Quantum Physics and Number Theory connected by the Riemann Zeta Function Dissertation zur Erlangung des Doktorgrades Dr. rer. nat. der Fakultät für Naturwissenschaften der Universität Ulm vorgelegt von Cornelia Feiler aus Ulm Institut für Quantenphysik Ulm, 2013 Amtierender Dekan: Prof. Dr. Joachim Ankerhold Erstgutachter: Prof. Dr. Wolfgang P. Schleich Zweitgutachter: Prof. Dr. Peter Reineker Tag der Prüfung: 18.02.2014 Contents List of figures vi List of tables vi List of symbols vii Introduction 1. The 1.1. 1.2. 1.3. 1.4. 1.5. 1.6. 1 Riemann zeta function: mathematical representations Dirichlet series . . . . . . . . . . . . . . . . . . . . . . . Alternating sum . . . . . . . . . . . . . . . . . . . . . . Analytical continuation . . . . . . . . . . . . . . . . . . Symmetric representation . . . . . . . . . . . . . . . . . Riemann-Siegel formula . . . . . . . . . . . . . . . . . . Extended Berry-Keating formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 3 3 4 5 6 6 2. The Riemann zeta function: topology 2.1. Continuous Newton method: essentials . . 2.2. Newton flow of the Riemann zeta function 2.2.1. Asymptotic for σ → +∞ . . . . . . 2.2.2. Asymptotic for σ → −∞ . . . . . . 2.2.3. Non-trivial separatrices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 9 10 10 12 14 3. Combined quantum systems 3.1. Time evolution . . . . . . . . . . . 3.2. Joint measurement . . . . . . . . . 3.3. Phase space representations . . . . 3.3.1. Wigner and Moyal function 3.3.2. Wigner and Moyal matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 15 16 17 17 18 4. Riemann states 4.1. Definition and properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2. Wigner functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3. Moyal functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 21 23 25 . . . . . . . . . . . . . . . . . . . . 5. Alternating sum and truncated Riemann states 29 5.1. Exact solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 5.1.1. Construction à la Riemann states . . . . . . . . . . . . . . . . . . . 29 i Contents 5.1.2. Entanglement . . . . . . . . 5.2. Truncated Riemann states . . . . . 5.3. Truncated alternating sum . . . . . 5.3.1. Reference state and overlap 5.3.2. Wigner representation . . . 5.3.3. Moyal representation . . . . 5.4. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Riemann-Siegel states 6.1. Definition of the states . . . . . . . . . . . . 6.1.1. Riemann-Siegel state . . . . . . . . . 6.1.2. Riemann-Siegel reference state . . . 6.2. Entanglement . . . . . . . . . . . . . . . . . 6.2.1. Riemann-Siegel state . . . . . . . . . 6.2.2. Riemann-Siegel reference state . . . 6.3. Wigner matrix of the Riemann-Siegel state 6.4. Wigner matrix of the reference state . . . . 6.5. Overlap in Wigner representation . . . . . . 6.6. Overlap in Moyal representation . . . . . . 6.7. Summary . . . . . . . . . . . . . . . . . . . 7. Berry-Keating reference states 7.1. Definition . . . . . . . . . . . . . 7.2. Wigner function of Berry state . 7.3. Wigner matrix . . . . . . . . . . 7.4. Overlap in Wigner representation 7.5. Overlap in Moyal representation 7.6. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8. Jaynes-Cummings-Paul approach to the Riemann 8.1. Jaynes-Cummings-Paul model . . . . . . . . . 8.2. Riemann Hamiltonian . . . . . . . . . . . . . 8.2.1. Exact solution . . . . . . . . . . . . . 8.2.2. Approximate solution . . . . . . . . . 8.3. Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 31 35 36 36 37 39 . . . . . . . . . . . 41 41 42 42 43 43 44 44 46 48 50 54 . . . . . . 57 57 58 60 60 62 68 . . . . . 69 69 71 73 73 77 Summary 79 A. Special functions A.1. The gamma function Γ . . . . . . . . . . . A.1.1. Definition and functional equation A.1.2. Asymptotics . . . . . . . . . . . . A.1.3. Formulae . . . . . . . . . . . . . . 83 83 83 83 84 ii . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Contents A.2. The function χ(s) . . . . . A.2.1. Definitions . . . . A.2.2. On the critical line A.2.3. Asymptotics in the A.3. The function ϑ(s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . critical strip . . . . . . . . . . . . . 84 85 85 86 86 B. Approximations in the critical strip B.1. Riemann-Siegel formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . B.2. Berry-Keating formula on the critical line . . . . . . . . . . . . . . . . . . B.3. Berry-Keating formula in the critical strip . . . . . . . . . . . . . . . . . . 87 87 88 92 C. Continuous Newton method C.1. Newton flow . . . . . . . . . . C.1.1. Sink and source . . . . C.1.2. Origin of separatrices C.2. Further examples . . . . . . . C.2.1. The function χ . . . . C.2.2. The function Fe . . . C.3. Technical details . . . . . . . C.4. Newton flow of ζ . . . . . . . C.5. Zeros of the derivative ζ 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 95 95 97 99 101 101 102 102 102 D. Phase space representations D.1. Properties of the Wigner function . . D.2. Moyal functions . . . . . . . . . . . . D.3. Wigner matrix . . . . . . . . . . . . D.4. Fock states . . . . . . . . . . . . . . D.4.1. Wigner and Moyal functions D.4.2. Sum over Wigner functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107 107 108 109 110 111 113 E. Entanglement E.1. Quick check . . . . . . . . . . . E.2. Schmidt decomposition . . . . . E.3. Wigner representation . . . . . E.4. Riemann-Siegel reference state E.5. Berry-Keating reference state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115 115 115 117 118 119 F. Normalization and proportionality factors F.1. Truncated Riemann state . . . . . . . F.2. The function E(n, τ ) . . . . . . . . . . F.3. Berry state . . . . . . . . . . . . . . . F.4. Truncated alternating sum . . . . . . . F.5. Riemann-Siegel reference state . . . . F.6. Berry-Keating reference state . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 121 122 123 124 125 125 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . iii Contents G. Effective Hamiltonian 129 G.1. Time-dependent Hamiltonian . . . . . . . . . . . . . . . . . . . . . . . . . 129 G.2. Interaction picture and effective Hamiltonian . . . . . . . . . . . . . . . . 130 H. Matrix elements 133 † H.1. Calculation with â and â . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 H.2. General expression . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134 H.3. Explicit expression up to fourth order . . . . . . . . . . . . . . . . . . . . 136 Bibliography 137 Zusammenfassung 143 Publikationen 147 Poster und Vorträge 149 Curiculum vitae 151 Danksagung 153 iv List of figures 1.1. Absolute value of Riemann zeta function ζ and sketch of its structure . . 1.2. Symmetric representation Z and its approximations on the critical line . . 1.3. Z and its approximations for different σ and large τ . . . . . . . . . . . . 4 7 8 2.1. Newton flow of the Riemann zeta function . . . . . . . . . . . . . . . . . . 11 2.2. Separatrices of ζ and χ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 4.1. Wigner function of the initial Riemann state W|σ,0i for different σ . . . . . 23 4.2. Time evolution of W|2,τ i and overlap |h2, 0|2, τ i|2 . . . . . . . . . . . . . . 24 4.3. Time evolution of the Moyal function W|2,τ ih2,0| . . . . . . . . . . . . . . . 26 5.1. Wigner function of the initial truncated Riemann state W|1/2,0iν for different ν . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2. Time evolution of W|2/3,τ i100 and overlap |100 hφ|2/3, τ i100 |2 . . . . . . . 5.3. Truncated alternating sum ζN for different σ and N for small τ . . . . . 5.4. Truncated alternating sum ζN for different σ and N for large τ . . . . . 5.5. Time evolution of the Moyal function W|2/3,τ i100 hφ| . . . . . . . . . . . . 6.1. Time evolution of Ŵ|ΨRSi of the Riemann-Siegel state 6.2. 6.3. 6.4. 6.5. 6.6. 6.7. . . . . . 31 33 34 35 38 . . . . . . . . . . . 45 Wigner matrix elements of Ŵ|ΦRSi of the Riemann-Siegel reference state Wigner representation KRS of overlap |hΦRS|ΨRSi|2 . . . . . . . . . . . . . Time evolution of the Moyal functions W|1/3,τ i10 h1/3,0| and W|1/3,−τ i10 h1,0| Moyal representation K RS of overlap hΦRS|ΨRSi . . . . . . . . . . . . . . . Riemann-Siegel formula ζRS for σ = 1/3 and σ = 1/2 . . . . . . . . . . . Argand diagram of ζ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.1. Difference W|φB(1/2,τ )in0 − W|1/2,0in0 between Wigner functions of Berry state |φB(σ, τ )in0 and initial truncated Riemann state |σ, 0in0 . . . . . . . 7.2. Wigner matrix elements of Ŵ|ΦBKi of the Berry-Keating reference state and difference to fig. 6.2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.3. Wigner representation KBK of overlap |hΦBK|ΨRSi|2 and difference to fig. 6.3 7.4. Moyal functions W|1/3,τ i10 hφB(1/3,τ )| and W|1/3,−τ i10 hφ (1,τ )| and difference B to fig. 6.4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.5. Moyal representation K BK of overlap hΦBK|ΨRSi and difference to fig. 6.5 . . 47 49 52 53 55 55 59 61 63 65 66 8.1. Two-level atom in cavity . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 8.2. ζ displayed by approximated Hamiltonian . . . . . . . . . . . . . . . . . . 76 v C.1. C.2. C.3. C.4. Newton Newton Newton Newton flow flow flow flow of of of of simple zero, simple pole and Möbius transformation . polynomials with two and three zeros . . . . . . . . . χ and Fe . . . . . . . . . . . . . . . . . . . . . . . . . ζ for imaginary parts up to τ = 300 . . . . . . . . . . . . . . . . . . . . . . 96 98 100 103 D.1. Wigner functions W|ni of Fock states . . . . . . . . . . . . . . . . . . . . . 111 D.2. Real part of Moyal function W|nihm| of Fock states . . . . . . . . . . . . . 112 D.3. Sum SN of Wigner functions W|ni . . . . . . . . . . . . . . . . . . . . . . 113 F.1. F.2. F.3. F.4. F.5. Square of normalization Nν of truncated Riemann states . . . Behavior of E(n, τ ) in dependence on n for fixed τ . . . . . . en of Berry states . . . . . . . . . . Square of normalization N B (N ) Proportionality factor Ma of the truncated alternating sum Factors in Ŵ|ΦRSi of the Riemann-Siegel reference state . . . . . . . . . . ζN . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 122 123 124 126 F.6. Factors in Ŵ|ΦBKi of the Berry-Keating reference state . . . . . . . . . . . 127 List of tables 2 . . . . . . . . . . . . . . 50 6.1. Analysis of fig. 6.3 – Wigner representation of ζRS 6.2. Analysis of fig. 6.5 – Moyal representation of ζRS . . . . . . . . . . . . . . 54 2 . . . . . . . . . . . . . . 62 7.1. Analysis of fig. 7.3– Wigner representation of ζBK 7.2. Analysis of fig. 7.5 – Moyal representation of ζBK . . . . . . . . . . . . . . 68 C.1. Trivial zeros of the derivatives ζ 0 and χ0 . . . . . . . . . . . . . . . . . . . 104 C.2. Non-trivial zeros of ζ 0 with τ < 100 and arg ζ at these points . . . . . . . 104 C.3. Non-trivial zeros of ζ 0 with imaginary part 100 < τ < 300 . . . . . . . . . 105 H.1. Matrix elements hn ± j| (â + ↠)µ |ni for n = 0, . . . , 4. . . . . . . . . . . . . 134 vi List of symbols |ψi |Ψi |σ, τ i |σ, τ iν |φBi |ΨRSi |φiN |ΦRSi |ΦBKi state of one system state of two systems Riemann state truncated Riemann state Berry state Riemann-Siegel state reference state for truncated alternating sum Riemann-Siegel reference state Berry-Keating reference state χ C C function, factor in functional equation of ζ overlap of time-evolved state and reference state set of complex numbers Erfc complementary error function γ γB Γ factor in Riemann-Siegel reference state factor in Berry-Keating reference state Gamma function ~ Ĥ0 ĤaJC Ĥa Ĥfield I Ĥint Ĥint ĤJC ĤR Planck constant free Hamiltonian anti-Jaynes-Cummings Hamiltonian free Hamiltonian of two level atom free Hamiltonian of cavity field interaction Hamiltonian in interaction picture interaction Hamiltonian in Schrödinger picture Jaynes-Cummings Hamiltonian Riemann Hamiltonian K KRS KBK K K RS K BK kernel kernel kernel kernel kernel kernel Ma (ν,N ) Ma MRS proportionality factor for alternating sum proportionality factor for truncated alternating sum proportionality factor for Riemann-Siegel formula of of of of of of Wigner representation Wigner representation for Riemann-Siegel formula Wigner representation for Berry-Keating formula Moyal representation Moyal representation for Riemann-Siegel formula Moyal representation for Berry-Keating formula vii List of symbols MBK proportionality factor for Berry-Keating formula n̂ n0 nB N Nν Nn0 en N B ~ N N N\{0} photon number operator summation limit in Riemann-Siegel formula summation limit in Berry-Keating formula norm of Riemann state |σ, τ i norm of truncated Riemann state |σ, τ iν norm of Riemann-Siege reference state norm of Berry-Keating reference state vector of the Newton flow set of natural numbers including 0 set of natural numbers without 0 ω Rabi frequency σ̂z σ s Pauli spin matrix real part of complex variable s complex variable τ ϑ imaginary part of complex variable s function, derived from χ W|ψi W|ψihφ| Ŵ|Ψi Wigner function of state |ψi Moyal function of |ψihφ| Wigner matrix of state |Ψi ζ ζBK ζN ζRS Z ZBK ZRS Z Riemann zeta function Berry-Keating formula truncated alternating sum of ζ Riemann-Siegel formula symmetric representation of ζ symmetric representation of ζBK symmetric representation of ζRS set of integers viii Introduction Number theory is not pure Mathematics. It is the Physics of the world of Numbers. Alf van der Poorten1 Number theory concentrates on the fundamentals of mathematics: integers or, more precisely, primes. Questions concerning the properties of primes influence all branches of mathematics and therefore also physics. Even our lives are affected by prime numbers, for example when we think of the security of codes, in bank transfer or communication, which are based on prime factorization. On the other hand, the distribution of primes [2] is intimately connected to the distribution of the non-trivial zeros of the Riemann zeta function ζ [3–5]. Since the location of these zeros, which is object to the Riemann hypothesis, is by now unconfirmed, many efforts are directed to the investigation of ζ, not only with mathematical but also with physical methods [6]. For example, the non-trivial zeros can be considered as eigenvalues of an Hermitian operator [7, 8], while the phase of ζ on the critical line is closely related to the scattering on an inverted harmonic oscillator [9]. Moreover, the Riemann zeta function emerges as a partition function of a quantum dynamical system [10], from wave packed dynamics in a potential with logarithmic energy spectrum2 [13], or a quantum Mellin transform [14]. In this thesis, we consider several representations of the Riemann zeta function to gain insight into its characteristics. Therefore, we shortly recall the properties and different representations derived from the definition in chapter 1, and introduce the continuous Newton method to visualize the topology of ζ by lines of constant phase in chapter 2. The corresponding Newton flow particularly illuminates the location of special points: the single pole as well as the zeros of ζ and its derivative ζ 0 . Chapter 3 proposes a physical approach to the Riemann zeta function (see also [15]). It is motivated by the similarity of the phases in the Dirichlet series of ζ to the time evolution of an entangled state in cavity quantum electrodynamics, described by the Jaynes-Cummings-Paul model [16,17]. Yet, in our system, we have to choose the dependence of the Hamiltonian on the number operator n̂ logarithmically instead of linear. The resulting states reproduce the different representations of the Riemann zeta function when we take the overlap of the time-evolved state with an appropriately chosen reference state, thus making the values of ζ accessible to measurement. Moreover, we remind of the Wigner and Moyal representations [18–21] which are used to describe the behavior of the states and the overlap in phase space. 1 2 quoted by M. Planat in [1] Recently was shown that two [11] or more [12] bosons in this potential allow us to factor numbers. 1 Introduction The following chapters are dedicated to the description of the different zeta states and their phase space representations, continuing the work of the diploma thesis [22]. In chapter 4, we define the Riemann state |σ, τ i of the cavity whose overlap with its initial state produces the Dirichlet representation of ζ(σ + iτ ). In this case, entanglement is not necessary. However, the states are restricted to the region of complex space where the real part σ of the argument s is larger than one, in agreement with the limitation of the Dirichlet sum. Mathematically, we can overcome the restriction σ > 1 with the alternating sum representation of ζ which is convergent for all positive values of σ. But the states producing the exact sum underlie the same limitation as the Riemann states, as we show in the first section of chapter 5. Nevertheless, we can use the alternating representation when we truncate the summation limit. In the physical picture, this yields the truncated Riemann states |σ, τ iν which are normalizable as long as the truncation parameter ν is finite. Even so, the normalization constant makes the probability amplitudes in |σ, τ iν small for large ν. Consequently, an improvement of the approximation by increasing the summation limit results in very small values for the overlap of the states. The approximations by the Riemann-Siegel and Berry-Keating formula finally allow the definition of states which can reach into the critical strip, as shown in the chapters 6 and 7. There, we take advantage of the similar structure of the formulae and use in both cases the Riemann-Siegel state |ΨRS(t)i as a time-evolved state. It consists of two counter-rotating truncated Riemann states coupled to the excited and ground state of a two-level atom. This structure is reminiscent of the Schrödinger cat state [23,24] realized experimentally in high-Q cavities [25,26] or with ions stored in a trap [27]. In contrast to the Schrödinger cat state, the phases of the Riemann-Siegel state depend logarithmically on the photon number n, due to the Riemann Hamiltonian ĤR . The differences of the formulae are reflected in the properties of the corresponding reference states. Although these differences are small, they result in the fact that the Berry-Keating reference state |ΦBKi is always entangled whereas the Riemann-Siegel reference state |ΦRSi becomes separable on the critical line σ = 1/2. The last chapter gives a short review of the Jaynes-Cummings-Paul model and suggests how we can modify this model to produce the Riemann zeta function with an approximated Hamiltonian. Finally, the appendices contain supplements to the chapters: Appendices A and B summarize the properties of the functions involved in the different representations of ζ, which are presented in chapter 1. The characteristics of the Newton flow are derived and illustrated by examples in appendix C. In appendix D, we recall the different phase space representations obtained from the familiar Wigner function and describe the phase space functions of the Fock states in more detail, since all zeta states are defined in this basis. Appendix E reminds of different tests for entanglement and contains proofs concerning the entanglement of the Riemann-Siegel and Berry-Keating reference state. The following appendix lists the properties of the normalization and proportionality factors occurring in the different states, while the last two appendices accomplish the analysis of chapter 8. 2 1. The Riemann zeta function: mathematical representations In this chapter, we give a short overview of the different representations of the Riemann zeta function ζ and their properties. Although not every representation is suited for our purpose to realize ζ as the outcome of a joint measurement of two quantum states, we nevertheless recall them for the sake of completeness. 1.1. Dirichlet series The definition of the Riemann zeta function [4] by the Dirichlet series ∞ X 1 ζ(s) ≡ ns (1.1) n=1 for a complex variable s was already known to Leonhard Euler in 1737. He proved the connection ∞ X Y 1 1 = (1.2) s n 1 − p−s n=1 p prime between ζ and the infinite product over all prime numbers p. However, both representations suffer from the restriction to the area σ ≡ Re s > 1. Needless to say, this restriction will also affect the quantum states we derive from the Dirichlet series in chapter 4. In this region, the zeta function has no zeros. This becomes clear when we recall that the product on the right hand side of eq. (1.2) consists only of non-zero factors. Moreover, for s = 1 the zeta functions has one simple pole since the Dirichlet sum merges into the harmonic series. 1.2. Alternating sum A simple way to expand the domain of definition is subtracting from the Dirichlet series twice the sum over all even terms ∞ ∞ ∞ X X X 1 1 (−1)n − 2 = − ns (2n)s ns n=1 n=1 n=1 which yields an alternating sum convergent for σ > 0. The left hand side of this equation is equal to (1 − 21−s ) ζ(s). Therefore, ζ can be represented by the alternating sum ζ(s) ≡ 1 21−s − 1 ∞ X (−1)n n=1 ns (1.3) 3 1. The Riemann zeta function: mathematical representations Figure 1.1.: The left picture shows |ζ(s)| with s = σ + iτ and the right one a sketch of its significant points. The only pole at s = 1, marked in the sketch by a dot, is cut on the left picture to show the structure of the zeta function around this point more clearly. The blue line denotes the real axis on which the trivial zeros are located at s = −2k, k ∈ N\{0}, whereas the red line indicates the critical line σ = 1/2, conjectured location of the non-trivial zeros [3]. All zeros are marked in the sketch by crosses and the critical strip 0 < σ < 1 is indicated by the shaded area. in the whole right of the complex plane s. We will see in chapter 5 that this representation is not as powerful in our physical picture as it is in mathematics, since the constructed states are only normalizable for σ > 1. For this reason, we need other representations to reach beyond the wall at σ = 1. 1.3. Analytical continuation In 1859, Bernhard Riemann published his seminal paper “Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse” [3]. Although he used the zeta function merely as a tool to determine the number of primes beneath a given value, he was the first who gave an analytical continuation of ζ valid in the whole complex plane. In this paper, he derived the integral representation Z∞ s/2 s−2 π 1 1 − s+1 2 + x 2 ζ(s) ≡ − + dx x w(x) (1.4) Γ( 2s ) s − 1 s 1 with w(x) ≡ from which the functional equation ∞ X 2x e −πn , n=1 ζ(s) = χ(s) ζ(1 − s) 4 (1.5) 1.4. Symmetric representation immediately follows. In literature [4,28,29], the function χ(s) is defined in different ways, whose equivalence we show in appendix A.2. In what follows, we use the representation [28] 1−s s− 12 Γ 2 χ(s) ≡ π , (1.6) Γ 2s where Γ(s) denotes the Gamma function. Recalling that the zeta function has no zeros in the region σ > 1, the functional equation (1.5) reveals that the only zeros for σ < 0 are given by χ(s) or precisely by the poles of Γ(s/2) in the denominator of definition (1.6). The poles are located at s = 0, −2, −4, . . . and describe the so-called trivial zeros of ζ at s = −2k, k ∈ N\{0}. At s = 0, we find a removable singularity, since the pole of Γ is compensated by the pole s−1 in eq. (1.4). Hence, all other zeros have to lie in the critical strip 0 < σ < 1, indicated in the right picture of fig. 1.1 by the shaded area. Due to the functional equation (1.5), they must be located symmetrically to the axis σ = 1/2, the so-called critical line. Moreover, they are symmetric to t = 0. This follows from the fact that if z0 ∈ C is a zero its complex conjugate z̄0 is also one. The famous Riemann hypothesis [3] states that all non-trivial zeros of ζ have real part 1/2. However, a rigorous proof of this claim still does not exist, but the hypothesis was already confirmed for the first 1.5 billion zeros [30]. 1.4. Symmetric representation Another useful expression provides the symmetric function Z(s) ≡ χ−1/2 (s) ζ(s) , (1.7) which with the help of χ(s) = χ−1 (1 − s) transforms the functional equation (1.5) into Z(s) = Z(1 − s) . (1.8) This expression reveals the convenient property that Z is real on the critical line. Moreover, the zeros of Z are given by the trivial and non-trivial zeros of ζ, but due to the symmetry, there are also zeros at s = 2n + 1 and an additional pole at s = 0. Needless to say, the functional equations (1.5) and (1.8) can only describe the zeta function in the critical strip when we use a representation of ζ valid there, for example the integral representation given by eq. (1.4). However, the complexity of this formula is quite difficult to realize with our physical approach. Therefore, we will use in this thesis the famous expressions of Riemann-Siegel [29] and Berry-Keating [31], introduced in the following sections, to find states for an approximate description of ζ in this region. 