ (X' - amen] 2 6 5 swim) _ 3 j1(k%lf2 — 433') = US$2.33) [pq(F1) + A1071”)? - 61%(771) l _. .. + $— 11309111?) [(02 — a + §)qu(7'l) - 79(r1) 3 .. + assets/2.91)] . 3“ — a kfix kqplfz — 553' 3j1(’€%1‘) = j1(kj{.|f2 —il"73|) 111.2: 191.);2'2 — :53) ’ L1(k%.’17) 1 —— m1 k": 2 L kq‘, L qu‘,. 1+ 1(p-13)+ 2( fl) zl+0((k%7)2). ”5(ki'viv) 222 (9.211) (9.212) (9.213) (9.214) (9.215) As a. can take any value between zero and one, it can be considered as an optimization parameter that is to be used to select the best DME-coordinates. Refer to the next. section for some detail on related works. Recovering previous DMEs of the scalar part of the OBDM At this point, it should be emphasized that the approximations that are used up to now are exactly the same as the ones that are used in section 5.3.3 when we derived PSA-DME for pq(f’1,f'2) using (11, 7") as the DME-coordinates. Hence, we can recover the PSA-DME of pq(I-f+ 772,11- 7‘72), i.e., Eq. (5.21) by setting 532 = —i"3 = F/2, F1 2 if and a = 1/2. In this case, X = 0, f = Fand Eq. (9.211) reduces to -+ F .. F _. 7‘2 1 —. _. p.92 + 5. R. — 5) 2 Hakim) p.119 + gnaw) [$qu) — TAR) + 3921141334 , (9.216) Q where US$21“) and 113(1324‘) are as given in Eqs. (9.212) and (9.213), with if being replaced with F. In obtaining Eq. (9.216), we considered only the leading order contribution to A2(k}:r). The only difference between that of Eq. (9.216) and the corresponding expression in Ref. [170] is the fact that flaky) = 105 j3(k%r)/(k}r)3 in the original DME. As can be seen from the series expansion, jl(k‘[]~‘7) ~ ,q , 2 My ~ 1+O((AF7) ), (9.217) j3(k%7‘) ~ .9 .. 2 105 ———(k%r)3 ~ 1+0((1,.,) ), (9.218) the two 7r—functions are similar in their leading order. Due to the (kg-r)2 prefac- tor that we have in the second-order correction, the difference between the second- order correction terms of PSA-DME and the original DME appears in terms beyond (DH/1121)"). As noted in Ref. [170], this difference in the higher-order terms should not 223 be surprising due to the the ambiguity of DME correction terms beyond O((k‘}r)2). Hence, we have effectively recovered the DME of Ref. [170]. In Ref. [202], the authors generalize the original DME of [170] as , 2 pq(R + bF, R - (1 — b) 7") 2 Ila/91.7“) pq(R) + %H§(k%r) [(b2 — b + 1/2) qu(R) _. 3 _. _ Tq(R) + gk%2pq(R)] 9 (9.219) with the same 7r—functions as given in Ref. [170]. Parameter b = 1/2 recovers the usual relative and center of mass coordinates. To obtain this expansion from our generalized PSA-DME, Eq. (9.211), one sets F1 = 12', =3, = bf", and 533 = —(1 —b)f’, (9.220) which implies a? = F and if = 0. Hence, Eq. (9.211) reduces exactly to Eq. (9.219) with parameter a playing the role of parameter b. Optimizing parameter b, the authors of Ref. [202] note that b = 0 which amounts to expanding about one of the particles, instead of the center of mass, seems to give the best accuracy for molecular systems. This further enforces the view that optimization of parameter a can result in increased accuracy of the DME. Further approximation with respect to if In the generalized PSA-DME of pq(7"'1 + f2, 7"} + :33), our angle averaging with respect to the orientation of if is a well-supported step in that the scalar part of the OBDM is known to have a weak angular dependence on the orientation of the relative co- ordinate [176]. If we stretch the argument and assume that the dependence on the orientation of the other non-locality coordinate, X, is weak, we can average over the 224 orientation of X . In this case, Eq. (9.211) reduces to _. _. _. _. _, .132 .. 1 X2 _, pq(r1 + 1:2, 71+ 1:3) 2 Hakim) pq('r1) + —6— H§(k}';;r) [(a2 — a + -2- + 7) qu(r1) —o 3 ( —o 4.00 + gb'f'2A2U'f-F)PATH]a (9221) whose simplicity makes investigating its accuracy a worthy step. As mentioned in the previous section, parameter a may be optimized to reduce some of the inaccuracy that may result from averaging over the orientation of X . 9.5.5 Generalized PSA-DME for the vector part of the OBDM The generalized PSA-DME for the vector component of the OBDM involves a signif- icantly less algebra than that of the scalar component as we stop at the linear order in the Taylor series expansion. it involves exactly the same approximations as the ones that we used in section 5.3.4. Extracting the non-locality operator and a phase space factor, followed by Taylor series expansion of the operator —o —¢ SqJ/(Fl + $2773 + £3) : eiE-(EZ—fg) 852-(62—ig)+53'(63+i§) Sq.u(f2a 773) 1V2=""3=""1 2:: eiE'(52_53) [I + :32 - (V72 — 2'19.) —0 ._. +113 ' (V3 + 1(1)] 8%,,(772, 713) (9.222) f'2‘-"*3='""1 where we truncated the expansion at first order. Since §q(f'1) = 0 for time-reversal invariant systems, the contribution from the leading term vanishes. Likewise, the contributions from the linear 2232/3 - 1: terms vanish. Using Eq. (9.112), the definition of the cartesian spin-orbit density, 114,”, given in Eq. (9.83), one writes Eq. (9.222) 225 as .s-,,_,,(r~'1 + .32. F1 + .133) z ie“" Z 2,. Jq.,.,,(r’-1). (9.223) p. The final step involves performing the PSA over a deformed sphere that characterizes the local momentum distribution. Let us start from a spheroid in momentum space defined by the equation = 1. (9.224) For ease of notation, we write (1(1-Z) as a and C(12) as c in the following. We constrain the position-dependent quantities a and c by requiring that the spheroid has a given volume and quadrupole moment, viz, 4 . 4 Vq E gash}? = §w3a2c, (9.225) _. 2 (—a2 + 02) P570?) 2 2a2 + c2 . (9.226) The II—function is obtained via the integration over the phase space of interest Hf ': —q3/ (”C 62.1“]; . i (9.227) Carrying out the integration over the volume V,, encompassed by the spheroid given in Eq. (9.224) can be done by using a stretched coordinate system from the transfor- mation ~ I? E (k,,ky,k,) _. 113' E (kg/($311.), (9.228) C 226 such that one finally obtains 3,901 + .132. F, + 2'3) 2 'i Hid-2.1;) Z 1:,,.1.,,,,.,(7‘-3), (9.229) p where ~~ .1133 "ma—q Hakim) E 3J‘1q1'l) = 3111 Pl] 3"), (9.230) kpr k; 132—rd and ~ 2 2P" " ”3 kg. 2 ( + 2(5)) 1:}. (9.231) 2-P:§’(R) Setting P303.) = 0, which consists of performing the PSA over INM phase-space, results in the same II—function with is}. replaced by k3,. For spherical systems, one can simplify the expression further by writing Jq,w,(ff) as a sum of pseudoscalar, vector and (antisymmetric) traceless tensor parts given in Eq. (9.165). Since in these systems, both the pseudoscalar and the tensor parts vanish, one obtains i 2 nfulgr) .13 x L071). (9.232) gq (F1 + f2,'1—‘1 + £3) 1" Hence, using Eq. (9.229), we have a DME for any type of nonlocal coordinate depen- dence. For example, 3(F1,1"'1 + 553) is obtained by simply setting :32 2: 0, while setting 7"} = If and f2 = —j:'3 = ”/2, Eq. (9.229) reduces to —o 9,,(1‘2' + 3,1? — g) 2 i n-;‘(1.s‘;.-r) Z r,,J,,,,,,(1‘?'), (9.233) p where 1'1“" 1.” r is as given in E . 9.230 , with :1? replaced with 1". Due to the possible 1 F Cl dependence of the accuracy of the DME on the specific coordinates used, one cannot 227 claim or expect the same accuracy in expanding, for instance, .§'q('f'1,f'1 + 1172) and am + 772,11 — 772). 9.5.6 Remarks on the generalized PSA-DME As can be seen from Eqs. (9.212), (9.213) and (9.230), the final results of PSA- DME of pq(7“'1 + :32, F1 + 553) and §',,(F1 + :32, F1 + 553) are not separable in :32 and :33. In contrast, all nonlocal densities that involve only two of the coordinates such as pq(7"'1, 772), pq(7"'1, F1 + £2), led to a completely separable expansions. This can leave the perception that the objective of having a separable approximation, which is what the DME proposes to achieve, is not yet met. However, pq(7"'1 + 52, F1 +53) and 52,071 +552, F1 +553) appear in the HF energy from the chiral EFT NNN interaction at NZLO where 1'32 and 53 are part of the interaction form factors. Refer to section 7.1. In fact, the interaction does not depend on F1. Thus, all terms that depend solely on 52 and :33, whether they are separable in these two coordinates or not, can in principle be integrated out with the interaction form factors. The actual direct analytical integration of such terms is very difficult, if not impossible. Refer to section 9.11 for details on how we solve this problem. Leaving the technicalities for the relevant sections, it should be clear at this point that a local EDF will rasult from the application of the generalized PSA-DME of pq(7“'1+a':'2,1"'1+53) and §q(7'"1 +f2,1"'1 +5233) to the HF energy of the chiral EFT NNN interaction at NzLO. 9.5.7 The modified-Taylor series expansion As discussed in section 5.3.6, the modified—Taylor series approach provides a formal framework to extend the applicability of the density matrix expansion to non-time reversal invariant systems. It is obtained by replacing the coefficients of the Taylor series expansion of the densities with 7r-functions. The dimensionless variable 9 is 228 used to denote the possible argument of the 7r—functions. The modified Taylor series expansion of the densities that. appear in the exact HF energy from a generic NN interaction reads _. 7" _. 1" _. .. pq(R i '2') N 778(9) 1’9”?) i “am 5 VIZ/MU?) 1 F 9 2 _. +5799) (5 - vi.) p.02) (9234) -o 4 4" 9 d 5' F d -° _. 9.,(12 i ) m we) sq(R) 1: M9) (5 v3 3.,(12) 1 q " .. 2 a +§n;(9) (a R) 5721(3), (9235) —o -o F . —' F -o —. -o “(R i 5) = «((9)309 i «((9) (5 - v3 Jq(R) 1 ~ F 2 2.. .. +57%» 5 . VR AR), (9236) -* -* 7:. d -‘ -‘ " 1? —o -o —o Jq(R i 5) “ 7r5(9) AR) i «Km-2- VR q(R) 1 f F 2 -+ _. +57% (Q) E V}; Jq(R). (9.237) _. F .. F p .. p F -o -' _. pq(R 2i: ‘2‘, R 3F '2‘) % H0(Q)pq(R) Ii: H1(Q)§ V1 — V2 pq(r1 T2) .. Fl="2=R 1 p F _. 2 F 9 2 fl _, 'i' §H2(§2) 5 V1) + (5 ‘ V2) pq(T1 , 7‘2) i=2=fi 77 .. F .. _, _, 415(9) (5 - v1) (5 v2)pq(r1.r2) , (9.238) 51:52:}? -0 " F " F 9 -* d s F a d " " "’ Sq(R It 5, R :F 5) = H0(Q)9q(R) Ii: H1(Q)§ ( 1 — V2)Sq( 1, 7‘2) _, "1=I-'2=R ~ F 2 4 2 ._n.m)[(§ .) + ( v2 )2(... , F.) _ ~1:‘2=R 229 and for the pairing densities ~ 77 , ~ _. - F _. ~ _. pq(Ri") z 7r(“SD/MU?)3'37Ti)(9)§ vhf/MU?) 1 , F - K s +§ng(o)(§ . v3) pqm) (9.240) 1 .. F - . F - _~_, .. 9.02%) ~ 7r.s.(mivrl(m(§ v1.)s.(12) 1 g F .. - q +§7r2(o) 5 - v squz) (9.241) ~ - F . F - 1 - F .. 2 F .. 2 + — 113(9) V] + — V2 fiq(Fl 3 F2) 2 2 2 -. _-. -R 1" 2— p F - F _. 1 _, _, — 113(0) - V1 — V2 pq(71 , 7‘2) , (9.242) 2 2 -' _-.° .1? 