Breit-Wigner to Gaussian transition in strength functions

Similar documents
Misleading signatures of quantum chaos

Spectral Fluctuations in A=32 Nuclei Using the Framework of the Nuclear Shell Model

Eigenvalue Density for Random Covariance Matrices from a 2 2 Partitioned GOE

Isospin symmetry breaking and the nuclear shell model

arxiv: v2 [cond-mat.stat-mech] 30 Mar 2012

arxiv:nucl-th/ v1 11 Mar 2005

arxiv: v2 [nucl-th] 24 Sep 2009

Physics Letters B 702 (2011) Contents lists available at ScienceDirect. Physics Letters B.

Recent results in quantum chaos and its applications to nuclei and particles

Is Quantum Mechanics Chaotic? Steven Anlage

LEVEL REPULSION IN INTEGRABLE SYSTEMS

arxiv: v1 [nucl-th] 5 Nov 2018

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 13 Mar 2003

Quantum Chaos as a Practical Tool in Many-Body Physics ESQGP Shuryak fest

Size Effect of Diagonal Random Matrices

Quantum Chaos as a Practical Tool in Many-Body Physics

arxiv: v1 [nucl-th] 4 Feb 2011

Universal theory of complex SYK models and extremal charged black holes

Statistical Behaviors of Quantum Spectra in Superheavy Nuclei

RANDOM MATRICES, POINT-GROUP SYMMETRIES, AND MANY-BODY SYSTEMS

arxiv: v1 [cond-mat.stat-mech] 21 Nov 2007

arxiv:cond-mat/ v1 29 Dec 1996

Experimental and theoretical aspects of quantum chaos

arxiv: v1 [nucl-th] 7 Nov 2018

arxiv: v1 [nucl-th] 5 Apr 2017

Statistical Approach to Nuclear Level Density

Chapter 29. Quantum Chaos

arxiv: v1 [math-ph] 16 Mar 2015

Spectral Statistics of Instantaneous Normal Modes in Liquids and Random Matrices

Effect of Unfolding on the Spectral Statistics of Adjacency Matrices of Complex Networks

Energy Level Statistics of the U(5) and O(6) Symmetries in the Interacting Boson Model

Spectra and Chaos in the SYK Model

Random matrix analysis of complex networks

UNIVERSITY OF OSLO CHAOS IN NUCLEI AND THE K QUANTUM NUMBER. J. Rekstad, T.S. Tveter and M. Guttormsen

Emergence of chaotic scattering in ultracold lanthanides.

Dynamical thermalization in isolated quantum dots and black holes

arxiv:nucl-th/ v1 5 Aug 1997

Recent Progress on the 0 + Dominance

Order-to-Chaos Transition in Rotational Nuclei

arxiv: v1 [cond-mat.stat-mech] 25 Jul 2008

Spectral properties of random geometric graphs

Electronic and Level Statistics Properties of Si/SiO 2 Quantum Dots

Random Matrix Theory and Cross-correlations in Global. Financial Indices and Local Stock Market Indices

Spin Cut-off Parameter of Nuclear Level Density and Effective Moment of Inertia

New T=1 effective interactions for the f 5/2 p 3/2 p 1/2 g 9/2 model space: Implications for valence-mirror symmetry and seniority isomers

arxiv: v1 [nucl-th] 4 Feb 2008

Revista Mexicana de Física ISSN: X Sociedad Mexicana de Física A.C. México

arxiv: v2 [physics.data-an] 18 Nov 2007

arxiv: v2 [cond-mat.stat-mech] 15 Apr 2016

arxiv: v1 [nucl-th] 8 Sep 2011

Universal conductance fluctuation of mesoscopic systems in the metal-insulator crossover regime

Angular-Momentum Projected Potential Energy Surfaces Based on a Combined Method. Jianzhong Gu. (China Institute of Atomic Energy, Beijing, China)

Shell model study of Neutron rich Oxygen Isotopes

Level density; occupancies; spectrum fluctuations; time reversal; nuclear inter-

Exact solution of the nuclear pairing problem

Robustness of edge states in graphene quantum dots

Quantum Confinement in Graphene

Eigenvectors under a generic perturbation: non-perturbative results from the random matrix approach