5 1. The Riemann zeta function: mathematical representations 1.5. Riemann-Siegel formula Bernhard Riemann mentioned in a letter to Karl T. W. Weierstraß in 1859 that he had found a new expansion of the zeta function, but he had not yet simplified its derivation enough to include the formula in his article on prime number theory. Fortunately, the unpublished representation was found amongst his ‘Nachlass’ and rederived by Carl L. Siegel [29]. Therefore, it was named Riemann-Siegel formula. The main part of this representation, independently rediscovered by Godfrey H. Hardy and John E. Littlewood [32] in 1920, is given by ζ(s) ∼ = ζRS(s) ≡ n0 n0 X X 1 1 + χ(s) , s 1−s n n n=1 (1.9) n=1 where χ(s) is defined by eq. (1.6) and the summation limit r τ n0 ≡ 2π (1.10) depends on the imaginary part τ of s. Here, the floor function (or Gauß bracket) [x] indicates the largest integer smaller or equal to x. The correction term to the main part ζRS is given in appendix B. The Riemann-Siegel formula, eq. (1.9) provides a reliable approximation for large τ , especially in the critical strip 0 < σ < 1. However, it suffers from the fact that, by construction, the function jumps at τ = 2π n20 . Before we introduce the Berry-Keating formula, which avoids these discontinuities, we use eq. (1.9) in eq. (1.7) to get n0 X 2 1 √ cos ϑ(s) + i s − ZRS(s) ≡ ln n (1.11) 2 n n=1 as Riemann-Siegel approximation of the symmetric representation. The function ϑ(s) is defined by e iϑ(s) ≡ χ−1/2 (s) . Equation (1.11) clearly shows that Z is real on the critical line. 1.6. Extended Berry-Keating formula Michael V. Berry and Jonathan P. Keating [31] derive a series of convergent sums as an approximation of the symmetric representation Z, eq. (1.7), on the critical line, which is, according to them, term by term more accurate than the Riemann-Siegel formula. The proof that the main term of this series already contains the Riemann-Siegel representation ZRS, eq. (1.11), and its first correction term is given in appendix B.2. The Berry-Keating formula in the critical strip (see appendix B.3) reads " # nX B −1 χ−1/2 (s) E(n, τ ) χ1/2 (s) E (n, τ ) ZBK(s) = + (1.12) (n + 1)s (n + 1)1−s n=0 6 1.6. Extended Berry-Keating formula Z(1/2 + iτ ) 4 2 20 30 40 50 60 τ -2 -4 Z(1/2 + iτ ) Z(1/2 + iτ ) 1.0 3 2 0.5 1 14 14.5 15 τ 0 56 57 58 59 60 τ -1 -0.5 -2 Figure 1.2.: The upper plot shows Z(s), eq. (1.7), and ZRS(s), eq. (1.11), indicated by the red and black dotted line, respectively, in comparison to ZBK(s), eq. (1.12), given by the blue dashed line, for σ = 1/2 and nB = 4. The Berry-Keating representation ZBK excellently fits the exact curve of Z. Only high resolution in the lower pictures reveals a tiny deviation for small values of τ . In contrast, the approximation by the Riemann-Siegel formula ZRS(s) is worse and jumps at τ = 2πn20 caused by the jumps of n0 , marked by the orange line. 7 1. The Riemann zeta function: mathematical representations Figure 1.3.: For large values of τ , one can barely distinguish between the exact curve Z(s) and ZBK(s) indicated by the red and blue dashed line, respectively. Even for σ 6= 1/2 ZBK is a better approximation than ZRS(s) (black dotted line), which even misses two zeros for τ around 220 (left picture). Here, we have used nB = 8 while n0 = 5. with τ ≡ Im s, nB ≥ n0 and the abbreviation # "r τ n+1 1 ln p τ . E(n, τ ) ≡ Erfc 2 1−i 2π (1.13) Here, Erfc denotes the complementary error function defined in eq. (B.8). In the original definition, Berry and Keating use an infinite sum, but cutting it at nB is convenient and has no influence on the results as convergence considerations prove [31]. Moreover, the truncation parameter nB does not depend on τ . Thus, the discontinuities of the Riemann-Siegel formula are eliminated. Figures 1.2 and 1.3 confirm that ZBK is a more accurate approximation of Z compared to the Riemann-Siegel formula ZRS, eq. (1.11). Even for small values of τ , the blue dashed curve of ZBK and the red one of Z fit quite perfectly. Only high resolution in the bottom pictures reveals small deviations. For large τ , one can barely distinguish between these two lines, whereas the black dotted line of ZRS even misses two zeros in the left picture of fig. 1.3. The orange line indicates the values of n0 and the jumps of ZRS at τ = 2π n20 . Before we use the different representations of ζ to construct quantum states reproducing – with an appropriately chosen reference state – the values of the Riemann zeta function by a joint measurement, we describe a method to find numerically the zeros of a complex function. This method is applied to the Riemann zeta function to give a flavor of its topology with respect to the phase of its values. 8 2. The Riemann zeta function: topology This chapter introduces the continuous Newton method, which is designed to find numerically the zeros (or poles) of a complex function. We demonstrate its benefits by the example of the Riemann zeta function. In contrast to the previous chapter, the topology of ζ is not described by contour lines, that is lines with constant absolute value, but with the help of the Newton flow consisting of lines with constant phase. Already in 1933, Eugene Jahnke and Fritz Emde [33] depicted complex functions by lines of constant phase. Later, Albert A. Utzinger, Andreas Speiser and Juan Arias-deReyna [34–36] concentrated on the description of real lines to get a better understanding of the behavior of the Riemann zeta function. Moreover, Kevin A. Broughan and A. Ross Barnett [37] considered the holomorphic flow of ζ. We emphasize that the Newton flow is quite different to the holomorphic flow, since it involves the reciprocal of the logarithmic derivative, the Newton quotient. 2.1. Continuous Newton method: essentials The continuous Newton method is a steepest descent method with respect to an appropriately chosen gradient, the so-called Sobolev gradient [38–40]. The solutions s(t) of the differential equation ds(t) F (s(t)) ṡ(t) ≡ =− 0 , dt F (s(t)) form the Newton flow of the complex function with derivative F 0 ≡ dF (s) . In appendix C, ds we elucidate the following properties: (i) The Newton flow consists of lines with constant phase. (ii) Zeros of F act like sinks for the Newton flow, whereas (iii) poles are sources and (iv) infinity can be both. (v) The flow lines can only cross each other in hyperbolic points [41], where the derivative F 0 vanishes. We call these lines separatrices. (vi) There are m separatrices crossing each other in the point s00 if the mth derivative of F at s00 is the first non-vanishing derivative. 9 2. The Riemann zeta function: topology Since the forward uniqueness of the trajectories gets lost in the hyperbolic point, we refer to the different parts as incoming or outgoing separatrices according to their direction with respect to the hyperbolic point. Incoming separatrices separate the flow between neighboring zeros or poles, whereas outgoing separatrices divide flows with different origin. Hence, the pictures of the Newton flow even reveal the location of hyperbolic points. 2.2. Newton flow of the Riemann zeta function Herein, we analyze the Newton flow of the Riemann zeta function. Figure 2.1 indicates that its behavior in the neighborhood of the critical strip is dictated by the behavior in the outer regions. Hence, we first explain the flow shown in the right and left panel and make the connection to the central part from each side. Moreover, we restrict our description to the upper half of the complex plane since the Newton flow of ζ is symmetric with respect to the real axis. 2.2.1. Asymptotic for σ → +∞ Recalling the definition of the Dirichlet series, eq. (4.1), it becomes obvious that the behavior of ζ for large positive real parts of s ≡ σ + iτ is governed by the function Fe (s) ≡ 1 + e −s ln 2 , (2.1) whose Newton flow is shown in appendix C.2.2. Indeed, the right panel of fig. 2.1 reveals separatrices with vanishing phase (solid green lines) which come from left and arrive at a zero of the derivative ζ 0 at σ → +∞ and τ = ±2k π/ ln 2, k ∈ N, marked by green circles. The only curves which start at +∞ are the outgoing separatrices of these points. They are located between the incoming ones at τ = ±(2k + 1) π/ ln 2 and indicated by green arrows. The flow lines near the outgoing separatrices point to the right and turn before they reach infinity, leading back to the critical strip in the neighborhood of the outgoing separatrices. When we examine the Newton flow at smaller values of σ, shown in the middle of fig. 2.1 and for a larger range in fig. C.4, the structures become more complicated. Nevertheless, the separatrices which continue in the right picture are clearly visible: The incoming separatrices with imaginary part larger than π/ ln 2 ∼ = 4.5 cross the complex plane from left to right forming ‘natural’ groups of non-trivial zeros. In contrast, the outgoing separatrices with τ > 10 have to end in a non-trivial zero of ζ since they alternate with the incoming separatrices. Moreover, it seems that all the curves coming from right do not cross the critical line. However, the separatrix at τ = π/ ln 2 and the flow around it acts differently. Although the separatrix comes from the right and can be considered as outgoing part with respect to the hyperbolic point at s00 = +∞, it is simultaneously an incoming separatrix for the first trivial zero of ζ 0 at s00 ∼ = −2.72. It encloses the flow between the pole sp = 1 and the first trivial zero s0 = −2. 10 2.2. Newton flow of the Riemann zeta function Figure 2.1.: The Newton flow of ζ reveals the separatrices, indicated by green lines, and the pole as well as the zeros and hyperbolic points, marked by black, red and green dots, respectively. The behavior of the Newton flow in the central panel is governed by the asymptotic shown on the left and right. From eq. (2.2) follows that the behavior of ζ for σ → −∞ is mainly determined by χ, eq. (1.6). Hence, in the left panel, the separatrices are real with phase 0 (solid lines) or π (dashed) in alternating sequence. They direct the flow into the trivial zeros of ζ 0 on the negative real axis, which are slightly shifted to the right of the corresponding zeros of χ0 . We note, that the behavior of the flow above the separatrix into the trivial zero of ζ 0 at σ ∼ = −4.94 is different but not shown here: only some real positive lines are separatrices, the other flow lines end in the non-trivial zeros of ζ. In the right panel, the asymptotic for σ → +∞ is given by Fe , eq. (2.1). Even though all separatrices are real and positive, only the incoming parts start at −∞, leading into the zeros of ζ 0 at σ → +∞ and τ = 2k π/ ln 2, k ∈ N/{0}, indicated by green circles. These lines form natural groups of the non-trivial zeros of ζ. The corresponding outgoing separatrices are the only lines starting at +∞. Their initial imaginary parts is τ = (2k + 1) π/ ln 2, k ∈ N/{0, 1}, and they end eventually in a non-trivial zero of ζ. However, the separatrix on the real line and the one starting at s = +∞ + iπ/ ln 2 are special. They enclose the Newton flow from the pole sp = 1 into the first trivial zero s0 = −2, as the central panel shows. Finally, there are also non-real separatrices determined by the non-trivial zeros of ζ 0 (dot-dashed green lines). Here, we have only depicted the first one which separates the flow into the second and third non-trivial zero of ζ. The phase on this line is ∼ 0.033. Hence, the curve is close to the real positive line (solid violet). Moreover, the dashed violet lines, indicating ordinary flow lines with phase π, end in the non-trivial zeros apparently without crossing the critical line (orange). In fig. C.4, we show the Newton flow up to τ = 300. 11 2. The Riemann zeta function: topology 2.2.2. Asymptotic for σ → −∞ Now, we concentrate on the left panel of fig. 2.1. There, the source of the flow is at σ → −∞, since (i) the flow from the only pole is restricted to disappear in the first trivial zero, and (ii) all lines starting at +∞ are real separatrices which end in the non-trivial zeros. To explain the inclination of the flow lines with respect to the real axis, we make use of the functional equation (1.5) and the approximation of the Dirichlet sum by Fe . This yields ζ(s) ∼ (2.2) = 1 + e −(1−s) ln 2 χ(s) showing that the asymptotic for σ → −∞ is mainly influenced by the function χ depicted in fig. C.3. The trivial zeros of ζ and χ are the same and the trivial zeros of both functions are located on the real axis, generating real separatrices with alternating signs, indicated by dashed lines if the phase is equal to π. However, the position of the trivial zeros of ζ 0 and χ0 do not entirely coincide due to the influence of the Dirichlet series in eq. (2.2). Of course, this influence gets weaker for σ → −∞, as shown in tab. C.1, and the pictures of the flow look the same. A detailed analysis of the Newton flow of χ is given in appendix C.2.1. Figure 2.2 shows the differences between the separatrices and real lines of ζ and χ. Here, we mark only the special points of the Riemann zeta function by dots: the pole (black) and the zeros of ζ (red) and ζ 0 (green). The gray and blue lines indicate the separatrices and real lines of χ, respectively, whereas the corresponding lines of ζ are depicted in green and violet. The different styles of the curves represent the phases: solid = 0, dashed = π and for the dot-dashed lines determined by the hyperbolic point in which they cross. We neglect to draw the lines of χ to the right of the critical strip if they are not necessary to describe the topology of ζ. As expected, the real separatrices of χ through the hyperbolic points with σ < −4 are almost the same as the separatrices of ζ through the trivial zeros of ζ 0 . To make the tiny deviations visible, we provide detailed views on the right side of fig. 2.2. Moreover, the separatrix of χ with phase ∼ π/4 (dot-dashed gray curve) through its non-trivial zero on the critical line at τ ∼ = 6.23 does not occur as separatrix in the Newton flow of ζ. Below this line, χ possesses two real separatrices enclosing the flow between its pole s = 3 and its zero s = −2 as well as the flow between its pole s = 1 and its zero s = 0. In the case of the Riemann zeta function, we have only one pole at sp = 1 which is connected by the flow to the first trivial zero s0 = −2. Above the dot-dashed gray curve, the function χ has no separatrices. But in the Newton flow of ζ some of the positive real lines of χ (solid blue curves) become the positive real separatrices of ζ (green solid lines). They cross the critical strip and continue as separatrices into the zeros of ζ 0 at +∞ familiar from the right panel in fig. 2.1. In contrast, each negative real line ends at a non-trivial zero, apparently without crossing the critical line. They are depicted as dashed violet curves. 12 2.2. Newton flow of the Riemann zeta function Figure 2.2.: The separatrices of χ and ζ show that the influence of the Dirichlet series in the approximation eq. (2.2) is quite small in the left half of the complex plane: For σ < −4, the difference between the real separatrices of χ and ζ – depicted by gray and green lines, dashing indicating the phase π – is only visible in the magnified views with higher resolution on the right. In contrast, the separatrix through the non-trivial zero of χ on the critical line with τ ∼ = 6.23 (dot-dashed gray line) does not appear as separatrix in the flow of ζ, as well as other separatrices to the right (see fig. C.3) which have been neglected in the graph. There are two real separatrices of χ displayed below its non-trivial zero. They enclose the flow between the poles s = 1 and s = 3 and the zeros s = 0 and s = −2 of χ. In contrast, the Riemann zeta function only possesses one real separatrix in this region, since ζ has only one pole at sp = 1. This separatrix starts at s00 = +∞ + iπ/ ln 2 and leads to the first trivial zero of ζ 0 at σ ∼ = −2.7. Above, we find different separatrices of ζ, but only the non-trivial one of χ (dot-dashed gray line). The real separatrices of ζ can be matched to some of the positive real lines of χ (solid blue curves), whereas the negative real lines (dashed blue) are similar to the negative curves of ζ, indicated by dashed violet lines which end in a non-trivial zero of ζ. 13 2. The Riemann zeta function: topology 2.2.3. Non-trivial separatrices So far, we have only described the real separatrices of ζ determined by the trivial zeros of ζ 0 located on the real axis or at σ → +∞. The outgoing parts of the last form natural groups of two or more non-trivial zeros of ζ for τ > 18. Hence, the flow into different zeros of one group must be separated by non-real curves through hyperbolic points. This property of the Newton flow indicates that there are n − 1 non-trivial zeros of the derivative ζ 0 in the region of a group with n non-trivial zeros of ζ, provided that the higher derivatives of ζ do not vanish at the hyperbolic points. Otherwise, there are less zeros of ζ 0 according to property (vi) listed in the section 2.1. Yet, fig. C.4 confirms that up to τ = 300 no hyperbolic point exists where the second derivative of ζ vanishes [40]. Several results concerning the zeros of the derivatives ζ (k) are presented in [42]. Figure 2.1 only contains the separatrix through the first non-trivial zero of ζ 0 (dot-dashed green line), which is close to the real lines of ζ (violet curves) since its phase is ∼ 0.033. We show in tab. C.2 that the phases of the non-trivial separatrices up to imaginary part τ = 100 are quite small. This explains the ‘attraction’ of the real lines by the hyperbolic points observed in [36]. Finally, we mark that the importance of the hyperbolic points becomes evident in the formulation of the Riemann hypothesis by Speiser [35]: The derivative ζ 0 has no zeros in the left half of the critical strip 0 < σ < 1/2. Hence, it is worthwhile to further investigate the Newton flow of ζ and its derivatives, since it would reveal if the Riemann hypothesis was violated. 14 3. Combined quantum systems Our physical approach towards the Riemann zeta function takes advantage of its similarity to the time evolution of a quantum state. When we choose the associated Hamiltonian appropriately, we can reproduce the phases of the summands in the different representations of ζ. The overlap with an adequate reference state then takes care of the amplitudes. In this chapter, we introduce the general form of these states and show how the phase space representations are connected to the overlap of the states. 3.1. Time evolution We consider the quantum state |Ψ(0)i ≡ |ψos i ⊗ |ψat i = ∞ X n=0 ψn |ni ⊗ ce |ei + cg |gi , (3.1) which is the direct product of the initial oscillator and atomic state, |ψos i and |ψat i, respectively. The harmonic oscillator can be represented by a single mode of a cavity field [43] or the motion of a trapped ion [44]. At time t = 0, it is given by a superposition of Fock states |ni with probability amplitudes ψn . Likewise, the atomic state is in a superposition of the ground and the excited state, |gi and |ei, of a two-level atom with probability amplitudes cg and ce . Both states have to be normalized to ensure the probability interpretation which implies ∞ X n=0 |ψn |2 = 1 and |ce |2 + |cg |2 = 1 . The time evolution arises due to the Riemann Hamiltonian1 ĤR ≡ ~ω ln(n̂ + 1) σ̂z (3.2) in the interaction picture. Here, ω denotes the Rabi frequency which establishes the coupling between the atom and the field mode and σ̂z ≡ |eihe| − |gihg| 1 (3.3) M. Planat et al. suspected in their paper [1] that it ‘may be that the logarithmic Hamiltonian [ln n̂] is realized in the field of quantum physiology, where the perception states are proportional to the logarithm of the excitation instead of being proportional to it.’ 15 3. Combined quantum systems is the Pauli spin matrix. The Hamiltonian ĤR is reminiscent of the effective Hamiltonian of the Jaynes-Cummings-Paul model [28, 43] ĤJC ≡ ~ω n̂ σ̂z , which is central to cavity quantum electrodynamics. Since this model has – despite of its simplicity – a remarkable power of prediction, we give a short review and propose some modifications to realize the Riemann Hamiltonian in chapter 8. With the Hamiltonian ĤR , eq. (3.2), we get |Ψ(t)i ≡ |ψe (t)i|ei + |ψg (t)i|gi (3.4) for the time-evolved state. It is crucial to keep in mind that the abbreviations |ψe (t)i ≡ ce ∞ X n=0 ψn e −iωt ln(n+1) |ni and |ψg (t)i ≡ cg ∞ X n=0 ψn e +iωt ln(n+1) |ni are in general not normalized to unity since the entangled state |Ψi has to be normalized. Due to the fact that ĤR contains σ̂z , the two states accumulate the opposite phases ±iωt ln(n + 1) which depend linearly on time but logarithmically on the photon number n. It is this superposition which creates the Riemann zeta function in the critical strip as we see in what follows. 3.2. Joint measurement We take the overlap C(t) = hΦ|Ψ(t)i between the time-evolved state |Ψ(t)i and the entangled reference state X |Φi ≡ |φa i|ai . (3.5) a Here, a ∈ {e, g} and the in general unnormalized states |φa i are given by the expansion |φa i ≡ ∞ X n=0 φna |ni in Fock states with probability amplitudes φna . Equations (3.4) and (3.5) then yield the expression C(t) = ∞ X n=0 φ ne ce ψn e −iωt ln(n+1) + φ ng cg ψn e iωt ln(n+1) . (3.6) The dependence of the phase factor exp(−iωt ln(n + 1)) = (n + 1)−iωt indicates an intimate connection between the overlap (3.6) and the Dirichlet representation (1.1) 16 3.3. Phase space representations when we recall that the summands are n−σ−iτ . The term with the opposite phase is reminiscent of the second sum in the Riemann-Siegel formula (1.9). Hence, we use eq. (3.6) to produce the different representations of the Riemann zeta function by quantum states in the next chapters. There, we also show that entanglement is crucial to reach into the critical strip, but is unnecessary in the case of the Dirichlet sum. Before we turn to the definition of the states, we give a short overview of the phase space representations used afterwards to illustrate their behavior. 3.3. Phase space representations For our purpose, the most suitable representation of a quantum state in phase space is given by the Wigner formalism [18] and the closely related Moyal representation [20]. We give now a short overview of the required expressions and refer for a more detailed description of the formalism and the properties of the functions to appendix D and to literature, e.g. [28, 45]. 3.3.1. Wigner and Moyal function The well-known definition [18] 1 W|ψi (x, p) ≡ 2π~ Z∞ i ξ ξ dξ e − ~ pξ hx + |ψihψ|x − i 2 2 (3.7) −∞ of the Wigner function provides us with the overlap 2 |hφ|ψi| = 2π~ Z∞ Z∞ dx −∞ dp W|ψi (x, p) W|φi (x, p) (3.8) −∞ of two pure states |φi and |ψi. It involves the product of the corresponding Wigner functions integrated over phase space, spanned by the position x and momentum p of the wave function. Since the Wigner function is real per definition, eq. (3.8) makes only the absolute value of a complex number available, but not its phase. Fortunately, we can directly calculate the scalar product hφ|ψi with the expression hφ|ψi = Z∞ −∞ dx Z∞ dp W|ψihφ| (x, p) , (3.9) −∞ where we have used the generalize definition 1 W|ψihφ| (x, p) ≡ 2π~ Z∞ i ξ ξ dξ e − ~ pξ hx + |ψihφ|x − i 2 2 (3.10) −∞ 17 3. Combined quantum systems of eq. (3.7) for the Moyal function of the two states. Equation (3.9) is quite powerful since it yields the explicit formula Z∞ Z∞ X X W|ψa ihφa | (x, p) C = hΦ|Ψi = hφa |ψa i = dx dp a −∞ −∞ (3.11) a for the scalar product of two entangled states |Φi and |Ψi of the form (3.5) in terms of Moyal functions of their oscillator parts |φa i and |ψa i. 3.3.2. Wigner and Moyal matrix Another representation of an entangled state in phase space is the Hermitian Wigner matrix, eq. (D.9), X Ŵ|Ψi = W|ψa ihψa0 | |aiha0 | (3.12) a,a0 with a, a0 ∈ {e, g}. This expression immediately follows from the definition of the Wigner function, eq. (3.7), for a pure entangled state |Ψi of the form (3.5). In appendix D.3, we give a generalization of this definition and its properties to arbitrary quantum states. The Moyal functions W|ψa ihψa0 | are given in terms of the Wigner and Moyal functions W|ni and W|nihm| , eqs. (D.20) and (D.17), of the Fock states by "∞ # ∞ n−1 X X X ψna ψ na0 W|ni + W|ψa ihψa0 | = ψna ψ ma0 W|nihm| + ψma ψ na0 W |nihm| . n=0 n=1 m=0 (3.13) Due to the Hermiticity of the Wigner matrix, the diagonal elements W|ψa ihψa | ≡ W|ψa i = ∞ X n=0 |ψna |2 W|ni + 2 ∞ n−1 X X n=1 m=0 Re ψna ψ ma W|nihm| (3.14) are the Wigner functions of the states |ψa i and therefore real. In contrast, the offdiagonal matrix elements are complex and connected via W|ψg ihψe | = W |ψe ihψg | . Moreover, the probability factors ψne and ψng merely scale the height of the Wigner and Moyal functions of the Fock states since they are independent of the phase space variables x and p. As shown in appendix D.3, we find the probability amplitude Z∞ Z∞ |C| ≡ |hΦ|Ψi| = 2π~ dx dp K 2 2 −∞ −∞ of two entangled states |Φi and |Ψi by integrating the kernel X K ≡ Tra Ŵ|Ψi Ŵ|Φi = W|ψa ihψa0 | W|φa0 ihφa | a,a0 18 (3.15) (3.16) 3.3. Phase space representations over the whole phase space. In analogy to eq. (3.8), K contains the product of the corresponding Wigner representation (3.12) of the states. However, since Ŵ|Φi and Ŵ|Ψi are operators, the product does not only include the Wigner functions W|ψa i ≡ W|ψa ihψa | , that is the diagonal elements of Ŵ|Ψi , but also the Moyal functions which represent the off-diagonal matrix elements. Just for the sake of completeness, we mention that the generalization of the Wigner matrix (3.12) is obviously the Moyal matrix X Ŵ|ΨihΦ| = W|ψa ihφa0 | |aiha0 | a,a0 of the states |Φi and |Ψi, emerging directly form the definition of the Moyal functions (3.10). This representation implies that the scalar product hΦ|Ψi, eq. (3.11), results from the trace X W|ψa ihφa | K ≡ Tra Ŵ|ΨihΦ| = a of the Moyal matrix integrated over phase space. In the following chapters, we use the Moyal representation (3.11) and the Wigner representation (3.15) of the overlap C(t), eq. (3.6), to describe the Riemann zeta function in phase space. 