7‘1—12— :. -o 7? -‘ F q _, “ 9 F _' T. :0 —o —o we i 5,12 :1: 5) 2 11.191941?) i H1(9)§ - ( 1 — v2)e.(n , r2) _. “1:72:12 1 .~ .. 2 F .. 2 ~ + - H"(9) - V1 + - V2 511(71 , F2) 2 2 - _- _fi 1"?— 3 F -. F ~ .. e —H3(Q) — ' V1 — V2 .Sq(71, 72) , (9.243) 2 2 1=Q=fi At this point, the modified-Taylor series expansions of the densities can not be written in terms of the local densities defined in appendix 9.2.3 and 9.2.5. Implementing the steps explained in section 5.3.6, one obtains an equivalent expansion, this time with explicit local densities. These are pqu'i i g > ~ «ampm 1 + :- ng’m) (g - 63)? pq(1'2') (9.244) 5.0% i g“) a WWW?) iwf 2+: mm 123-M) + 54-11392) (qu0?) _. F ~ I" /)q(Ri§.R:F§) ~ 7.2 -2T(l(fi)) — -1-2-H§(§2)TQ(R), (9.246) 2 F 2 F ., 2 , é. _. 2 g. 2 3MB: :1: 5,12 ; 5) s 113(0) equu?) :tzH‘1(Q) erWuz) + in: (Q) (Aswm) . 2 _3 47343)) — 1’15 §(Q)Tq‘,,(R), (9.247) _. 17 1' ... .. v F _. ._. _. JQ(R i i) z 7r(J)(Q).7(1(R) i “{(Q)§ VRJq(R) 1 -.- “ .. 29 _. +§W$(9)(%'VR) AR), (9248) -o -o F " -° -' " 7:. -0 -o -o Jq(R 1*: 5) % 7r91(9)Jq(R) i ”“535 (VR)Jq(R) 1 ~ ‘* ~ 2~ -+ +§7F2J(Q)(‘;"VR) AR), (9249) and for the pailing densmeb ., " F - ~ -‘ ~ F —0 ~ —o Pq(R i 5) ”1 778(9) (09(3) i WW5 VRPq(R) 1 p F 2 K r +57%”) §°Vn Pq(R) (9.200) :. 7" ~ .3 7" ~ 2. ~ 69(3 i 5) % «062)de i7T1(Q)§'VRSq(R) 1 “ .. 2 9 +§7r§(Q) (avg) 93(3), (9.251) ~ “ F ‘* F -+ . _, " .. 7‘2 ~ MR 2: 5,12 2: 5) z 118(9) MR) iznflmr MR) + firm) (qum) 2 2 r ~ 2 ~ 4902)) fingqum), (9252) ~ ~ 7" ~ 7" I: ~ ; ~ 7‘2 % ~ ~ Sq.u(R i 5: R :F 5) % “MUS“ q.u AR) i7H1.(Q)7‘quw/(R) + EH25?) (Asun?) ~ ... 7‘2 ~ _9 4233412)) 1511.362)?" ,,,(R) (9.253) 231 9.6 Derivation of EDF from HF energy of local N N interaction In this section, we give detailed derivation of the HF energy from a generic local two-nucleon interaction and the EDF that results after the application of the DME. 9.6.1 Central contribution We demonstrate the derivation of the contribution to the HF energy from the central part of N N interaction by deriving the corresponding expression for the spin-triplet, isospin-triplet channel. Making use of the spin/isospin projection operators given in Table 1.2, the projection of a local central interaction in this channel reads V51 = (1+13102)(1+ “Ian/0%). (9.254) tb-IH Plugging this in Eq. (6.2), using the definition of the OBDM (Eq. (9.70)) and its scalar/isoscalar—vector/isovector decomposition, we obtain 1 _. .. .. .. .. _. .. .. Emu] = g / dn .5. V519) [90(71)P0(7‘2) — po(v-1,r2)po(r2,n) 1 .. _. .. _. + Z (5 00(7‘1) 60102 + 250(71) 00102) ”1‘72 1 _, 1_, _, _, X (5 pO( 2) 60102 + 530(7‘2) 00201) .. .. 1s .. a _. - (”202(5/9001113) 60102 + 250(7‘1,7‘2) 00102) l _, _, - 1_, _, _, _, X (5 p0(7"2, Tl) 00102 + 58003: T1)'00201) 1 _, 1 _ _, 1 r 1 Lg .. + [(5 p051) + E(—1)2 P1(T1))(§Po(7'2)+ 5(4)? p.99) 1 fl _, 1 1.., 1 _. X (5p0(r2, 7‘1) + §(_1)2 pl(r297‘l)) 1 .. _. _. .. _. _. _. .. + ‘13 Z Z (/)()(l‘1)/)()(I‘2) 60102 + 3()(r1)oaalo2 X 8()(7‘2)°00201 q 0102 +101 (r1) ,0] (F2) 60102 + S—l(Fl)-50102 X 31(71‘2)-50‘201) 1 _, _, _. .. _ E Z Z (100(7‘177‘2) P0("'2a T1) 50102 ‘1 ”1‘72 +s‘6(fi,r2).50152 X ‘57)(F23F1l5‘7201 + p1(F1’3F2)p1(F2’F1)60102 +3-l(F17 71:2).60102 X 51(F21F1)50201) ] ' (9'255) To find the final, most simplified form of the above expression, we make repeated use of the relations listed in Eqs. (9.12)-(9.13) to obtain 1 _, _, 9 _, _, 9 _, _, 3 _. .. EgNllll = g/drl (1T2 V610) [190(7‘1)PO(7'2)— 1| 100(7‘11T2)|2+1P1(7‘1)P1(7‘2) 3 _, _, . 3_, _, _, _, 3_, _, _, _, _, _, — 1| 101(7‘1~.7‘-2)|2 + 130(7‘1).30(7‘2)— 130(7'1,7‘2)-30(7'2a7”1) 1_, _, _, 4 l_, _, _, _, _, _, r + 181(7‘1).81(T2) — 181(7'1,7'2).81(I‘2,7'1):l . (9.206) In terms of the proton and neutron densities 1 _. _. _. _. .. _. .. _. ECNNllll : gZ/drl dV‘2 VC1'1(7')I:3pq(rl)pq(r2) _ 3pq(rla7'2)pq(r217'l) ‘7 +55%) - 9(3) — 5.15.52) - W()] (9257) 1 -o —o 3 -0 “ +g E fdrl drz V8.10?) I: pq(T1)pq(7‘2) “ pq(7'1 72)pq(72 T1) 1 a —o —o 1 -o -o —-o -o —v -o r +-2—sq(r1) - 9.,(1‘2) — §Sq(7‘1 7‘2) 5‘(’2.T1) (9 208) Similar derivations can be done for the other three channels. 233 9.6.2 Spin-orbit contribution For the spin-orbit interaction, we demonstrate the derivation for the spin-triplet, isospin-singlet channel. Starting with Eq. (6.2) and the projection of the spin-orbit part of the interaction in this channel, we can write the Hartree contribution as i - - .. a ~ _. s _. Efslyullol = ‘1: [6171 d7'2 V133 ('7‘) T' [Pq(7'2)V1X3c7(7‘1,7‘2')|r~2/=r~1 q + §q(:.l) X 62p¢i(772, 772,)IF2,=172 ] i a _, a _, l-o a a .~ a = 7:; [drldy~2vgg(r)r. [pq(7‘2)(§VX 'q('r1)+z.]q(r1)) _, a la _. .7 a +8.,(7‘1) X (“V/Jam) + WWW] VP‘IIH Z / (1771 (1742 Vgg(r)7='. [mm/5(a) + -r'-'- 5.,(7‘1) 69 32% q i 2 v x awn/2.03) + gqum) x am] (9259) Noting that / cm die V1.30") .7. v a [mm mm] = o, (9.260) where 6 = 61 — 62, the Hartree contribution from the spin-orbit interaction in the spin-triplet, isospin-singlet channel reads ENN _1 l.-'1-:V10.. -+ "-* 1' 1* -' 7.7 21 LS.d2'r[10] — Z Z ”1072 15(7)?" q("1)/)(i(72) + v5q(7‘1) X L702) - (9- 6 l q The Fock contribution in the spin-triplet, isospin singlet channel reads Efiéi’pllm = §(V5 — V. + v7 — v8), (9.262) 234 Where expressing the V,- s in terms of the density matrix Vs, = Z Z: / (17"1d7"2ng(7 )7 x V2pq(7",202 F101) -(01|0|03) pq( F103 F202) 010203 q V6 = Z Z/dfldflVfig )I">< §1p(;(7"103,I"202)-(01|5|0:3)pq(1"202,7"101) 010203 q V7 2 Z Z/drl d7‘2VL5( )rxV2pq (7204,7101) (a2lala4)p —(7"101,7"2a2) 010204 q V8 2 Z Z / d7"1d7"2 V112- (7‘) 7"x {71/75 (F101, F202) - (02IEIU4) pq (7"204,7"101). 010204 q (9.263) The manipulation of the above four expressions involves repeated application of spin-traces. Finally, one obtains i _. _. .. _. _. _. -‘ .. _. ELS,e;rci10i = .51 Z/ dT1d1‘2 V23”) [80(7‘177‘2) ' 7' ® V2pq(7'2,7'1) q —o +Pq(7_‘ia7"2)"? ‘33 V2 ‘ -§q(F2,F1)] - (9264) Spin-orbit contribution in time-reversal invariant systems To recover the expression given in [170] for the contribution from the Hartree-Fock energy in spin-orbit interaction, it suffices to show that _ 2:q f dFl d7"2 VngO‘) pq(I“‘1,f'2)I'" X {72 ‘ 55(F1, 772) can be simplified as = " Z/ (1'71(1F2V£2(7‘)pq(7"1,772)77x V2 ' $707143) ‘1 Z Z] dFl d7"2 V359“) [$70717le ' 'FX fi2Pa(7-‘ie7—"2) q —- pq-(7‘1,F2).§'q(r1,7‘2) - 62 x 7‘" 1 dVL _, _. .. _, _. _. _. " Z] drld; 2 — rd“: )‘ 97(T1J2)3q(7‘127'2) ‘7‘ X 7‘ 235 = Z/ (1F1(lT-::2 V320") 87,071, F2) ‘ 7? X figqu-"l, F2) . (9.265) r1 Thus taking the above result, adding it to the terms coming from the spin-triplet isospin-triplet one and assuming 1032(7) = VL‘;(7‘) 1 one obtains -O _ 1 a 4 ~ ~ ~ Egg ._. E /d7‘1dr‘2VLs(r)[7‘- (mm) +7.2 2; gm» 7x mama], ‘1 (9.266) which is exactly the same expression as given in [170] for time-reversal invariant systems. Note that there is a factor of two difference between our expression and the expression in Ref. [170] which is due to a factor of two difference between the spin-orbit interaction used in our derivation and in Ref. [170]. 9.6.3 Tensor contribution The derivation of the Hartree contribution from the tensor part of the NN interaction is trivial due to the specific operator structure, viz, (0102IEI. ‘ é;- 52 ' grl030'4> = <01|61|03> ' gr <02|52|04> ' 57-, (9.267) (0102' 51 ‘ 32 [0304) = (01'61IO'3> ‘ (02'32'0’4> . (9.268) For example, in the spin—triplet, isospin—singlet channel, 35“},(7‘3) ' 7: "(7(772) ‘ f — 537071) ° $7072) (9269) , 1 _. _. 57%?”10] = Z 2/ (17‘1 dT‘Q VIEOU‘) q 1This assumption is essential to obtain the form given in [170] as the authors use the same 7‘ dependence in both channels. 236 The derivation of the Fock contribution involves a significant number of spin-traces. Due to the similarity with the derivation given in section 9.6.1 for the central piece, we do not repeat the derivation. 9.6.4 Leading-order pairing contribution Restricting the derivation to a central interaction and to the spin-singlet, isospin- triplet channel, the leading-order pairing contribution reads (<1>|Vc°‘l>pm-r = g: 2: KM (z'CIJ'qlVO1 quzqw (9270) q ijkl Thus, 1 _, _, _ (cplvgl|q>)p,,, = {3‘2 122 / (mam/31m 1 .. ..- ><2(:,1 —Pq (7202,7"100‘3—Pq (730277101) 1 ~ 1 _2__p‘l* (r202,r101)?;pq (T1019T202) 1 1 _E; ”q (1302,7103)20—gfiq(rq2.0’117:lo_'2) 1 , i i_ 1 .. i i_ +—Pq 0202,7101)_Pq(7‘102,7"201) - (9271) 20' 0'1 20'1 After resolving into scalar/isoscalar, vector/isovector parts and simplifying one gets <IV°ll> £12] an (172 )Ipqm. 1w. (9272) 9.6.5 The resulting EDF: EDF-NN-DME In section 5.3.6, we discussed the basic steps that are involved in the derivation of a local EDF from the exact HF energy of a generic NN interaction through the 237 application of the modified-Taylor series (or any other DME). These steps are best exemplified by the simplifications that we carried out to obtain Eqs. (5.46) and (5.47). \Ve apply essentially the same step on the exact HF B given in section 6.1 and arrive at a local EDF which we call EDF-NN-DME. It has three components: the particle-hole EDF (given in Eq. (4.11)), the particle-particle EDF (given in Eq. (6.35)) and the additional terms of the EDF (given in Eq. (6.36)). The couplings of the particle-hole part of the EDF, in terms of the notation defined in Table 1.2, are given by AP = 3a3P1[21 -afPP[(«s>2—(na21 AP = (— 3i3a2PPP+f —3a§“>[n5(ns+n§)1 APPP = (3—1—3a + 332321133533] + (3.1233 — gamma] APPPP = —<3—13a§*PP+3-33a2PPP)[(vrr>2] APP = gamma (1132] +3af“[<«§)2— (1132] APW = :aé‘su[7r57r{— H3113] AVPP = —:a3222[«f«J—n3n31 APAP = —3—12a§°1[7r37r3+n3r13] +(géagn—éagll)[7rgww§— H3113] +333“ [23333 _ 113113] 3% = —3a3“[(«a21+§a3“[n5( §+na1 AVsoVs : (3__12aC01_ 3_12 acn 8+éa§11)[(flf)2] APP = (— 1—13a3PP—3—13a2P“+3a3P{newsman-s1 A” = —g-a§‘11[113‘113] BPP = (7323 33+ 1—13a1PPP>[(«3>2 +21 + (332—3311 + 713a?2°)[(vr8)2 - (11:92] B"T = (— —3—32( 1010—§1§(z€01+%a€“+3—12ag'00 )[H8(H‘°+H§)] 238 B233” (3—33a E33 + 33-30333 + 3—330 5” + 33—1309” )[5553] + (535610 + 61133301 63433011 _ 6143200 )[ HGHP] BPPP —(3—33a5PP + 333a? + 3335“ + 333(15PP )[(+5)2) BPP (333a5PP — 333+ 5PP)[<+5)2 +< n52 1 +(3—33c5 - 3—13af°°)[(7r5)3—(n5)3] BPW %a3510[7r33 7:3 —ngnP]+ +513 3511533 7r1 ngnf] BPPP _; 33103731 753 HPHf] _ {335351153537 + H3113] BPPP (334550 — 3—33a5PP)[75 5+ H5115) +(333-a5“ - 313M — H5115) +<— 3—33a5PP + 35° - 3135“ + 333331333333] +(— 3—33a + 3—33 50 + 3—33a — 15633333133333.3333 BPPPP -(Z a5PP+ 3 3a5PPn< 1)) +(—1—3—6 a3T104+ 136 aT“) [I13 (II; + HP 3)] BVsoVs (—&a€30 +16_14acr01 _ 6-1—4 (1011 + 611(33700 +323a5PP3 +— 35% 1)) B“ ($125310 3—1:ag01 _ 3335511 + $5500 _ gag“) + 35310 +305“ -3331 )[n5(n5+n5)) 3” (3a5PP —3a5 “)[H5H51 and for the partlcle-partlcle part of the EDF AP = 3—33-a5PP[n5)2, APP = I:730,3P“’1[I13n.3], A” = iag’lflngnfl. (9.273) In the EDF that results from the additional terms, given in Eq. (6.36), only the couplings from the Coulomb piece are unspecified. These are given by CPP = 03‘ «(3)2, (9.274) CPAP = 0Vpr where in these equation a? refers to using the Coulomb interaction to compute these couplings. 