Random Wave Model in theory and experiment

Spectral and Parametric Averaging for Integrable Systems

Coupling of giant resonances to soft E1 and E2 modes in 8 B

Central density. Consider nuclear charge density. Frois & Papanicolas, Ann. Rev. Nucl. Part. Sci. 37, 133 (1987) QMPT 540

Collapse and the Tritium Endpoint Pileup

arxiv:cond-mat/ v1 [cond-mat.stat-mech] 20 Apr 2004

Validity of Born Approximation for Nuclear Scattering in Path Integral Representation

Quantum instability for a wave packet kicked periodically by a quadratic potential

Moment (and Fermi gas) methods for modeling nuclear state densities

Transverse wobbling. F. Dönau 1 and S. Frauendorf 2 1 XXX 2 Department of Physics, University of Notre Dame, South Bend, Indiana 46556

arxiv: v2 [nucl-th] 8 May 2014

Anderson Localization Looking Forward

Multiple Multi-Orbit Fermionic and Bosonic Pairing and Rotational SU(3) Algebras

Proxy-SU(3) Symmetry in Heavy Nuclei: Prolate Dominance and Prolate-Oblate Shape Transition

Continuum Shell Model

A Multi-Level Lorentzian Analysis of the Basic Structures of the Daily DJIA

The shell model Monte Carlo approach to level densities: recent developments and perspectives

Random unitary matrices

QUANTUM CHAOS IN NUCLEAR PHYSICS

Statistical distributions of quasi-optimal paths in the traveling salesman problem: the role of the initial distribution of cities

Chaotic Scattering of Microwaves in Billiards: Induced Time-Reversal Symmetry Breaking and Fluctuations in GOE and GUE Systems 2008

Giant Enhancement of Quantum Decoherence by Frustrated Environments

Quantum Theory of Many-Particle Systems, Phys. 540

arxiv:nucl-th/ v1 18 Jan 2006

From the SYK model, to a theory of the strange metal, and of quantum gravity in two spacetime dimensions

arxiv: v1 [nucl-th] 17 Nov 2015

arxiv:physics/ v1 [physics.med-ph] 8 Jan 2003

Dynamical localization and partial-barrier localization in the Paul trap

Experimental evidence of wave chaos signature in a microwave cavity with several singular perturbations

Nuclear Level Density with Non-zero Angular Momentum

Dynamical Localization and Delocalization in a Quasiperiodic Driven System

Introduction to Theory of Mesoscopic Systems

2008 Brooks/Cole 2. Frequency (Hz)

TitleQuantum Chaos in Generic Systems.

arxiv: v2 [cond-mat.dis-nn] 9 Feb 2011

Phases in an Algebraic Shell Model of Atomic Nuclei

arxiv: v4 [nlin.cd] 2 May 2017

Eigenvalue statistics and lattice points

Universal Impedance Fluctuations in Wave Chaotic Systems. Steven M. Anlage 1,3

Quantum chaos. University of Ljubljana Faculty of Mathematics and Physics. Seminar - 1st year, 2nd cycle degree. Author: Ana Flack

arxiv: v1 [nucl-th] 23 Jul 2018

arxiv: v1 [math-ph] 25 Aug 2016

Transcription:

Breit-Wigner to Gaussian transition in strength functions V.K.B. Kota a and R. Sahu a,b a Physical Research Laboratory, Ahmedabad 380 009, India b Physics Department, Berhampur University, Berhampur 760 007, India ABSTRACT: Employing hamiltonians defined by two-body embedded Gaussian orthogonal ensemble of random matrices(egoe(2)) plus a mean-field producing onebody part, strength functions (for states defined by the one-body part) are constructed for various values of the strength of the chaos generating two-body part. Numerical calculations for six and seven fermion systems clearly demonstrate Breit-Wigner to Gaussian transition, in the chaotic domain, in strength functions as found earlier in nuclear shell model and Lipkin-Meshkov-Glick model calculations. PACS number(s): 21.10.Pc, 24.60.Lz, 21.60.Cs 1