19 4. Riemann states In this chapter, we construct states which reproduce the Dirichlet series (1.1) ∞ X 1 ζ(s) ≡ ns (4.1) n=1 of the zeta function by a joint measurement. It is crucial to keep in mind that this representation is only valid for Re s ≡ σ > 1, since this fact restricts the quantum states to the same region. Therefore, we name these states Riemann states. Moreover, we analyze the behavior of their phase space representations. 4.1. Definition and properties When we rewrite the Dirichlet sum (4.1) by using s ≡ σ + iτ and shifting the summation index, it becomes evident that ζ(s) = ∞ X n=0 1 e −iτ ln(n+1) (n + 1)σ (4.2) holds some similarities with the quantum mechanical time evolution, eq. (3.4). Indeed, choosing ce = 1 and cg = 0, the overlap C(t), eq. (3.6), between the time-evolved state |Ψ(t)i and the entangled reference state |Φi yields C(t) = ∞ X φ ne ψn e −iωt ln(n+1) . (4.3) n=0 We can reproduce the Dirichlet representation (4.2) by eq. (4.3) if the probability amplitudes of the initial oscillator state and the reference state read ψn (σ) ≡ N (σ) (n + 1)σ/2 and φne ≡ ψn (σ) , respectively. For the sake of simplicity, we choose φng = 0, since it does not contribute to eq. (4.3). The overlap C(t) = |N (σ)|2 ∞ X n=0 1 e −iωt ln(n+1) = |N (σ)|2 ζ(σ + iωt) (n + 1)σ 21 4. Riemann states emerges then from the scalar product hΦ|Ψi of the state |Ψ(t)i = N (σ) ∞ X 1 e −iωt ln(n+1) |n, ei σ/2 (n + 1) n=0 (4.4) and its initial state |Φi ≡ |Ψ(0)i used as reference. The normalization is given by !−1/2 ∞ X 1 N (σ) = = ζ −1/2 (σ) (4.5) (n + 1)σ n=0 which is only convergent for σ > 1. Hence, with these states we can only describe the zeta function in the region where the Dirichlet series (4.2) is convergent. Equation (4.4) clearly shows that the time-evolved state |Ψ(t)i remains a product state of the excited state |ei and the state |σ, τ i ≡ N (σ) ∞ X 1 e −iτ ln(n+1) |ni σ/2 (n + 1) n=0 at the rescaled time τ = ωt. Thus, the overlap ζ(σ + iωt) C(t) = he|hσ, 0| |σ, ωti|ei = hσ, 0|σ, ωti = ζ(σ) (4.6) (4.7) reduces to the scalar product of the time-evolved state |σ, ωti with its initial state, revealing that entanglement is not necessary for the description of the zeta function in the region σ > 1. Therefore, we call |σ, τ i Riemann state. Before we turn to the analysis of the Wigner function of the Riemann state, it is worthwhile to mention that |σ, τ i, eq. (4.6), is a generalization of the thermal phase state [46] ∞ X |ψp i ≡ Np e −σn/2 |ni . (4.8) n=0 Both states are coherent superpositions of photon number states with real expansion coefficients which decay with n. Their similarities and differences become evident, when we compare the corresponding photon distributions: In the case of the thermal phase state (4.8), we obtain |hn|ψp i|2 = |Np |2 e −σn , (4.9) which decays exponentially with the photon number n and has a maximum at n = 0, that is for the vacuum state |0i. For the Riemann state, the maximum of |hn|σ, 0i|2 = |N (σ)|2 = |N (σ)|2 e −σ ln(n+1) (n + 1)σ (4.10) is also at n = 0. However, the photon statistics (4.10) only decays polynomially with n, which is much slower than the exponential decay. In this sense, we have replaced n by ln(n + 1) in eq. (4.9), in agreement with the construction of the Hamiltonian ĤR , eq. (3.2). 22 4.2. Wigner functions Figure 4.1.: Wigner function1 W|σ,0i , eq. (4.11), of the initial Riemann state |σ, 0i for different values of σ. For small values we get a tail aligned along the x-axis with interference fringes on both sides created by the contribution of the Moyal functions. The interference fringes as well as the tail vanish for larger σ due to the photon statistics, eq. (4.10). 4.2. Wigner functions The definition of the Wigner function, eq. (3.7), yields for the Riemann state |σ, τ i the expression W|σ,τ i n o n+1 −iτ ln m+1 ∞ n−1 Re e W X X |nihm| W 1 |ni , +2 = σ/2 ζ(σ) (n + 1)σ [(n + 1)(m + 1)] n=0 n=1 m=0 ∞ X (4.11) where τ = ωt is the rescaled time. Of course, W|σ,τ i underlies the same restriction σ > 1 as the state |σ, τ i itself. Like in the photon statistics (4.10), the Wigner function of the vacuum state |n = 0i gives the largest contribution to the first sum. Although the Gauss shape of the Wigner function W|0i is superimposed by the wells of the Wigner functions W|ni with larger n, the whole sum remains positive in the entire phase space producing a structure which is rotational symmetric with a maximum at the origin. For a more detailed description of the shapes of W|ni we refer to appendix D. Because of the factor [(n+1)(m+1)]−σ/2 in the double sum, the Moyal function W|nihm| of the smallest energy eigenstates gives the highest contribution to this sum. For τ = 0, the superposition of both sums produces a tail aligned along the x-axis which, due to the denominator, gets longer for σ → 1 since more and more terms become significant in the double sum. Figure 4.1 shows this characteristics. Moreover, the Moyal functions shift the maximum of the first sum to the right and produce interference fringes besides the tail. These properties become clear when we examine the Wigner function of the so-called truncated Riemann state |σ, τ iν in section 5.2. To understand the behavior of W|σ,τ i for τ > 0, we recall the explicit expressions of the Wigner and Moyal functions of the Fock states, eqs. (D.20) and (D.21), which transform 1 All plots are created with Mathematica [47] and we use ~ = M = κ = 1 in all phase space plots. 23 4. Riemann states Figure 4.2.: According to eq. (4.15) the Riemann zeta function given by the Dirichlet sum is proportional to |C(t)|2 , depicted on top, which describes also the scalar product between the initial Riemann state |σ, 0i and the time-evolved state |σ, ωti. In the Wigner formulation of quantum mechanics this scalar product is given by the overlap in phase space between the corresponding Wigner functions W|σ,τ i , eq. (4.11), presented in the middle row for σ = 2 and three values of τ = ωt. The bottom row presents the product of these Wigner functions with the Wigner function of the initial state W|2,0i shown in the left picture of fig. 4.1. Integration of this product over phase space produces the values marked in the top picture by blue dots. 24 4.3. Moyal functions eq. (4.11) into W|σ,τ i 1 e −2|α| = ζ(σ) π~ 2 " ∞ X n=0 wnn (n + 1)σ # wnm |α|n−m n+1 . +2 cos (n − m)β + τ ln m+1 [(n + 1)(m + 1)] σ/2 n=1 m=0 ∞ n−1 X X (4.12) For the sake of simplicity, we have omitted the explicit notation of the dependence on α = α(x, p), eq. (D.16), in its phase β ≡ arg α and in the real function wnm , eq. (D.18). But we keep in mind that α is connected to the phase space variables via Re α ∼ x and Im α ∼ p. Equation (4.12) shows that the time-dependent factor in the double sum produces a clockwise rotation of the Moyal functions 2 W|nihm| e −2|α| = wnm |α|n−m e −i(n−m)β π~ (4.13) for n > m by the angle n+1 . (4.14) m+1 This causes a rotation of W|σ,τ i for τ > 0 which results in a curling of the tail around the origin, as depicted in the middle row of fig. 4.2 for σ = 2. Due to the different rotation angles, eq. (4.14), the structure of the tail splits up and the initial symmetry in p gets lost. ϕnm (τ ) ≡ τ ln Finally, we make the connection between the Wigner functions of the states and their overlap C(t), eq. (4.7), by recalling eq. (3.8), that is Z∞ Z∞ ζ(σ + iωt) 2 . |C(t)| = |hσ, 0|σ, ωti| = 2π~ dx dp W|σ, 0i W|σ,ωti = ζ(σ) 2 2 −∞ (4.15) −∞ The bottom row of fig. 4.2 shows the product W|σ, 0i W|σ,τ i at different times τ = ωt revealing that a small region of phase space around the central maximum produces the main contribution to the overlap. The corresponding values of the overlap (4.15) are marked by dots in the picture on top. Since the Wigner formalism only produces the absolute value of ζ, we analyze in the next section the representation of the scalar product hφ|ψi based on Moyal functions. 4.3. Moyal functions The scalar product hσ, 0|σ, ωti of the initial and time-evolved Riemann state can be evaluated directly by eq. (3.9), that is Z∞ Z∞ ζ(σ + iωt) C(t) = hσ, 0|σ, ωti = dx dp W|σ,ωtihσ, 0| = , ζ(σ) −∞ (4.16) −∞ 25 4. Riemann states Figure 4.3.: The lower pictures illustrate the time evolution of the Moyal function W|2,τ ih2,0| , eq. (4.17), in dependence on τ = ωt. The second row shows that the initial real part, which is equal to W|σ,0i , becomes a short positive wedge with a negative ridge for τ = 1, while the imaginary part displays a mainly negative wedge. For larger times, both wedges curl more and more around the origin. The pictures of the absolute value (fourth row) are almost the same as the ones of the real part since |ζ(2 + iτ )| ∼ = Re {ζ(2 + iτ )}, as the curves in the top picture indicate. The values marked there by dots emerge from the Moyal functions by integration over phase space according to eq. (4.16). The last row depicts the phase of W|2,τ ih2,0| which changes between −π and π for τ > 0. 26 4.3. Moyal functions in terms of the complex Moyal functions, eq. (3.10), "∞ X e −iτ ln(n+1) 1 W|σ,τ ihσ,0| = W|ni ζ(σ) (n + 1)σ + n=0 ∞ X n−1 X n=1 m=0 e −iτ ln(n+1) W|nihm| + e −iτ ln(m+1) [(n + 1)(m + 1)]σ/2 W |nihm| # (4.17) providing us with the absolute value and the phase of ζ. In contrast to the Wigner function W|σ,τ i of the Riemann zeta state |σ, τ i, eq. (4.11), both sums in eq. (4.17) contain time-dependent factors and therefore produce complex values for τ = ωt > 0. Figure 4.3 illustrates the time evolution of the Moyal function W|σ,τ ihσ,0| for σ = 2. Apart from the initial Moyal function, which is of course equal to the real-valued Wigner function W|σ,0i depicted in the left picture of fig. 4.1, the real part of W|2,τ ih2,0| develops a central maximum with a negative tail curling around the origin for τ > 0. In contrast, the imaginary part possesses mainly a negative wedge at τ = 1, which curls in the same way as the real part when time evolves. The shapes of absolute value |W|2,τ ih2,0| | are almost the same as the ones of Re W|2,τ ih2,0| since |ζ(2 + iτ )| ∼ = Re {ζ(2 + iτ )}, confirmed by the upper picture. Of course, the negative regions are positive in the case of the absolute value. The bottom row depicts the phase of W|2,τ ih2,0| . The first picture only shows the negative interference fringes of W|2,0i in blue, while for larger times the phase changes around the tail from −π to π, indicated by the colors of the thermometer on the right. Finally, when we integrate the Moyal functions according to eq. (4.16) over phase space, we get the values of the corresponding part of ζ marked in the top picture by dots. We conclude this chapter by reminding that the Riemann states are only suitable to describe the Riemann zeta function for σ > 1. In the following chapters, we show that we can penetrate the wall at σ = 1 with the help of the truncated Riemann states |σ, τ iν , which will be defined in section 5.2. 27 5. Alternating sum and truncated Riemann states One way to push the limitation σ > 1 to smaller values of σ is to use the alternating sum representation (1.3) of the Riemann zeta function. In our physical picture, however, this approach does not work. In this chapter, we show that all states which represent the alternating sum exactly suffer from the same restriction σ > 1 as the Riemann states. The only possibility to reach into the critical strip is to approximate ζ in this region, which leads to the definition of truncated Riemann states. 5.1. Exact solution A comparison between the overlap C(t), eq. (3.6), and the alternating sum of ζ, eq. (1.3), yields the condition ∞ ∞ X X Ma (−1)n+1 −iτ ln(n+1) ! ψn ce φ ne e −iτ ln(n+1) + cg φ ng e iτ ln(n+1) = 1−s e 2 −1 (n + 1)σ n=0 n=0 for the probability amplitudes. Here, we have already used the abbreviation τ = ωt. The proportionality factor Ma contains the normalization constants of the states. Without loss of generality, we choose real normalization constants from now on. There are two possibilities to choose the probability amplitudes: (i) omitting entanglement by choosing cg = 0 like in the case of the Riemann state, or (ii) allowing entanglement. We consider both ways separately in what follows. 5.1.1. Construction à la Riemann states In the case of the Dirichlet sum, it is sufficient to consider the Riemann state which represents only the oscillatory system. Since the alternating sum is similar to the Dirichlet representation, both describe ζ with one infinite sum, it is intuitive to choose the amplitudes ce ≡ 1 , cg ≡ 0 , ψn ≡ N (σ) (n + 1)σ/2 and φne ≡ N (σ) (−1)n+1 . (n + 1)σ/2 (5.1) Here, the normalization N (σ) is given by eq. (4.5) which restricts the states again to the region σ > 1. When we use other distributions of the phase (−1)n+1 in the probability amplitudes, e.g. ψn ≡ N (σ) e iϕ(n+1) (n + 1)σ/2 and φne ≡ N (σ) e i(π−ϕ)(n+1) (n + 1)σ/2 29 5. Alternating sum and truncated Riemann states with ϕ ∈ [0, 2π), the same normalization occurs. Therefore, we cannot reach into the critical strip with these states. Now, the question arises if we can extend the region of definition by changing the photon statistics of the states, that is using (a) ψn ≡ Nψ (σ) with a > 0 . (n + 1)σa (5.2) Indeed, for this state the normalization is given by (a) Nψ (σ) ≡ ∞ X n=0 1 (n + 1)2σa !−1/2 = ζ −1/2 (2σa) for σ > 1 2a (5.3) and we arrive at states with σ < 1 by choosing a > 1/2. However, the probability amplitudes of the corresponding reference state have to be (a) φne ≡ Nφ (−1)n+1 (n + 1)(1−a)σ with the normalization (a) Nφ ≡ ∞ X 1 (n + 1)2σ(1−a) n=0 !−1/2 = ζ −1/2 (2σ(1 − a)) convergent for 2σ(1 − a) > 1. Hence, we need a < 1/2 for values σ < 1 which is not (a) allowed in the normalization Nψ , eq. (5.3). Only for a = 1/2 it is possible to get both states normalized, but then σ has to be larger than 1 as we have shown before. Needless to say, we can choose again other distributions of the phase (−1)n+1 in the amplitudes of the state and reference state, but this does not change the normalizations. 5.1.2. Entanglement It remains to show that even entanglement cannot√help to overcome the wall at σ = 1. Indeed, when we choose for example ce = cg = 1/ 2 and the photon statistics ψn according to eq. (5.1), we can distribute the phases (−1)n+1 to both probability amplitudes φne and φng of the reference state |Φi, eq. (3.5). For example, splitting the alternating sum in an odd and an even part yields for the probability amplitudes φne ∼ − δn,2k+1 (2k + 2)σ/2 and φng ∼ δn,2k (2k + 1)σ/2 with δn,k denoting the Kronecker delta. In this case, the normalization of the reference state is again given by N (σ), eq. (4.5), which restricts the states to the plane with σ > 1. Hence, the alternating sum is mathematically a powerful tool to reach into the critical strip, but there are no physical states which could realize the infinite alternating sum by the joint measurement described in section 3.2. 30 5.2. Truncated Riemann states Figure 5.1.: The Wigner function of the truncated Riemann state |1/2, 0iν has ν − 1 minima surrounding the central maximum which is shifted to the right of the origin. Both properties result from the Moyal functions in the double sum of eq. (5.6). For larger summation limits ν, the surrounding minima become a ring of interference fringes which faints further away from the positive x-axis. 5.2. Truncated Riemann states The simplest way to get normalizable states with probability amplitudes similar to eq. (5.1) is to use only a finite number of Fock states. Therefore, we define in analogy to the Riemann state (4.6) the truncated Riemann state |σ, τ iν ≡ Nν (σ) ν−1 X 1 e −iτ ln(n+1) |ni σ/2 (n + 1) n=0 with its normalization Nν (σ) ≡ ν−1 X n=0 1 (n + 1)σ !−1/2 (5.4) (5.5) valid for all σ > 0 if ν is finite. Such truncated states can be realized with methods of state engineering [48] and state synthesis [49]. For ν → ∞ the expression (5.4) becomes equal to the definition (4.6) of the Riemann state |σ, τ i∞ ≡ |σ, τ i and, due to the normalization, again restricted to σ > 1. Nevertheless, the truncated Riemann states provide a deeper insight into the structure of the Riemann states as we show now. 31 5. Alternating sum and truncated Riemann states The Wigner function of the truncated Riemann state n o n+1 −iτ ln m+1 ν−1 n−1 Re e W X X |nihm| W |ni +2 W|σ,τ iν (α) ≡ Nν2 (σ) σ/2 (n + 1)σ [(n + 1)(m + 1)] n=1 m=0 n=0 ν−1 X (5.6) is the truncated version of W|σ,τ i , eq. (4.11). It is obvious that the photon statistics |hn|σ, τ iν |2 = Nν2 (σ) (n + 1)σ for n<ν enhances the contributions of the Fock states with small photon number. Hence, the highest contribution to the first sum in eq. (5.6) is given by the Wigner function of the vacuum state W|0i , which is positive in the whole phase space with a maximum at the origin. The contributions of the higher Fock states do not change these properties. Even the sum over the Wigner functions without the factors (n + 1)−σ is positive, as we show in appendix D.4.1. In the double sum the Moyal function W|1ih0| gives the largest contribution, which – due to the shape of its real part – enhances the values at x > 0 and reduces them for x < 0 when superposed with the first sum. This becomes evident in the second picture of fig. 5.1 where W|1ih0| is the only contribution of the double sum. The real parts of the higher Moyal functions are always zero at the origin and the contributions with odd m have a minimum following on the positive x-axis (see fig. D.2). Hence, the maximum of the first sum at the origin gets shifted to the right as we see in the other pictures of fig. 5.1. Since W|ni and W|nihm| , eqs. (D.20) and (D.21), contain the factor 2 e −2|α| , acting as an envelop, the whole structure flattens with increasing distance to the origin. Moreover, the star-shaped patterns of Re W|nihm| , depicted in fig. D.2, lead to ν − 1 minima around the origin. They create for larger summation limits ν a ring with increasing radius in which the structure distant to the positive x-axis faints. Thus, they develop the tail aligned along the x-axis which is known from the Wigner function of the Riemann state |σ, τ i. Although the structure of the truncated Riemann state resembles the one of |σ, τ i, we have to keep in mind that, due to the normalization, they can only become equal for ν → ∞ if σ > 1. For times τ 6= 0 the real parts of the Moyal functions W|nihm| are additionally rotated clockwise around the origin with angles ϕnm (τ ) given by eq. (4.14). Since the rotation angle changes for different combinations of n and m the initial symmetry in p gets destroyed. Hence, the whole structure of the Wigner function W|σ,τ iν becomes even more complicated for large τ and, of course, with increasing ν. The left pictures in the bottom part of fig. 5.2 show this behavior for W|2/3,τ i100 at different times τ . We postpone the description of the other pictures shown in fig. 5.2 to section 5.3.2 and analyze the influence of the truncation on the alternating sum before. 32 5.3. Truncated alternating sum Figure 5.2.: The absolute value squared of the overlap C, eq. (5.8), which gives the values of the truncated alternating sum, is shown in blue in the upper right picture for σ = 2/3 and (100) 2 ν = N = 100. The red dashed line indicates |Ma ζ| , emphasizing the good agreement between the approximation (5.8) and the exact values for large N . The values marked by dots can be calculated from the products W|φiN W|σ,τ iN , depicted below in the right pictures, by integration over phase space according to eq. (5.12). In the pictures of the products the values at different points are much smaller than the ones of the Wigner functions W|σ,τ iN , eq. (5.6), shown on the left. The clockwise rotation of the Wigner function curls the tail around the origin. The interference fringes besides the tail fade away with larger distance to the x-axis. The upper left picture confirms that W|φiN is W|σ,0iN mirrored on the p-axis, in agreement with eq. (5.11). 33 5. Alternating sum and truncated Riemann states Figure 5.3.: The absolute value of the truncated alternating sum ζN , eq. (5.7), is depicted for σ = 1/2 (red) and σ = 2/3 (blue) for the truncation parameters N = 100 and N = 200 in the first row. As expected, for both N the red line is close to zero at the zeros of ζ, marked by the vertical dashed lines at τ ∼ = 14.135, τ ∼ = 21.022, τ ∼ = 25.011 and τ ∼ = 30.425. However, the magnified pictures of the difference |ζN | − |ζ| in the bottom show that the values do not vanish entirely at these points. Nevertheless, the quality of the approximation ζN increases for larger N as well as for increasing σ, as the pictures of the difference in the second row confirm. 34 5.3. Truncated alternating sum σ = 1/2 σ = 2/3 |ζN | 6 4 2 680 682 684 680 682 684 τ Figure 5.4.: The absolute value of the truncated alternating sum ζN , eq. (5.7), is depicted for σ = 1/2 (red) and σ = 2/3 (blue) for the truncation parameters N = 200 and N = 230 by dashed and dotted lines, respectively. The solid lines indicate the exact curves ζ ≡ ζ∞ . The pictures show that we need higher truncation parameters N for larger imaginary parts τ to get an adequate approximation of ζ ≡ ζ∞ . The dashed vertical lines indicate the zeros of ζ at τ∼ = 681.895, τ ∼ = 682.603 and τ ∼ = 684.014. = 679.742, τ ∼ 5.3. Truncated alternating sum In analogy to the truncated Riemann states, we define the truncated alternating sum ζN (s) ≡ 1 21−s − 1 N −1 X n=0 (−1)n+1 (n + 1)s (5.7) by cutting the sum in eq. (1.3) at a finite truncation parameter N . Figures 5.3 and 5.4 show that the quality of the approximation ζN strongly depends on the chosen parameters and can always be improved by enlarging the truncation N . The first row in fig. 5.3 presents the absolute value of ζN for σ = 1/2 (red) and σ = 2/3 (blue) as well as N = 100 and N = 200 for small imaginary parts τ . As expected, the red curve is close to zero at the zeros of ζ, marked by the vertical dashed lines. However, the magnified pictures of the difference |ζN | − |ζ| in the bottom rows confirm that the values do not vanish entirely at these points. The zeros are only reached in the limit N → ∞. For σ = 2/3, the difference to the exact values is smaller, since the influence of the summands in eq. (1.3) with large n fades for increasing σ. Figure 5.4 demonstrates that we need larger truncation parameters N for larger imaginary parts τ to get a reliable approximation of the Riemann zeta function ζ ≡ ζ∞ , indicated by the solid curves. The dashed and dotted lines represent N = 200 and N = 230, respectively. In what follows, we use σ = 2/3 in the phase-space pictures of the states and the overlap since the deviation of ζN to the exact values is smaller than for σ = 1/2. The description of ζ for smaller σ is subject to chapter 6 and 7, where we reproduce the Riemann-Siegel and Berry-Keating formula which are more suited to approximate ζ in the critical strip. 35 5. Alternating sum and truncated Riemann states 5.3.1. Reference state and overlap Now, we turn to the calculation of the overlap ) C(t) = N hφ|σ, ωtiν = M(ν,N (σ, ωt) ζN (σ + iωt) . a (5.8) With the definition (5.4) of the truncated Riemann state |σ, τ iν and the truncated alternating sum, eq. (5.7), we find the reference state |φiN ≡ NN (σ) N −1 X n=0 (−1)n+1 |ni (n + 1)σ/2 (5.9) with N ≤ ν and the proportionality factor ) M(ν,N (σ, τ ) ≡ NN (σ) Nν (σ) 21−σ−iτ − 1 . a (5.10) In appendix F.1, we show that the normalization NN , eq. (5.5), decreases for increasing truncation limit N and σ in the critical strip. This behavior makes the proportionality (N ) (N,N ) factor Ma ≡ Ma , analyzed in appendix F.4, as well as the overlap C small for large N . Nevertheless, we examine the Wigner and Moyal representations of ζN in the remaining section. 5.3.2. Wigner representation Equation (3.14) yields the Wigner function W|φiN = NN2 (σ) " N −1 X n=0 # N −1 n−1 X X (−1)n+m Re W|nihm| W|ni +2 (n + 1) σ [(n + 1)(m + 1)] σ/2 n=1 m=0 of the reference state |φiN defined by eq. (5.9). This expression is reminiscent of the Wigner function W|σ,0iN of the initial truncated Riemann state. Indeed, the first part is for ν = N the same as in eq. (5.6) and therefore symmetric in x and p. With the help of eq. (4.13), we can transform the numerator in the double sum into n o (−1)n+m Re W|nihm| (α) = fnm (|α|) Re e −i(n−m)(β+π) = Re W|nihm| (−ᾱ) . Here, the abbreviation fnm (|α|) denotes the absolute value of W|nihm| , which only depends on |α|. When we now use the p-symmetry of the Moyal functions and the symmetries of the first sum, we see that the Wigner function of the reference state is given by W|φiN (x, p) = W|σ,0iN (−ᾱ) = W|σ,0iN (−x, p) . (5.11) Hence, W|φiN is the opposite hand of the Wigner function of the initial truncated Riemann state |σ, 0iN with respect to the p-axis, as the upper left picture of fig. 5.2 confirms. 