9.6.6 Analytical couplings from chiral EFT NN interaction at N2LO In this section, we derive the analytical couplings of the particle—hole EDF (given in Eq. (4.11)) for time-reversal invariant systems starting from the finite-range part of chiral EFT NN interaction at N2LO (Eq. (6.5)). In line with the exact treatment of the direct part advocated in section 5.3.7, the contributions to the couplings that come from the Hartree part of the HF energy are not included. Furthermore, in conformance with the notation used in the NNN case, we use isoscalar/isovector notation instead of proton/ neutron notation. The starting point for the derivation is the expression for the couplings expressed as a functional of the 7r—functions. Note that the Fock contributions that we are interested in correspond to those terms that contain solely HE, i.e. no 7rf, for any density (. We illustrate the derivation taking the calculation of App as an example APP = 3 fdrr2[V(91(r)(H8(I-cpr))2—3V313+1(r)(1'18(kpr))2)],(9.275) where the interaction vertices V3910) and V3310) are given by Eq. (6.10) and (6.11). 240 The starting chiral interaction is in momentum space and hence 1 . ._._. App : fdr 7,2 diezq-r [I/(Ql(q) (H8(kF7.))2 16n2 — 3 V519,) (H3(L-,.~7-))2)] . (9.276) The subsequent step requires specifying the 7r—functions which can be fixed according to any viable DME approach. In our case, we use PSA-DME with the 7r—functions given by Eqs. (5.22), (5.23) and (5.36). Next, we perform the integration with respect to r first. This is actually an important step to see that the integrals do not actually diverge. In contrast, in the NN N case, it is easier to perform the integrations first with respect to the momentum coordinates. Refer to appendix 9.11.4 for a related discussion. In performing the r integrals, we define 11(9) 2 [Mme(1159))? (9277) 12(9) 2 /r4clv~jo(qr)H8(r)H3(r), (9.278) 13(9) 2 /r2drj0(qr)(nf(r))2, (9.279) [4(a) s /r4drj0(q-r)(nf(r))2, (9.280) where (i = q / kp. Upon inserting the PSA-DME 7r—functions, these integrals become 119;) = 1293:1392) = 2207“ — 12g + 16) 9(2-9), (9281) 14(9) = 2% (2 — r22) 8(2 — (i), (9.282) where 9 denotes the unit step function. What remain are one-dimensional integrals with respect to the momentum coordinate, q. At this point, the couplings take the 241 form pp_ Ct — pAp _ C, — JJ _ Ct — where q2dq Fi’m) [MW 1....) + 212(9/kr)] — .3 7d,. F 2 k. [cqu r:'(q)12(q/k.~) 7T F 1 . ._ch I 4wk% / q2dq 13(q/kp) (1+ 295%) I1W9) — M; f (M; r1391) lug/a), 1“?"(0) = W(Q) i=0 = warm) t=1 with 2' E {C, S, T, LS}. These exchange-force form factors are given by Vé(q) W501) V5101) st (q) Vflq) W235 (9) = VC(q) + u«’c(q) + 3Vs((1) + 3WS(‘I) +q2VT(q) + (12WT((1), = wc(q) + 3W3(q) + <12WT((1), = VC((1) + Wc(q) — Vs(q) — Ws(q), = Wc(q) - I'll/5(0), : VT(q) + I’VTW): = — WT(q). (9.283) (9.284) (9.285) (9.286) (9.287) (9.288) (9.289) (9.299) (9.291) (9.292) (9.293) The remaining one-dimensional integrals are calculated after plugging the chiral EFT NN interaction form factors (at NQLO) given in Ref. [12]. The complete expressions for the couplings are too lengthy to reproduce here. Consult the Mathematica files 242 of Ref. [161]. Here, we list the contribution to the couplings from the LO finite-range piece. As given in Ref. [12]. the only LO finite-range piece is a one-pion exchange term. Therefore we have I 2 + 772.27 W7?” = —( 9‘“ )2(] (9.294) M where all other components (VC, etc.) are zero. With 21. E lap/mm the non-zero couplings from the finite-range LO potential are: 2 _ 9A 2 , 4 6 _ 2 Ag) — —W{(—21 + 498a + 64a — 16a ) — 12u(35 + 4n )arctan(2u) 3 +472 (7 + 16u2(8 — 913)) log(1 + 41(2)}, (9.295) 856; = 2A???) (9.296) Amp — — 35934 {(—3 + 72u2 + 4224) — 60 u arctan(2u) <0) " 3072 fgmguf’ ‘ ' 1 + 4—u_2 (3 + 54112 - 72124) log(1 + 4122)} (9.297) B53" = 2.4%)” (9.298) A Ag) 2 —4Af0)p (9.299) Bg’g, = 2A5, (9.300) 2 g3, 5 + 12a2 4 2 A1 = { —1 4 } . (0) 48f1¥m12r (1 + 421.2)2 + U2 Og(1 + u ) (9 301) 2 2 13(0) = 221(0) (9.302) which can easily be put under the form of Eq. (6.39). 9.7 HFB equations from EDF-NN-DME The general formalism of HFB equations is discussed in Ref. [81]. Just like HF equations, they are solved self—consistently. In coordinate space, HFB equations for 243 —0 general nonlocal “mean field”, h"(f'a, F’ 0’), and pairing field, A4070, r’ 0'), take the form h'q(Fa, F'a’) A"(F(7, 7"" 0') Ug(f"a'q) U307}; q) / F"; = E: o’ —A"*(FU, F’a') —h’q*('f'0, F’a’) VL"(7""’0’q) V’f’(f"0q) (9.303) where U30” 0’ q) and Vflf‘" 0’ q) represent the upper and lower components of the quasi-particle wave functions. E3 is the quasi-particle energy and h’q(f' a, 1" ’ 0’ ) is defined as h.'q(Fa, 1“" 0') E h.q(Fa, F'a') — X’ (500/ 5(F— f"). (9.304) In Eq. (9.304), X1 is the chemical potential which is calculated from particle number constraint at each stage of the self-consistent iteration [205]. In configuration space, the mean and pairing fields are given by 68 65 1] Starting from a local HFB energy density, the mean and pairing fields become 10- cal in coordinate space. This is shown explicitly in the next sections where we de- rive hq(f'0, 1"" 0’), and pairing field, A"(Fa, F’ 0’) for EDF—NN-DME discussed in the previous section. These derivations involve just repeated applications of functional derivative which is briefly discussed in appendix 9.1.6. 244 T‘l 9.7 .1 The mean field from EDF-NN-DME The derivation of h"(f'0,f"' 0’) in (F,a,q) space proceeds by taking the functional derivative 6 s”. + 5"" + 5’?” hj-l- E ( A 6pq (DUI-+1771) ’ (9.307) if where from the coordinate representation of the mean field, one has the configuration representation h}.- = Z jdfidf’g 99;('F101(1)hq(F101:71202)99i(7-‘202(1)a (9308) “102 with 99,-(Faq) denoting the spin up/ down components of the basis 92,:(Fq). Since the energy functional is quasi-local, it results in a local field of the form llq(F10'1, F202) E 60?] — F2) hq(F1, 0'1, 02). (9.309) This field acts on the spin up and spin down components of the wave function through Z h"(F1;0102)¢,-('F102q) = [11" 99,-](77101q), (9.310) (72 and it is given by hwm) = [4‘7 - 3,07)? + Um + §,(F).a — 2mm 9 + (7' Km] —\‘7’- [C'.,(~F)-5]V7 — 2 [mm a W + W 8 mm] —%[6.D,(F)afi + 5.61347?) - 6] 2,021). (9.311) where a shorthand notation for the tensor product has been used A ® B = 2 AW, By“. The various components of the are given by u [.1 , _ (A—l) 2 p7 pT , 87m — 2A... h +A ,).,(mB 72.0). (9312) 0,0) = 2APPpQ(F) + A” mm + APAPA/qu + A [ApAppM] —26- [vam + AP“? 3 13m — fi-[AVP’LM] +2Bpppqm + Bcrqm + BPAPqum + A [WM/2m] -22 [3mm] + 8”“? - far) — fi-[BVPJ £702] +62/d7—y P120“) _ 82(§_)1/3p1/3(,:'), (9.313) lF—F’l 71' p 8.7m = 281835.22) — 274/1““?- 22.09] — AJJT....<7=) —2AJ~7F.,,.,(F) + A‘“A3Asq,u(F) + A [A3A85.,..,(1=)] — QAVSOVS A8q_u(7:) + 2 61/023 Va [APVJ jq/3(F)] OB — 2 51/06 [141va Vajq.t3(f)] 00 +28“ $77,110:) — 2VV[BVWS6 ' 52(5)] — 3'” Tau/("7) —QBJqu'J/(7_') + BSASASQ,”(F) + A [BSA886,U(7_‘)] — QBVSOVS AquUT) + 2 61/03 Va [BPVJ $74303] 03 — Z Ema [BVPJ Vajti.fi(fl] 1 (9-314) 06 Am) = 47427.37) + quaatAPV’VowU‘71 GB _ 2 61/03 Va [AVPJ Sq,i3(7‘.’)] m3 _2BP7' jqr/(F') + 2 gym; [BPVJ Va-Sci.i3(f')] 03 _ 2 6m); Va- [BVPJ Sad?” OH ’22 .., . _.I I. Am fdr 2,: Jq’-V(r )’ (9310) q 246 Cqm = —A“sq.u(r*> — B'”.s-q...(0. (9.316) I)... = 2A“J.,.,,..(0 — Z a... v. [MW/1m] + z [22122.09] +22” [J.,,,,,,(r*) + 6.4.2400] + 2B""JWW(F) -— Z ..,... V. [BM/2.7m] + Z ..,... [va vapqm] + 2;” [Jmm + 6,,” Z Jq,.,..]a, (9.317) Dam = —2AJ-is...m — zBJJ's....(o. (9.318) 9.7 .2 The Pairing field from EDF-NN-DME The derivation of the pairing field proceeds by starting with the variational A3]. = W‘ (9.319) U The pairing field A” in coordinate space is defined through [81] A?) E Z [(17-31dF299:(F101(I)993(77202Q)Aq(77101s77202)- (9-320) ”102 It is local A(I(F10'1, 77202) = 6(771— 7'72) Aq(1—"1;0’1.0'2). (9.321) and has the structure ~ -o Aq(f';01,02) = U,,(F;01,02) + 6 - Dq('f";01,02)6 + Aq('f';01,02) - 6, (9.322) 247 where the field components read 0.2.01.0.) = 27123230 5.5.1.2 + 2A!” (A209 — 42(7)) 3.3.1., + 2 A [AW (3,0)] 5160,32 +3 2 (62(aglaulal) + 01(61IO’VI5’2)) V7,,[Ajj jq,,,,,(f')] , W (9.323) 15,07; 0., 02) = 812323.41?) a. 5.152, (9.324) Aq‘p(77;0'1,0’2) = ZZ<52<02|0V|01) + 01(51|0V|0’2>) Ajjjq,uu(7_‘) . (9.325) V 9.8 Numerical solution of EDF-HF equations in spherical systems For the preliminary self-consistent tests of the DME discussed in section 5.4.6, we performed self-consistent calculation of the HF equations. This calculation was done with the assumption of spherical symmetry, which also implies time-reversal invari- ance. As the starting EDF, we took two different cases: (i) EDF-NN-DME with the Bogoliubov contributions turned off. This is what we call full-DME. (ii) In the second case, EDF-NN-DME is changed in such a way that the Hartree contributions to EDF—NN-DME are replaced with their exact counterparts, with the Bogoliubov contribution still turned off. This is what we call exchange-only-DME. Since the Bo- goliubov contribution is turned off in both calculations, we refer to both calculations as EDF-HF calculation. In both full-DME and exchange-only-DME, the spherical self-consistent HF equa- tions take the form hqwfifl = fiq99z'(77(1)a (9-326) 248 where h" is the single particle Hamiltonian given by ... ... m = —V-Bq(r)V + Uq(r) — trig-6 x a. (9.327) The only difference between the two is in the actual values of the field components: 8.09. Um and Wm 9.8.1 Bill-DME in spherical systems The components of h" for the case of full-DME are given by (A — 1) 'r qu = my + Am PW?) + Bp pm, (9328) mm = 2.4/”pm + A" mm + ApApqum + A [AW/Mm] -29. [AV/Wm] + AW - fqm - WWW HEW/m + Emma + mam/2.7m + A [BPAppm] —2€7- [BVPVPV'pq] + BM? - Jim — {HEW £109] +62/dw ”PM — e2(g)‘/3p:/3(o. (9329) IF- F’l W. = AJJLm — V [Ammm + Awfipm — .4”qu +3“ng — 6 [BM/am] + BVPJW.