Strength functions (also called local density of states (LDOS) in literature) of simple modes are basic ingredients of many particle systems such as atomic nuclei [1]. In the last few years, with developments in quantum chaos [2], nature of strength functions for finite isolated interacting many particle systems are being investigated using a variety of models: (i) Zelevinsky et al using nuclear shell model [3]; (ii) Wang et al using the three-orbital Lipkin-Meshkov-Glick model [4]; (iii) Borgonovi et al using a symmetrized coupled two-rotor model [5]; (iv) Benet et al using a chaotic model of two coupled quartic oscillators [6]. Results from the shell model and Lipkin-Meshkov- Glick model studies clearly showed that strength functions exhibit, for interacting particle systems, Breit-Wigner to Gaussian transition. Moreover it is seen that the transition takes place much after the onset of chaos in energy levels (i.e. much after level fluctuations start following GOE). In order to establish that this transition is generic to interacting particle systems, we carried out one plus two-body embedded Gaussian orthogonal ensemble (EGOE(1+2)) calculations and the results are reported in this short communication (see also [7, 8]). Given a compound state φ k, the probability of its decay into stationary states ψ E (generated by H) isgivenby φ k ψ E 2. Then the strength function F k (E) is, φ k = E C E k ψ E F k (E) = E C E k 2 δ(e E )= δ(h E) k (1) The standard form [1], normally employed in many applications in nuclear physics, for strength functions is the Breit-Wigner (BW) form characterized by a spreading width Γ k, F k:bw (E) = 1 2π Γ k (E E k ) 2 +Γ 2 k /4 (2) where E k = k H k = F k(e)e de. Starting with H = h(1) + λv (2) (a one plus two-body interaction as in the shell model) with h(1) defining the φ k states, it is easily seen that the assumptions that give F k:bw (E) will break down when the mixing is strong. Thus it is expected that the form of F k (E) will be different from 2

BW for large λ. In order to find the form of strength functions in large λ limit, it is useful to think of φ k as a compound state generated by the action of a transition operator O on a state ψ Ek (for example ground state), O ψ Ek φ Ek = ; φ [ ψ Ek O O ψ Ek ] 1/2 Ek φ Ek =1 φ Ek = O δ(h E)Oδ(H E CE E k ) k ψ E F k;o (E) = E O Oδ(H E k ) (3) The numerator in the expression for F k (E) in (3) is nothing but the bivariate strength density generated by the operator O [9] and then it is easy to recognize that F k (E) is the conditional density of this bivariate strength density. It is known [9] that EGOE(k), the embedded GOE of k-body interactions (see Ref. [10] for the definition of EGOE(k)), in general gives bivariate Gaussian strength densities. Therefore, for EGOE(2), which models a generic two-body hamiltonian, the strength functions take Gaussian form (conditional density of a bivariate Gaussian is a Gaussian). However, for generic interacting particle systems, it is more appropriate to consider one plus two-body EGOE(1+2) hamiltonians {H} = h(1) + λ{v (2)} where h(1) is the onebody mean-field producing part and {V (2)} is the chaos generating two-body part; note that {V (2)} is GOE in two-particle space with unit matrix elements variance and λ is the interaction strength. Starting with EGOE(1+2) it is to be expected that for sufficiently large values of λ, EGOE(2) description should be valid and therefore (3) gives the shape of the strength function to be Gaussian for large λ; this result is indeed seen in numerical calculations. Now the important questions are: (i) how F k (E) changes as λ is varied and (ii) from which value of λ strength functions take Gaussian form. For further understanding of the nature of F k (E), we performed EGOE(1+2) calculations in 924 dimensional space generated by six fermions (m = 6)intwelve single particle states (N = 12) and similarly in the 3432 dimensional N = 14and m = 7 space. The single particle energies employed are ɛ i = i+(1/i), i =1, 2,...,N as 3