36 5.3. Truncated alternating sum Finally, we can represent the overlap (5.8) in terms of the Wigner functions by the expression 2 2 |C(t)| = |N hφ|σ, ωtiν | = 2π~ Z∞ −∞ dx Z∞ dp W|σ,ωtiν (x, p) W|σ,0iN (−x, p) . (5.12) −∞ It is depicted in blue in the upper right picture of fig. 5.2 for τ = ωt. The red dashed (100) line indicates |Ma (2/3, τ ) ζ(2/3 + iτ )|2 to emphasize the good agreement between the approximation by the truncated sum ζ100 and the exact values of ζ. The right pictures below show the product W|φiN W|σ,τ iN of the Wigner functions for three different times. They have a central structure which is shifted to the left due to the mirrored shape of W|φiN . The surrounding interference fringes are created by the tails of the Wigner functions. When we integrate the data over phase space according to eq. (5.12), we get the values of |C|2 marked in the upper right picture by dots. 5.3.3. Moyal representation According to eq. (3.9), the overlap C(t) = N hφ|σ, ωtiν , eq. (5.8), is equal to the Moyal function, eq. (3.13), "N −1 X (−1)n+1 e −iτ ln(n+1) W|ni W|σ,τ iν N hφ| = Nν (σ) NN (σ) (n + 1)σ n=0 # N −1 n−1 X X (−1)m+1 e −iτ ln(n+1) W|nihm| + (−1)n+1 e −iτ ln(m+1) W |nihm| + [(n + 1)(m + 1)]σ/2 n=1 m=0 (5.13) of the truncated Riemann state |σ, τ iν at τ = ωt and the reference state |φiN integrated over the whole phase space. Here, the choice N ≤ ν again truncates the sums over n at N − 1. As in the case of the Riemann states, both sums of the Moyal function contain time-dependent factors, but now additionally multiplied by a phase factor ±1. We show in the bottom of fig. 5.5 the real and imaginary part as well as the absolute value of the Moyal function W|σ,τ iν hφN | for σ = 2/3, N = ν = 100 and different values of τ . The shapes for τ = 0 are all symmetric in x and additionally symmetric in p in the case of the real part and the absolute value, while the imaginary part is antisymmetric in p. For larger times, these symmetries get lost since the truncated Riemann state |σ, τ iν starts to rotate. The picture on top presents the dependence of the overlap C(t), eq. (5.8), on τ . The values marked by dots emerge from the pictures by integration over phase space as defined in eq. (3.9). 37 5. Alternating sum and truncated Riemann states Figure 5.5.: In the bottom pictures the Moyal function W|2/3,τ i100 hφ| , eq. (5.13), is depicted for ν = N = 100 and different values of τ by its real and imaginary part as well as by its absolute value. For τ = 0, we find structures which are symmetric in x and in the case of the real part and absolute value additionally symmetric in p, while the initial imaginary part is antisymmetric in p. All symmetries get lost for τ > 0 due to the rotation of the truncated Riemann state |2/3, τ i100 . To make the main structure more visible, we have cut the outer regions where only small interference fringes occur. Integration of W|2/3,τ i100 hφ| over phase space gives the values marked in the upper picture due to eq. (3.9) which connects the overlap of the states with the Moyal representation. 38 5.4. Summary 5.4. Summary This chapter has demonstrated that the exact version of the alternating sum can only be produced by quantum states with σ > 1. Indeed, if we use only one quantum system, the normalization of the states forbids photon statistics with σ ≤ 1. Even combinations of two entangled states fail, since one of them is, due to its normalization, restricted to the same region. To circumvent this problem, we have truncated the alternating sum. The resulting states are even for σ ≤ 1 normalizable. However, the quality of the approximation with the truncated alternating sum ζN depends strongly on the used parameters. Especially in the critical strip, we need for large imaginary parts τ large truncation limits N to get adequate results. But large values of N produce small normalization constants for the (ν,N ) states which reduce the value of the proportionality factor Ma and therefore make the overlap C tend to zero. Hence, we dedicate the next chapters to the investigation of the more suitable approximations of ζ to enter the critical strip: the Riemann-Siegel and Berry-Keating formula. 39 6. Riemann-Siegel states In the previous chapters, we have shown that the Dirichlet representation of the Riemann zeta function leads to states only normalizable for σ > 1. Even the states which produce the alternating sum exactly suffer from the same restriction in the physical picture. Only the truncated Riemann states are able to enter the critical strip. However, the accuracy of the corresponding approximation of ζ by the truncated alternating sum strongly depends on the chosen parameters. For large imaginary parts τ , we need quite large truncation limits N , which result in small probability amplitudes and therefore in a very small overlap of the states. Besides, the functional equation (1.5) is of no use either, since the above expressions only expand the region to σ < 0, missing again the critical strip. Hence, we employ now the Riemann-Siegel formula ζRS which provides a good approximation of ζ in the critical strip, especially for large imaginary parts τ . In contrast to the alternating sum, the summation limit n0 is quite small and consequently, the resulting amplitudes of the states are much larger. Yet, we will see that the states have to be entangled to reproduce the correct phase dependence of the summands in ζRS. In the following, we restrict our investigation to positive σ. The values of ζ for negative σ can be found with the functional equation (1.5). 6.1. Definition of the states We start from the main part of the Riemann-Siegel formula [29] n0 n0 X X 1 1 ζRS(s) ≡ + χ(s) , s 1−s n n n=1 (6.1) n=1 where χ(s) and n0 are defined by eqs. (1.6) and (1.10), respectively. When we use again s = σ + iτ and shift the index of the sum in eq. (6.1), we arrive at ζRS(σ + iτ ) = nX 0 −1 n=0 " e iτ ln(n+1) e −iτ ln(n+1) + χ(σ + iτ ) (n + 1)σ (n + 1)1−σ # . (6.2) The second sum accumulates phases with opposite sign with respect to the first sum. As we mentioned in chapter 3, the same combination of phases occurs in the JaynesCummings-Paul model which describes the interaction of a two-level atom with a single mode of the cavity field. Following this example, we can construct the states producing the Riemann-Siegel formula. 41 6. Riemann-Siegel states 6.1.1. Riemann-Siegel state First, we choose as oscillator state the truncated Riemann state |σ, τ iν , eq. (5.4), and√an equal distribution between the atomic probability amplitudes, that is ce = cg = 1/ 2. The initial state then reads 1 |ΨRS(σ, 0)i ≡ |σ, 0iν ⊗ √ |ei + |gi 2 according to eq. (3.1). Time evolution with ĤR yields the state 1 |ΨRS(σ, τ )i ≡ √ |σ, τ iν |ei + |σ, −τ iν |gi 2 (6.3) which consists of two counter-rotating truncated Riemann states, |σ, τ iν and |σ, −τ iν , each entangled to one part of the two-level atom. Since this state contains the desired phases ±iτ ln(n + 1), we call it Riemann-Siegel state. Our next task is to choose the appropriate reference state. 6.1.2. Riemann-Siegel reference state Now, we turn to the examination of the scalar product CRS(t) ≡ hΦRS|ΨRS(σ, ωt)i = MRS ζRS(σ + iωt) (6.4) to identify the probability amplitudes of the reference state |ΦRSi defined by eq. (3.5). The proportionality factor MRS contains the normalizations of the states. With ζRS and |ΨRSi given by eqs. (6.2) and (6.3), respectively, we find ν−1 Nν (σ) X CRS(t) = √ 2 n=0 = MRS φ ne n0 (τ )−1 X n=0 1 1 e −iωt ln(n+1) + φ ng e iωt ln(n+1) σ/2 (n + 1) (n + 1)σ/2 1 1 e −iτ ln(n+1) + χ(σ + iτ ) e iτ ln(n+1) σ (n + 1) (n + 1)1−σ . (6.5) At first, we note that the range of the summation index n is different on both sides of eq. (6.5). The sum on the left-hand side can only become equal to the one on the right if n0 (τ ) ≤ ν , (6.6) since then we can choose for the probability amplitudes of the reference state φne = φng = 0 for n ≥ n0 , which truncates the sum on the left at n0 . It is crucial to keep in mind that the right-hand side of eq. (6.5) and, therefore, especially n0 depends on τ ≡ Im s. Then, we compare the exponential functions on both sides and find additionally the condition τ = ωt, which, in contrast to the previous chapters, cannot be identified as 42 6.2. Entanglement a rescaled time. Indeed, due to eq. (6.6), the identification of τ as rescaled time would imply that the truncation parameter ν of the Riemann-Siegel state is time-dependent. However, eq. (6.6) only requires that ν is not smaller than the constant n0 chosen for the parameter τ ≡ Im {σ + iωt}. Finally, we find that the reference state |ΦRS(σ, τ )i = Nn0 (σ, τ ) |σ, 0in0 |ei + γ (σ, τ ) |2 − 3σ, 0in0 |gi (6.7) with fulfills eq. (6.4) if 1 Nn0 (σ, τ ) ≡ q 1 + |γ(σ, τ )|2 γ(σ, τ ) ≡ and thus MRS ≡ (6.8) Nn0 (σ) χ(σ + iτ ) Nn0 (2 − 3σ) (6.9) Nν (σ) √ Nn0 (σ) Nn0 (σ, τ ) . 2 6.2. Entanglement Before we turn to the investigation of the Wigner matrix elements, we analyze if the states |ΨRSi and |ΦRSi are entangled as we assume by their construction. The methods of testing for entanglement are described in appendix E. 6.2.1. Riemann-Siegel state For the Riemann-Siegel state, eq. (6.3), the probability amplitudes fulfill ψng = ψ ne , which simplifies eq. (E.7) to 2 ν−1 X 2 ψne ≤ n=0 ν−1 X n=0 |ψne |2 !2 . Moreover, ψne is of the form rn e −iθn leading to ν−1 ν−1 X X 2 −2iθn ≤ rn2 . r e n n=0 n=0 Hence, the sum on the left-hand side can only be the same as the right one if the phases fulfill ! θn ≡ τ ln(n + 1) = 2πk with k∈N ∀ n = 0, ..., ν . This is of course the case for τ = 0, since the initial state is a product state, or if ν = 1, but it cannot be fulfilled for ν ≥ 2 and τ 6= 0 for all n at once. 43 6. Riemann-Siegel states 6.2.2. Riemann-Siegel reference state Now, we use the overlap of the field states, eq. (E.3), to analyze the entanglement of the reference state |ΦRSi directly. From representation (6.7) follows that 2 nX 0 −1 1 |n0 hσ, 0|2 − 3σ, 0in0 |2 = Nn0 (σ) Nn0 (2 − 3σ) ≤1 (n + 1)1−σ n=0 or nX 0 −1 n=0 1 (n + 1)1−σ !2 ≤ nX 0 −1 n=0 nX 0 −1 1 1 σ (n + 1) (m + 1)2−3σ (6.10) m=0 when we use the definition (5.5) of the normalization Nn0 . In appendix E.4, we show that for n0 ≥ 2 equality in eq. (6.10) is only given if σ = 1/2. Hence, the reference state |ΦRSi is only for σ = 1/2 disentangled, hinting the peculiarity of the critical line. We note that we can also investigate the entanglement by determining the Schmidt decomposition of the density matrix ρ̂ ≡ |ΦRSihΦRS| as described in appendix E.2. However, there is no obvious connection between the eigenvalues of the reference state and the location of the zeros of the Riemann zeta function. 6.3. Wigner matrix of the Riemann-Siegel state As we have seen in section 3.3.2, the Wigner representation of an entangled state is given by the Hermitian Wigner matrix, eq. (3.12), which reads for the Riemann-Siegel state, eq. (6.3), 1 Ŵ|ΨRSi ≡ W|σ,τ iν |eihe| + W|σ,−τ iν |gihg| + W|σ,τ iν hσ,−τ | |eihg| + W|σ,−τ iν hσ,τ | |gihe| . 2 (6.11) Hence, the diagonal matrix elements are given by the Wigner functions of the truncated Riemann state divided by two, shown in fig. 5.1 and on the left in fig. 5.2. As expected, the diagonal matrix element connected to the excited state |ei rotates clockwise whereas the one connected to the ground state |gi rotates counter-clockwise in phase space according to W|σ,−τ iν (x, p) = W|σ,τ iν (x, −p) . (6.12) This behavior is shown in the first and second row of fig. 6.1 for σ = 1/2 and ν = 10. Needless to say, for τ = 0 the diagonal matrix elements are equal to each other and symmetric in p. Now, we turn to the off-diagonal matrix elements in eq. (6.11) which are given by the 44 6.3. Wigner matrix of the Riemann-Siegel state Figure 6.1.: Time evolution manifests itself in the matrix elements of the Wigner matrix Ŵ|ΨRS i of the Riemann-Siegel state in different ways. Caused by the rotation of the corresponding Wigner function W|σ,±τ iν of the truncated Riemann state, the diagonal matrix elements, shown in the two upper rows, rotate clockwise or counter-clockwise. In contrast, the off-diagonal matrix element, which is proportional to the Moyal function W|σ,τ iν hσ,−τ | , remains symmetric in p but changes its inner structure. The pictures display the matrix elements for σ = 1/2 and ν = 10. At τ = 0, all matrix elements are the same and proportional to W|1/2,0i10 already shown in fig. 5.1. 45 6. Riemann-Siegel states Moyal function W|σ,τ iν hσ,−τ | = Nν2 (σ) " ν−1 X e −2iτ ln(n+1) (n + 1) σ W|ni n=0 ν−1 X n−1 X # e −iτ ln[(n+1)(m+1)] Re W|nihm| . +2 [(n + 1)(m + 1)] σ/2 n=1 m=0 (6.13) In contrast to the Wigner functions W|σ,τ iν , eq. (5.6), the time-dependence acts in both sums as factor and the Moyal functions of the Fock states are not rotated. Hence, the initial symmetry of W|σ,0iν hσ,0| ≡ W|σ,0iν in p is conserved for all times τ , that is W|σ,τ iν hσ,−τ | (x, −p) = W|σ,τ iν hσ,−τ | (x, p) . (6.14) Moreover, for τ 6= 0 and ν > 1, the complex factors in both sums suppress or even produce negative contributions of W|ni and W|nihm| to the real or imaginary part of the Moyal function W|σ,τ iν hσ,−τ | . This leads to different distributions of positive and negative regions in the patterns shown in the two bottom rows of fig. 6.1 compared to the structures of the Wigner functions depicted above. Here, σ = 1/2 and ν = 10. It remains to mention that for larger values of σ the internal structures of the patterns of the Wigner matrix elements are less distinct since, due to the photon statistics, the contributions for higher n are smaller. 6.4. Wigner matrix of the reference state The Wigner matrix, eq. (3.12), of the Riemann-Siegel reference state |ΦRSi, eq. (6.7), is given by Ŵ|ΦRSi ≡ Nn20 W|σ,0in0 |eihe| + |γ|2 W|2−3σ,0in0 |gihg| +γ W|σ,0in0 h2−3σ,0| |eihg| + γ W|2−3σ,0in0 hσ,0| |gihe| . (6.15) Since the diagonal matrix elements are proportional to the Wigner functions of the initial truncated Riemann states |σ, 0in0 and |2 − 3σ, 0in0 , eq. (4.11), they are always symmetric in p. Moreover, the shapes of W|σ,0in0 and W|2−3σ,0in0 differ only in height as long as the condition n0 (τ1 ) = n0 (τ2 ) is fulfilled for different values τ1 and τ2 , because then the dependence on τ only appears in the factors Nn20 and |γ|2 Nn20 . The two upper rows of fig. 6.2 show the diagonal matrix elements of Ŵ|ΦRSi for τ = 700 (n0 = 10) at σ = 1/3 and σ = 1/2. Due to |γ| = 1 for σ = 1/2, the diagonal matrix elements are the same on the critical line. In contrast, the structures for σ = 1/3 are only similar to each other. Although the values of the Wigner function W|1,0in0 are smaller than the ones of W|1/3,0in0 , the matrix element connected to the ground state |gi is more pronounces because of the larger factor |γ|2 Nn20 . A detailed analysis of the normalization factors is given in appendix F.5. 46 6.4. Wigner matrix of the reference state Figure 6.2.: Wigner matrix elements of the Riemann-Siegel reference state at τ = 700 (n0 = 10) for σ = 1/3 and σ = 1/2. The patterns of the diagonal matrix elements depicted in the first and second row are always symmetric in p. For each σ, their shapes change only in height if n0 is the same for different τ . However, due to the factor |γ|2 Nn20 , the diagonal matrix element connected to the ground state is more pronounced for σ < 1/2 than the one connected to the excited state. The pictures in the two bottom rows show the real and imaginary part of the off-diagonal Wigner matrix element which is connected to the Moyal functions W|σ,0in0 h2−3σ,0| . The patterns are only on the critical line symmetric in p. For σ = 1/3 they are nearly symmetric or anti-symmetric, respectively, since the phase of γ is equal to arg χ ≈ 0.89 π. 47 6. Riemann-Siegel states The two bottom rows of fig. 6.2 present the real and imaginary part of the off-diagonal matrix elements of Ŵ|ΦRSi , which are connected to the Moyal functions, eq. (3.10), W|σ,0in0 h2−3σ,0| ≡ Nn0 (σ) Nn0 (2 − 3σ) nX 0 −1 W|nihm| (n + 1)σ/2 (m + 1)(2−3σ)/2 n,m=0 (6.16) via the factor |γ| e ±i arg γ Nn20 . Hence, on the critical line they are – apart from the phase of the factor – the same as the diagonal matrix elements and therefore symmetric in p. Thus, all matrix elements differ for σ = 1/2 only in height but not in shape. However, for σ 6= 1/2, the shapes of the off-diagonal matrix elements have in general no symmetry in p since the phase ± arg γ = ± arg[χ(σ + iτ )], defined by eq. (6.9), rotates the Moyal functions W|nihm| of the Fock states. Only for the special case arg χ = πk with k ∈ Z, the real part of the off-diagonal matrix elements is symmetric in p whereas the imaginary part is antisymmetric. These properties are reversed for arg χ = (2k + 1)π/2. Since arg χ(1/3 + 700 i) ≈ 0.89 π, the bottom pictures of fig. 6.2 for σ = 1/3 are nearly symmetric or antisymmetric, respectively. 6.5. Overlap in Wigner representation Now, we have all ingredients to determine the overlap Z∞ Z∞ 2 |CRS| = 2π~ dx dp KRS = MRS |ζRS(σ + iωt)|2 2 −∞ (6.17) −∞ in Wigner representation from eq. (3.15) and, with the help of eq. (6.4), the absolute value of the Riemann-Siegel representation ζRS. The kernel, eq. (3.16), KRS ≡ Nn20 h W|σ,τ iν W|σ,0in0 + |γ|2 W|σ,−τ iν W|2−3σ,0in0 2 n oi + 2 Re γ W|σ,τ iν hσ,−τ | W |σ,0in0 h2−3σ,0| (6.18) follows from the definitions (6.3) and (6.7) of the Riemann-Siegel state and its reference state. Here, we have combined the off-diagonal matrix elements using that the Wigner matrix is Hermitian. A closer look on KRS reveals that it is only on the critical strip symmetric in p. This becomes evident when we recall the properties of the matrix elements of the RiemannSiegel state and its reference state: The diagonal matrix elements of the Riemann-Siegel state are connected via the relation (6.12) while the diagonal matrix elements of the reference state are symmetric in p and even the same for σ = 1/2. Since the prefactors are also the same on the critical line, the sum in the first row of eq. (6.18) is symmetric in p for σ = 1/2. The off-diagonal matrix elements of the Riemann-Siegel state are always symmetric in p, eq. (6.14), whereas this holds only for σ = 1/2 in the case of the reference state. Hence, KRS is only on the critical line symmetric in p. 48 6.5. Overlap in Wigner representation Figure 6.3.: The upper pictures show |CRS|2 , eq. (6.17), as black solid line for σ = 1/3 and σ = 1/2 with dots marking the values at τ1 ∼ = 681.21 and τ2 ∼ = 681.89. Here, τ2 is chosen at a zero of ζ. 2 The red dashed line indicates the exact curve MRS |ζ|2 . In the bottom pictures KRS, eq. (6.18), is depicted for the different combinations of σ = 1/3 and σ = 1/2 with τ1 and τ2 . The shapes are for σ = 1/2 (two right pictures) symmetric in p and the positive areas clearly dominate for τ1 , whereas at τ2 the positive and negative contributions cancel each other when integrated over phase space, indicating the zero of ζ. Although the structures for σ = 1/3 (two left pictures) are less pronounced than for σ = 1/2, the positive contributions dominate in both pictures as tab. 6.1 confirms. For the sake of simplicity, we have chosen ν = n0 = 10. In the bottom row of fig. 6.3, we present KRS for σ = 1/3 and σ = 1/2 with ν = n0 = 10 for two different values of τ . Since τ2 is chosen at a zero of ζ, the positive and negative areas in the fourth picture cancel each other quite well when we integrate over phase space. This is also confirmed by the analysis of the data in tab. 6.1. In the other pictures, the positive areas dominate. Although the structures for σ = 1/3 are less pronounced, the values calculated in tab. 6.1 for τ1 are nearly the same for both σ, but for τ2 the value for σ = 1/3 is about 25-times larger. This strongly indicates that ζ(1/3 + iτ2 ) cannot be zero. The overlap |CRS|2 , defined by eq. (6.17), is depicted as black solid line in the pictures above with dots marking the values at τ1 and τ2 . Due to the approximate character of the Riemann-Siegel formula, the black line differs from the red dashed line which 2 |ζ(σ + iτ )|2 . indicates the exact curve MRS 49 6. Riemann-Siegel states σ τ1 τ2 1/3 R R 2π dx dp KRS 2 MRS |ζRS|2 |ζ|2 0.151 0.00398 37.80 39.99 1/2 0.154 0.00992 15.54 16.53 1/3 0.00374 0.00398 0.9397 0.8267 1/2 0.000149 0.00992 0.01499 0 Table 6.1.: The values in the middle columns result from the data shown in the phase space pictures of fig. 6.3. The square of the absolute value |ζRS|2 can either be calculated from |CRS|2 , eq. (6.17), or directly with the Riemann-Siegel formula (6.1). The integration over KRS results for τ1 in similar values for both σ. However, the value at τ2 is for σ = 1/3 about 25-times larger than for σ = 1/2, coinciding with the fact that the negative and positive contributions cancel each other quite well in the phase space picture of KRS for σ = 1/2 at τ2 . Due to the approximate character of the Riemann-Siegel formula, the integral does not vanish entirely, causing the deviations to the exact values |ζ|2 . In the next section, we apply the definition of the Moyal functions to get a complexvalued expression for the overlap CRS itself. 6.6. Overlap in Moyal representation We have shown in section 3.3.1 that we can express the overlap (3.11) of two entangled states of the form (3.5) in terms of the Moyal functions of the oscillator parts. Hence, the complex-valued expression (6.4) is given by the phase space integration Nn CRS = √ 0 2 Z∞ −∞ dx Z∞ −∞ dp K RS = MRS ζRS (6.19) over the kernel K RS ≡ i Nν (σ) h W|σ,τ in0 hσ,0| + γ(σ, τ ) W|σ,−τ in0 h2−3σ,0| Nn0 (σ) (6.20) of two Moyal functions of the form W|σ1 ,τ in0 hσ2 ,0| = Nn0 (σ1 ) Nn0 (σ2 ) nX 0 −1 e −iτ ln(n+1) W|nihm| . σ1 /2 (m + 1)σ2 /2 (n + 1) n,m=0 (6.21) Here, we have used the property Z∞ −∞ dx Z∞ −∞ dp W|σ1 ,τ iν n0 hσ2 ,0| Nν (σ1 ) = Nn0 (σ1 ) Z∞ −∞ dx Z∞ dp W|σ1 ,τ in0 hσ2 ,0| (6.22) −∞ of the Moyal function of two truncated Riemann states with n0 ≤ ν. This equation reflects the fact that the Fock states with ν > n0 do not contribute to the overlap 50 6.6. Overlap in Moyal representation n0 hσ2 , 0|σ1 , τ iν . The truncation parameter ν only appears in the normalization Nν (σ1 ) of the state |σ1 , τ iν and therefore acts as a factor. Moreover, when we rewrite eq. (6.21) by "n −1 0 X e −iτ ln(n+1) W|ni W|σ1 ,τ in0 hσ2 ,0| = Nn0 (σ1 ) Nn0 (σ2 ) (n + 1)(σ1 +σ2 )/2 n=0 # nX 0 −1 n−1 X e −iτ ln(n+1) W|nihm| + e −iτ ln(m+1) W |nihm| + , σ1 /2 (m + 1)σ2 /2 (n + 1) n=1 m=0 we find that the Wigner functions W|ni in the first sum are multiplied by a timedependent factor, like in the case of the Moyal function W|σ,τ iν hσ,−τ | , eq. (6.13), connected to the off-diagonal Wigner matrix elements of the Riemann-Siegel state. In contrast to eq. (6.13), the time-dependent factors in the double sum result here in a rotation of the Moyal functions W|nihm| . This behavior is well-known from the rotation of the Wigner function W|σ,τ iν , eq. (5.6), but now the rotation angle is ϕn0 = τ ln(n + 1), eq. (4.14), which depends only on one summation index. Since W|nihm| and its complex conjugated are rotated by different angles, W|σ1 ,τ in0 hσ2 ,0| has in general no symmetries. We show in fig. 6.4 the real and imaginary part of the two contributions W|σ,τ in0 hσ,0| and W|σ,−τ in0 h2−3σ,0| of K RS, eq. (6.20), for σ = 1/3 and n0 = 10 at four different times. Besides the times τ1 and τ2 , which are the same as in fig. 6.3, we depict also the Moyal functions at τ = 680 and τ = 683 to give a flavor of the rotation. As expected, W|1/3,τ i10 h1/3,0| rotates clockwise whereas W|1/3,−τ i10 h1,0| rotates counter-clockwise due to the different signs of τ . With the help of eq. (6.9) and eq. (6.21), the kernel (6.20) reads " # nX 0 −1 W|nihm| e −iτ ln(n+1) χ(σ + iτ ) e iτ ln(n+1) K RS = Nν (σ) Nn0 (σ) + (6.23) σ/2 (2−3σ)/2 (m + 1) (m + 1) (n + 1)σ/2 n,m=0 in terms of Fock states. Although each contributing Wigner function is in general not symmetric in p, we find for σ = 1/2 the connection K RS (x, −p) = e i arg{χ(1/2+iτ )} K RS (x, p) (6.24) since W|nihm| (x, −p) = W|nihm| (x, p), eq. (4.13). Hence, the absolute value of K RS is on the critical line symmetric in p, as fig. 6.5 confirms, where we depict K RS by its real and imaginary part (upper rows) as well as its absolute value and phase (bottom rows) for the same values of σ and τ as in fig. 6.3. The dominant positive or negative areas are produced by the factor γ, eq. (6.9), in front of the Moyal function W|σ,0in0 h2−3σ| . Its real part is at τ1 for both values of σ negative while the imaginary part almost vanishes. For τ2 , the real part is positive and the imaginary part negative and about 10-times smaller than the real one. Finally, we get from eq. (6.19) the representation 1 ζRS = 2 Nn0 (σ) Z∞ −∞ dx Z∞ −∞ h i dp W|σ,τ in0 hσ,0| + γ(σ, τ ) W|σ,−τ in0 h2−3σ,0| (6.25) 51 6. Riemann-Siegel states Figure 6.4.: Real and imaginary part of the Moyal functions W|σ,τ in0 hσ,0| and W|σ,−τ in0 h2−3σ,0| , eq. (6.21), for σ = 1/3, n0 = 10 and different values of τ . Here, τ1 and τ2 are the same as in fig. 6.3. As expected, W|σ,τ in0 hσ,0| (first and third row) rotates clockwise whereas the other Moyal function rotates counter-clockwise. 