<0 - BJJJEW, (9.330) where 6 and A operators that occur in the fields probe only the radial part as we are dealing with spherical systems. It should be noted that (i) all the local densities depend only on the magnitude of 7" (ii) all the derivative operators are not meant to act on the wave functions, they only act on the densities. All the couplings such as A” and 8"” are as defined in section 9.6.5. 249 9.8.2 Exchange-only-DME in spherical systems In this case, the field components read _ (‘14— 12) pT pT B.(r9 — 2.4... n +A p.,<79+B p.09 (9331) v.09 = [dc [2V5"“”’(IF—F’I)pq(7"’)+2Vé5‘7'””(IF—F’l)pa(f") +/d1‘~" [vggppur —r'))(r'-7=').j;(r') + VZ‘E‘WUF- F’|)(F’ — B) 47.06) ‘l‘ 2AM) 900:) + APT 79(7?) + APAP AM?) +3 [ApAppqm] _ 2:7. [216mm,] + we . m —6 [AVPJJ 7,99] + 23%.,(29 + B‘" 7,01) + 8””qu (F) A [épAppq(f)] — 26 [BVPV‘Ofi/y] + BPVJV “q(.) —6O[BW jam] + ere/did I:p( ‘3) T e2(%)%p§(fl. (9.332) d?” VZ§””( 7‘ - 7" ’l)pq(7" )+ Vi’fé’ppflr- 7609.70”) 7'“ +A’“J. 1. In order to make use of this relation, one writes (prllfi - Bq(r)§|qb,,) as (9,207 - 3.109691.) = (cm/IWBQUH .69)..) + <¢.I|3q(r)A|¢p>, (9.338) where one can write (¢“/|Bq(r)A|gb,,) as (éyllBNMIép) = Z<¢>flIIBq(7‘)|¢./)(MAW) = ——Z<¢,.IIBq(r)I<2u)<¢ulT|¢u>- (9339) V Finally, using the exact kinetic relation in Eq.(9.339) and plugging in Eq.(9.338), one obtains a simplified formula for the matrix element. This is exact in the ideal 252 case of both no truncation of the basis states and a box of infinite size. In practical calculations, one has to truncate the number of basis states and also use a finite— sized box. These truncations make the use of Eq.(9.339) numerically unstable and erroneous? This is the case especially when the inverse-effective mass term, Bq(r), is very different from 1 inside the nucleus. For cases where the inverse-effective mass remains more or less the same as the inverse bare nucleon mass, using Eq.(9.337) or Eqs.(9.338) and (9.339) give the same results for the matrix elements. Matrix elements of the central potential part The matrix element for the central potential part of the single particle field, Uq(r), reads Ru! 5”! (SD-I 5 Id I (9.340) mm 11' ’ <¢,./|Uq(r)l = / 717-72_§£U,(.~) where again where )1’ = (n’l’j’m’T’) and p = (nljmr). Matrix elements of the spin-orbit part The matrix element of the spin-orbit part of the single particle field, 3W - 6 x ('7', reads 3' " _. . . . 3 R ' R" x 6”! (51.1.1 5mm, 5”! , (9.341) where again 21’ = (n’l’j’m’r’) and p = (7107717). 2This must be due to the practical violation of the completeness relation and the use of a finite box size. It can be shown numerically (by increasing the box size) that effect of the later is minimal. 253 9.8.4 Self-consistent iterations and convergence As can be seen from the results of the matrix elements for the three parts of the single-particle fields, the hamiltonian couples basis states only within a single I — j block. Hence, in the actual numerical solution of the HF equations, one diagonalizes each I — j block independently. Of course, the other parts of the calculation will involve all the relevant l — j blocks. To drive the calculation towards convergence, we implemented both Broyden’s method [194] and Imaginary-time method in separate calculations. After convergence, the results of the two methods usually agree to three decimal points, and hence the results reported in section 5.4.6 have been obtained using both methods. 9.9 The HF energy of chiral EFT NNN interaction at N2LO Here, we give a few remarks on the symbolic derivation of the HF energy of chiral EFT NNN interaction at N 2L0 and give the complete expression for non time-reversal invariant systems. The corresponding simplified expressions for IN M, PNM and time- reversal invariant systems are also stated. 9.9.1 Remarks on the symbolic implementation The details of the symbolic derivation of the HF energy from the chiral EFT is discussed in Ref. [156]. In addition to automating a tremendous amount of spin— isospin and other algebraic steps, we have demonstrated that the approach can be generalized to treat nonlocal interactions such as the quasi-local Skyrme interactions. There are several extensions of the symbolic derivation that can be made in the future: (i) One can envision expanding the work in such a way that first-order pairing 254 correlations (due to the N N N interaction) are treated along with the HF part, viz, performing HFB (Hartree-Fock—Bogoliubov) calculations. Combining this extension with proton-neutron mixing, one can have a start- ing Skyrme-like functional that can be used to handle proton-neutron pairing correlations as discussed in Ref. [207]. (iv) Implementing a similar scheme to treat four-nucleon interactions can also be one area of extension. 9.9.2 HF energy from the E-term Direct part The direct part, which comes from the E—term, reads . ~ - 1 (V3 '3‘“) = 5 E / drpo09 7909. (9.342) Single-exchange part The contribution from the single-exchange part, which originates from the E-term, reads (vamp = 2B [07(37239 + 32.09pm + 3po09 Bum-Bum —po09 B109 - 3109 + 4p109st'o09 - 51(9) (9343) Double-exchange part The contribution from double-exchange part of the E-term reads ., 1 2 . 3 _. 043?“) = 1—6E / .1. [371309 + 72.097139 + 9p..<9so(9-so<9 +1000") 51(0-5'1m + 2p1(79§1(7‘)'«?o(7‘) . (9-344) 255 E—term contribution for specific systems In symmetric IN M, the HF energy from the E-term reduces to = — 1% B / mm. (9.345) In unpolarized PNM (pure neutron matter), the HF energy from the E—term vanishes which is due to Pauli exclusion principle. In time-reversal invariant systems, the HF energy contribution from the E-term takes the form wit”) = —-,% E / 0(7):;09 — pom/£09 . (9.347) 9.9.3 HF energy from the D-term Direct part The contribution from the direct part of the D-term reads urban- —9A CD 1 iq (r —r ) __‘I___23‘13 , , : ___ 2 3 3 2 (V337 ) 4f2 f2A -2- /(172dr3 [1) (277) 3d(13( (12 +772? X72002) 3302) 109). (9.348) In symmetric IN M, there is no contribution from this term. Likewise, for time-reversal invariant systems where .§'0/1(F) = 0, the contribution from this term vanishes. 256 Single-exchange part The contribution from the single-exchange part of the D-term reads /HF.D,l;r __ —!}A CD 1 3., .-(r -r) Q3343, (lm ) _ mm4/d12dr3/(2i) ——3dq3 c. 3 3 2 m x[— p0(r3,r2)51(72)31(12,71) ‘750’1730(F3 F2)31('F21F3)3?(F2) —p1(F2.F3)30(F3 F2)81(F2) - 91(F3 F2)30(F2.F3)31(F2) +i€“”7 33(7‘},f};)s’l‘(f'3,f'2)3‘§3(F2) - p1(F2.F3)33(F31F2lsflF2) 1 1 —o -o -o -o -o _ 590(F2)31(F3.F2)31(F2aF3) — 5637 PO(T2)P1(7'3a7'2)P1(7'zs7'3) 1 a w'u -* a -' " W “ '* +56 W6 3 P(1(7‘2)-91(7'317'2)31(F27T3) 3 _. _. 71131339333. F2) 3(7‘217‘3) _. 3 _. .. '3 .. +po(7'2)3[1 (‘7'317‘2)3'17(7'2.7‘3) 2 .37 90(72)P0(F31F2)00(F21F3) + (5137- P0(F2)/)1(F3. F2)fl1(F2. F3) 3 . _. _. _. . .. _. + 5 60"” (““3" p0(r2) 83(1'3, 7'2) 3502, 7‘3) — €07” CW” 100(F2) 3?('F3. F2)3i)(F2a F3l] - (9349) For symmetric IN M, the expression simplifies to .‘ — CD 1 q? VHI,D.1:1¢,INM : 914 /d7‘ “d /_<_2_1__) ___d iQ3. .(r3— r2) <31» ) 4f2 f2A——;r 4 7‘2 T3 (13 e q3+m,2, X [—3 5 p0(7‘2)p()(7'3, 7‘2)/)()(7‘2, 733)] , (9.350) while for unpolarized PNM, one has 2. , — CD 1 1 ._. _. .. (1.2 VHI,D.1:.PNM : 9.4 —/d"3d“’-/ d" lQ3.(r3—r2) 3 < J” l 41.2 f3A.4 ’2 ’3 (2733 7‘38 q§+m3 3 x [— 2 p.. 4f? f2A1: 4 ”(7‘ (271)3“136 (73—sz >< [—3 E511 00(72)P0(7:1 772)/}o(7"'2 11.73) 1 _. _. _. _. _. + 5 5117 p0(7‘2)/)1(7‘3, 7’2)p1(7‘2, 7‘3) + _ CW 518” p0(1='2)sg(r3,112113031125) 1217121911713,712191012211) . (9.352) Double-exchange part The contribution from the double-exchange part of the D-term reads — CD 1 qfi q VIIF,D,21' : 9A /d I—o [__2_1_) __gd iq3. .(r3— r2)______ 3 < 3N ) 4f; LEA: T6 T2( T3 (13 6 q, + 7112 x [3 (5,37 po('r‘72)po(7_"21‘7:31)P0(77317:2) — 5111/11(772)Po(7-"21F3)P1(7731772) + 3 p(1(7"2)s€(f72, 773193973: 7‘2) — 3 cm" 6MB" P0(7"2)33(7_"21773)33;(F31772) - 10007219130721 73)«917(7‘_:117—"2) +e"7”€wiiupo(7—"2)S 1"(7217‘3)1'(7'317‘2) +6/70(7_"217‘31)-‘>’g(712)33(7711”7:72) " 35137{70(73):773)~‘7g(7_‘:2)33(7?3v71:2) +3511 0003 72)93(7‘2)- 0(73 7‘2) — 2101072»'773)-9‘g(7-"72)51’1(F3a71:2) +5131P1(F217"?1)Sg(772)81y(7731772) — 137P1(511F2)8g(772)81(F217‘31) _. 23 (~12! 612113114 CW 331 (121932 (113, 1121933 (1;, 13,) 258 —23 ("13" 6"?" ”‘4 6"4W291;1(F22 F3)9o2(F2)~9g3(’-'32 F2) +i6 ("1‘63" "2’3“ ("47” 3,17‘71(24)9g2(F5-, F2)-9:)‘3(F22F5) + 2' 6‘1"" e"2“3"4 6"47”531(F2)-9l1‘2(F3-.F2)'9‘1‘3(F2=F5) + 2'. ("1‘3" e"‘2“3“4 ("47" 531(F2 F3) 962(F2) 903(F3a F2) 22 #13" "2*‘3Hd‘47" .“1( r2) ’1‘2(r3,r.-2) "3(‘F22'F3) + 3557 p1(F2)p0(F32 F2)/)1(F2e F3) — 5m p1(‘Fz)/)1(F32 F2)PD(F22 F3) —P1(F' )Sg(7‘227‘3)5i(7'3772) + (071/ fwdu p1(F2)33(‘F22 F3).9:’(F3,F2) + 3 P1(F2)‘9g(F32 F2)5il(F2a 713) — 3 6M" 6w" p1(F2)83(F32 Fz)'9‘i'°(F22 F3) — 2po(F2a'F3)9ij(F2)32(F3~F2) + 6x37 00(F2 7 39) i3('F2)31(F32F2) + 363,. p0(r3. r2)9’13(F-2)81(F3 F2) + 6101(F2a F3)5f(F2)33(F32 F2) — 3 63,, p1(F2, F3)3'[13(F2)33(F3, F2) — 5,37 201(F3, F2)«9'?(F2)33(F2a F5) + i 621193” €u2213l‘4 6‘4"" 5:1(F2)s’;2(f'3, 213.933 (F2, F3) + 2: #19" €“2"3"4 6‘47" 351 (7‘2, 7%)8’1‘2 (F2)8i'3 (F3, F2) _ 22 fulfill 6222324 2242'” s’;1(2='2)s’;2(r3, 2333332, 53) _ 2'3 6‘15" €p2p3p4 3‘4?" 3‘1” (@352 (F3, F2)s’1'3(f‘2, F3) — 2'3 6‘1"?” e"2"3"4 6‘47" 3:1(F22 F3)3i‘2(F2)=9$3(F32 F2) +26 21116” (#22324 2242" 3’1‘1(r2)sp2(773 r2) $303,273) . (9.353) In symmetric IN M, the contribution reduces to ‘ ‘ 'x ' ’92! CD 1 (13 VHP’D'Z ,IMI : l (11" jd; ciq3-(r3-12) __ (W > 4!: LEA-.216 ”2 3 1‘ q§+méi 259 X I: 3 [)()(7—':2)/)0(7-':2. F3)p0(7-':5, 7;) :l , (9.354) while for unpolarized PNM, one has ‘ . ) V _ CD . (1:2 VHP.D.2.r.I NM :- 941/ l—o d /_(_2_1__)1 iq3. .(F3—r2) < 3N > 4f? PA 16 ( r2 T3 3( (13 e (13 + m2 X [ 3pyz(‘f72)p1.('72,F3)pn(7:35,7::2)] - (9355) For time-reversal invariant s ’steins, usin the relations 5’0 1 r = 0 and 0 1 772,173 = / / p0/1(f"3, F2), one obtains — CD 1 1 VHF.D,2.r.,TRI = 9A _ fd—o d” f < 3N > 4f; LEA; 16 r2 r3 (27r 3 x [35m puma/Mg, F3)Po(7-"3, F2) B 7 £153 653.6342) 93% (1% + mgr + 5137' P1(F2)P0(F2, 7130/9101?» 7.32) +3p.(2=r2)s€(r='2, mswm) — 3 6m" 6”" po(7—"2)83(‘F2a fi;)8‘6(7::3, F2) —P0(’2)3 a[(73 7‘3)9‘1("3~7‘2) + 6m" 6w" po(F2)3?(F2a F3)8°1"(773, 72'2) +2p1(F2)-95(F2,738193,?» — 2 607” CLUB” p1(F2)83('F2, F3)8T (F3, F2) . (9.356) 260 D-term contribution for specific systems Combining the results obtained for the D-term, the contribution to the HF energy of symmetric INM reads ', — CD 1 1 -- .. ~ (1.2 ( 3” l 4 f3 133A. 16 ””3 (2703 W (13+mgr X [ ‘3po(772).00(7::29F3)PO(F3aF2)]- (9-357) Combining the results obtained for the D-term, the contribution to the HF energy of unpolarized PNM reads ___ _QA CD i fdf. (17.3/ 1 drf ei-(73.