in [11]. For various values of λ in h(1)+λ{v (2)}, strength functions are constructed by choosing the φ k states to be the mean-field h(1) states defined by the distribution of m particles in the N single particle states; their energies E k are E k = k h(1) + λv(2) k. In the calculations E and E k are zero centered for each member and scaled by the spectrum (E s) width σ; Êk =(E k ɛ)/σ and Ê =(E ɛ)/σ. For each member CE k 2 are summed over the basis states k in the energy window Êk ± and then ensemble averaged FÊk (Ê) vsê is constructed as a histogram; the value of is chosen to be 25 for λ and beyond this =. Results for Êk = 0 are shown for λ = 5,, 5,, and in Fig. 1; in the plots F Ek (Ê) dê =1. Fromthe figures it is clearly seen that there is BW to Gaussian transition in F k (E). A measure for this transition is, R(λ) = i [ F (λ) k (E i ) F k:bw (E i ) ] 2 [ F(k:BW) (E i ) F k:ed (E i ) ] 2 (4) i In (4) E i are defined by the center of each bin in the histogram representing F (λ) k (E) and ED corresponds to Gaussian with Edgeworth corrections; the ED incorporates [12] skewness (γ 1 ) and excess (γ 2 ) corrections. Strength function F (λ) k (E) is BW for R = 0 and Gaussian (or ED) for R = 1. The value of the interpolating parameter λ = λ Fk for onset of the transition from BW to Gaussian is taken to be R(λ Fk )=0.7. For λ = 5,, 5,,, the values of R are 3, 6, 1, 0.7, 0.9, 3 for the six particle example. For the seven particle example the corresponding numbers are 5, 2, 5, 0.75, 0.93, 3 respectively. Thus, with the measure R in (4), λ Fk for both the 6 and 7 particle cases. In Fig. 2 shown are EGOE(1+2) results for various values of λ for the nearest neighbour spacing dustribution and 3 (L), 0 L 40 statistic. The λ Fk deduced from the results in Fig. 1 should be compared with λ = λ c 8 and 6, derived via Fig. 2, for six and seven particle examples repectively for onset of chaos in level 4

fluctuations. The λ c numbers are consistant with (Nm 2 ) 1 variation as expected from the theory given in [13]. It should be noted that in the (2s, 1d) m=12,jπ T =0 + 0 shell model example considered in [3], λ c andλ Fk. Thus the BW form for F k (E), which begins some what before λ approches λ c (for λ<<λ c the strength functions are delta functions with perturbative corrections), extends much into the chaotic domain (defined by λ>λ c ) and the transition to Gaussian shape takes place in the second layer defined by λ Fk (λ Fk >> λ c ). As the λ Fk in our two examples did not show much variation with (N,m), it appears that λ Fk will have weak dependence on m unlike λ c. Perhaps for sufficiently large m, the transition to Gaussian form takes place in the thermalization regime discussed by Flambaum and Izrailev [7] (see also [14]). It should be remarked that EGOE(2) gives Gaussian form for state densities for sufficiently large m and its extension to partial state densities (which are sums of strength functions) is indeed the basis of statistical nuclear spectrocopy [9, 15]. In conclusion, the results of the present EGOE(1+2) study (Figs. 1,2) and the earlier shell model [3] and Lipkin-Meshkov-Glick model [4] analysis establish firmly the Breit-Wigner to Gaussian transition for strength functions, in the chaotic domain, in interacting particle systems such as atomic nuclei. Acknowledgements This work has been partially supported by DST(India). References [1] A. Bohr and B. Mottelson, Nuclear Structure, Volume 1 (Benjamin, New York, 1969). [2] T. Guhr, A. Müller-Groeling, and H.A. Weidenmüller, Phys. Rep. 299, 189 (1998). [3] V. Zelevinsky, B.A. Brown, N. Frazier, and M. Horoi, Phys. Rep. 276, 85 (1996); N. Frazier, B.A. Brown, and V. Zelevinsky, Phys. Rev. C 54, 1665 (1996). 5