52 6.6. Overlap in Moyal representation Figure 6.5.: The kernel K RS of the Moyal representation, eq. (6.23), is depicted by its real and imaginary part (top rows) as well as absolute value and argument (bottom rows) for ν = n0 = 10 and the same values of σ and τ used in fig. 6.3. In general, the structures of K RS have no symmetries, but the absolute value is for σ = 1/2 symmetric in p, as eq. (6.24) confirms. The structures here have more dominant positive or negative areas than the corresponding Moyal functions in the pictures of fig. 6.4, since W|σ,−τ in0 h2−3σ,0| is for τ1 additionally multiplied with the (almost purely) negative factors γ(1/3, τ1 ) ∼ = −1.5 and γ(1/2, τ1 ) ∼ = −1 and for τ2 by the ∼ complex values γ(1/3, τ2 ) = 1.5 − 0.13 i and γ(1/2, τ2 ) ∼ = 1 − 0.09 i. Moreover, K RS is connected to the Riemann-Siegel formula via eq. (6.25). The explicit values calculated from the data are given in tab. 6.2. 53 6. Riemann-Siegel states σ τ1 1/3 R R dx dp K RS −0.2568 − 0.9548 i Nn−2 (σ) 0 6.220 0.9887 e −1.8335 i 1/2 1/3 −8 · 10−6 − 0.7854 i 5.021 0.0707 + 0.1388 i 6.220 0.1558 1/2 0.0242 e −0.04321 i −1.597 − 5.939 i −1.597 − 6.119 i −4 · 10−5 − 3.944 i −4 · 10−5 − 4.066 i 0.4398 + 0.8631 i 0.2607 + 0.8710 i 3.944 e −1.5708 i e 1.0995 i 0.0241 − 0.00104 i ζ 6.150 e −1.8335 i 0.7854 e −1.5708 i τ2 ζRS 0.9687 5.021 e 1.0995 i 0.1212 − 0.00524 i 0.1214 e −0.04321 i 6.324 e −1.8261 i 4.066e −1.5708 i 0.9092 e 1.2800 i 0 0 Table 6.2.: The values in the middle columns are calculated from the data displayed in the phase space pictures of fig. 6.5. The explicit connection between ζRS and the integral over K RS is given by eq. (6.25). Due to the approximate character of ζRS, the calculated values deviate from the exact ones given in the last column. for the Riemann-Siegel formula in terms of the Moyal functions. The dependence on the arbitrary chosen summation limit ν vanishes completely. Thus, the limit ν = n0 for the Riemann-Siegel state suffices to produce ζ. In tab. 6.2, we present the values of ζRS calculated by the data shown in the phase space pictures of fig. 6.5. Additionally, we depict the behavior of the real and imaginary part as well as the absolute value and phase of ζRS in fig. 6.6. The blue dashed curves for σ = 1/3 exceed the curves of σ = 1/2 given in red. Moreover, the phase is discontinuous at the zeros of ζ on the critical line, confirmed by the Argand diagram [9] shown in fig. 6.7. In both pictures, we have marked the values calculated in tab. 6.2 by dots. 6.7. Summary We have shown in this chapter that the Riemann-Siegel formula (6.1) can be expressed by the scalar product (6.4) of the Riemann-Siegel state |ΨRSi, eq. (6.3), and its reference state |ΦRSi, eq. (6.7). Due to time evolution, the Riemann-Siegel state becomes entangled while the reference state is entangled for all σ 6= 1/2. Hence, only σ = 1/2 produces p-symmetric phase space pictures for the kernel KRS of the Wigner representation, eq. (6.18), and quasi-symmetric structures for the Moyal representation K RS, eq. (6.24). The physical picture therefore emphasizes the importance of the critical line. To enhance the accuracy of the calculated values, we discuss the Berry-Keating formula, which already contains the first correction term to ζRS, in the next chapter. 54 6.7. Summary Im {ζRS} Re {ζRS} 8 6 4 2 680 682 680 684 682 τ 684 -2 -4 -6 |ζRS| arg{ζRS} Π 8 Π 2 6 4 680 2 - 680 682 τ 684 682 684 τ Π 2 -Π Figure 6.6.: Dependence of ζRS, eq. (6.1), on τ shown by its real and imaginary part as well as absolute value and argument for σ = 1/3 (blue dashed line) and σ = 1/2 (red solid line). The argument is discontinuous at the zeros of ζ, as the Argand diagram in fig. 6.7 shows. The values marked by dots are the ones calculated in tab. 6.2. Im ζ 6 4 2 -2 2 4 6 8 Re ζ -2 -4 -6 Figure 6.7.: The Argand diagram of ζ confirms that there is no zero for σ = 1/3 (blue dashed line) but one for σ = 1/2 (red curve) in the interval 680 ≤ τ ≤ 683.5. The curves start at values with positive imaginary part and proceed in clockwise direction, indicated by the arrows. The dots mark the values calculated in tab. 6.2. 55 7. Berry-Keating reference states In section 1.6, we have shown that the approximation of the Riemann zeta function can be improved by using the Berry-Keating formula ζBK. Since it is similar to the RiemannSiegel formula, we can produce ζBK from the overlap of the Riemann-Siegel state with the Berry-Keating reference state, which we derive in this chapter. The analysis of the corresponding Wigner and Moyal representations confirm the enhanced accuracy of the approximation. 7.1. Definition With definition (1.7) of the symmetric representation, the Berry-Keating representation of Z, eq. (1.12), yields " # nX B −1 e −iτ ln(n+1) e iτ ln(n+1) ζBK(σ + iτ ) = E(n, τ ) + χ(σ + iτ ) E (n, τ ) (7.1) (n + 1)σ (n + 1)1−σ n=0 the Berry-Keating formula of the Riemann zeta function. The abbreviation E(n, τ ), given by eq. (1.13), contains the complementary error function [50] while the summation limit fulfills nB ≥ n0 . For the sake of simplicity, we choose nB = n0 (τ ) whenever we compare the Berry-Keating representations to the Riemann-Siegel ones. Since eq. (7.1) is reminiscent of the Riemann-Siegel formula (6.2), we use the RiemannSiegel state |ΨRS(σ, τ )i, eq. (6.3), to calculate the overlap CBK ≡ hΦBK(σ, τ )|ΨRS(σ, τ )i = MBK ζBK(σ + iτ ) . Comparison of both sides leads to the definition en (σ, τ ) |φ (σ, τ )in |ei + γ (σ, τ ) |φ (2 − 3σ, τ )in |gi |ΦBK(σ, τ )i ≡ N B B B B B B (7.2) (7.3) of the Berry-Keating reference state, where we have introduced the Berry states |φB(σ, τ )inB en (σ, τ ) ≡N B nX B −1 n=0 E (n, τ ) |ni (n + 1)σ/2 (7.4) with n0 ≤ nB ≤ ν and the normalization en (σ, τ ) ≡ N B "n −1 #−1/2 B X |E(n, τ )|2 n=0 (n + 1)σ . (7.5) 57 7. Berry-Keating reference states The Berry states |φBi are given by eq. (7.4) with E replaced by its complex conjugate E. In analogy to the normalization of the Riemann-Siegel reference state, eq. (6.8), the normalization of the Berry-Keating reference state reads but now the factor 1 en (σ, τ ) ≡ q , N B 2 1 + |γB(σ, τ )| γB(σ, τ ) ≡ en (σ, τ ) N B en (2 − 3σ, τ ) N B χ(σ + iτ ) (7.6) (7.7) en of the Berry states, eq. (7.5). As is involved, which depends on the normalizations N B in the case of the Riemann-Siegel reference state, we have to interpret τ in the definition (7.3) of the reference state |ΦBKi as the imaginary part of s and not as time. Finally, we identify the proportionality factor MBK ≡ Nν (σ) e en (σ, τ ) . √ NnB (σ, τ ) N B 2 In contrast to the last chapter where the Riemann-Siegel reference state consists of the truncated Riemann state |σ, 0in0 and |2−3σ, 0in0 defined by eq. (5.4), the Berry-Keating reference state |ΦBKi depends on the Berry states |φB(σ, τ )inB and |φB(2−3σ, τ )inB which contain additionally the n-dependent factors E(n, τ ) and E(n, τ ) in their probability amplitudes. Due to these factors, the Berry-Keating reference state is always entangled while the Riemann-Siegel reference state, eq. (6.7), becomes separable on the critical line, as we show in the appendices E.4 and E.5. Since we already gave a detailed analysis of the Wigner matrix Ŵ|ΨRSi of the RiemannSiegel state in section 6.3, we restrict ourselves to the investigation of the Wigner representations of the Berry states and the Berry-Keating reference state in the next sections. Afterwards, we describe the overlap in Wigner and Moyal representation which produces the Berry-Keating formula (7.1). 7.2. Wigner function of Berry state According to eq. (3.14), the Wigner functions of the Berry states, eq. (7.4), are given by "n −1 B X |E(n, τ )|2 W|ni 2 e W|φB(σ,τ )inB ≡ NnB (σ, τ ) (n + 1)σ n=0 # nX B −1 n−1 X Re E (n, τ ) E(m, τ ) W|nihm| +2 . (7.8) [(n + 1)(m + 1)]σ/2 n=1 m=0 Since the values of E are equal to one for almost all n < n0 , as shown in appendix F.2, the shapes are reminiscent of the ones for the Wigner functions of the initial truncated Riemann state |σ, 0inB defined by eq. (5.6). Nevertheless, the τ -dependence of E creates 58 7.2. Wigner function of Berry state Figure 7.1.: Difference W|φB(σ,τ )in0 −W|σ,0in0 between the Wigner functions, eqs. (7.8) and (5.6), of the Berry state |φBin0 and the initial truncated Riemann state |σ, 0in0 for σ = 1/2 and different τ determining the summation limit n0 = n0 (τ ). At the jumps 2πn20 of the RiemannSiegel formula, we find the largest differences between the two states. They become smaller in the interval 2π n20 ≤ τ < 2π(n0 + 1)2 and rotate due to the τ -dependent factors E(n, τ ) in W|φB(σ,τ )in0 . Moreover, the deviations fade for increasing summation limit. small deviations which make W|φBinB asymmetric in p. To emphasize the inequality of the two states, we depict the difference W|φB(1/2,τ )in0 −W|1/2,0in0 between the Wigner function of the Berry state |φBin0 and the initial truncated Riemann state |σ, 0in0 , eqs. (7.8) and (5.6), in fig. 7.1 for different values τ and summation limits n0 (τ ) up to four. The pictures show that the deviations are distinct at the jumps τ = 2π n20 of the RiemannSiegel formula and fade in the interval 2πn20 ≤ τ < 2π(n0 + 1)2 before the next jump. Of course, increasing summation limits result in more complicated, but less pronounced structures. The same properties hold for the Berry state |φBinB . Indeed, when we recall that W|nihm| (x, −p) = W|nihm| (x, p), a closer look at the Wigner function W|φ (σ,τ )inB B ≡ en2 (σ, τ ) N B "n −1 B X |E(n, τ )|2 W|ni (n + 1)σ n=0 # nX B −1 n−1 X Re E(n, τ ) E (m, τ ) W|nihm| +2 [(n + 1)(m + 1)]σ/2 n=1 m=0 reveals that it is the opposite hand of W|φB(σ,τ )inB with respect to the x-axis, that is W|φ (σ,τ )inB (x, p) B = W|φB(σ,τ )inB (x, −p) . (7.9) This relation makes the Wigner matrix of |ΦBKi special on the critical line, as we will see in the next section. 59 7. Berry-Keating reference states 7.3. Wigner matrix Equation (3.12) yields the Wigner matrix 2 e Ŵ|ΦBKi ≡ NnB W|φB(σ,τ )inB |eihe| + |γB|2 W|φ (2−3σ,τ )inB B + γB W|φ (σ,τ )inB hφB(2−3σ,τ )| |eihg| B + γ B W|φ |gihg| (σ,τ )| |gihe| (7.10) (2−3σ,τ )inB hφB B for the Berry-Keating reference state |ΦBKi, eq. (7.3). The diagonal matrix element connected to the excited state |ei is proportional to the Wigner function of the Berry state |φB(σ, τ )inB , eq. (7.8), whereas the one connected to the ground state |gi is the opposite hand of W|φB(2−3σ,τ )inB , due to eq. (7.9). Hence, on the critical line, the diagonal matrix elements are the mirror images√of each other, en = 1/ 2, eq. (7.6), is since for σ = 1/2 eq. (7.7) simplifies to γB = 1 and, therefore, N B independent of τ . In fig. 7.2, we depict the Wigner matrix elements of the Berry-Keating reference state |ΦBKi for σ = 1/3, τ = τ1 ∼ = 681.21 and nB = 10 on the left. As expected, the shapes resemble the ones of the Riemann-Siegel reference state shown in fig. 6.2, but the pictures on the right reveal the small deviations calculated from the difference Ŵ|ΦBKi − Ŵ|ΦRSi ≡ W|ei |eihe| + W|gi |gihg| + W|eihg| |eihg| + W|gihe| |gihe| (7.11) of the Wigner matrices, eqs. (7.10) and (6.15), with nB = n0 (τ ). The two bottom rows present the real and imaginary part of the off-diagonal matrix element connected to |eihg|. It is proportional to the Moyal function W|φ (σ,τ )in hφ (2−3σ,τ )| B B B defined by W|φ (σ ,τ )inB hφB(σ2 ,τ )| B 1 en (σ1 , τ ) N en (σ2 , τ ) ≡ N B B nX B −1 n,m=0 E (n, τ ) E (m, τ ) W|nihm| (n + 1)σ1 /2 (m + 1)σ2 /2 . (7.12) This expression is only symmetric in p on the critical line, just like the Moyal function W|σ,0in0 h2−3σ,0| , eq. (6.16), in the case of the Riemann-Siegel reference state. For σ 6= 1/2, asymmetries appear, caused by the factors E in the Moyal function and additionally by the factor γB containing the function χ. Thus, the shapes of the off-diagonal matrix element in fig. 7.2 again only seem to be antisymmetric. 7.4. Overlap in Wigner representation Finally, the Wigner representations Ŵ|ΨRSi and Ŵ|ΦBKi , eqs. (6.11) and (7.10), of the Riemann-Siegel state and the Berry-Keating reference state yield the absolute value of the overlap (7.2) Z∞ Z∞ 2 |CBK| = 2π~ dx dp KBK = MBK |ζBK(σ + iτ )|2 , 2 −∞ 60 −∞ (7.13) 7.4. Overlap in Wigner representation Figure 7.2.: The Wigner matrix elements of the Berry-Keating reference state |ΦBKi are shown on the left for σ = 1/3, τ = τ1 ∼ = 681.21 and nB = 10. To avoid cumbersome notation, we have used the abbreviations |φBi ≡ |φB(σ, τ )inB and |φBi ≡ |φB(2 − 3σ, τ )inB . Although the shapes all look like the ones of the Riemann-Siegel reference state |ΦRSi for the same values (fig. 6.2), there is a small difference defined by eq. (7.11) and depicted on the right. The differences of the diagonal matrix elements (two upper right pictures) emphasize that the ones of the BerryKeating reference state are not symmetric in p, in contrast to the ones of the Riemann-Siegel state. Moreover, the shapes of the off-diagonal matrix element (two bottom rows) are only nearly antisymmetric, since the τ -dependence of E in eq. (7.12) and the multiplication with γB makes them asymmetric with respect to p. 61 7. Berry-Keating reference states 1/3 R R 2π dx dp KBK 0.158 2 MBK 0.00401 |ζBK|2 39.31 37.80 39.99 1/2 0.163 0.01003 16.22 15.54 16.53 1/3 0.003 0.00401 0.8552 0.9397 0.8267 1/2 1.5 · 10−5 0.01003 0.00152 0.01499 0 σ τ1 τ2 |ζRS|2 |ζ|2 Table 7.1.: The values in the middle columns result from the data shown in the phase space pictures of fig. 7.3. For |ζBK|2 they can either be calculated from eq. (7.13) or directly with the Berry-Keating formula (7.1). The values confirm the improvement of the Berry-Keating formula compared to the Riemann-Siegel formula, eq. (6.1). with the help of eq. (3.15). Here, the kernel reads KBK ≡ e2 h N nB W|σ,τ iν W|φB(σ,τ )inB + |γB|2 W|σ,−τ iν W|φ (2−3σ,τ )in B B 2 n oi + 2 Re γB W|σ,τ iν hσ,−τ | W|φ (σ,τ )inB hφB(2−3σ,τ )| B (7.14) according to eq. (3.16). Like the kernel KRS, eq. (6.18), producing the Riemann-Siegel formula, the phase space pictures of KBK are symmetric in p only if σ = 1/2, due to the symmetries of the Wigner and Moyal functions given by eqs. (6.12) and (7.9) as well as eqs. (6.14) and (7.12), respectively. Figure 7.3 shows KBK for nB = 10 and the combinations of σ = 1/3 and σ = 1/2 with τ1 ∼ = 681.21 and τ2 ∼ = 681.89 in the second row. The pictures are more pronounced for σ = 1/2 and, especially for τ2 , the shapes differ from the ones of KRS which are repeated in the third row. The bottom pictures present the quite large difference KBK − KRS for σ = 1/2. For σ = 1/3, the values only deviate by |KBK − KRS| < 0.0006. Finally, the first row displays the absolute value of the overlap |CBK|2 , eq. (7.13), in blue with dots marking the values we get from integration over the phase space pictures 2 |ζ|2 is plotted as red dashed line and M2 |ζ |2 of KBK. Moreover, the exact curve MBK RS BK as black dotted line. As expected, we get more accurate values from the Berry-Keating formula than from the Riemann-Siegel approximation. This is also confirmed by tab. 7.1 which shows the values calculated from the phase space pictures in comparison to the ones evaluated by the formulae. 7.5. Overlap in Moyal representation The definition (3.11) of the overlap hΦBK|ΨRSi leads to the Moyal representation ∞ Z∞ en Z N B CBK = √ dx dp K BK = MBK ζBK 2 −∞ 62 −∞ 7.5. Overlap in Moyal representation Figure 7.3.: We depict in the second row the kernel KBK, eq. (7.14), for nB = 10 and the same values of σ and τ as used in fig. 6.3 for the kernel KRS (pictures repeated in the third row). In contrast to KRS, the pictures of KBK for σ = 1/2 are less pronounced. Moreover, the distribution of the positive and negative regions in phase space is quite different for τ2 . The bottom pictures show the difference KBK − KRS for σ = 1/2. For σ = 1/3, we get only very small differences with |KBK − KRS| < 0.0006. However, integration of KBK over phase space according to eq. (7.13) yields 2 more accurate values compared to the exact ones MBK |ζ|2 , as tab. 7.1 confirms. In the upper 2 2 pictures, the red dashed curves indicate MBK |ζ| while the blue curves represent |CBK|2 with dots marking the values for τ1 and τ2 . To emphasize the improvement of the approximation by the 2 Berry-Keating formula, we indicate MBK |ζRS|2 as black dotted line. 63 7. Berry-Keating reference states of the Berry-Keating formula. Here, the kernel of the phase space integral is given by i Nν (σ) h K BK ≡ W|σ,τ inB hφB(σ,τ )| + γB(σ, τ ) W|σ,−τ in hφ (2−3σ,τ )| (7.15) B B NnB (σ) with the Moyal functions en (σ2 , τ2 ) W|σ1 ,τ1 inB hφB(σ2 ,τ2 )| = NnB (σ1 ) N B The corresponding Moyal function W|σ1 ,τ1 in nX B −1 e −iτ1 ln(n+1) E(m, τ2 ) W|nihm| . σ1 /2 (m + 1)σ2 /2 (n + 1) n,m=0 B hφB(σ2 ,τ2 )| (7.16) emerges from eq. (7.16) by sub- stituting E with E . In definition (7.15) of the kernel, the prefactor Nν /NnB results form the property Z∞ −∞ dx Z∞ −∞ dp W|σ1 ,τ1 iν nB hφB(σ2 ,τ2 )| Nν (σ1 ) = NnB (σ1 ) Z∞ dx −∞ Z∞ dp W|σ1 ,τ1 inB hφB(σ2 ,τ2 )| −∞ for nB ≤ ν which also holds if we replace |φBinB by |φBinB . Analogous to eq. (6.22), this expression reflects the fact that the Fock states with ν > nB do not influence the overlap nB hφB(σ2 , τ2 )|σ1 , τ1 iν . The first and second row of fig. 7.4 depict the real and imaginary part of the Moyal functions W|σ,τ inB hφB(σ,τ )| and W|σ,−τ in hφ (2−3σ,τ )| appearing in eq. (7.15) of B B the kernel K BK for σ = 1/3, τ = τ1 and nB = 10. The pictures below reveal the tiny difference between these Moyal functions and the Moyal functions W|1/3,τ1 i10 h1/3,0| and W|1/3,−τ1 i10 h1,0| of the truncated Riemann states, defined by eq. (6.21) and shown in fig. 6.4. The additional factor E in the Moyal functions containing the Berry states, eq. (7.16), result in a larger rotation of some of the Moyal functions W|nihm| . The direction of rotation depends, as in the Riemann-Siegel case, on the sign of τ in the truncated Riemann state |σ, τ iν . Finally, we express the kernel, eq. (7.15), in terms of Fock states and arrive at " # nX 0 −1 −iτ ln(n+1) E(m, τ ) iτ ln(n+1) E (m, τ ) W|nihm| e χ(σ + iτ ) e en K BK = Nν N + . B σ/2 (2−3σ)/2 σ/2 (m + 1) (m + 1) (n + 1) n,m=0 On the critical line, this yields the relation K BK (x, −p) = e i arg{χ(1/2+iτ )} K BK (x, p) analogous to eq. (6.24). Hence, the absolute value of K BK is symmetric in p. In fig. 7.5, we show only the real and imaginary part of K BK and its connection 1 ζBK = e Nν (σ) NnB (σ, τ ) 64 Z∞ −∞ dx Z∞ −∞ dp K BK (7.17) 7.5. Overlap in Moyal representation Figure 7.4.: The real and imaginary part of the Moyal functions W|σ,τ inB hφB(σ,τ )| and W|σ,−τ in hφ (2−3σ,τ )| are depicted in the two top rows for σ = 1/3, τ = τ1 and nB = 10. In B B the bottom rows, we present the differences ∆WφB ≡ W|1/3,τ1 i10 hφ(1/3,τ1 )| − W|1/3,τ1 i10 h1/3,0| and ∆Wφ ≡ W|1/3,−τ1 i10 hφ (1,τ1 )| − W|1/3,−τ1 i10 h1,0| between the Moyal functions shown above and B the Moyal functions displayed in fig. 6.4. The differences are quite small, but they illustrate that the factor E in eq. (7.16) results in an additional rotation of some contributing W|nihm| . 65 7. Berry-Keating reference states Figure 7.5.: Here and on the next page, we show the real and imaginary part of ζ BK for σ = 1/3 and σ = 1/2 in the first row, with dots marking the values at τ1 ∼ = 681.21 and τ2 ∼ = 681.89. In the insets, the differences ∆ζ of ζBK − ζ and ζRS − ζ are depicted as blue and as black dotted line, respectively. The red dashed vertical lines indicate τ1 and τ2 . The row below presents the real or 66 7.5. Overlap in Moyal representation imaginary part of K BK, eq. (7.15), as phase space pictures which are connected to ζBK via eq. (7.17). They are reminiscent of the ones for K RS, fig. 6.5, of the Riemann-Siegel representation, repeated in the third row. However, the kernels are not the same as the bottom row confirms, where the difference K BK − K RS is depicted. 67 7. Berry-Keating reference states σ τ1 1/3 1/2 τ2 1/3 R R dx dp K BK −0.2584 − 0.9813 i −8 · 10−6 − 0.807 i 0.0513 + 0.1405 i en−1 Nn−1 N B B 6.180 4.994 6.181 ζBK ζ and (ζRS) −1.597 − 6.064 i −1.597 − 6.119 i −4 · 10−5 − 4.029 i 0.3182 + 0.8683 i (−1.597 − 5.939 i) −4 · 10−5 − 4.066 i (−4 · 10−5 − 3.944 i) 0.2607 + 0.8710 i (0.4398 + 0.8631 i) 1/2 0.0076 − 0.0003 i 4.995 0.0382 − 0.0017 i 0 (0.1212 − 0.00524 i) Table 7.2.: The values in the middle columns are calculated from the data displayed in the phase space pictures of fig. 7.5 and connected to ζBK via eq. (7.17). The right column shows that the values ζBK are much closer to the exact values of ζ than the values of ζRS given in brackets. to the Berry-Keating representation ζBK, eq. (7.1). The corresponding values are displayed in tab. 7.2. To emphasize the difference of the shapes to the ones depicted in fig. 6.5 for the Riemann-Siegel representation, we repeat K RS in the third row and present the difference K BK − K RS beneath. Although the difference of the kernels is small compared to the phase space pictures above, the insets in the first row emphasizes that the deviation ζBK − ζ (blue curve) is smaller than ζRS − ζ (black dotted line), confirming the improvement of the approximation by the Berry-Keating formula. 7.6. Summary Taking into account that the function E effectively cuts the sums at n0 and E(n, τ ) = 1 for almost all n < n0 , the modifications of the Berry-Keating formula in contrast to the Riemann-Siegel formula are small. However, these little changes enhance the accuracy of the approximation by including the first order correction term of the Riemann-Siegel approximation. The price we have to pay for this improvement in the physical picture is that the Berry-Keating reference state is always entangled. Nevertheless, we can identify the critical line: the Wigner matrix elements as well as the absolute value of the kernel K BK in the Moyal representation are only symmetric in p for σ = 1/2. We conclude the analysis of zeta states by mentioning that a further enhancement of the accuracy in the physical description can be achieved by inclusion of higher order correction terms. The investigation of the resulting states is left for future work. In the next chapter, we suggest a modification of the Jaynes-Cummings-Paul model to create the Riemann Hamiltonian which is essential in the realization of the zeta states. 68 8. Jaynes-Cummings-Paul approach to the Riemann Hamiltonian The eye-catching similarity between the phase of the Riemann zeta function and the time evolution produced by the effective Hamiltonian of the Jaynes-Cummings-Paul (JCP) model suggests that modification of this outstanding approach should indicate the physical realization of the Riemann Hamiltonian ĤR . Hence, we give in this chapter a short review of the JCP model and then show how we have to change the interaction to generate the Riemann Hamiltonian. 8.1. Jaynes-Cummings-Paul model In 1963, Edwin T. Jaynes together with Fred W. Cummings [16] and, at the same time independently, Harry Paul [17] introduced a model describing the interaction of an atom with a quantized light field. Although it is the most elementary model for the interaction – it considers only a fixed two-level atom and a single mode of the radiation field – the JCP model is able to describe most phenomena in cavity quantum electrodynamics. Thus, ‘it has become the drosophila of quantum optics.’ [28] The JCP model makes use of some simplifications [45]: (i) The internal structure of the atom is depicted by only two energy levels, the excited state |ei and the ground state |gi, which are separated by an energy gap of ~ωa , see fig. 8.1. Hence, the free Hamiltonian1 of the atom is given by Ĥa ≡ ~ωa σ̂z , 2 (8.1) where σ̂z , eq. (3.3), is the Pauli spin operator. (ii) We consider only a single mode of the radiation field ~ˆ ≡ E~0 (â + ↠) E (8.2) with E~0 containing the vacuum electric field and the mode function of the cavity C. This yields the free Hamiltonian of the field Ĥfield ≡ ~Ω â† â . 1 (8.3) There is, of course, an additional contribution in Ĥa which is proportional to the identity operator 1̂. This contribution yields overall phases which can be neglected since they cancel themselves in the interaction picture which we use to determine the effective Hamiltonian. We neglect the vacuum energy of the cavity field for the same reason. 