(F3—F2) (I; 3” 4f: LEA. 16 2 3 (2703 “ (1% +m£ X [ _3 pn(772)pn(772» 7.;?3)p11(7?3v7?2)] a (9358) where pn(F) and pn(F, 1"" ) refer to the local and non-local parts of neutron matter density. In time-reversal invariant systems, the HF energy contribution from the D—term takes the form — CD 1 1 -~ ~ ~ (1%.7 MHF,D,TRI : gA _ [d-od-o / d-O zq3.(r3—r2) 3 3 < 3” l 4f: ffiA. 16 T2 ’"3 (2w)3 q” q§+m£ x l46.5.po(2='2>po<fi2,r3>po(fa,F2) + 2 5m {10(F2)101(F2, 773)!)1073: F2) + 5,37 p1(F2)p0(F2, 71'3)/)1(”F3a 7:32) — 3 [2007983012,ers-3%. F2) + 3 6m" 6W3” p0(-F2)38(F2. f3)s‘(j( 173, F2) + p0('f'2)s‘13(7"'2, mews, F2) 071/ (yd-£311 _ 6 p0(F2)S?(F2, F3)S‘i)(7737 772) 261 +wreaks-”('52.fis>sY(W2> - 2 6m” 6”" P1(Fiz)83(7‘72, F3)s‘f(F3, 'le - (9359) 9.9.4 HF energy from the C-term The HF energy from the C-term of the chiral EFT N2LO 3N F can be grouped into two groups: a D-like term and remaining terms (which we call R-part). This grouping originates from the operator structure of F3: given in Eq.(2.20). The D-like term is associated with 605 [ —4 fig + 2 igrcj; (13] whereas the R—part relates to 2260377,: Ek- ((f,- + (L). In the following, the HF energy from the various parts of the C-term are given. Direct part The contribution from the D-like piece of the direct part is 2 _ ~ -. l 1 ,s _. s .- s .. <%Ig‘,CD,dn> : (g; ) 5 deIdF2(l7?3 [quadfiezq21r2—rl) ezq3.(r3—r1) 1r '7 2 c —4 +2— . ‘ (r13 +m2xq§ +m§> l L? Le"? ‘13 XPO(F1)3?(F2)SY(F3)- (9.360) For both symmetric INM and time-reversal invariant systems, this contribution van- ishes. The R-part contribution from the direct part vanishes (Vgfim’i’dir) = 0. (9.361) Hence, the direct part vanishes for spin-unpolarized INM/PNM and time-reversal invariant systems. 262 Single-exchange part The contribution from the D-like piece of the single—exchange part reads 2 1 1 _-~ 7 .~ ~ ~ = (2%) Z/ ”1“”1’3/ Wdedaewm-rnelm-m 1f 7 2 —4——+2— -. (q%+m2.>(q§+m2.>l f: 3‘” ‘13 X [—Po(773»7'"'1)3i}(772)3i(7?1f3) _z'em 8‘13(1'"'2)s’0‘(7"3, F1)s‘f(f"1,773) — P1071, F3) sf('f’2)83(f'3, F1) — 101073: F1) 3?(F2)33(F1a 7:3) —i 6"” 8f(7’2)-95(F3, 771)8‘i’(771.7"3) - 91071, 773)«9lj(7-"2)83(F3a F1) 1 ,3 1 —‘2'00(771)31(7733F2)3i(72,f35) - 55.87 P0(F1)pl('F3’F2)p1(F2’F3) 1 V w u -* -' ‘* w " 4 +56” 6 7 po(7‘1)8’1‘(7‘3a7‘2)31(7‘2ar3) 3 —O _ 5 90(7‘1)86 (7’3, 7‘2)33(T22 7‘3) _. .. _. _. _. 3 .. _. .. _. _. +00(7'1)3?(7‘337‘2)3i(7‘2a7‘3) — 55m Po(T1)Po(T3aT2)/)0(T2a7"3) + 6x37 pO(F1)p1(F3a F2)Pl (772’ F3) 3 _. _, _, _. .. + 5 wwem" p0(r1)sf)‘(r3, r2)8‘6’(7‘2: 7‘3) — Gym/6M7” p0(F1)8lll(T-:'3, F2)SLIU(F2, 7.33)] . (9.362) For INM, this expression reduces to 2 ‘ ‘ I 2 1 —o —o —o 1 —-o —o 1". '7 —F ._. F —-F (VB’S'CD’IXJNM) = (79;) X fd‘l‘1dT-2d7‘13/(2 )6dQ2dq3e‘l2-('2 1)e'Q3~(3 1) 1r 71' fl 7 2 (1293 617". 03 3 - . . -4 +2—r '0 (q§+m2,) l f3 L!” “l 3 .. _. .. .. _. >< [— 5 537- P0(7‘1)Po(7‘3a 7‘2)l00(7‘2a 73)] 9 (9-363) 263 while for unpolarized PNM, one has 1 1 -~ 7 ~ ~ ~ ~ (LrLIIF, ,CD 1)(. .PNM) : (g; )24 _/drld7—: 2(17-03 / (2 )6 déTquij elq2.(12—r1) ezq3.(r3—r1) 7r 71’ . ‘7 2 (12(13 4017",. 203 _. _. X . —( (13+ ".3331. n23.) l 7.3— + I: ’2 “l 3 _, _. _. _. .. X [— 5 6L”? pn('rl)pn(7'3v T"2)pn(7'2e 7.3)] - (9364) For time-reversal invariant systems, it becomes 21 1 ,3 _, _, ._. .4 _. < 3N > Q—fn 71 7‘2 7‘3 (27rls (12 Q3? 8 2 +2—(I2'Q3 X(c1§ +m {Wig +7713) 4f2 f3 36 x[- 5 3.3.3.1 (7'1)100(73172)100(r2ar3) 1 + 2 5m PO(T1)P1(737 7‘2)p1(7‘2, 7 3) +51000031(7‘337‘2)31(7‘2,7‘3) — §PO(7‘1)30(7‘33T2)"’0(r2’r3) 3 1"” w -y -o —o —o u; —o a +§€ld 6 ‘7 100(7‘1)Sg(7‘377‘2)30(7‘2,7'3) 1 u w u _. .. -' «2 ~ -' — 53" e 3 posl (23,731] . (9.365) The contribution from the R—part of the single-exchange piece reads 1 ._. _. _. ._. _. .. (Mg/Foam) = (29;)2 ;—:2- : /df'1dr2dr3/qufidfie'qz-(rr’l)8333-03—71) 31 B2 71 20. " —0 ' I B2 o_ 7' .7. (-i T ,7" .z....m.[a‘ii(7-'33 F2) —533232 /)0(7_':237T3)301(7-"33772) + 53232 Po("-"33F2)P1(F23773) ] - (9.366) The contribution of the R-part of the single-exchange piece vanishes for both spin— unpolarized INM/PNM and time-reversal invariant systems. Double-exchange part The contribution from the D-like piece of the double-exchange part reads HF,CD,2 (V333 x) 2 1 1 ~-o co -o ---o :0 --0 (9A) [df'ldfizdig / qu‘gdq‘g 6'02-(r2-T1)c'03-('3-’1) 7T ‘27; IE 5 7 2 (12(13 Clmw C3 —o ... . . - +2— « ~(- (33 +m£> 1 #1 _. --30 (71,7:3)8#2(7‘2271)313(F-‘59'F2)J 3 266 X [_Ep3plp4(Li/147262,;72V + €;3/13;146;42;r4u€7p1u __ €13p3u4 6114;11u6'w2u _ 6.7/13Il46/11114V6311211] ] . (9367) For symmetric INM, this reduces to 2 .‘( v, 1 _, __, _, 1 —o —o 1" 1“. —F I... F —F (lGESCD'Z‘JM’) = (#29?) 1—6[drldmdrg/-——(27r)6dq2dq3€q2-(2 1)(/1343 1) 7r 13 “r 2 x —4—,—- + 2—(12-q3] (2% + mime? + mi) f: f3 x [35732PolF1,FilP0(F22F1)PO(7-:3,73)]2 (9.368) while for unpolarized PNM, one has 2 ‘ , 1 —' —o —o 1 4 —-0 '_° _' -_. 'F 7 _-' (VJLF‘CDJX‘PNM) = (Ty?) -1_6 /d7'1(17‘2dr3 f—(2 )sdq2dq3 e'q2'lr2 ’lle"13-(’3 r1) 1r 7f 7 2 (1503 €1m1r 63.. .. . -4 +2—. ' . (22+m2)(q§+m3,>l f3 f£q2 “l X I: 36137 7012(771, 71:3)pn(7?2a Fl)pn(773$ 772)] ° (9-369) For time-reversal invariant systems, the expression given in Eq.(9.367) can be simpli- fied only slightly. The reason is the phase factor eii2'lf2"F1)ei§3'(F3‘Fll prevents one from treating F1 on equal footing as F2 and F3, i.e. even though one can interchange F2 and F3 and recover the same expression, the same can not be said of F1 and F2 or F1 and F3. This is further compounded by the fact that the HF energy of the C-term’s double exchange involves invariably three non-local scalar / vector densities. Another simple interpretation of this is, most if not all of the available symmetries in the co- ordinates have already been utilized in Eqs. (7.13)-(7.15). Hence, the corresponding expression reads ‘2 1 1 .2 2 2 .2 2 2 Vl‘if2CD.2X.TRl : 9.4 _ [die (1F (117 / (Ii-rd“ ezq2.(r2—TI)GIQ3.(T3-r1) ( 33. ) 2f 16 1 2 .3 ___—(2mg 12 (Is 267 I3 "7' '2 (1.2 (13 01777.7r c3 _, _, >< . —'—++2—qz-qa (61-3 + "1%)(73 +7712?) f: f3 X 35,132 90(7-‘22 F1)P0(F3-, 772)!)0071: F3) — 6,22 [7002, F1)p1(F32 F2)p1(F1~ Fa) + 9 po(Fz. Fi)b€(Fs2 F2)33(F1« 71’3) _ gfuBVEW p003, 7‘} 223073, @2301, F2) — 3po(F2,F1)29?(F32F2)5i(7712F3) + €pl3l/6w‘7-y poll-".27 F1)8’1‘(F3, F2)s‘f(F1, F3) — 35737 700071, F3)Sg(F22 7-"l)*9il(7?3v F2) + 35237 po(F3, 7"*2)-""g(7?22 7‘31)330‘9 773) — 2p1(F1, F2)SB(F22 F1)8l(F3» F2) + 6237 p1(F1, 508372, Fll3il'F3w 7:72) - 5m P1('F3, F2)33(F22F1)3i(fivfii) + 357% P0073: F2)p1 (F22 F1)p1 (F1? F3) - ,32 pow—"12F3)P1(F22F1)P1(F3aF2) + 3p1('F2,F1)Sg(F32 F2)3i(7712 7:23) — 36"5"6“’7" p1(F2, F1)86‘(F3,772)3‘f(7712f3) -701(‘F22F1)~9?(F32F2)83(F1,F3) + 6 P1071,773)3113(‘F2a771)37(7_‘32F2) — 35,32 p1(F],fi,)s’13(F2,F1)83(F3,F2) —o —‘ ,3 -o -o ' -o -o — 5,232 [)1(7‘3-. 7‘2)31 (7'22 T1)83(7'12 7'3) 268 "L L. + 2:3 ( 1m 12323202, 77088302 F2) _ % S[11101371090202 F1)5’1‘3(1‘{,f72) +.2;‘1(a,>-:,>.:‘2(r2 22233222» $an 23> "232 ml- <>> X [_ffl3fllu463p4l/E‘,;12u + 6t3I‘3H46l‘2M4l/67p1u _6511373467147111262'73231 __ €7'u371467t1714vfx331213] . (9.370) The contribution from the R-part of the double-exchange of the C-term is composed of (‘éfllvF,CR,2X> : <%fLF,CR1.2X> + (%}kF,CR2,2X) + (M’IIJVF.CR3,2X)+ (v3HF “CR4 2X) (9.371) VHF.CR1,‘2X> <‘/HF,CR‘2.2x), (VlllflCR3,2x>, and (VHF,CR4.2x where ( 3N 3N M M ). Denoting CR1 E (v3lIIVF,CI{l,2X> , CR2 E (VHF CR2 2x) , CR3 E (V3 VHF CR3 2X) , CR4 E <%ILF,CR4.2X> , these are given by 922527 1 CR1 = (2—fn)fi It: 8 fd‘FldFQd’FEs fwddlzd‘zs 82(12‘0'2—7'1)e"’3'("3-r1) '72 Q21 Q22 (131 Q3 ((1% + "7%)(613 + mi) 3 7 x l6 1 “‘1 (”5132726712313 + 6’321‘2672l‘3 + 652%672’9) 269 , we have CR2 x [ 67317172 (6 It _.I‘..—.II _. H ..II 31174.. (911(12.I‘1)sl‘2(7‘3 72)S‘03(Il.7'3) — 811(7‘,2 r1)902(13 7'2)31‘7(r1,7'3)) 1 A,‘ ‘ I —o —o —o --o —o —o —o —o —o —o + 677 ‘1”1 011232 3111(7‘217‘1)(/)1(7‘3-. 7'2)/)0(7'197‘3) ‘ P0031 7'2),01(7‘1>7'3)) ' 1' ‘1' .‘c‘W t It -' -0 H 4.115311771632272 (311(7'2-7‘1)512(71 7‘3)100(7'3 7‘2) I‘ —o —o H I1 —o —0 fl —0 —' -o -' ‘311(7'2e7‘1)- 902(7711 7'3)P1(7"3> 7"2) + 511(T2>Tl)'912(r32r2)/70(T1=r13) — 371710217‘1)902(7_'3>7‘72)p1(77’1>773)> ] - (57-372) 2 . C 2 -' -o d 1 1 r r 1 r r (g;—:) f—;8 fdrldrgdr3/(——2W)6dq2dq3602°(72‘ ~1)€q3'(3- 1) B 7 ‘7 (12B1q22qglq.2 ((12 +777 720013 +7772 7r) fl —0 —o I1 -o -o [1‘ —~o -o 6I72fl3 _ 6132;136’72172) 511(7‘39T2)502(T21771)313(T11T3) 13272 13 '7' 6 171711 ('6'3272ép2“3 + 6’32”2672”3 _ 652“3672"2) X8710] 7'3)90 2(7'217'1)913(773 T2) 3 + 6 171711 (—6332.326“2“3 + 6fi2p2672113 + 632I136‘72772) X31103, 7‘2)312(7'2> 717303011 T3) fl + 6 171711 (5132726112113 - 6732772572773 + 6527‘36727‘2) X 90101,T3)912(72171)313(T31T2) - fl -. -. I1 — 63171“10132u26'12#3 301(7"3> 7‘2)312(7"2> 7"1)31 (7711773) +if‘131'711/EB2’Y2V p1('I-::33 F2) p1(F2-> Fl )/)0('Fla E3) — p0(F21F1)/)1(F11 773)) 5,7 l [J] _, _, _. .. ... ._. +5 171 1033 2.72 (31 (7‘1, r3)p0('r2,1‘1)p1(7‘39 7‘2) #1 a -0 -0 4 "’ "' —81 (7'317'2)P0(7'2>7“1)Pl(7'1>7'3) +37171077317:72)p1(‘7?2>771)00(7—‘71>7‘73) — 901(7'1 7“3)Pl(’"2 r1)p1(r3,r2)) +i€.{31’71171(63212796311131 + 6152736322,, _ 62327363211 — 6172771652,?) 270 CR3 11 _, _, V _. _. _. —. x 311(7‘3, r2)812(7'12 7'3)P0("‘22 7'1) _ . 63171316323232 2’51 (F2, F2)st,‘3(F1. 71:32.32. F1) — l fi'31771u63'32;11u (— 831(F3F2)8:2(F1, 'FEg)p1(F'2,f'1) + 30 (’1, 72031 (7‘3, 7'2)/)1 (T21 7'1) [1 '7 l1 4 —o '7 —o —o —o —o ‘311(7 7'3 72) 02(72 7‘1)PI(7‘1~7‘3) + 311(7‘227‘1)812(7'127100003272) —sll(r",2 r1).s(,2(r1, -r )p1(r3.r2) + 811(7‘3.7“2)812(7‘227‘1)po(7‘127"3)) — z' 65171"672”1U (32121 (7‘3. 772%? (771» 7:3)701 (772» F1) 52 u .. .. _. .. _. .. +311(7"127'3)30 (T2,r1)p1(r3,7‘2) #1 _. _. _. _. _. _. u 13 +Sol(7‘1,72)~9f2(73272)pl(71273) — 501(71273)912(7°227‘1)101(7"327‘2) l. 3 _, _, __, _, —3:10—22F1)912(7‘32T2)Po(7‘127‘3))J- (9.373) 2 . ‘ C I 1 ,4 .. _. ._. _. ~ (29—__fA)B f128 fdfldfédfii f_(27r)6d(72(1625€"12‘(r2—r1)6103-(7‘3—1'1) 1r ’ *1 “r (15 (12 (131712 (92 + m12rXQ3 '7‘ mgr) x [2313131 (652.726 #2733 _ 5772732532733 — 632773532772) 1 p _. _. _. _. x (331(72,r1)812(r3,T2)S‘;3(7‘1~7‘3) [‘1 -o H _I‘ -o 1 -o l7 -o -0 —3so (12.232332 :2) 3(F1. F2.s>+2-1(r2.r1>’3sg3< 2, >) + 63171“ 5.3222 351522771) (3P0(F327::2)PO(F12 F33) " 91(fi33'F2)pl(F13F3)) _ 63131771 53272 s’:1(f'2,7"'1)2p1(7:§5,72:2)po(7_"1272:3) + 2‘ 5731717‘15‘32727‘2 ( — 3. ”1 (12 r1)sg2(71,F3)p1)(fi32F2) + 301(1‘2 77'31):2(F1,fi3)p1(173,f'2) [12( +3801(7‘2,T1)«S‘02(7‘3272)p0(7"1,7‘3) — 80102270817327‘2)Pl(7‘1 7‘3) 271 CR4 2 . 