[4] W. Wang, F.M. Izrailev, and G. Casati, Phys. Rev. E 57, 323 (1998). [5] F. Borgonovi, I. Guarneri, and F.M. Izrailev, Phys. Rev. E 57, 5291 (1998). [6] L. Benet, F.M. Izrailev, T.H. Seligman, and A. Suarez-Moreno, chao-dyn/9912035 (2000). [7] V.V. Flambaum and F.M. Izrailev, Phys. Rev. E 56, 5144 (1997). [8] V.K.B. Kota and R. Sahu, nlin-cd/0006003 (2000); Phys. Rev. E (in press); V.K.B. Kota, R. Sahu, K.Kar, J.M.G. Gómez, and J. Retamosa, nucl-th/0005066 (2000). [9] J.B. French, V.K.B. Kota, A. Pandey, and S. Tomsovic, Ann. Phys. (N.Y.) 181, 235 (1988). [10] T.A. Brody, J. Flores, J.B. French, P.A. Mello, A. Pandey, and S.S.M. Wong, Rev. Mod. Phys. 53, 385 (1981). [11] V.V. Flambaum, G.F. Gribakin, and F.M. Izrailev, Phys. Rev. E 53, 5729 (1996). [12] M.G. Kendall and A. Stuart, Advanced Theory of Statistics, Volume 1, 3rd edition (Hafner Publishing Company, New York, 1969). [13] Ph. Jacquod and D.L. Shepelyansky, Phys. Rev. Lett. 79, 1837 (1997). [14] M. Horoi, V. Zelevinsky, and B.A. Brown, Phys. Rev. Lett. 74, 5194 (1995). [15] J.B. French and V.K.B. Kota, Ann. Rev. Nucl. Part. Sci. 32, 35 (1982), S.S.M. Wong, Nuclear Statistical Spectroscopy (Oxford University Press, New York, 1986); V.K.B. Kota and K. Kar, Pramana - J. Phys. 32, 647 (1989). 6

FIGURE CAPTIONS Fig. 1 Strength functions for EGOE(1+2) for various values of the interaction strength λ: (i) for a system of 6 fermions in 12 single particle states with 25 members; (ii) for a system of 7 fermions in 14 single particle states (due to computational constraints, here only one member is considered just as in [11]). In the figure, the histograms are EGOE(1+2) results and continuous curves are BW fit. For the 6 fermions case, the dotted curves are Gaussian for λ 5 and Edgeworth corrected Gaussians (ED) for λ>5. Similarly for the 7 fermions case, the dotted curves are Gaussian for λ and Edgeworth corrected Gaussians (ED) for λ>. See text for further details. Fig. 2 Nearest neighbour spacing distribution P (S) vss and Dyson-Mehta 3 (L) statistic for 0 L 40 for various values of the interaction strength λ in EGOE(1+2). For P (S), histograms are EGOE(1+2) results, dashed curves are Poisson and continuous curves are Wigner distribution. For 3 (L), filled circles are EGOE(1+2) results, dashed curves are Poisson and continuous curves are for GOE. Results are shown for the six and seven particle examples considered in Fig. 1. 7

6 EGOEè1+2è : N=12,m=6 2.5 6 EGOEè1+2è : N=14,m=7 2.5 4 1.5 4 1.5 2 2 0 1.5 1.5 0 1.5 1.5 Strength Function 0.9 5 2 1 0 1 2 λ=.2 Strength Function 0.9 5 2 1 0 1 2 λ=.2 λ=.3 λ=.5 λ=.3 λ=.5 (E ε)/σ (E ε)/σ

P(S) EGOE(1+2) : N=12,m=6 λ=.01 λ=.03 λ=.08 3 (L) 8.0 6.0 4.0 λ=.01 EGOE(1+2) : N=12,m=6 4.0 3.0 λ=.03 λ=.08 5 5 S L EGOE(1+2) : N=14,m=7 8.0 EGOE(1+2) : N=14,m=7 4.0 λ=.01 λ=.03 6.0 4.0 λ=.01 3.0 λ=.03 P(S) λ=.06 3 (L) λ=.06 λ=.08 S λ=.08 L