69 8. Jaynes-Cummings-Paul approach to the Riemann Hamiltonian ~ωa ei ~ωa gi |ei |gi ~ E C Figure 8.1.: The two-level atom in the cavity C has an energy gap of ~ωa between the excited state |ei and the ground state |gi. (iii) The interaction between atom and field is approximated by the dipole interaction ~ˆ Ĥint ≡ −e~rˆ E (8.4) of the two level atom with the quantized field, eq. (8.2), in the cavity C in which (iv) the motion of the atom is neglected. In terms of the Pauli operator σ̂ ≡ |gihe| the dipole moment is e~rˆ = p~ σ̂ † + p~ ∗ σ̂ and eq. (8.4) then reads ~ Ĥint = ~g σ̂ + σ̂ † â + ↠with the constant g = |~p ~E0 | coupling the dipole moment to the electric field. Both vectors are chosen to be anti-parallel to achieve the sum of the Pauli operators as well as the sum of the annihilation and creation operator â and ↠, respectively. A transformation of the interaction Hamiltonian into the interaction picture via the free Hamiltonian Ĥ0 = Ĥa + Ĥfield , eqs. (8.1) and (8.3), leads to i i I Ĥint ≡ e ~ Ĥ0 t Ĥint e − ~ Ĥ0 t = ĤJC + ĤaJC . (8.5) The first part, the so-called Jaynes-Cummings Hamiltonian ĤJC = ~g ↠σ̂ e −i(ωa −Ω)t + h.c. , consists of the energy conserving terms whereas the second contribution, the anti-JaynesCummings Hamiltonian ĤaJC = ~g â σ̂ e −i(ωa +Ω)t + h.c. , annihilates a photon while transferring the atom into the ground state. This ‘unphysical’ behavior of the second term is often neglected based on the rotating wave approximation [28]: Due to the larger frequency ωa + Ω, the anti-Jaynes-Cummings part rotates faster 70 8.2. Riemann Hamiltonian than the Jaynes-Cummings contribution with frequency |ωa − Ω|, leading to negligible contributions. We mention here that Zheng-Hong Li et al. [51] propose in their paper a transformation of the interaction Hamiltonian quite similar to the transformation into the interaction picture with the advantage that it erases one of the contributions completely without using approximations. However, we will not follow this approach, but use second order perturbation theory to derive the effective Hamiltonian. I , eq. (8.5), is of the form The interaction Hamiltonian Ĥint X I Ĥint (t) ≡ Ĥj σ̂ e −iνj t + Ĥj† σ̂ † e iνj t (8.6) j which, in the far off-resonant case, produces the effective Hamiltonian i X 1 h † Ĥeff ≡ Ĥj Ĥj − Ĥj Ĥj† 1̂ + Ĥj† Ĥj + Ĥj Ĥj† σ̂z 2~νj (8.7) j as we show in appendix G.2. Hence, we get from eq. (8.5) 1 ~g̃˜ ĤJC,eff = ~g̃ n̂ + σ̂z + 1̂ , 2 2 h i h i where g̃ ≡ g 2 (ωa − Ω)−1 + (ωa + Ω)−1 and g̃˜ ≡ g 2 (ωa − Ω)−1 − (ωa + Ω)−1 . When we now use the rotating wave approximation to neglect the terms with (ωa + Ω)−1 we arrive at the well-known result [28]. Since constant terms produce overall phases, ĤJC,eff = ~ g2 n̂ σ̂z ωa − Ω is frequently referred to as effective Jaynes-Cummings Hamiltonian. In the next section, we derive in the same way an effective Hamiltonian which reflects, at least approximately, the logarithmic behavior of the Riemann Hamiltonian. 8.2. Riemann Hamiltonian To get the Riemann Hamiltonian ĤR , eq. (3.2), as effective Hamiltonian of a slightly changed Jaynes-Cummings-Paul model, we start again in the Schrödinger picture, but now with the Hamiltonian ĤS ≡ Ĥ0 + fˆ · σ̂ + σ̂ † , (8.8) where once more Ĥ0 ≡ ~Ω n̂ + ~ω2 a σ̂z . The Hermitian operator fˆ only acts in the Hilbert space of the field. Transformation into the interaction picture yields I Ĥint = ∞ X hm|fˆ|ni e −iΩt(n−m) |mihn| n,m=0 σ̂ e −iωa t + σ̂ † e iωa t . 71 8. Jaynes-Cummings-Paul approach to the Riemann Hamiltonian Here, we have used the completeness relation of the Fock states to transform fˆ. Substituting j = n − m, we find that the interaction Hamiltonian I Ĥint = ∞ X Ĥ−j σ̂ e −iν−j t j=0 + ∞ X Ĥj σ̂ e −iνj t + h.c. j=1 is of the form (8.6) with frequencies νj ≡ ωa + jΩ and the approximations Ĥj ≡ ∞ X n=j hn − j|fˆ|ni |n − jihn| Ĥ−j ≡ Ĥj† . as well as (8.9) Hence, the effective Hamiltonian, eq. (8.7), simplifies to ∞ ∞ X X Ĥeff = εj(+) Ĥj† Ĥj + Ĥj Ĥj† + ε0 Ĥ02 σ̂z + ε(j−) Ĥj† Ĥj − Ĥj Ĥj† 1̂ , j=1 j=1 where we have introduced the inverse energies 1 1 1 (±) ± and εj ≡ 2~ νj ν−j ε0 ≡ 1 . ~ν0 With eq. (8.9), we arrive at ∞ X ∞ ∞ 2 2 X X ˆ Ĥeff = ε(+) hn − j|fˆ|ni |nihn| + |n − jihn − j| + ε0 hn|f |ni |nihn| σ̂z j n=0 j=1 n=j ∞ X ∞ 2 X (−) ˆ + εj hn − j|f |ni |nihn| − |n − jihn − j| 1̂ . j=1 n=j Before we can compare this result to the Riemann Hamiltonian, eq. (3.2), we have to substitute n − j in the bras and kets |n − jihn − j| by n and then interchange the summation over j and n. This yields ∞ X Ĥeff = Fn σ̂z + Cn 1̂ |nihn| (8.10) n=0 with the abbreviations n ∞ 2 2 X ˆ 2 X (+) (+) ˆ ˆ Fn ≡ ε0 hn|f |ni + εj hn − j|f |ni + εj hn|f |n + ji j=1 (8.11) j=1 and Cn ≡ 72 n X j=1 ∞ 2 X 2 ε(j−) hn − j|fˆ|ni − ε(j−) hn|fˆ|n + ji . j=1 (8.12) 8.2. Riemann Hamiltonian So far, the only approximation we have used is ν −2 ν −1 to get the form (8.7) of the effective Hamiltonian. When we now compare eq. (8.10) to the Riemann Hamiltonian ĤR = ~ω ∞ X ln(n + 1)|nihn| σ̂z , n=0 we find that the coefficients of σ̂z must fulfill the equation ! Fn = ~ω ln(n + 1) ≡ F (n) (8.13) in which the frequency ω can be chosen freely. Moreover, the coefficients Cn , eq. (8.12), must be constant or negligible. We will answer the question how to choose fˆ in what follows. 8.2.1. Exact solution The simplest way to solve eq. (8.13) is to choose an operator fˆ = f (n̂) which is a function of the number operator n̂. For this choice all matrix elements hm|f (n̂)|ni with m 6= n vanish, that is, 2 Fn ≡ ε0 hn|fˆ|ni and Cn ≡ 0 . (8.14) Hence, eq. (8.13) determines the exact solution p fˆ ≡ f (n̂) ≡ ~ ν0 ω ln(n̂ + 1) which coincides with the definition of the Riemann Hamiltonian, eq. (3.2). The answer to the question how one could realize this Hamiltonian experimentally would exceed the scope of this thesis. Thus, we only mention here that Ruynet L. de Matos Filho and Werner Vogel propose in [52] the ‘engineering of the Hamiltonian of a trapped atom’, which generates Hamiltonians reminiscent of eq. (8.8). In the remainder of this section, we show that we get a good approximation of the Riemann zeta function without the exact Riemann Hamiltonian. 8.2.2. Approximate solution Inspired by the non-linear Jaynes-Cummings-Paul model [53], where fˆ ∼ â + ↠, we now choose fˆ as a function of the position operator κ x̂ ≡ √ â + ↠2 or, more precisely, as a polynomial fˆ ≡ f (x̂) = µX max fµ x̂µ (8.15) µ=0 with the inverse length κ introduced in eq. (D.15). 73 8. Jaynes-Cummings-Paul approach to the Riemann Hamiltonian We immediately see that in eq. (8.11) the absolute square of the matrix elements |hn ± j|x̂µ |ni|2 only contain the elements with j ≤ µ, producing a polynomial in n with degree µ (see appendix H.1 and H.2 for a general expression of the matrix elements). This yields (µmax ) Fn µmax 2 2 ˆ 2 X (+) ˆ ˆ εj Θ(n − j) hn − j|f |ni + hn|f |n + ji ≡ ε0 hn|f |ni + (8.16) j=1 and likewise for eq. (8.12) (µmax ) Cn ≡ µX max (−) εj j=1 2 2 ˆ ˆ Θ(n − j) hn − j|f |ni − hn|f |n + ji . (8.17) Both coefficients are polynomials in n of the order µmax . Here, the Heaviside function Θ is defined by ( 1,x≥0 Θ(x) = . 0,x<0 Since on the one hand there is no polynomial series expansion of the logarithm and, on the other, a logarithmic dependence is always smaller than a linear one, one is inveigled to assume that the best approximation of eq. (8.13) is given by the lowest terms of a polynomial. However, our goal is to approximate the Riemann zeta function, that is to compare the exact overlap C(t) ≡ hΦ|Ψ(t)i, eq. (3.6), with the approximation hΦ|e − ~i Ĥeff t |Ψ(0)i = ∞ X n=0 φ ne ce ψn e − it Fn + Cn ~ + φ ng cg ψn e it ~ Fn − Cn . (8.18) Here, the effective Hamiltonian Ĥeff , eq. (8.10), is used to calculate the time evolution of the initial state |Ψ(0)i. This task benefits of three properties of the overlap: (i) The arguments of the exponential functions of both equations must only be equal by modulo 2π, that is t ! Fn ± Cn mod 2π = ωt ln(n + 1) mod 2π . ~ (ii) The factors φ ne ψn and φ ng ψn are in the case of the zeta function proportional to n−σ , which makes deviations of Ĥeff from ĤR for larger n negligible, and (iii) deviations for one special n can be compensated by the others since only the sum over all n matters. Hence, even if the effective Hamiltonian does not provide a good approximation to the Riemann Hamiltonian, the overlap can nevertheless approximate the zeta function quite well. We show this behavior now exemplarily for the Riemann state |σ, τ i and a roughly approximated Hamiltonian. 74 8.2. Riemann Hamiltonian Riemann state First, we recall that the Riemann state |σ, τ = ωti, eq. (4.6), is only defined for the region σ > 1, where the zeta function is given by the Dirichlet series (4.2). Moreover, the overlap reads C(t) = hσ, 0|σ, τ i = |N (σ)|2 ζ(σ + iωt) according to eq. (4.7). Then, we derive from eq. (8.18) the overlap C(t) = |N (σ)|2 ∞ X n=0 it 1 e− ~ σ (n + 1) Fn + Cn . Now, we have to choose the coefficients fµ in eq. (8.15) in such a way that F≡ ∞ X n=0 it 1 e− ~ (n + 1)σ Fn + Cn ! = ∞ X n=0 1 e −iωt ln(n+1) = ζ(σ + iωt) (n + 1)σ (8.19) is approximately fulfilled. To simplify this task slightly more, we can reduce the number of coefficients by choosing small atom frequencies. Indeed, for ωa Ω the largest contribution ˆ 2 ε hn| f |ni (8.20) Fn(µmax ) ∼ = 0 is given by the inverse energy ε0 . A closer examination of the matrix elements (see appendix H.1) shows, that only the even coefficients occur in hn|fˆ|ni. Hence, we choose for the odd coefficients f2µ+1 = 0. This choice is also justified by the fact that all contributions with odd coefficients are polynomials of the order 2µmax − 1, whereas the even contributions are of order 2µmax in n. Although eq. (8.20) is similar to the exact solution eq. (8.14), the coefficients Cn(µmax ) do not vanish here. On the contrary, Cn(µmax ) is also given by the even contributions with order 2µmax in n, but it is smaller than Fn(µmax ) due to the missing inverse energy ε0 . In fig. 8.2, we show eq. (8.19) for µmax = 4 and the coefficients f1 = f3 = 0, f0 = 0.2, f2 = 0.3 and f4 = −0.4 for the Riemann state with σ = 2. The explicit expressions of Fn(4) and Cn(4) are given in appendix H.3. In the upper row of fig. 8.2, we have depicted the real and imaginary part of F in blue, eq. (8.19), and ζ(σ + iωt) in red. The overall behavior fits quite well even though F fluctuates strongly. This becomes more evident when we compare |F| to the approximation of the Dirichlet series ζ(σ + iωt) ≈ 1 + cos(ωt ln 2) sin(ωt ln 2) −i ≡ ζ̄ , 2σ 2σ (8.21) indicated by the yellow curve in the picture below, to emphasize the periodic structure of F. In the third row, we illustrate the deviation ∆≡ |F| |ζ(σ + iωt)| (8.22) 75 8. Jaynes-Cummings-Paul approach to the Riemann Hamiltonian Re F Im F 1.6 0.4 1.4 0.2 1.2 20 1.0 40 60 80 100 t -0.2 0.8 0.6 20 40 60 80 t 100 -0.4 |F| 1.6 1.4 1.2 1.0 0.8 0.6 ∆ 1.6 1.4 1.2 1.0 0.8 0.6 ∆(n+) 3.0 2.5 2.0 1.5 1.0 0.5 20 40 60 80 100 t Figure 8.2.: The real and imaginary part of F, eq. (8.19), is shown in the first row for Fn(4) in blue. The parameters are chosen as follows: f0 = 0.2, f1 = f3 = 0, f2 = 0.3, f4 = −0.4, σ = 2, ω = ωa = 1, Ω = 100, ~ = 1, κ = 1 (sum cut at N = 100). Although F fluctuates around the values of ζ(σ + iωt) (red curve) the overall behavior fits quite well. This becomes evident in the second row, where the periodic structure of |F| (blue) is given by the approximation of |ζ(2+ it)|, eq. (8.21), indicated in yellow. Furthermore, the deviation ∆, eq. (8.22), (third row) between F and the zeta function itself is, as expected, smaller than the deviation of the exponents ∆(n+) , eq. (8.23), shown in the bottom row. 76 8.3. Summary of the approximation to the exact values. It is smaller than the deviation ∆n(+) ≡ (Fn + Cn ) mod 2π ~ω ln(n + 1) (8.23) of the exponents in eq. (8.19) depicted below. Hence, F provides a better approximation of the zeta function than the exponent Fn + Cn of the logarithm. Of course, the approximation can be improved by using higher order polynomials and optimizing the choice of the coefficients as well as the detuning between the atom frequency ωa and the field frequency Ω. 8.3. Summary In this chapter, we have shown that it is possible to get an approximation of the Riemann zeta function by using a Jaynes-Cummings-Paul-like model. Even if the effective Hamiltonian resulting from this model does not match the Riemann Hamiltonian the resulting overlap fits the overall behavior of the zeta function quite well, as the rough approximation with a polynomial of fourth order for fˆ confirms. A further investigation of approximations with higher order polynomials or other functional dependence as well as the application of this method to other zeta states is left for future work. 77 Summary Due to its intimate connection to the distribution of the primes, the Riemann zeta function ζ is famous beyond the borders of mathematics. Therefore, we have discussed different representation of ζ in this thesis, adopting mathematical as well as physical methods: For the illustration of ζ in complex space, we have used the continuous Newton method, which depicts a complex function F by lines of constant phase. These lines start at the poles of the function and lead into its zeros. Hence, we can easily locate the zeros and poles in the pictures of the Newton flow. Moreover, the zeros of the derivative F 0 are visible since they generate a crossing of the flow lines which act as separatrices for the flow into the different zeros. In case of the Riemann zeta function, the pictures of the Newton flow would therefore reveal if the Riemann hypothesis was violated. Our physical approach is inspired by the interaction of a cavity field with a twolevel atom, described by the Jaynes-Cummings-Paul model. We realize the Riemann zeta function by the overlap hΦ|Ψ(t)i of two quantum states, but with a time evolution governed by the Riemann Hamiltonian ĤR which depends logarithmically on the number operator n̂. The properties of the states depend on the formula we want to reproduce: In the region where the real part σ of the argument s is larger than one, the cavity system suffices to produce ζ. However, entanglement is crucial for most representations in the critical strip. We have shown in particular that: (i) The Dirichlet sum, defining ζ for σ > 1, is proportional to the overlap of the Riemann state |σ, τ i with its initial state |σ, 0i. But the limitation of the Dirichlet sum to the right of the critical strip is transferred to the states, due to conservation of probability. (ii) In contrast to the Dirichlet series, the alternating sum converges already for σ > 0. However, our physical approach fails to enter the critical strip as long as we try to reproduce the exact formula of the infinite alternating sum. The normalization again restricts the states to σ > 1. Even entanglement is useless in this case. (iii) Only the truncated version of the alternating sum allows the realization of ζ for σ < 1 by a single system. The resulting truncated Riemann states |σ, τ iν are normalizable as long as the truncation ν is finite. But when we increase the truncation to improve the approximation of ζ by the truncated alternating sum, the normalization of the states tends to zero and decreases the value of the overlap as well. 79 Summary (iv) Nevertheless, entanglement is the key to overcome the wall at σ = 1: Unlike the Dirichlet series and the alternating sum, the approximations of ζ by the Riemann-Siegel and the Berry-Keating formula involve not only the phases −iτ ln(n + 1), but also the conjugate phases iτ ln(n + 1), which are available through the entanglement with the atom. Moreover, the sums in these representations are truncated at relatively small limits n0 ≤ nB . Consequently, we need only a few Fock states to produce these representations by the overlaps hΦRS|ΨRS(t)i and hΦBK|ΨRS(t)i, respectively. The similarities and differences of the formulae are reflected in the following properties of the states: • The time-evolved entangled Riemann-Siegel state |ΨRS(t)i is used for both representations. It consists of two truncated Riemann states coupled to the excited and ground state of the atom. Thus, it is reminiscent of the Schrödinger cat state, but with phases logarithmic instead of linear in n. Due to the entanglement, the time evolution acts in different ways on the matrix elements of the Wigner representation Ŵ|ΨRSi : The diagonal matrix elements rotate in opposite directions depending on the state of the atom. In contrast, the off-diagonal matrix elements only change their internal structure, but remain symmetric in the phase space variable p. • The Riemann-Siegel reference state |ΦRSi involves the initial truncated Riemann states |σ, 0in0 and |2 − 3σ, 0in0 coupled to the atom. It is entangled for σ 6= 1/2 and disentangled for σ = 1/2. Hence, the phase space pictures of the state reveal a symmetry in p on the critical line. • In contrast, the Berry-Keating reference state |ΦBKi contains the Berry states |φB(σ, τ )inB and |φB(2 − 3σ, τ )inB and is entangled for all σ even though the probability amplitudes differ only slightly from the ones of the initial truncated Riemann states used in |ΦRSi. Anyway, the off-diagonal matrix elements of Ŵ|ΦRSi are symmetric in p for σ = 1/2. Moreover, the special role of the critical line reveals itself in the phase-space representations of both approximations: The kernel K of the Wigner representation as well as the absolute value of the corresponding kernel of the Moyal representation K is p-symmetric for σ = 1/2. (v) Additionally, we have shown that the Riemann zeta function can be simulated by the time evolution with a non-logarithmic effective Hamiltonian: Already a polynomial of fourth order produces a rough approximation of ζ reflecting the overall behavior quite well although the Hamiltonians do not match. We conclude this thesis by noting that our physical approach can be generalized to other complex functions if they can be represented by the series ∞ X an (σ, τ ) bn (σ, τ ) f (σ + iτ ) ≡ + . (n + 1) iτ (n + 1)−iτ n=0 80 As in the case of the Riemann zeta function, the factors (n + 1)±iτ can be interpreted as time evolution of a quantum state |Ψ(t)i governed by the interaction Hamiltonian ĤR and the rescaled time τ = ωt. The factors an and bn are distributed to the probability amplitudes of the state |Ψ(t)i and the appropriately chosen reference state |Φi. But it is crucial to keep in mind that the amplitudes of the initial state |Ψ(0)i must be independent of τ . Otherwise we cannot reproduce the phases (n + 1)±iωt in |Ψ(t)i by this simple time evolution. As a consequence, all τ -dependencies of an and bn must be included in the probability amplitudes of the reference state, since there τ = Im s is simply a constant determining the imaginary part of the point s in complex space. In this way, we can transfer the characteristics of the complex function into properties of quantum states and make them available by measurement. 81 A. Special functions In this appendix, we bring back to mind some properties of the functions Γ, χ and ϑ, which appear in the different representations of the Riemann zeta function. A.1. The gamma function Γ Already in 1729, different representations of the gamma function were given by Daniel Bernoulli and Leonhard Euler in letters to Christian Goldbach [54]. We recall here only the more familiar representations and properties of Γ. A.1.1. Definition and functional equation The gamma function Γ is defined by [50, 55] Γ(s) ≡ Z∞ dx xs−1 e −x (A.1) 0 for Re s > 0. Integration of this equation by parts yields the functional equation Γ(s) = 1 Γ(s + 1) s (A.2) which is used to define the analytical continuation of Γ in the rest of the complex plane. From eq. (A.1) easily follows that Γ (s) = Γ(s̄) as well as the special values Γ(1) = 1 and Γ(1/2) = π 1/2 . With the help of eq. (A.2), we find the asymptotic expression Γ(s) ∼ = s−1 for s → 0. Another expression for the gamma function is the Gauß representation Γ(s) = lim n→∞ n! ns . s · (s + 1) · ... · (s + n) It clearly shows that Γ has poles at zero and all negative integer numbers. This property determines the location of the trivial zeros of the Riemann zeta function. A.1.2. Asymptotics For large τ we can approximate Γ with the method of stationary phase by Γ(s) = Z∞ 0 dx e −S(x) ∼ = e −S(xs ) Z∞ 1 e− 2 S 00 (x s) x 2 , −∞ 83 A. Special functions where we have used the Taylor expansion of the phase S(x) = x − (s − 1) ln x around the stationary point xs = s − 1 given by S 0 (xs ) = 0. Integration yields the Stirling formula p Γ(s) ∼ = 2π(s − 1) e −(s−1)[1−ln(s−1)] for Re s > 1 which we can expand to smaller real parts with the help of the functional equation (A.2). This results in the expression √ 1 Γ(s) ∼ (A.3) = 2π e −s+(s− 2 ) ln s for Re s > 0. Moreover, we find from the series expansions [50] x2 ln(1 + x) ∼ + O(x3 ) −1 < x ≤ 1 =x− 2 ∼ π − 1 + O x−3 x>1 arctan x = 2 x for s = σ + iτ and large τ that π σ ln s ∼ − + O τ −2 . = ln τ + i 2 τ (A.4) Hence, eq. (A.3) can be expressed by √ πτ π 1 1 −1 ∼ ln τ − +i σ− − τ + τ ln τ + O τ . Γ(s) = 2π exp σ− 2 2 2 2 (A.5) A.1.3. Formulae The gamma function is connected to the trigonometric functions [56] via π Γ(s) Γ(1 − s) = sin(πs) and (A.6) π . cos(πs) (A.7) 22s−1 Γ(2s) = √ Γ(s) Γ (s + 12 ) . π (A.8) Γ (s + 12 ) Γ (s − 12 ) = For twice the argument we get All three equations are used in the next section to show the equivalence of the different representations of χ found in literature. A.2. The function χ(s) The function χ plays a crucial role in the analytic continuation of the Riemann zeta function. It connects the values of ζ for σ > 1/2 to the region where σ < 1/2 via the functional equation (1.5). 84 A.2. The function χ(s) A.2.1. Definitions In literature, χ is defined in many ways: While Siegel [29] used the formula χ(s) ≡ (2π)s 2Γ(s) cos πs 2 we find in [28] the representation χ(s) ≡ π and additionally in [4] s− 21 Γ 1−s 2 Γ 2s , (A.9) χ(s) ≡ 2s π s−1 Γ(1 − s) sin (A.10) πs 2 . (A.11) The equivalence of representation (A.11) and (A.9) becomes evident when we use eq. (A.6) to replace the sine. To arrive at definition (A.10), we substitute Γ(s) in eq. (A.9) using the doubling formula (A.8) and apply eq. (A.7). The function χ inherits the property χ (s) = χ(s̄) of the Gamma function, as eq. (A.10) easily confirms. However, the functional equation reads χ(1 − s) = χ−1 (s) . (A.12) Moreover, eq. (A.10) shows that there exist only ‘trivial’ zeros of χ, which are located at s = 0, −2, −4, ... due to the poles of the denominator Γ(s/2), and poles at s = 1, 3, 5, ... created by the poles of the numerator Γ((1 − s)/2). A.2.2. On the critical line Already the functional equation (A.12) of χ indicates the special role of the critical line. Indeed, there we find 1 iτ χ ( 12 + iτ ) = exp i τ ln π − 2 arg Γ + , 4 2 that is |χ ( 21 + iτ ) | = 1 for all τ . With the help of eq. (A.3), we get the approximation n τ π o χ ( 21 + iτ ) ∼ −τ − + O(τ −1 ) = exp −i τ ln 2π 4 (A.13) for large τ . 