9A 047 _._._, l _‘dw «_q ‘.- ?_~ dr dr (17‘ —d d- 33232 317633-33 ’17 (an) f2 8/ l 2 3/(27T)6 (I2 (138 B '1 '7 (12 fi1‘12243193 2 ((12 + m%)(q3 + mi) 3 [ 6 171"1 (—652725,12,13 + 5132172672113 + 5132173672‘9) X (3 3:1(F3, F2)sg2 (F2, F1).933(F1, F3) —3T1(F3,F2)532(F 72 F1)513(F12F3) _ 2311(F3,F2)s;‘2(F2,F1).933(F1,F;)) + 533171“ (532225112113 — 53217257233 + 57723363272) x (3 331(771, F3).9$2(F2, F1)33‘3(F’3. F2) —-sin(Fl,F3)sg2(F2,F1)81l3(F32F2) — 2331(F12F3)3u2(F22F1)313(F32F2)) + 66171”155272(- 3 831(F12 F3)PO(F22 F1)P0(F32 F2) + 3 3310732 F2)PO(F22 F1).00(F1: F3) +311(F1,F3)p0(F2,F1)p1(F3, F2) - 911(F3,F2)p0(F2,F1)p1(F1,F3) — 2 3:10:32 F2)P1(F22 F1)PO(F12 F3) + 2 331(F12F3)P1(F22 F1)p1(F3, FD) + z’ 61317112652721» (@0032,F1)p0(F1,F3)po(F3,F2) — 2 p(1(F12F3)/11(F32F2)P1(F2a F1) —Po(F2~F1)P1(F32F2)/71(F1~'33)) + 22313131632327? (31‘103,F2)s-’1"2(F1,fis)po(v‘2.fi) — 3331073, F2)Sg2(F1,F3)po(F22F1) 272 I" —o —o I”! —I —0 -' -* + 2311(7‘3, r2)302(7‘1.7‘3)101(7°2a 7‘1) ) _ If 5731311’5327‘1" ( — 5.7171(771, 7‘35)S:2(F32F2)p0(F22F1) + 3 831 (FL @532 (F3: 772)P0(F2v F1) — 2 531071, F3)s:2(7_‘§;, F2)P1(F22 F1) — 3 331(1‘3, Fs)832(F22F1)p0(F3aF2) +3‘1‘1(-F1,733;)332(F2,F1)p1(F3,F2) + 3 331 (F3, F2).932(F2, F1)P0(F12 F3) — 371073, F2)832(F22 F1)/)1(F1, F3) —287102,F2)si’3(F2.F1)po = _ (5%) $55; [(171617‘2d'r3 / qumqa 273 3 ‘1 7’2 31 32 x(2772477241)(3373-53-31) ‘72 ‘12 Q3 713 (Q2 + "1222)(Q3 + mgr) x [#3171" 63272” 00(F12 F3)/)0(F22 F1 WOO—‘32 772)] 2 (9.376) while for unpolarized PNM, one has (LSHFWCRhPNM) = .Q_A 204 3 fair dr (1F [31/1 —dq dq 2_f2 f28 123 69:36 23 x 8122.(F2—F1),2q3.(F3—F1) Q21Q22Q31Q32 ‘ (q; +m3>x3 - V] 1th) 1 . ~ 2 u 1 p 2 “1 -' 9 385 + i] H2(k1‘3)$3 (2 (T1)- ( ' ) Hence, in the case where g“1 2 po /1, then (ill'ul = y'all, i.e. the current density. Since we are dealing with time-reversal invariant systems, 3.0/1 : 6. However, we did not set it to zero in Eqs. (9.382)-(9.383) as we wrote the equation to hold for both scalar and vector densities, and when ("I = Fig/1, then cf 1~”1 = J0/1~I‘1"1’ i.e. cartesian tensor spin-current density. As to €51 in Eq (9.385) and in general, it refers to a second order correction term in the expansion. It is analogous3 to iAp — T + 3/ 5 16% p of the DME discussed in section 5.3.3. Obviously, €51 = O in the case where c” 1 = 5'0/1. The same notation applies to (”2 and (”3. Even though all notations and conventions have been explicitly given in Ta- ble 1.2, we recap the ones used for the 7r-functions. Taking (”1 as an example, .0 .0 ,0 . . #51 mill ,wgl refer to the 7r—functlons of the local part of the non—local densrty ,1 .1 ,1 . (“1 and n31 ,7r’1‘1 ,ngl refer to the zeroth, first and second 7r—functlons for the local density that appears at first order with respect to the relative gradient opera- ” Q g tor. Hence, these 7r—functions are equal to 7T6), 71f, n5, #3, mi, 71'; or 7r3, 7rf, n3, 7rd, 7r1J , 71"; when §“1 = 100/1 or §“2 = .§'0/1 respectively. The example given in Eq.(9.385) il- 3It does not mean they are the same. 279 lustrates this statement explicitly. Since we are dealing with time-reversal invariant ~ — ~ - r - . u- 1r systems, 713,7ri',7r§,7r3.71{.7r§ are Irrelevant. HO’ and I‘ll2 can be seen as some common . ll.“ . . . prefactors while H2’ is the 7r—function of the second order correction. Comments on the DME ansatz The following comments are at play concerning the DME ansatz: (i) The ansatz is designed to be a general template on which all the known (and perhaps future) analytical/parameterized DMEs can be mapped. This can be done by setting the various n—functions to the values dictated by the analytical/parameterized DME at hand. This allows for a minimal effort to adopt the symbolic machinery to specific cases. (ii) As mentioned at the beginning of this section, any discrepancy between the EDF that results after the application of the DME, and the exact NNN HF energy is solely due to the DME ansatz of the nonlocal densities and the n—functions. This is a trivial statement in the case of two-body interactions. However, it is not so for three-body interactions as unless one makes a convenient coordinate choice, it is not trivial to treat even the non-DME part exactly. Thus, by improving the n—functions, one can hope to get better and better accuracy. 9.10.2 The G-tensors and their analytical forms In sections 721- 7.2.1, we identified the three generic forms (Eqs. (9.406)-(7.2.1)) of the terms that occur in the HF energy from the chiral EFT NNN interaction at N2LO. We refer to the interaction form factors that enter these equations as G-tensors. They are of the form 51 11 B2 72 Glfll'yli3272($2,$3,q2’(137w) E [(19 d9 eid’2i2eirf3f3 Q2 Q3 (12 (I3 , (’2 Q3 (<13 + ""%)(q§ + mi) (9.386) 280 13 7' '* r '1 ”- (1293 03703,.173,q-2,(13.w) E [(19, (I9. e'q2'126'q3'r3 . , * ’2 “3 ((15 + "18%)(93 +7713.) (9.387) qflql an,» 1.. _ , i“? .1 3 3 C11 (1'3,(13) : [(lflq3 (3 ((31.13 m. (9.388) L 71’ where (19,12 and qu3 refer to the differential solid angles of the two vectors, and w is the angle between :32 and 173. In the following, we derive the analytical forms of these tensors. Indices 51,71, 132,72 can take values {1,2,3} corresponding to the cartesian labels {.1:, y, 2}. To obtain the analytical form for 03171527261», 11:3, q2, q3, w), we define m -: 2'31}: (1697 F (:L,q) dflqe -———— (9.389) (12+m73' In the case where the vector 55 is along the 2? direction, we denote the F ”(ii, q) tensor as F ffli’, q), which equals ,Q‘ 5 3 Ff‘r(;i:',q) = 5),, /(que'q"'-q—q—q—. (9.390) 2 + 771.3, Denoting F,11 E Ff‘,Ff2 E Fzyy,Ff3 E Ff“, we obtain 2 . q 31(qrr) . = —4 , 9.391 F. (liq) 7rq2+m72r (gm) ( ) 2 . V (I 310151?) FW .. = —4 9.392 2 . . q . J1(Q~’L‘) 1W~ . = —— 2 .. —4 . 9.393 z ($.61) (12 +771; ( Mom) 7r (qr) ) ( ) Next, we need to define a convenient coordinate system. If one could define both 5132 and 5:3 to be along the 2 direction of the coordinate system, the G-tensor could be calculated very easily. But in the actual case, one cannot, in general, define both 932 and 5:3 to be along the :2 direction. Let :32 be such that it is along the 2 direction and 281 .73 be in the .1: — 2 plane. with angle (12 between .82 and 1'3. Next, write Gdlll'd2l2(332e$3a(12103-93) = F1‘31*~’2(f2,qe)F7172(f3,q3) = 651132 F 5131 (4132, (I2) Fi’17'2(1’39113-,w) » (9394) where due to the specific coordinate system chosen, the G-tensor will be nonzero only when {31 = 32. The same cannot be said about 71 and 72. In the next step, we concentrate on F 71 72(1‘3, q3,w). One can calculate this quantity by performing a few steps involving rotation of the coordinate system. Hence, rotate the coordinate system with respect to the y-axis such that 5:3 aligns with the new 2’ axis. The rotation matrix for this operation reads cos(w) 0 —sin(w) R(w) = 0 1 0 sin(w) 0 cos(w) The transformation of the various terms of F7172 is (obviously only the vectors get affected) (1 => RACE, (9.395) where we have left out other trivial terms. Using these intermediate results ' . , _ _ —1 —1 , fl “ , F7172 (1'37 (13' W) — 6Itll‘2 12111171 (w) RI12'72 (“2) F21 2(1‘3, (13) ' (9‘396) 282 Plugging this result into Eq. (9.394), one obtains the analytical form 131/31 0‘3111‘:’2"2(:r2.:173,(12.93.92) = 5131132R_l (WlR—l (to) F2 (£732,02)F§‘"(-’1332(13)- ml Mg (9.397) To calculate the related but different G tensor, C(37(1*2,173,q2,q3,w), we define We represent the case where the vector i" is along the z axis by F f (f, q) 4.. 3 F501", 4) = 33,. f 49.6“ q 7 q2 + 771?. where 63,3 E 623 and representing F,1 E F}, F,2 E F3, F23 E Ff, we obtain F:(.1:,q) = 0, Fzy(;l7,q) = 0, . q . Fz ;, = 4 ————,—' , . .(r q) I flq2+m§ 11(93‘) Plugging this result Gfi”(:r2,$3,qz,qa.w) = 638 63,.R;,1(w)Ff(:v2,q2) Ff(rvs,q3)— Finally, we have a trivial tensor G defined in Eq.(9.388) is given by (137(i’, q) E Ffll(.ii, q). 283 (9.398) (9.399) (9.400) (9.401) (9.402) (9.403) (9.404) (9.405) 9.10.3 Sample DME simplification After the specification of the DME ansatz and the analytical calculation of the G- tensors, the next logical step is to plug them into the respective exact HF energy terms to obtain the EDF. We discuss the DME simpification by considering a term that has the form of Generic-Form-l. Consider g:;.i.(kx3) ems)", (9.410) n max [mnza.c(n) 11:0 m=0 where ,j',f,(k;r2) and gf,‘,¥,,(k;1:3) are unspecified scalar functions and the number of terms in the inner summation depends on the value of n, expressed as (mmax(n)). Note that the angular integrations (with respect to 552 and 553) do not require the values of ,j‘,:,(k:r2) and gfi'f,,(ka:3) to be specified. In addition, the special symbolic technique that we developed, Ref. [161], helped us avoid specifying mmax(n) which is not known anyways. As to mum we found that nma,r = 5 suffices in the practical implementation [161]. In section 9.11.3, we discuss how these scalar functions are obtained for Ilifr derived from the generalized PSA-DME. In addition, we show that at nmam Z 5, Eq. (9.410) becomes practically exact for these Tr—functions. Actually, one can increase rim,t for any Hip if there is a need. The only problem with increasing um“;c to a much higher value is the rapid increase in the time-complexity of the symbolic computation. Similar to the application of the DME (ansatz) to the exact HF ener , the ex ansion of II? introduces et another increase in the number of gy p z,fr y terms to simplify. With this expansion at hand, we are able to perform the angular integrations with respect to :32 and :33. At this point, due to the complexity of the problem, we introduce the assumption of spherical symmetry. This implies that all the local densities depend only on the magnitude of the radial vector. In other words, we now have only three independent directions: f2, :33 and 7"}. Appendix 9.10.5 discusses how we can relax this assumption. The generic form of the required angular integrations 286 for the case of spherical symmetry read /(IQI2dS2I3 (i3 . 6)1(;1‘:2 . i3)1nli‘2 8' i73ln (i2 . 6)!) (5:2 ® fl)"1 "1 (.132 59 73):; (5:3 99 f1):g(.i:3 59 73):: , (9.411) where 1,771.,72. and p are integers. The maximum power of gradient in the DME ansatz for any density is fixed at two, hence, I e {0, 1,2} and also pc {0, 1,2}. Due to the specific form of the ansatz, mc{0,1,2,3,4,5,6,7,8,9, 10} and n6 {0, l, 2}. The ex- ponents 111, 712,113,714 e{0,1}. Even though the generic angular dependence given in Eq. (9.411) is very complex, all these terms do not occur at the same time. The origin of the various angular dependencies is (i) 3.6 and 332.6 are due to the DME ansatz, (ii) $2.123 originates from the DME ansatz (Eqs. (9.382)-(9.384)), the ro- tation matrix of the G tensors (Eq. (9.102)) and the expansion of 11:2}, (Eq. (9.410)) (iii) |:i‘:2 ® :i:3l comes from the rotation matrix in the G tensors and (iv) (:22 (8) 73):: where is {1, 2, 3, 4} originate from the directional coupling in the DME of the vector density, 32,“. Remember that in spherical symmetry, only the vector component of the cartesian spin-current tensor density is nonzero. The exponents n1, n2, 71.3 and 714 can not be one at the same time i.e, at most only three of the exponents can be one at the same time. This is due to the fact that only three local/non-local densities (be it vector/ scalar) densities appear in all terms of the exact 3NF HF energy. Due to the huge number of terms generated by the DME expansion, direct multi- dimensional (four dimensions) angular integrations is both impossible and not re- quired. Rather we developed a Mathematica rule-based technique to replace the multi-dimensional integrals with four independent single dimensional integrals. The merit of this technique is that one can calculate the single dimensional integrals once and use their stored values in the whole computation. The angular integrations is 287 followed by several symbolic manipulation techniques to obtain the final EDF. The details of these techniques can be found in Ref. [161]. We remark that the symbolic automation enabled us to keep all higher-order terms in the final EDF (up to sixth order) for whenever necessary. 9.10.4 Contributions to EDF-NNN-DME In the following, we list the EDF, truncated at fourth-order, that results from the application of the DME on specific contributions to the HF energy (of time-reversal invariant systems) from chiral EFT NNN interaction at N2LO. Due to the assumption of spherical symmetry that we imposed in the previous angular integration step, the given expressions hold only for spherical systems. The actual values of the couplings as a functional of the w—functions is found in the Mathematica files of Ref. [161]. First let us define the auxiliary quantity (3 /1(7“') as «kl/10") = "I" ° firm/10*). (9.412) Also note that Cfi/i (7") stands for the second-order correction density which in the case of generalized PSA-DME reads 1 3 C0310:.) = 5400/15.) — 70/10:) + 3ka 100/10“). (9413) Fourth order EDF from the D-term D .. P3 '3 PO”? 2 ”8‘8 2 2 8 = / dr C. me + 6. pummm + 6. powwow pi‘fi 2 2 ”OFF? 2 +6’1 101071900) + Cl 90(7‘11110'79‘10’) J2 .. .. _. _. +Cf0 0 Pom J09“) ' Jom + CfIJOJl 10107) J00?) -J1(r”) 288 J2 ._. .4 I. ’ d a ‘ +ch 1pm .1199 - .119?) + 6.wa pom [v - Jomli +ci’0"°’“° pom J13 0 is a real parameter and n is an integer. Conver- gence of the integral requires that n+u+u Z 0. Note that in our problem, u = kp/mfl. The application of Gegenbaur’s expansion introduces higher order spherical Bessel functions. I.e. if the expansion is truncated at p = 5, up to sixth-order (35(1)) spher- ical Bessel functions are introduced. However, we are not aware of any analytical or symbolic integration technique that works for any values of the indices: {11, u, ,u} which satisfy the convergence criterion. Furthermore, we need to perform calcula- tion of hundreds these integrals. In Ref. [157], we discuss how we solve this problem with the Mathematica package that we developed. In the package, we designed and implemented a symbolic integration algorithm that can calculate these integrals for any values of {71, V, p}, which satisfy the convergence criterion, and gives the exact analytical expression. In principle, we have analytical expressions for the couplings at this point. How- ever, the actual Mathematica implementation of the DME starting from the exact 302 HF energy from the chiral EFT N NN interaction at N2LO has not been discussed. In Ref. [161], we give a detailed presentation of the symbolic implementation. We remark that the implementation is done in such a way that it can be adapted to deformed time-reversal invariant systems as discussed in appendix 9.10.5 with only a modest amount of change. The implementation for deformed time-reversal invariant systems will be the subject of a subsequent publication [162]. A detailed analysis and discussion of these couplings can be found in Ref. [160]. Finally, the lengths of the coupling expressions prevent us from reproducing even a single coupling in here. Consult the Mathematica files of Ref. [161]. 303 BIBLIOGRAPHY [1] W. Zuo, I. Bombaci and U. Lombardo. Phys. Rev. C 60 (1999) 024605. [2] A. Lejeune, U. Lombardo and W. Zuo. Phys. Lett. B 477 (2000) 45. [3] W. Zuo, A. Lejeune, U. Lombardo and J. F. Mathiot. Nucl. Phys. A 706 (2002) 418-430. [4] W. Zuo, A. Lejeune, U. Lombardo and J. F. Mathiot. Eur. Phys. J. A 14 (2002) 469-475. [5] F. Coester, S. Cohen, B. Day and C. M. Vincent. Phys. Rev. C l (1970) 769-776. [6] R. Brockmann and R. Machleidt. Nuclear Matter. Phys. Rev. C 42 (1990) 1965- 1980. [7] R. Brockmann and R. Machleidt. Finite nuclei. Phys. Rev. C 42 (1990) 1981- 1988. [8] A. Nogga, S. K. Bogner and A. Schwenk. Phys. Rev. C 70 (2004) 061002. [9] S. Weinberg. Nucl. Phys. B 363 (1991) 3-18. [10] C. Ordonez, L. Ray and U. van Kolk. Phys. Rev. C 53 (1996) 2086-2105. [11] U. van Kolk. Prog. Part. Nucl. Phys. 43 (1999) 337-418. [12] E. Epelbaum, W. Glockle and U.-G. Meijner. The two-nucleon system. Nucl. Phys. A 671 (2000) 295-331. [13] T. T. S. Kuo and E. Osnes. Springer Lect. Notes. Phys. 364 (1990) 100170. [14] K. Suzuki and S. Y. Lee. Prog. Theor. Phys. 64 (1980) 2091-2106. [15] S. K. Bogner, A. Schwenk, T. T. S. Kuo and G. E. Brown. ArXiv:0111042[nucl- th]. [16] R. B. Wiringa, V. G. J. Stoks and R. Schiavilla. Phys. Rev. C 51 (1995) 38-51. 304 [17] R. B. W'iringa, R. A. Smith and T. L. Ainsworth. Phys. Rev. C 29 (1984) 1207- 1221. [18] I. E. Lagaris and V. R. Pandharipande. N ucl. Phys. A 359 (1981) 331-348. [19] R. Machleidt. (Wiley, New York, NY, 1986), pp. 71-173. [20] R. Machleidt. vol. 19 (Plenum, New York, 1989). [21] R. Machleidt, K. Holinde and C. Elster. Phys. Rept. 149 (1987) 1-89. [22] R. Machleidt. Nucl. Phys. A 790 17—23, 2007. [23] S. C. Pieper and R. B. Wiringa. Annu. Rev. Nucl. Part. Sci. 51, 53 (2001) and references therein. [24] R. Roth, J. R. Gour and P. Piecuch. Phys. Rev. C 79 (2009) 054325. [25] R. Roth, J. R. Gour and P. Piecuch. Phys. Lett. B 679 (2009) 334-339. [26] M. Bender, P.-H. Heenen, P.-G. Reinhard. Rev. Mod. Phys. 75 (2003) 121 and references therein [27] SK. Bogner, R.J. Furnstahl, A. Schwenk. ArXiv:0912.3688[nucl-th] and refer- ences therein. [28] SK. Bogner, R.J. Furnstahl, A. Nogga, A. Schwenk. ArXiv:0903.3366[nucl-th]. [29] R. Roth, P. Papakonstantinou, N. Paar, H. Hergert, T. Nek, and H. Feldmeier. Phys. Rev. C 73, 044312 (2006). [30] NM. Hugenholtz. Physica 23 (1957) 533. [31] T. Lesinski, K. Bennaceur, T. Duguet, and J. Meyer. Phys. Rev. C74, 044315 (2006). [32] T. Lesinski, M. Bender, K. Bennaceur, T. Duguet, and J. Meyer. Phys. Rev. C76, 014312 (2007). [33] J. Margueron, H. Sagawa, and K. Hagino, Phys. Rev. C77, 054309 (2008). [34] T. Niksic, D. Vretenar, and P. Ring. Phys. Rev. C78, 034318 (2008). [35] B. G. Carlsson, J. Dobaczewski, and M. Kortelainen. Phys. Rev. C78, 044326 (2009). [36] S. Goriely, S. Hilaire, M. Girod, and 8. Peru. Phys. Rev. Lett. 102, 242501 (2009). [37] H. Yukawa. Proc. Phys. Math. Soc. Japan 17, 48 (1935). [38] Peter Ring, Peter Schuck. The nuclear many-body problem. (Springer, 2004). 305 [39] C. Q. Li and R. hv’lacllleidt. Phys. Rev. C 58 (1998) 1393-1402. [40] V. G. J. Stoks, P. C. van Campen, T. A. Rijken and J. J. De Swart. Phys. Rev. Lett. 61 (1988) 1702—1705. [41] C. Q. Li and R. Machleidt. Phys. Rev. C 58 (1998) 3153-3162. [42] E. M. Henley and G. A. Miller. (North-Holland, Amsterdam, 1979), pp. 405-434. [43] S. Okubo and R. E. Marshak. Ann. Phys. 4 (1958) 166-179. [44] H. Wita la, W. Glockle, D. Huber, J. Golak and H. Kamada. Phys. Rev. Lett. 81 (1998) 1183-1186. [45] S. Nemoto, K. Chmielewski, S. Oryu and P. U. Sauer. Phys. Rev. C 58 (1998) 2599-2602. [46] S. Kistryn et al. Phys. Rev. C 72 (2005) 044006. [47] A. Nogga, H. Kamada and W. Glockle. Phys. Rev. Lett. 85 (2000) 944-947. [48] A. Faessler, S. Krewald and G. J. Wagner. Phys. Rev. C 11 (1975) 2069-2072. [49] J. Fujita and H. Miyazawa. Prog. Theor. Phys. 17 (1957) 360-365. [50] B. Loiseau and Y. Nogami. Nucl. Phys. B2 (1967) 470-478. [51] B. S. Pudliner, V. R. Pandharipande, J. Carlson, S. C. Pieper and R. B. Wiringa. Phys. Rev. C 56 (1997) 1720-1750. [52] H. T. Coelho, T. K. Das and M. R. Robilotta. Phys. Rev. C 28 (1983) 1812-1828. [53] S. C. Pieper, V. R. Pandharipande, R. B. Wiringa and J. Carlson. Phys. Rev. C 64 (2001) 014001. [54] S. A. Coon and H. K. Han. Few Body Syst. 30 (2001) 131-141. [55] R. Bieber et al. Phys. Rev. Lett. 84 (2000) 606-609. [56] K. Ermisch et al. Phys. Rev. C 68 (2003) 051001. [57] N. Kalantar-Nayestanaki and E. Epelbaum. ArXiv:0703089[nucl-th]. [58] D. R. Entem and R. Machleidt. Phys. Rev. C 68 (2003) 041001. arXiv:0304018[nucl-th] [59] E. Epelbaum, W. Glockle and U.-G. Meissner. Nucl. Phys. A747 (2005) 362-424. [60] E. Epelbaum, H. Krebs, D. Lee, U. -G. Meissner. Eur. Phys. J. A 41, (2009), 125-139. 306 [61] M. C. M. Rentmeester, R. G. E. Timmermans and J. J. (10 Swart. Phys. Rev. C 67, (2003) 044001. [62] D. R. Entem and R. Machleidt. Phys. Rev. C 66 (2002) 014002. [63] E. Epelbaum, W. Glockle and U.-G. Meijner. Eur. Phys. J. A 19 (2004) 125-137. [64] E. Epelbaum, W. Glockle and U.-G. Meijner. Eur. Phys. J. A 19 (2004) 401-412. [65] E. Epelbaum. Eur. Phys. J. A 34 (2007) 197-214. [66] J. Fujita and H. Miyazawa. Prog. Theor. Phys. 17 (1957) 360-365. [67] E. Epelbaum, H. Krebs and U.-G. Meijner. ArXiv:0712.1969[nucl—th]. [68] E. Epelbaum, H. Krebs and U.