85 A. Special functions A.2.3. Asymptotics in the critical strip In the critical strip 0 < Re s < 1, we find that both gamma functions in the definition (A.10) of χ have arguments with real part between 0 and 1/2. Hence, we can approximate them for τ 1 with the help of eq. (A.3) and the expressions s 2 1−s ln 2 ln τ iπ iσ ∼ − + O(τ −2 ) = ln + 2 2 τ τ iπ i(1 − σ) ∼ − + O(τ −2 ) = ln − 2 2 τ for the logarithms. Eventually, we arrive at [4, 5] σ− 1 2 2π t π −1 χ(s) ∼ + O(τ ) exp −i τ ln −τ − = τ 2π 4 (A.14) which is according to eq. (A.13) χ(s) ∼ = 2π τ σ− 1 2 χ( 21 + iτ ) . (A.15) A.3. The function ϑ(s) In the derivations of the Riemann-Siegel and Berry-Keating formula in appendix B, we use the definition e i ϑ(s) ≡ χ−1/2 (s) (A.16) to express the symmetric representation Z, eq. (1.7). This yields " 1−s # h s i Γ 1 1−s i 1 ϑ(s) ≡ arg Γ − arg Γ + s− ln π + ln 2s 2 2 2 2 2 Γ 2 and for σ = 1/2 τ 1 iτ θ(τ ) ≡ ϑ( 21 + iτ ) = arg Γ + − ln π 4 2 2 (A.17) which is real. Since Z is also real for s = 1/2 + iτ , θ(τ ) describes the phase of ζ on the critical line. This phase is, according to [9], closely related to the phase shift produced by scattering on an inverted harmonic oscillator. Besides the critical line, the complex conjugate of ϑ fulfills ϑ (s) = −ϑ(s̄) since the property Γ (s) e −i arg [ Γ(s) ] = Γ (s) = Γ(s̄) = |Γ(s̄)| e i arg [ Γ(s̄) ] results in 86 Γ (s) = |Γ(s̄)| and arg [ Γ(s) ] = − arg [ Γ(s̄) ] . B. Approximations in the critical strip The exact formulae of the zeta function are either not capable of producing states that could enter the critical strip, chapter 4 and 5, or are difficult to realize experimentally. Thus, we make use of the Riemann-Siegel and the Berry-Keating formula which are quite good approximations for large imaginary parts τ and can be conveniently casted in our physical picture. In this appendix, we give at first a short review of the Riemann-Siegel formula with focus on the remainder to provide the necessary expressions for the following sections. Then, we show that the Berry-Keating formula on the critical line is an improved version of the Riemann-Siegel formula since it contains additionally its first correction term. Finally, we give the extension of the Berry-Keating formula to the critical strip. B.1. Riemann-Siegel formula As mentioned in section 1.5, Carl L. Siegel rederived the semi-convergent representation of the zeta function [29] ζ(s) = ζRS(s) + RS (B.1) which contains in the main part ζRS(s) ≡ n0 n0 X X 1 1 + χ(s) s 1−s n n n=1 (B.2) n=1 the function χ(s) defined by eq. (1.6). Due to its dependence on τ , the cut-off n0 , eq. (1.10), leads to discontinuities. For the sake of completeness, we note that the remainder is according to [29] s+1 (−1)n0 −1 (2π) 2 τ RS ≡ (1 − e 2πis ) Γ(s) with SN ≡ N −1 X s−1 2 e i[ πs − τ2 − π8 2 ]S N N ! X 2−k i r−k k! 6 3N ak F (k−2r) (δ) + O . r! (k − 2r)! τ k=0 r≤k/2 The summation limit is N ≤ 2 · 10−8 τ and the derivatives F (k) of cos u2 + 3π 8 √ F (u) ≡ cos 2πu 87 B. Approximations in the critical strip are evaluated at δ ≡ recurrence formula √ τ− √ 2π n0 + 1 2 . Moreover, the coefficients ak in SN obey the √ (n + 1) τ an+1 = −(n + 1 − σ) an + i an−2 with a−2 = a−1 = 0 and a0 = 1. In literature [4, 29, 31], we find various representations of the whole remainder, which differ slightly due to their derivation. We will concern ourselves in what follows only with the first correction term s+1 (0) RS (s) with (−1)n0 −1 (2π) 2 τ ≡ (1 − e 2πis ) Γ(s) s−1 2 e i[ πs − τ2 − π8 2 √ cos τ − (2n0 + 1) 2πτ − π8 √ F ≡ cos( 2πτ ) ]F (B.3) given by Siegel [29] for τ > 0. The remainder for negative imaginary part τ follows from the symmetric form Z(s), eq. (1.7). When we take the Riemann-Siegel representation (B.1) with the remainder approxi(0) mated by its first correction term RS to evaluate Z, we arrive at n0 X 2 1 (0) (0) √ cos ϑ(s) + i s − ZRS (s) ≡ ln n + χ−1/2 (s) RS (s) , (B.4) 2 n n=1 where ϑ is defined by eq. (A.16). Since Z is symmetric to the critical line, eq. (1.8), the remainder fulfills (0) (0) RS (1 − s) = χ−1 (s) RS (s) in analogy to the functional equation of the Riemann zeta function, eq. (1.5). Hence, (0) R(0) (s) ≡ χ−1/2 (s) RS (s) has to be real on the critical line. Moreover, 1/4 2π (0) R (s) ∼ (−1)n0 −1 F = τ (B.5) holds in the whole critical strip, as the approximations of Γ and χ, eqs. (A.5) and (A.14), for τ 1 reveal. In the next section, we show that the representation of Z given by Berry and Keating combines the main term and the first correction R(0) in the first summand of a series expansion. B.2. Berry-Keating formula on the critical line Michael V. Berry and Jonathan P. Keating show [31] that one can cast the symmetric representation on the critical line Z ( 12 + iτ ) = e iθ(τ ) ζ( 12 + iτ ) , 88 B.2. Berry-Keating formula on the critical line with θ given by eq. (A.17), into a series of convergent sums Z ( 12 + iτ ) = Z0 (τ, K) + Z3 (τ, K) + Z4 (τ, K) + . . . . Here, K is an arbitrary constant and the main part Z0 can be expressed by (∞ ) r X exp {i(θ(τ ) − τ ln n)} 1 τ √ Erfc ξ(n, τ ) Z0 (τ, K) ≡ 2 Re 2 2Q2 (K, τ ) n n=1 (B.6) with the abbreviations ξ(n, τ ) ≡ ln n − θ0 (τ ) and Q2 (K, τ ) ≡ K 2 − iτ θ00 (τ ) (B.7) as well as the complementary error function (eq. (7.1.2) in Abramowitz & Stegun [56]) 2 Erfc (x) ≡ √ π Z∞ 2 dt e −t . (B.8) x Following the paper, we first show that the sum over n is convergent and then prove (0) that the main term Z0 is approximately ZRS , eq. (B.4). In eq. (B.6), the convergence of the sum is based on the property that Erfc (x) is for 2 large arguments proportional to e −x . Indeed, with r τ (B.9) y(n, τ, K) ≡ ξ(n, τ ) 2 2Q (K, τ ) we find !2 !2 2 n n τK τ p p ln ln = exp − Erfc (y) ∼ exp − τ τ 2(K 4 + 14 ) 2K 2 − i 2π 2π which becomes exponentially small for large τ . Hence, we can choose > 0 to fulfill the inequality !2 2 τK n p exp − ln ≤ e − . τ 2(K 4 + 14 ) 2π This yields n≤ r τ exp 2π ( s ) 1 1+ π 4K 4 2π K pτ (B.10) √ for the summation p index. For K = 1/ 2, the expression on the right is minimal and the exponent 2ε/τ vanishes for large τ . Therefore, it suffices to cut the sum at n0 , eq. (1.10), to get the main contribution. Now, we use the approximation (A.13) of χ to get the approximations for h i τ τ τ π θ(τ ) = arg χ−1/2 (1/2 + iτ ) ∼ − − + O(τ −1 ) (B.11) = ln 2 2π 2 8 89 B. Approximations in the critical strip and for its derivatives 1 τ θ0 (τ ) ∼ + O(τ −2 ) = ln 2 2π and 1 θ00 (τ ) ∼ + O(τ −3 ) , = 2τ which transform eq. (B.7) into n ξ∼ = ln p τ and 2π i Q2 ∼ = K2 − . 2 (B.12) Hence, eq. (B.9) is approximately !r i i n τ arctan 1 2 arctan 1 2 2 2 2K 2K y∼ e ≡ y e . = ln p τ 0 2 2|Q| 2π Taking into account that the complementary error function with complex argument is roughly given by1 (B.13) Erfc (y0 e iφ ) ≈ Erfc (y0 ) as long as |φ| ≡ 21 arctan 2K1 2 ≤ π4 , which holds for all K, we arrive at the approximation (∞ ) X exp {i(θ(τ ) − τ ln n)} 1 √ Z0 (τ, K) ∼ Erfc(y0 ) = 2 Re 2 n n=1 for the main part, eq. (B.6). Furthermore, the relation 1 1 Erfc (y0 ) = Θ(−y0 ) + Erfc (|y0 |) sgn(y0 ) , 2 2 where the Heaviside function and the signum are defined by 0 , x < 0 −1 , x < 0 1 Θ(x) ≡ and sgn(x) ≡ 0, x=0 2 , x=0 1, x>0 1, x>0 produces Z0 ∼ =2 1 , (B.14) n0 X cos[θ(τ ) − τ ln n] √ +R. n n=1 When we calculate the integral in definition (B.8) along the path t = r e iφ , we arrive at 2 Erfc(y0 e iφ ) = √ e iφ π Z∞ dr exp{−r2 cos(2φ) − ir2 sin(2φ)} y0 which converges only if |φ| ≤ π/4. eq. (B.13). √ The rough approximation cos x 1≈ 1 produces arctan 1 2 = π/8 and At the end, we choose K = 1/ 2 for which the phase is |φ| = 2 2K |Erfc (y0 e iπ/8 )| − Erfc (y0 ) ≤ 0.07. 90 B.2. Berry-Keating formula on the critical line Here, we have already used that the Heaviside function Θ(−y0 ) cuts the sum at n0 to reveal the main part in eq. (B.4). The remainder (∞ ) X exp {i(θ(τ ) − τ ln n)} √ R ≡ Re Erfc(|y0 |) sgn(y0 ) (B.15) n n=1 becomes quite small at the jumps τ = 2πn20 of the Riemann-Siegel formula. When we substitute n = n0 + k and τ = 2π(n0 + p)2 in eq. (B.12) and approximate the logarithm in ξ according to eq. (A.4), we get n k−p ∼ k−p 1 k−p 2 p ξ∼ ln = ln 1 + − = = τ n0 + p n0 + p 2 n0 + p 2π and can estimate the argument of the exponential function in eq. (B.15) by " # h i n τ π π θ(τ ) − τ ln n mod 2π ∼ mod 2π ∼ = −τ ln p τ − − = πn0 + 2πp2 − + πk − 4πpk , 2 8 8 2π using additionally the approximation of θ, eq.√ (B.11). From the above substitution of τ π and the approximation of ξ follows that y0 ∼ = |Q| (k −p). Therefore, eq. (B.15) transforms into ( 1 4 1 2π 2 R∼ (−1)n0 Re e 2πi(p − 16 ) = τ √ √ ∞ X π|k − p| π(k − p) (B.16) sgn · (−1)k e −i4πpk Erfc |Q| |Q| k=1−n0 −1/2 ∼ when we additionally apply n−1/2 ∼ = (2π/τ )1/4 . Since the error function vanishes = n0 for large k, we can extend the sum to −∞ and split it into positive and negative k. For p > 0, that is between the discontinuities at τ = 2πn20 , eq. (B.16) yields ∞ X |k − p| ∼ k −1 − 2i (−1) sin(4πpk) Erfc = −1 + i tan(2πp) . |Q| k=1 Here, we have used that the signum, eq. (B.14), becomes +1 for k > 0 and −1 for k ≤ 0. In the last step, we have assumed that the error function can be estimated by 1. This rough approximation is justified by the fact that 0 < p < 1 and the overall factor τ −1/4 makes R small for large τ . Finally, after some calculation and with the help of eqs. (B.3) and (B.5), we get that the remainder 1 1/4 1 2 2π 4 2π n0 −1 cos[2π(p − p − 16 )] ∼ R= (−1)n0 −1 F ∼ (−1) = = R(0) (1/2 + iτ ) τ cos(2πp) τ is approximately the first correction term R(0) of the Riemann-Siegel formula. Additional to the improvement of the approximation, the Berry-Keating representation benefits from several more properties [31]: 91 B. Approximations in the critical strip (i) The discontinuities of the Riemann-Siegel formula are eliminated since all terms are analytic functions of τ . (ii) The other terms Zi , i ≥ 3, ‘can be estimated explicitly’ and (iii) ‘numerical studies suggest that term by term the new series is more accurate than’ the Riemann-Siegel expression. Therefore, we present the extension of the Berry-Keating formula to σ 6= 1/2 in the following section. B.3. Berry-Keating formula in the critical strip Since we are also interested in a description of the zeta function besides the critical line, we now extend the Berry-Keating formula to the critical strip. Therefore, we first bring eq. (B.6) into the form Z0 = ∞ X e i(θ(τ )−τ ln n) Erfc(y) + e −i(θ(τ )−τ ln n) Erfc (y) √ 2 n n=1 with y ≡ y(n, τ, K) given by eq. (B.9). Then, we introduce the argument s ≡ σ + iτ by 1 substituting the function θ(τ ) with ϑ(s), eq. (A.16), and τ itself with T ≡ i 2 − s : Z0 (s) = ∞ X e i(ϑ(s)−T n=1 ln n) Erfc(y(n, T , K)) + e −i(ϑ(s)−T √ 2 n ln n) Erfc (y(n, T , K)) . We recall the approximation (A.15) of χ in the critical strip and arrive with the definition (A.16) at 1 σ 2π ϑ(s) ∼ θ(τ ) − i − ln , ϑ0 (s) ∼ = = θ0 (τ ) and ϑ00 (s) ∼ = θ00 (τ ) , 4 2 τ where we have neglected all terms smaller or equal to O(τ −1 ). These approximations yield ξ(n, T ) ≡ ln n − ϑ0 (s) ∼ = ξ(n, τ ) as well as Q2 (K, T ) ≡ K 2 − iT ϑ00 (s) ∼ = K 2 − it θ00 (τ ) + O(τ −1 ) ∼ = Q2 (K, τ ) for eq. (B.7), and thus y(n, T , K) ∼ = y(n, τ, K) for the argument of the error function. Hence, the same steps as in the previous section (0) (s), eq. (B.4). can be applies to proof that ZBK(s) is approximately ZRS 92 B.3. Berry-Keating formula in the critical strip √ Finally, we take advantage of eq. (B.10) by choosing K = 1/ 2 as well as the truncation parameter nB ≥ n0 and use the approximations in eq. (B.12) to get " # nB X χ−1/2 (s) E(n − 1, τ ) χ1/2 (s) E (n − 1, τ ) ZBK(s) ≡ + , (B.17) ns n1−s n=1 where 1 E(n, τ ) ≡ Erfc 2 "r τ n+1 ln p τ 1−i 2π # . Since ZBK provides an excellent approximation to the symmetric representation Z (see figures 1.2 and 1.3), we use eq. (B.17) to derive the Berry-Keating reference state in chapter 7. 93 C. Continuous Newton method Partial differential equations govern many physical phenomena. Their solution can often be found by casting the problem into a search for the minimum of a function. Hence, continuous descent methods [39, 40], like the continuous Newton method, have found wide application in physics. In this appendix, we illustrate the building blocks of the Newton flow by means of several examples. The last sections are dedicated to technical details needed for the pictures and to describe the Newton flow of ζ. C.1. Newton flow As already mentioned in section 2, the continuous Newton method consists of finding the solution s(t) for the differential equation ṡ(t) ≡ ds(t) F (s(t)) =− 0 . dt F (s(t)) (C.1) Provided that F 0 (s) ≡ dF (s) 6= 0, eq. (C.1) yields ds F 0 (s(t)) · ṡ(t) = dF (s(t)) = −F (s(t)) dt for a holomorphic function F . Hence, the trajectories s(t) are defined by F (s(t)) = F (s(0)) e −t (C.2) and represent lines of constant phase. The totality of the Newton trajectories s(t) form the Newton flow. We investigate the structure of the flow by some simple examples in the next sections. C.1.1. Sink and source In the limit t → ∞, eq. (C.2) becomes zero. Therefore, all trajectories of the Newton flow end eventually at a zero of F , which makes the continuous Newton method a perfect tool to find the zeros of a function. Moreover, the limit t → −∞ indicates that the starting point of the trajectories is a pole. We have to mention here that the direction of the Newton flow results from the choice of the minus sign in the differential equation (C.1), which is a mere convention. Choosing plus instead of minus would reverse the direction of the flow, but not change the location of the curves in complex space. 95 C. Continuous Newton method Figure C.1.: Newton flow of the functions F0 , Fp and the Möbius transformation FM defined in eq. (C.3) (from left to right). The zero (red dot) acts as sink while the pole (black dot) is source of the Newton flow. Infinity can be source and sink. In fig. C.1 from left to right, we present the Newton flow of the functions F0 (s) ≡ s , Fp (s) ≡ s−1 and FM (s) ≡ s+1 s−1 (C.3) in complex space s = σ + iτ . Here, F0 and Fp display a single zero and a single pole, respectively, while the Möbius transformation FM possesses both. As expected, the zero acts as a sink and the pole as a source for the Newton flow. Moreover, infinity can be source and sink. Here and in all other pictures of the Newton flow, zeros and poles are marked by red and black dots, respectively. We note that although these pictures are reminiscent of the field of charges, there is no one-to-one correspondence between the Newton flow and electrostatics [57]. Indeed, if we consider the components N1 ≡ −Re {F/F 0 } and N2 ≡ −Im {F/F 0 } of the Newton ~ as components of an electric field, the component vector N ~ ≡ ∂N2 − ∂N1 curl N ∂σ ∂τ (C.4) of the curl does in general not vanish due to the Cauchy-Riemann differential equations, which are fulfilled by a holomorphic function. We give a short proof of this statement before we turn to further examples in the next section. Proof. Since F is holomorphic, the function −F/F 0 is also holomorphic and therefore fulfills the Cauchy-Riemann differential equations [50] ∂N1 (σ, τ ) ∂N2 (σ, τ ) = ∂σ ∂τ and ∂N1 (σ, τ ) ∂N2 (σ, τ ) =− . ∂τ ∂σ ~ to simplify the notation. Here, we have used the components of the Newton vector N Hence, we get ~ ≡ 2 ∂N2 curl N (C.5) ∂σ from eq. (C.4), which in general does not vanish. 96 C.1. Newton flow It is worthwhile to represent the curl of the Newton flow in terms of the real and imaginary part u(σ, τ ) and v(σ, τ ) of the complex function F (s) = F (σ + iτ ): With the help of the identity dF (s) ∂F dσ F 0 (s) = = = uσ + ivσ , ds ∂σ ds where fσ denotes the partial derivative of the function f with respect to σ, the Newton quotient reads F uuσ + vvσ uvσ − uσ v − 0 =− 2 −i 2 = N1 + iN2 . F uσ + vσ2 uσ + vσ2 Hence, the curl of the Newton flow, eq. (C.5), is given by 2u uσ vσ + v(vσ2 − u2σ ) uσσ + −2v uσ vσ + u(vσ2 − u2σ ) vσσ ~ curl N = −2 (u2σ + vσ2 )2 (C.6) which illuminates the fact that the curl vanishes particularly at the zeros of F or of the ~ vanishes on the whole real axis if F (σ) = u(σ, 0) or on derivative F 0 . Moreover, curl N the whole imaginary axis if F (iτ ) = iv(0, τ ). We give now a short proof of the last two statements. Proof. The proofs are based on the property that the partial derivative of a function f (x, y) with respect to one of the independent variables fulfills ∂f (x, y) ∂f (x, 0) = . (C.7) ∂x y=0 ∂x (i) If F (σ) = u(σ, 0), then follows that the imaginary part v(σ, 0) = 0 and therefore the derivatives vσ (σ, 0) and vσσ (σ, 0) are zero due to eq. (C.7). Hence, all terms in the numerator in eq. (C.6) vanish on the real axis. (ii) For F (iτ ) = iv(τ ), eq. (C.7) yields uτ (0, τ ) = 0 and uτ τ (0, τ ) = 0. Additionally, we get uτ = −vσ and uτ τ = −uσσ from the Cauchy-Riemann differential equations. ~ vanishes on the imaginary axis. Thus, curl N C.1.2. Origin of separatrices The structure of the Newton flow in the examples above can also explain the flow of more complicated functions like higher order polynomials [41] or the Riemann zeta function. Indeed, the flow lines of the polynomials Fd (s) ≡ s2 − 1 and Ft (s) ≡ s3 − 1 (C.8) shown in fig. C.2 all start at infinity and end in one of the zeros sd1/2 = ±1 or stk = e 2πi k/3 , k = 0, 1, 2, of Fd or Ft , respectively. Yet, there have to be lines separating the flow towards the different zeros. In the case of the function Fd (left picture 97 C. Continuous Newton method Figure C.2.: Newton flow of the functions Fd (left) and Ft (right), eq. (C.8). The separatrices, indicated by the green dashed lines, separate the flow into the zeros sd1/2 = ±1 or stk = e 2π i k/3 , k = 0, 1, 2, of Fd or Ft , respectively. They cross each other in the origin, marked by a green point, where the first derivative of the functions vanishes. Since Fd (0) = −1 = Ft (0), the phase on the separatrices is π in these examples. of fig. C.2), these lines are obviously given by the imaginary axis on which the flow is always directed to the origin. There, it looses its forward uniqueness and can either continue to the zero at +1 or to the zero at −1. Hence, we call such a crossing point hyperbolic point [41] and refer to the crossing trajectories as separatrices. We proof now: (i) The Newton flow lines can only cross at points s0 where the first derivative F 0 (s0 ) vanishes, and (ii) the number of separatrices through one of these hyperbolic points is determined by the degree of the first non-vanishing derivative of the function. Proof (i). Suppose, s0 is a hyperbolic point and F 00 (s0 ) 6= 0. Then, 1 2 F (s) ∼ = F (s0 ) + F 0 (s0 ) (s − s0 ) + F 00 (s0 ) (s − s0 ) 2 is the Taylor expansion up to second order at this point. Using eq. (C.2) on the left hand side yields the approximate behavior s 0 (s ) F F 0 (s0 ) 2 F (s0 ) 0 ∼ s1/2 (t) = s0 − 00 ± − 2 00 (1 − e −t ) 00 F (s0 ) F (s0 ) F (s0 ) of two trajectories around s0 ≡ s(t0 ). Since s0 is the crossing point of the two trajectories, s1 (t0 ) = s2 (t0 ) = s0 must be fulfilled which requires that the square root and the fraction F 0 /F 00 vanish independently at t = t0 . Hence, the first derivative F 0 (s0 ) must be zero and t0 = 0. Proof (ii). Suppose, that m > 1 is the order of the first non-vanishing derivative of F . Then, the Taylor expansion around the hyperbolic point s0 reduces to 1 (m) F (s) ∼ F (s0 ) (s − s0 )m = F (s0 ) + m! 98 C.2. Further examples and the approximate behavior of the separatrices is given by 2πi k/m ∼ m! sk (t) = s0 + e F (s0 ) F (m) (s0 ) 1−e −t 1 m (C.9) with k = 0, . . . , m − 1. Hence, there are m separatrices crossing in s0 . Since the point s0 with F 0 (s0 ) = 0 in eq. (C.9) is reached at t = t0 = 0, negative t reflect the incoming part of the separatrix whereas positive t mark the outgoing part. Investigating the near neighborhood of the point s0 , that is |t| 1, we find that the separatrices, eq. (C.9), are approximately given by where F (s0 ) 1/m 1/m iφ 2πi k/m ∼ m! (m) sk (t) = s0 + e · e ·t , F (s0 ) Im F (s0 )/F (m) (s0 ) 1 . φ≡ arctan m Re F (s0 )/F (m) (s0 ) Thus, the outgoing parts of the separatrices leave s0 under the angles φ + 2πk/m while the incoming parts enter with φ + 2πk/m + π/m, because t = −|t| there. In fig. C.2, the separatrices are straight lines, confirming the properties of separatrices discussed before. Since the origin (green dot) is the only point where the derivatives Fd0 and Ft0 vanish, the phase of the separatrices is π, indicated by dashed green lines. Moreover, we see that the incoming parts of the separatrices divide the flows into different zeros whereas the outgoing parts separate flows from different directions. We emphasize that separatrices are not necessarily real. Recalling that Newton trajectories are lines of constant phase, it is obvious that the phase of the separatrix is defined by the phase of the function F at the hyperbolic point on this line. Therefore, we mark hyperbolic points and the corresponding separatrices in green. It remains to mention that zeros with multiplicity larger than one act like simple zeros. Although the first derivative vanishes at the zero, it does not create a separatrix. Indeed, the solution of s0 (t) = s(0) e −t of eq. (C.1) for F0 ≡ s differs only by a factor 1/2 in the exponential from the solution sd0 (t) = s(0) e −t/2 for Fd0 ≡ s2 . Hence, the Newton flow approaches a zero with multiplicity 2 with half of the speed. C.2. Further examples This section is dedicated to the description of the Newton flow of the functions χ and Fe which govern the asymptotic behavior of the Newton flow of ζ shown in section 2.2. Since the Newton flow of both functions is symmetric to the real axis, we restrict the description to the upper half of the complex plane. 99 C. Continuous Newton method Figure C.3.: The left picture shows the Newton flow of the function χ, eq. (C.10). The separatrices through the hyperbolic points on the real axis have phase 0 (solid green lines) and π (dashed green lines) in alternating sequence. The separatrix through the non-trivial zero of χ0 at σ = 1/2 and τ ∼ = 6.23 consists of curves with phase ∼ π/4 (dot-dashed green lines) and divides the complex space into the different regions listed in section C.2.1. The violet lines are additional real lines with positive (solid) or negative (dashed) values and the orange line indicates σ = 1/2. The right picture shows the Newton flow of the function Fe , eq. (C.11). There, all flow lines start at −∞ and end eventually at a zero with s = ±i(2k + 1)π/ ln 2. The incoming parts of the separatrices cross the complex plane parallel to the real axis, ending in the zeros of Fe0 at +∞ with τ = ±i2k π/ ln 2, marked by green circles. They are positive real lines since Fe = 1 at the hyperbolic points. The fact that the only lines which start at +∞ are also real and positive suggest that they are the outgoing trajectories of the hyperbolic points. Therefore, we have marked them in green with green arrows emphasizing the exceptional starting point at +∞. We show only the upper part of the complex plane in the pictures, since the Newton flow is in both examples symmetric with respect to the real axis. 100 C.2. Further examples C.2.1. The function χ As recalled in appendix A.2, the function 1 π s− 2 Γ 1−s 2 χ(s) ≡ Γ 2s (C.10) has zeros at the origin and the negative even integers whereas the poles are located at the positive odd integers. Moreover, a hyperbolic point appears between every two neighboring zeros or poles on the real axis and an additional hyperbolic point is located on the critical line σ = 1/2 at τ ∼ = 6.23, which we call non-trivial zero of χ0 in analogy to the non-trivial zeros of ζ. The left picture of fig. C.3 shows the Newton flow of χ. There, the dot-dashed green line marks the non-trivial separatrices with phase ∼ π/4 through the non-trivial zero of χ0 . They divide the Newton flow in four regions: (i) On the left of the non-trivial zero, the flow lines from −∞ end in the trivial zeros with s0 ≤ −4, while (ii) on the right, the ones ending at +∞ start at the poles with sp ≥ 5. (iii) The flow below the outgoing non-trivial separatrices, that is around the origin, connects the pole sp = 1 to the zero s0 = 0 as well as sp = 3 to s0 = −2 and sp = 5 to s0 = −4. (iv) Above the non-trivial separatrix, the flow crosses the complex plane from −∞ to +∞. Additionally, the separatrices through the hyperbolic points on the real axis are indicated by solid and dashed green lines for phase 0 or π, respectively. They direct the flow in region (i) into the different zeros and separate the flow from the different poles in region (ii). In region (iii), the separatrices connect the hyperbolic points σ ∼ = 1.45 and σ∼ = −0.45 (phase π) as well as σ ∼ = 3.75 and σ ∼ = −2.75 (phase 0). Hence, they enclose the flow between the pole and the corresponding zero sp = 1 and s0 = 0 or sp = 3 and s0 = −2. Finally, we note that there are no separatrices in region (iv). We have marked there the real lines in violet (dashing indicates again the phase π) to emphasize their difference to the separatrices of ζ discussed in section 2.2.2. C.2.2. The function Fe The function Fe (s) ≡ 1 + e −s ln 2 (C.11) contains the first two summands of the Dirichlet series of the Riemann zeta function, eq. (4.1). It possesses zeros at ±i(2k + 1) π/ ln 2, k ∈ N, and zeros of Fe0 for σ → +∞ and τ = ±i2k π/ ln 2, indicated by green circles in the right picture of fig. C.3. 101 C. Continuous Newton method The picture shows that all Newton flow lines start at −∞ and eventually end in a zero of Fe . However, it seems that the green lines are exceptions since they are parallel to the real axis. Yet, we can identify them as incoming and outgoing parts of the separatrices: The incoming parts are obviously the green lines starting at −∞. They lead straight into a zero of Fe0 at +∞ and separate the flows into neighboring zeros. Since Fe = 1 in the hyperbolic points, the phase of the separatrices is zero. The other green lines start at +∞, but already at the same height as the zero of Fe in which they end. Nevertheless, the phase vanishes on these lines and they separate the flow above the zero from the one below. Hence, we refer to them as the outgoing part of the separatrices. Since they are the only lines which come from +∞, we have marked them by green arrows. C.3. Technical details The pictures of the Newton flow are produced with Mathematica [47] and the commands [41]: F [s ] := ... N 1[σ , τ ] := −Re[F [σ + Iτ ]/F 0 [σ + Iτ ]] N 2[σ , τ ] := −Im[F [σ + Iτ ]/F 0 [σ + Iτ ]] StreamPlot[{N 1[σ, τ ], N 2[σ, τ ]}, {σ, σ0 , σ1 }, {τ, τ0 , τ1 }] for the intervals σ ∈ [σ0 , σ1 ] and τ ∈ [τ0 , τ1 ]. Here, I defines the imaginary unit i. The zeros, poles and hyperbolic points are marked afterwards. C.4. Newton flow of ζ Figure C.4 shows the Newton flow of the Riemann zeta function up to an imaginary part of τ = 300. The separatrices are depicted in green and additional real curves in violet to emphasize their difference to the (real) ‘bones’ analyzed by Arias-de Reyna in his ‘x-ray’ [36]. Solid and dashed lines indicate phase 0 and π, respectively, whereas the phase of dot-dashed curves is determined by the hyperbolic point in which they cross. The red dots mark the zeros of ζ [58, 59] and green dots the hyperbolic points s00 listed in tables C.1 - C.3. All values have been calculated with Mathematica [47]. C.5. Zeros of the derivative ζ 0 In tab. C.1, we present the trivial zeros of the derivatives ζ 0 and χ0 . The difference between the values decreases for zeros with smaller real parts. Table C.2 contains the hyperbolic points s00 of the Riemann zeta function and the phase at these points. Since the phase is quite small – the absolute values are smaller 102 C.5. Zeros of the derivative ζ 0 Figure C.4.: Newton flow of ζ for imaginary parts up to τ = 300. 