-G. Meijner. ArXiv:0801.1299[nucl-th]. [69] E. Epelbaum et al. Phys. Rev. C 66 (2002) 064001. [70] B. D. Day. Rev. Mod. Phys. 39 (1967) 719-744. [71] M. Cell-Mann and F. B. Low. Phys. Rev. 84 (1951) 350-354. [72] A. Fetter and J. D. Walecka. Quantum Theory of Many-Particle Systems (McGraw-Hill, New-York, 1971). [73] K. A. Brueckner. Phys. Rev. 100 (1955) 36-45. [74] K. A. Brueckner. Phys. Rev. 100 (1955) 36-45. [75] K. A. Brueckner, C. A. Levinson and H. M. Mahmoud. Phys. Rev. 95 (1954) 217-228. [76] K. A. Brueckner and C. A. Levinson. Phys. Rev. 97 (1954) 1344-1352. [77] [178] K. A. Brueckner, J. L. Gammel and H. Weitzner. Phys. Rev. 110 (1958) 431-445. [78] [179] K. A. Brueckner, A. M. Lockett and M. Rotenberg. Phys. Rev. 121 (1958) 255-269. [79] J. Goldstone. Proc. R. Soc. London Ser. A 239 (1957) 267-279. [80] B. H. Brandow. Rev. Mod. Phys. 39 (1967) 771-828. [81] V. Rotival. PhD thesis. 2009. [82] H. Q. Song, M. Baldo, G. Giansiracusa and U. Lombardo. Phys. Rev. Lett. 81 (1998) 1582-1587. [83] B. H. Brandow. Phys. Rev. 152 (1966) 863—882. 307 [84] R. L. Becker, A. D. Mackellar and B. M. Morris. Phys. Rev. 172 (1968) 1264- 1290. I [85] S. Kohler and R. J. McCarthy. Nucl. Phys. A106 (1968) 313-322. [86] R. K. Tripathi, A. Faessler and A. D. MacKellar. Phys. Rev. C 8 (1979) 129-134. [87] R. D. Mattuck. A guide to Feynman Diagrams in the Many-Body problem. 1975. [88] R. G. Paar and W. Yang. Density Functional Theory of Atoms and Molecules (Oxford University Press, Oxford, 1989). [89] J. Perdew and S. Kurth. Density functionals for non-relativistic coulomb systems in the new century, A Primer in Density Functional Theory, pp. 155, Springer, 2003. [90] R. M. Dreizler and E. K. U. Gross. Density Functional Theory (Springer, Berlin, 1990) [91] W. Kohn: Nobel Lecture. Rev. Mod. Phys. 71 (1999) 1253-1266. [92] M. Head-Gordon and E. Artacho, Phys. Today 61 (2008), 0809.2041. [93] W. Kutzelnigg, J. Mol. Struct. 768 (2006) 163. [94] J .E. Drut, R.J. Furnstahl, L. Platter. ArXiv:0906.1463v2[nucl—th]. [95] P. Hohenberg and W. Kohn. Phys. Rev. 136 (1964) B864-B871. [96] W. Kohn and L. J. Sham. Phys. Rev. 137 (1964) A1697—A1705. [97] W. Kohn and L. Sham. Phys. Rev. 140 (1965) A1133-A1138. [98] J. Engel, Phys. Rev. C75 (2007) 014306. [99] RC. Giraud, B.K. Jennings and RR. Barrett, Phys. Rev. A78 (2008) 032507. [100] N. Barnea, Phys. Rev. C76 (2007) 067302. [101] BC. Giraud, Phys. Rev. C77 (2008) 014311. [102] BC. Giraud, Phys. Rev. C78 (2008) 014307. [103] J. Messud, M. Bender and E. Suraud, (2009), ArXiv:0904.0162[nucl-th]. [104] L. N. Oliveira, E. K. Gross and W. Kohn. Phys. Rev. Lett. 60 (1988) 2430-2433. [105] D. Lacroix, T. Duguet, M. Bender. Phys. Rev. C 79:044318, 2009. [106] J. R. Stone, P.-G. Reinhard. Prog. Part. Nucl. Phys. 58:587-657, 2007 and references therein. 308 [107] B. Gebremariam, T. Duguet and S. K. Bogner. ArXiv:0910.4979v1 [nucl-th]. [108] T. H. R. Skyrme. Phil. Mag. 1 (1956) 1043-1054. [109] M. Kortelainen, J. Dobaczewski, K. Mizuyama, J. Toivanen. Phys. Rev. C 772064307, 2008. [110] T. Lesinski, T. Duguet, K. Bennaceur, and J. Meyer, Eur. Phys. J. A40, 121 (2009). [111] J. E. Drut, R. J. Furnstahl, and L. Platter (2009), Arxiv:0906.1463 [nucl-th]. [112] T. Duguet and T. Lesinski, AIP Conf. Proc. 1165, 243 (2009). [113] S. K. Bogner, R. J. Furnstahl, and L. Platter, Eur. Phys. J. A39, 219 (2009). [114] N. Kaiser, Phys. Rev. C68, 014323 (2003). [115] S. Hirata, S. Ivanov, I. Grabowski, R.J. Bartlett, K. Burke and J. Talman. J. Chem. Phys. 115, (2001), 1635. [116] R.J. Bartlett , I.Grabowski, S.Hirata and S.Ivanov. J. Chem. Phys. 122, (2005), 034104. [117] D. J. Dean and M. Hjorth—Jensen. Rev. Mod. Phys. 75 (2003) 607-656. [118] J. Terasaki, R. Wyss and P. H. Heenen. Phys. Lett. B 437 (1998) 1-6. [119] A. L. Goodman. Phys. Rev. C 60 (1999) 014311. [120] J. R. Stone, J. C. Miller, R. Koncewicz, P. D. Stevenson and M. R. Strayer. Phys. Rev. C 68 (2003) 034324. [121] M. Samyn, S. Goriely and J. M. Pearson. Nucl. Phys. A 725 (2002) 69-81. [122] S. Goriely, M. Samyn, M. Bender and J. M. Pearson. Phys. Rev. C 68 (2003) 054325. [123] M. Samyn, S. Goriely, M. Bender and J. M. Pearson. Phys. Rev. C 70 (2004) 044309. [124] J. A. Sheikh and P. Ring. Nucl. Phys. A 665 (2000) 71-91. [125] J. L. Egido and P. Ring. Nucl. Phys. A383 (1982) 189-204. [126] J. Dobaczewski and J. Dudek. Acta Phys. Pol. B 27 (1996) 45-58. [127] F. Tondeur. Nucl. Phys. A315 (1979) 353-369. [128] S. J. Krieger, P. Bonche, H. Flocard, P. Quentin and M. S. Weiss. Nucl. Phys. A517 (1990) 275-286. 309 [129] G. F. Bertsch and H. Esbensen. Ann. Phys. (NY) 209 (1991) 327-363. [130] N. Tajima, P. Bonche, H. Flocard, P.-H. Heenen and M. S. Weiss. Nucl. Phys. A551 (1993) 434-450. [131] J. Terasaki, P.-H. Heenen, P. Bonche, J. Dobaczewski and H. F locard. Nucl. Phys. A593 (1995) 1-20. [132] M. Honma, T. Otsuka, B. A. Brown and T. Mizusaki. Phys. Rev. C 65 (2002) 061301. [133] M. Honma, T. Otsuka, B. A. Brown and T. Mizusaki. Phys. Rev. C 69 (2004) 034335. [134] A. F. Lisetskiy, B. A. Brown, M. Horoi and H. Grawe. Phys. Rev. C 70 (2004) 044314. [135] E. Chabanat, J. Meyer, P. Bonche, R. Schaefer and P. Haensel. Nucl. Phys. A627 (1997) 710-746. [136] J. C. Slater. Phys. Rev. 85 (1951) 385-390. [137] J. S. Bell and T. H. R. Skyrme. Phil. Mag. 1 (1956) 1055-1068. [138] T. H. R. Skyrme. Phys. 9 (1959) 615-634. [139] T. H. R. Skyrme. Nucl. Phys. 9 (1959) 635-640. [140] E. Perlinska, S. G. Rohozinski, J. Dobaczewski and W. Nazarewicz. Phys. Rev. C 69 (2004) 014316. [141] J. R. Stone, J. C. Miller, R. Koncewicz, P. D. Stevenson and M. R. Strayer. Phys. Rev. C 68 (2003) 034324. [142] T. Lesinski, M. Bender, K. Bennaceur, T. Duguet and J. Meyer. Phys. Rev. C 76 (2007) 014312. [143] T. Lesinski, K. Bennaceur, T. Duguet and J. Meyer. Phys. Rev. C 74 (2006) 044315. [144] T. Lesinski. Priv. comm. [145] E. Chabanat, P. Bonche, P. Haensel, J. Meyer and R. Schaefer. Nucl. Phys. A635 (1998) 231-256. [146] M. Beiner, H. Flocard, N. Van Giai and P. Quentin. Phys. A238 (1975) 29-69. [147] F. Tondeur, M. Brack, M. Farine and J. M. Pearson. Nucl. Phys. A420 (1984) 297-319. [148] S. Kohler. Nucl. Phys. A258 (1976) 301-316. 310 [149] F. Tondeur. Nucl. Phys. A315 (1979) 353-369. [150] J. Dobaczewski and W. Nazarewicz. Prog. Theor. Phys. Suppl. 146 (2002) 70-83. [151] S. Wolfram. The Mathematica. Book. Wolfram Media/Cambridge University Press, Cambridge, 1996. [152] M. Abramowitz and I. A. Stegun. Handbook of Mathematical thctions with Formulas, Graphs, and Mathematical Tables. Applied Mathematics Series. vol. 55, National Bureau of Standards, Washington, DC. ( 1964). [153] B. Gebremariam, T. Duguet and S. K. Bogner. ArXiv:1003.5210v1 [nucl—th]. [154] N. Kaiser. Phys. Rev. C68, 014323 (2003). [155] J. C. Slater. Phys. Rev. 81 (1951) 385. [156] B. Gebremariam, S. K. Bogner, T. Duguet. ArXiv:0912.3086[physics.comp—ph]. [157] B. Gebremariam, S. K. Bogner, T. Duguet. ArXiv:0910.4993v2 [physics.comp- ph]. [158] M. Bender, K. Bennaceur, T. Duguet, P.-H. Heenen, T. Lesinski, J. Meyer. Phys. Rev. C 80:064302, 2009. [159] T. Lesinski, M. Bender, K. Bennaceur, T. Duguet, J. Meyer. Phys. Rev. C 76, 014312,2007. [160] B. Gebremariam, S. K. Bogner, T. Duguet. In preparation. [161] B. Gebremariam, S. K. Bogner, T. Duguet. In preparation. [162] B. Gebremariam, S. K. Bogner, T. Duguet. In preparation. [163] R. J. Furnstahl, James C. Hackworth. Phys. Rev. C 56 2875, 1997. [164] M. V. Stoitsov et. al. In preparation. [165] D. Rozpedzik, J. Golak, R. Skibinski, H. Witala, W. Gloeckle, E. Epelbaum, A. Nogga, H. Kamada. Acta Phys. Polon. B37 (2006) 2889-2904. [166] V. Rotival et. al. In preparation. [167] S. A. Moskzowksi. Phys. Rev. C 2, 402 (1970). [168] J. Dobaczewski. Interactions, Symmetry breaking and Effective fields from quarks to nuclie. 2003. [169] H.W.Lee, Phys. Rep. 259, 147 (1995). [170] J .W. Negele and D. Vautherin. Phys. Rev. C 5, 1472, 1972. 311 [171] X. Campi and A. Bouyssy. Phys. Lett. 73B, 263, 1973. [172] J. Meyer and J. Bartel and M. Barack and P. Quentin and S. Aicher. Phys. Lett. B. 172, 122-128, 1986. [173] Soubbotin, V. B. and Vinas, X. ArXiv:9902039[nucl—th], 1999. [174] F. G. Perey and D. S. Saxon. Phys. Lett., 10, ”107”, 1964. [175] E. Wigner. Phys. Rev. 40, 749 (1932). [176] J. Martorell and E. Moya De Guerra. Ann. Phys., 158, 1-30, 1984. [177] M. Durand, V.S.Ramamurthy and P. Schuck. Phys. Lett, 113B, 116—118, 1982. [178] A. Bulgac, Jeffrey M. Thompson. Phys. Lett. B, 383, 127-132, 1996. [179] M. Casas, J. Martorell and E. Moya De Guerra. Phys. Lett. 167B, 263-267, 1986. [180] M. Prakash a, S. Shlomo and V.M.Kolomietz. Nucl. Phys, A370, 30-46, 1981. [181] T. Lesinski. Unpublished. [182] R. M. Koehl, Gregory K. Odom and Gustavo E. Scuseria. Mole. Phys, 87, 835, 1995. [183] J.W. Negele and D. Vautherin. Phys. Rev. C 11, 1031, 1975. [184] D. Sprung, M. Vallieres, X. Campi, and C.-M. K0. Nucl. Phys. A253, 1, 1975. [185] R. K. Bhaduri and D. W. L. Sprung, Nucl. Phys. A 297, 365 (1978). [186] K. Hagino, H. Sagawa, P. Schuck. ArXiv:0912.4792v1 [nucl-th]. [187] N. Pillet, N. Sandulescu, P. Schuck. ArXiv:0701086v1[nucl-th]. [188] A. B. Pippard. Proc. Roy. Soc. A 216, 547 (1953). [189] W. Satula. Phys. Scripta T125 (2006) 82-86. [190] SS. Pankratov, E.E. Saperstein, M.V. Zverev, M. Baldo, U. Lombardo. Phys. Rev. C 79: 024309, 2009. [191] Ricardo A. Broglia and Aage Winther. Phys. Lett. B. 124, 11-13, 1983. [192] G.P.Lepage. ArXiv:9706029v1[nucl-th], 1997. [193] D. M. Brink and E. Boeker. Nucl. Phys. A 91 (1967). [194] . C. G. Broyden. AMS 19 (92): 577593, 1965. 312 [195] G. N. Watson. A Treatise 011 the Theory of Bessel Functions, 211d ed. 1966. [196] M. Bender and P.-H. Heenen and P.-G. Reinhard. Rev. Mod. Phys. 75, 121, 2003. [197] Y. Engel, D. Brink, K. Goeke, S. Krieger, and D. Vautherin. Nucl. Phys. A249, 215 (1975). [198] K. O. Friedrich. Mathematical Aspects of the Quantum Theory of Fields, 1953. [199] Donald H. Kobe. Phys. Rev. A, 19, 1878-1885, 1979. [200] J. Dobaczewski and J. Dudek. Phys. Rev. C 52, 1827, 1995. [201] D. Vautherin and D. M. Brink. Phys. Rev. C5, 626, 1972. [202] R. M. Koehl, Gregory K. Odom and Gustavo E. Scuseria. Mole. Phys, 87, 835, 1995. [203] K. Husimi, Prog. Phys. Math. Soc. Japan 22, 264 (1940). [204] M.H.S. Amin, E.V. Bezuglyi, A.S. Kijko, A.N. Omelyanchouk. ArXiv:0404401[cond-mat]. [205] J. Dobaczewski and H. F locard and J. 'IYeiner. Nucl. Phys. A422, 103, 1984. [206] S. Levit, J. W. Negele, and Z. Paltiel. Phys. Rev. C 22, 19791995, 1980. [207] E. Perlinska and S. G. Rohozinski and J. Dobaczewski and W. Nazarewicz. Phys. Rev. C 69, 014316, 2004. [208] Brandon P. van Zyl. Phys. Rev. A. 68. 033601, 2003. 313