103 C. Continuous Newton method than 0.37 < 0.12 π for s00 with imaginary part up to τ = 100 – the non-real separatrices of ζ are close to its real lines. This explains the ‘attraction’ of the real lines by the hyperbolic points observed in [36]. Finally, tab. C.3 lists the hyperbolic points s00 with imaginary parts 100 < τ < 300, used in fig. C.4. s00 s0χ -2.7173 -2.7482 -4.9368 -4.9425 -7.0746 -7.0758 -9.1705 -9.1707 Table C.1.: The trivial zeros s00 of the derivative ζ 0 of the Riemann zeta function are located slightly to the right of the trivial zeros s0χ of χ0 . However, the difference between the values vanishes for σ → −∞. s00 arg{ζ(s00 )} s00 arg{ζ(s00 )} 2.4632 + i 23.298 0.03317 1.7743 + i 71.528 0.13360 1.2865 + i 31.708 0.01551 0.8646 + i 76.363 -0.26841 2.3076 + i 38.490 -0.09364 1.3285 + i 78.662 0.32082 1.3828 + i 42.291 0.15913 1.2036 + i 83.669 -0.23365 0.9647 + i 48.847 -0.15736 2.3940 + i 85.802 -0.02503 2.1017 + i 52.432 0.11157 0.8641 + i 88.178 0.30930 1.8960 + i 57.135 -0.15093 1.3041 + i 93.086 -0.36793 0.8487 + i 60.141 0.25713 0.7806 + i 95.293 0.17932 1.2073 + i 65.920 -0.25909 1.7984 + i 98.827 0.04971 1.8330 + i 68.611 0.07686 Table C.2.: Zeros s00 of the derivative ζ 0 and phase of ζ at these points. The absolute value of the phase is smaller than 0.37 < 0.12 π for s00 with imaginary part up to τ = 100. 104 C.5. Zeros of the derivative ζ 0 s00 s00 s00 s00 1.7671 + i 101.715 1.7758 + i 161.022 1.2142 + i 212.197 0.9150 + i 259.116 1.1606 + i 104.503 1.5504 + i 162.666 0.8772 + i 213.967 1.2240 + i 263.814 1.0925 + i 106.561 1.2010 + i 166.088 1.0786 + i 215.738 0.7917 + i 266.015 0.6356 + i 111.431 1.4141 + i 167.894 1.0282 + i 219.499 0.9555 + i 267.419 1.9060 + i 113.631 0.6455 + i 169.538 0.6210 + i 221.083 1.7119 + i 269.937 1.4473 + i 115.583 0.9288 + i 173.927 2.5097 + i 223.403 1.4125 + i 271.067 1.6571 + i 118.908 1.3442 + i 175.667 0.7463 + i 224.514 1.3726 + i 273.576 1.0226 + i 121.999 1.8222 + i 177.609 1.3869 + i 227.605 0.6955 + i 275.993 0.8478 + i 123.715 1.1700 + i 179.331 0.9281 + i 243.589 1.4412 + i 277.527 1.3792 + i 127.965 1.4466 + i 182.222 1.2864 + i 229.739 0.7509 + i 278.819 1.0570 + i 130.322 0.6160 + i 185.215 0.6305 + i 231.618 0.6408 + i 282.802 2.3285 + i 132.486 1.0499 + i 186.713 1.0561 + i 233.285 1.3150 + i 284.268 0.8563 + i 134.194 1.3982 + i 189.240 0.9255 + i 237.016 0.9281 + i 243.589 1.0433 + i 138.659 0.7739 + i 192.517 1.6714 + i 238.512 2.0256 + i 285.766 0.9438 + i 140.470 2.3054 + i 194.119 1.2642 + i 240.308 0.9911 + i 287.326 1.3056 + i 142.650 1.2421 + i 195.912 1.6496 + i 242.120 1.2213 + i 289.253 1.0188 + i 146.631 0.8109 + i 197.546 0.7282 + i 247.541 1.1848 + i 292.170 2.4238 + i 147.875 0.8654 + i 201.743 1.1059 + i 248.920 0.9719 + i 294.038 0.6629 + i 150.486 1.3098 + i 203.282 1.0648 + i 250.524 0.5921 + i 295.290 1.2660 + i 152.613 0.8653 + i 204.888 1.3697 + i 252.989 1.3321 + i 297.979 0.9670 + i 156.633 1.4953 + i 208.247 0.8258 + i 255.813 0.8634 + i 158.283 1.7404 + i 209.544 2.3571 + i 256.901 Table C.3.: Non-trivial zeros of ζ 0 with imaginary part 100 < τ < 300. 105 D. Phase space representations From the variety of possible functions describing a quantum mechanical state in phase space [20, 21, 28, 60–63], we concentrate on the function Z∞ i 1 ξ ξ dξ e − ~ pξ hx + | ρ̂ |x − i Wρ̂ (x, p) ≡ (D.1) 2π~ 2 2 −∞ introduced by Wigner in 1932, which he had ‘chosen from all possible expressions, because it seems to be the simplest’ [18]. It allows a quasi-probabilistic interpretation of the density operator ρ̂ and can be reconstructed from experiment, e.g. [64–66]. However, this definition can be generalized to non-Hermitian operators and even to coupled systems leading to Moyal functions and Wigner matrices, respectively. We give in this appendix a short overview over the properties of the different representations and finish with a detailed description of the Wigner and Moyal functions of the Fock states which are the basis of the states defined in this work. D.1. Properties of the Wigner function A special property of the Wigner function (D.1) is that it allows to calculate the probability distribution in one of the conjugate variables x and p by integrating over the other, that is Z∞ hx|ρ̂|xi = dp Wρ̂ (x, p) (D.2) −∞ for the position distribution and Z∞ hp|ρ̂|pi = dx Wρ̂ (x, p) (D.3) −∞ for the momentum distribution of the operator ρ̂. Hence, the trace of this operator Z∞ Z∞ Tr ρ̂ = dx dp Wρ̂ (x, p) (D.4) −∞ −∞ is given by integration over the whole phase space, as easily follows from eq. (D.2) as well as from eq. (D.3). The overlap Z∞ Z∞ Tr (ρ̂1 ρ̂2 ) = 2π~ dx dp Wρ̂1 (x, p) Wρ̂2 (x, p) (D.5) −∞ −∞ 107 D. Phase space representations of two operators ρ̂1 and ρ̂2 can be calculated from the product of their Wigner representations Wρ̂i , i = 1, 2. Obviously, the Wigner function is not a classical probability distribution. Indeed, if two quantum states are orthogonal to each other the left hand side of eq. (D.5) becomes zero. But positive definite Wigner functions on the right hand side would only allow strictly positive outcomes. Hence, the Wigner functions must become negative in some parts of the phase space. Therefore, they are called quasi probability functions to emphasize their non-classical behavior. The Wigner function is normally used to describe a quantum state in phase space, that is the density operator ρ̂ is Hermitian, ρ̂ = ρ̂† , and normalized to unity, Tr ρ̂ = 1. In this case, the Wigner function (D.1) becomes real and normalized to unity, eq. (D.4). However, equations (D.2) – (D.5) hold for any ρ̂ since their proofs do not involve the properties of the operator (see for example [28]). We will see in the next section that it is useful to investigate Wigner functions of non-Hermitian operators, which are called Moyal functions. D.2. Moyal functions We define the Moyal function [20, 21] 1 W|ψihφ| (x, p) ≡ 2π~ Z∞ i ξ ξ dξ e − ~ pξ hx + |ψihφ|x − i 2 2 (D.6) −∞ as the generalization of the Wigner function (D.1) for two arbitrary states |ψi and |φi. Hence, the Moyal function is complex, W|φihψ| = W|ψihφ| and reduces for |φi = |ψi to the Wigner function W|ψihψ| ≡ W|ψi of the state |ψi. As mentioned already in the previous section, eqs. (D.2) and (D.5) hold for any operator ρ̂. Thus, we find for ρ̂ ≡ |ψihφ| that the normalization (D.4) simplifies to Z∞ Z∞ dx dp W|ψihφ| = hφ|ψi , −∞ −∞ the scalar product of the two states – called ‘self-orthogonality’ in [20] – while the overlap, eq. (D.5), is given by the expression Z∞ Z∞ 2π~ dx dp W|ψihα| W|γihφ| = Tr (|ψihα| |γihφ|) = hφ|ψihα|γi . −∞ (D.7) −∞ Since we have not used any information about the states, these equations are even valid for unnormalized states. Needless to say, in the case of pure density operators, that is for |φi = |ψi and |γi = |αi, the formulae reduce to the familiar ones of the Wigner functions. Until now, the density operator describes only one system, for example a harmonic oscillator, but not a combination of two interacting systems. But with the help of 108 D.3. Wigner matrix definition (D.1), we can even generalize the formalism to such systems, needed to describe the states derived in this thesis. D.3. Wigner matrix Following the example of Wallentowitz [19], we define the Wigner matrix for the density operator of two interacting systems A and B. Therefore, we use definition (D.1) on the whole density operator with the modification that position and momentum apply only to system B. Lets consider the density operator XX %̂AB ≡ %ab,a0 b0 |a, biha0 , b0 | (D.8) a,a0 b,b0 expanded in the orthonormal basis {|ai} and {|bi} of the systems A and B, respectively. Definition (D.1) now leads to a Wigner matrix X Waa0 |aiha0 | Ŵ%̂AB ≡ (D.9) a,a0 with the Wigner matrix elements Waa0 ≡ X %ab,a0 b0 W|bihb0 | , (D.10) b,b0 where the Moyal functions W|bihb0 | are given by eq. (D.6). It contains all information of the motion of system B depending on system A’s degrees of freedom. Tracing over system A leads to the Wigner function of the reduced density operator ρ̂B ≡ TrA %̂AB and therefore a loss of all information about system A. Indeed, the Wigner function of the reduced system ρ̂B is given by X Wρ̂B = TrA Ŵ%̂AB = Waa , (D.11) a which only contains the diagonal matrix elements of the Wigner operator (D.9). This representation clearly shows that information about the entanglement between the two systems in %̂AB is contained in the off-diagonal matrix elements of the Wigner matrix. Since the density operator of the two systems %̂AB is Hermitian, the Wigner matrix is Hermitian, too, and therefore Waa0 = Wa0a . Moreover, the normalization TrAB %̂AB = TrA ρ̂A = TrB ρ̂B = 1 of the density operator and the fact that TrA Ŵ%̂AB = Wρ̂B, eq. (D.11), leads with eq. (D.4) to the normalization condition Z∞ Z∞ Z∞ Z∞ X Waa = 1 dx dp TrA Ŵ%̂AB = dx dp −∞ −∞ −∞ −∞ a for the Wigner matrix. 109 D. Phase space representations Furthermore, we find the product of two density operators %̂AB and σ̂AB from Z∞ Z∞ TrAB (%̂AB σ̂AB) = 2π~ dx dp TrA Ŵ%̂AB Ŵσ̂AB −∞ Z∞ = 2π~ −∞ −∞ Z∞ dx −∞ dp X Waa0 Ω aa0 . (D.12) a,a0 Here, Waa0 and Ωaa0 denote the matrix elements of the Wigner operators Ŵ%̂AB and Ŵσ̂AB, defined by eq. (D.10), of the density operators %̂AB and σ̂AB , respectively. In the last step, we made use of definition (D.9) to evaluate the trace. We now show the accuracy of this equation. Proof. The left hand side of eq. (D.12) reads with definition (D.8) X X TrAB (%̂AB σ̂AB) = TrAB %ab,a0 b0 %αβ,α0 β 0 |a, biha0 , b0 |α, βihα0 , β 0 | a,a0 b,b0 α,α0 β,β 0 = XX a,a0 %ab,a0 b0 %a0 b0 ,ab . (D.13) b,b0 On the right, we arrive with eq. (D.10) at X a,a0 Z∞ Z∞ Z∞ Z∞ XX %ab,a0 b0 % aβ,a0 β 0 2π~ dx dp W|bihb0 | W|βihβ 0 | . 2π~ dx dp Waa0 Ω aa0 = −∞ a,a0 b,b0 β,β 0 −∞ −∞ −∞ Recalling that W|βihβ 0 | = W|β 0 ihβ| and that the Moyal functions fulfill eq. (D.7), this expression becomes XX XX %ab,a0 b0 %̄aβ,a0 β 0 δβ 0 b0 δβb = %ab,a0 b0 %̄ab,a0 b0 . (D.14) a,a0 b,b0 β,β 0 a,a0 b,b0 Since %̂AB is Hermitian, that is %̄ab,a0 b0 = %a0 b0 ,ab , eq. (D.14) is equal to eq. (D.13). Needless to say, tracing over system A after multiplication of the Wigner operators, eq. (D.12), is not the same as if we multiply the Wigner functions of the reduced systems, eq. (D.11), since this product would only consist of the diagonal matrix elements of the Wigner matrices. D.4. Fock states Since in this work all states are given as expansions in the Fock basis, we now have a closer look on their phase space distributions [28]. 110 D.4. Fock states Figure D.1.: Since the Wigner functions of the Fock states are rotational symmetric to the origin, we have cut the pictures at p = 0 to emphasize the radial structure. The Wigner functions W|ni , eq. (D.20), with even n have a maximum at the origin while the ones with odd n have a minimum, each followed by n zeros in radial direction. The height of the wells decreases due to the exponential envelope. D.4.1. Wigner and Moyal functions With the help of the position representation r 2 κ − κ2 x2 √ H (κx) e hx|ni = n 2 n n! π (D.15) of thepFock states |ni, where Hn denote the Hermite polynomials and the inverse length κ ≡ (M ω)/~ contains the mass M and frequency ω of the harmonic oscillator, and the substitution ip 1 κx + α≡ √ (D.16) ~κ 2 we get from eq. (D.6) the Moyal functions W|nihm| e −2|α| = π~ 2 ( wnm (α) ᾱn−m wmn (α) The real function wnm (α) ≡ (−1)m r αm−n for n ≥ m for n < m . m! n−m n−m 2 Lm 4|α|2 n! (D.17) (D.18) contains the generalized Laguerre polynomials [50] Ln−m (x) m = m X k=0 n m−k (−x)k . k! (D.19) Hence, the shape of the phase space functions is governed by two contributions: (i) the envelope exp{−2|α|2 } and (ii) the real function wnm given by eq. (D.18). 111 D. Phase space representations Figure D.2.: The real parts of the Moyal functions of the Fock states, eq. (D.21), are symmetric in p if n − m is even, or antisymmetric for odd n − m. The star-shaped patterns are caused by the real part of the complex factor exp −i(n − m)β, the zero circles are determined by the shape of wnm . The imaginary parts of the Moyal functions look like the real ones but rotated by −π/2 n−m . Since wnm only depends on |α|, it is rotational symmetric around the origin of phase space. At the origin, it is positive for even m and negative for odd m, determined only by the term (−1)m since the generalized Laguerre polynomials, eq. (D.19), are always positive there. Moreover, the Laguerre polynomials have m different zeros for positive values x which create circles around the origin in phase space. Needless to say, the increase of the Laguerre polynomial to (±) infinity for x → ∞ is in W|nihm| suppressed by the envelope. Based on the previous observations about the function wnm , we see that the Wigner functions 2 e −2|α| W|ni (α) = wnn (α) (D.20) π~ are symmetric around the origin, where – depending on n – we have a maximum (n even) or minimum (n odd). Around that, we find the ridges and valleys caused by the minima and maxima of Ln and shaped by the envelope, as we see in fig. D.1. The zero circles around the origin caused by the zeros of the Laguerre polynomials also occur in the Moyal functions. However, when we rewrite eq. (D.17) for n > m 2 W|nihm| e −2|α| = wnm (α) |α|n−m e −i(n−m)β(α) π~ (D.21) we find an additional zero at the origin, due to the factor |α|n−m . Moreover, the phase factor exp{−i(n − m)β}, depending on the angle β ≡ arg α, scales the values of the real or imaginary part of W|nihm| with the factor cos[(n − m)β] or − sin[(n − m)β]. This π/2 π creates radial zero lines at β = (2k + 1) (n−m) and β = k n−m , k ∈ N, respectively. Hence, Re W|nihm| is symmetric in p while Im W|nihm| is antisymmetric. The star-shaped patterns of the real part of the Moyal functions are shown in fig. D.2. The shapes of the imaginary part are equal to the real part patterns rotated by −π/2 n−m . 112 D.4. Fock states N =2 N =3 N = 20 0.3 0.2 0.1 -4 -2 2 4 -4 2 -2 4 -8 -4 4 8 x Figure D.3.: Behavior along the x-axis of the absolute value of the Wigner function |W|N i |, eq. (D.20), compared to the sum SN −1 , eq. (D.22), for different N , indicated in red and blue, respectively. The shaded areas mark the regions where W|N i is negative. Since |W|N i | in these regions is always smaller than SN −1 , the sum SN is positive in the whole phase space. D.4.2. Sum over Wigner functions Due to the normalization of the Wigner function, the absolute value of W|1i around the origin is smaller than the Gauß shape of W|0i . This holds for the whole region where W|1i < 0. Thus, the sum of the two Wigner functions, W|0i + W|1i , is positive in the whole phase space. When we continue to sum up the Wigner functions W|ni , we find the same result: N X SN ≡ W|ni ≥ 0 . (D.22) n=0 Indeed, when we examine the absolute value of the Wigner function W|N i , given by the red line in fig. D.3, we see that it is smaller than the sum SN −1 (blue line) in the region where W|N i is negative (indicated by the shaded areas). Hence, SN ≡ SN −1 + W|N i is positive in the whole phase space. This is, of course, no mathematical proof, but there are hints which strongly suggest that eq. (D.22) holds for all N : (i) The value of the Wigner functions at the origin is +1 for even n and −1 for odd leading to S2N = 1 and S2N +1 = 0. (ii) The peak around the origin of |W|n+1i | is narrower than the peak of |W|ni |, which leaves a positive contribution from each summand W|2ni + W|2n+1i . Thus, SN is always positive around the origin. (iii) The higher order peaks of |W|N i | decrease exponentially in height for increasing |x| and become less important in comparison to the relatively large value of SN −1 as the pictures in fig. D.3 confirm. (iv) The last peak of W|N i for x → ∞ is always positive. From the positivity of the sum SN follows that the contribution of the sum N X n=0 W|ni (n + 1)σ in the Wigner function of the truncated Riemann state, eq. (5.6), is positive, too. 113 E. Entanglement In this appendix, we will show different ways to distinguish, if a pure state of the form X ψne |n, ei + ψng |n, gi (E.1) |ΨiAB ≡ n is entangled or not. Since we have used this compact notation for the state |Ψi to describe the Riemann zeta function, we return to it as often as possible. But for the sake of simplicity, we switch notation if it is convenient. E.1. Quick check A quantum state is entangled, if we cannot factorize it in a product of two states of the independent systems. The pure state |Ψi ≡ N c1 |α1 i|ei + c2 |α2 i|gi , (E.2) consisting of the atomic states |ei and |gi (of system A) and the normalized states |α1 i and |α2 i of system B, can only be written as a product state if |α1 i and |α2 i are linearly dependent, that is |α2 i = e iϕ |α1 i. Hence, the overlap fulfills |hα1 |α2 i|2 ≤ 1 (E.3) with equality only given for product states. The information how strong |Ψi is entangled can be extracted from so-called entanglement measures. For an introduction to entanglement measures see for example [67]. Since we are only interested in the question whether a state is entangled or not, we leave the question of how much for further investigations. E.2. Schmidt decomposition It can be shown [68] that every composite pure state |ΨiAB ≡ dA X dB X α β cαβ |αi|βi can be written in its Schmidt decomposition |ΨiAB ≡ k p X λi |ai i|bi i , (E.4) i=1 115 E. Entanglement where λi are the non-zero eigenvalues of the reduced density matrix ρ̂A. The states |ai i are the corresponding eigenstates and |bi i are orthonormal states, too. From the properties of the density matrix follows λi > 0 and k X λi = 1 . i=1 Moreover, we can show from the Schmidt decomposition, eq. (E.4), that the reduced density matrix ρ̂B has the same eigenvalues λi as ρ̂A. Another consequence of the Schmidt decomposition is that if and only if |ΨiAB and thus ρ̂AB is not entangled, the Schmidt decomposition has only one term with eigenvalue λ1 = 1 and the reduced density operators are pure. Hence, the Schmidt decomposition is a powerful tool to distinguish entangled states from product states. Now, we use eq. (E.1) to determine the eigenvalues of the reduced density matrix ρ̂A, with atom states belonging to system A and photon states to system B. Tracing over the photon states yields T ρ̂A ≡ TrB ρ̂AB = |ei, |gi A he|, hg| with the Hermitian matrix P A ≡ Pn n |ψne |2 ψne ψng ψne ψng n ≡ P 2 |ψng | P n a b b̄ 1 − a ! . Here, Tr A = 1 results from the normalization of |Ψi. Moreover, Hermiticity implies that the eigenvalues r 1 1 λ1/2 ≡ ± + |b|2 − a(1 − a) (E.5) 2 4 of the matrix A have to be real, leading to the inequality a(1 − a) ≤ |b|2 + 1 . 4 We note that in this expression equivalence produces degenerated eigenvalues λ1/2 = 1/2. However, as a density operator ρ̂A is positive semi-definite. Thus, the non-negativity of the eigenvalues, eq. (E.5), yields |b|2 ≤ a(1 − a) . (E.6) In the case of equality, we get λ2 = 0 and the Schmidt decomposition consists of a single non-zero term. Therefore, ρ̂A describes a pure state and ρ̂AB is disentangled. Needless to say, eq. (E.6) is equivalent to the condition Det A ≥ 0. 116 E.3. Wigner representation In terms of the probability amplitudes ψne and ψng , eq. (E.6) reads 2 X X X |ψmg |2 ψne ψng ≤ |ψne |2 n n (E.7) m which we use in section E.4 and E.5 to investigate the entanglement of the RiemannSiegel and Berry-Keating reference state. By using the correspondence X ψne/ng |ni ≡ N c1/2 |α1/2 i , n one can show that eq. (E.7) is equivalent to eq. (E.3). E.3. Wigner representation For the sake of completeness, we answer the question how entanglement effects the Wigner matrix of the state |Ψi, eq. (E.2). Therefore, we recombine the Wigner matrix elements of Ŵ|Ψi in W≡ Wee Weg Weg Wgg ! = N2 |c1 |2 W|α1 i c̄1 c2 W|α2 ihα1 | c1 c̄2 W|α1 ihα2 | |c2 |2 W|α2 i ! and integrate the trace of this matrix over the whole phase space. The result is always (2π~)−1 , since the Wigner functions W|α1 i and W|α2 i are normalized to unity. However, with the help of eqs. (D.5) and (D.7), we find that integration of the determinant of the matrix yields Z∞ Z∞ 2 2 2 2 2π~ dx dp Det W = N |c1 | |c2 | |hα1 |α2 i| − 1 ≤ 0 . −∞ (E.8) −∞ As in eq. (E.3), equality indicates a product state and negative values represent entanglement. Moreover, a direct evaluation of Det W for a product state results in a vanishing determinant, which can never occur for an entangled state since eq. (E.8) must be fulfilled. Thus, we get the equivalent formulations Det W = 0 ⇔ |α2 i = e iϕ |α1 i for the state |Ψi. 117 E. Entanglement E.4. Riemann-Siegel reference state We note in chapter 6 that the Riemann-Siegel reference state is only entangled on the critical line which is equivalent to the postulate that equality in eq. (6.10) is only given for n0 ≥ 2 if σ = 1/2. We now proof this statement. Proof. Lets substitute a ≡ σ − 1, b ≡ 1 − 2σ and xn ≡ n + 1 in eq. (6.10), that is nX 0 −1 xan n=0 !2 ≤ nX 0 −1 xa+b xa−b n m . (E.9) n,m=0 Obviously, both sides are equal if n0 = 1. For n0 = 2 we get a+b a−b a−b ≤ xa+b + x x + x 0 1 0 1 " −b # b x x0 a a 0 2a 2a x2a ≤ x2a + 0 + x1 + 2 x0 x1 0 + x1 + x0 x1 x1 x1 x0 . 1 ≤ cosh b ln x1 ⇔ ⇔ xa0 + xa1 2 The right hand side is only equal to one if the argument of cosh is zero, which is the case if b = 0 and therefore σ = 1/2. Otherwise, that is beside the critical line, cosh is larger than one. Lets now assume that equality in eq. (E.9) holds for a definite n0 > 2 only if σ = 1/2. Then follows for n0 + 1 nX 0 −1 xan + xan0 n=0 ⇔ ⇔ nX 0 −1 xan !2 xan !2 n=0 nX 0 −1 n=0 + 2 xan0 nX 0 −1 !2 xan n=0 + 2 xan0 nX 0 −1 n=0 xan ≤ ≤ ≤ nX 0 −1 xa+b n + xa+b n0 n=0 nX 0 −1 a−b xa+b n xm + n,m=0 nX 0 −1 n,m=0 ! xan0 nX 0 −1 xa−b m m=0 nX 0 −1 xan n=0 a−b xa+b n xm + 2 xan0 nX 0 −1 n=0 + " xa−b n0 xn0 xn xan cosh b ! + b ln xn0 xn xn0 xn −b # . Again, the cosh-term is larger than 1 for σ 6= 1/2 which makes the second term on the right larger than the second one on the left. Since the first terms already fulfill eq. (E.9), the statement that equality is only given for σ = 1/2 also holds for n0 + 1. 118 E.5. Berry-Keating reference state E.5. Berry-Keating reference state We show now that – in contrast to the Riemann-Siegel reference state – the BerryKeating reference state eq. (7.3) is always entangled. For this task, we use again eq. (E.7) which yields for the probability amplitudes en (σ, τ ) N en (σ, τ ) φne ≡ N B B and the inequality E (n, τ ) (n + 1)σ/2 en (σ, τ ) N en (2 − 3σ, τ ) γ (σ, τ ) φng ≡ N B B B E(n, τ ) (n + 1)(2−3σ)/2 n −1 2 nX B B −1 X E 2 (n, τ ) |E(n, τ )|2 |E(m, τ )|2 ≤ . (n + 1)1−σ (n + 1)σ (m + 1)2−3σ n=0 (E.10) n,m=0 e 4 (σ, τ ) N e 2 (σ, τ ) N e 2 (2−3σ, τ ) |γ (σ, τ )|2 Here, we have already neglected the factors N nB nB nB B on both sides. Applying the triangle inequality on the left hand side, that is n −1 2 n −1 B B X X E 2 (n, τ ) |E(n, τ )|4 ≤ , (n + 1)1−σ (n + 1)2−2σ n=0 n=0 and splitting the double sum on the right of eq. (E.10) into nX B −1 n,m=0 nX nX B −1 B −1 |E(n, τ )|4 |E(n, τ )|2 |E(m, τ )|2 |E(n, τ )|2 |E(m, τ )|2 = + (n + 1)σ (m + 1)2−3σ (n + 1)2−2σ (n + 1)σ (m + 1)2−3σ n=0 n,m=0 n6=m immediately reveals that the right hand side must always be larger than the left due to the terms with n 6= m. Hence, the Berry-Keating reference state is entangled for all σ and τ . 119 F. Normalization and proportionality factors We give here a short investigation of the normalization and proportionality factors appearing throughout this work. F.1. Truncated Riemann state The normalization, eq. (5.5), Nν (σ) ≡ ν−1 X n=0 1 (n + 1)σ !−1/2 (F.1) of the truncated Riemann states |σ, τ iν is involved in the description of the zeta function by the truncated alternating sum and the Riemann-Siegel formula in chapter 5 and 6. Moreover, the limit N∞ (σ) ≡ N (σ) = ζ −1/2 (σ) causes the restriction to σ > 1 of the Riemann states |σ, τ i in chapter 4. In fig. F.1, we show the behavior of Nν2 for different values ν in dependence on σ. The curves start at Nν (0) = 1/ν and increase until they reach the limit Nν (∞) = 1. Hence, the curve of Nν with larger summation limit is always below the ones with smaller ν. The lowest curve, indicated in black, displays the normalization N of the Riemann states which is only defined outside the critical strip. The dashed and the dotted gray line mark the values for σ = 1/2 and for σ = 1, respectively. Nν2 1.0 0.8 0.6 0.4 0.2 2 4 6 8 10 σ Figure F.1.: The curves of the normalization factor Nν2 , eq. (F.1), given for the values ν = 1 (orange), ν = 2 (blue), ν = 3 (purple) and ν = 20 (red), approach the limit one for large σ. Since they start at Nν2 (0) = 1/ν, the curves with larger ν are always below the ones with smaller ν. The lowest curve is given by the normalization N (black) which is only defined for σ > 1. The dashed and the dotted gray line mark σ = 1/2 and σ = 1, respectively. 121 F. Normalization and proportionality factors F.2. The function E(n, τ ) For a better understanding of the Berry-Keating formula, eq. (7.1), and the Berry states defined in chapter 7, we give here a short analysis of the behavior of the function # "r 1 τ n+1 E(n, τ ) ≡ Erfc (F.2) ln p τ 2 1−i 2π which involves the complementary error function Erfc, eq. (B.8) [56]. Since the behavior of E is the same for all τ , we show in fig. F.2 only a sketch of its absolute value and argument for a fixed value τ and the continuous p variable n with dots marking integer n. The vertical dashed orange line indicates n = τ /(2π) − 1, where E = 1/2. Thus, the first dot on the right of this line marks n0 . Both pictures show that for almost all n ≤ n0 the absolute value and the argument of E is constant, that is equal to one or zero, respectively. However, for n > n0 the absolute value is zero while the argument rotates clockwise in the complex plane. Only the values near the orange line deviate from this behavior. The property |E| = 0 for n > n0 results from the behavior of the complementary error function and was used by Berry and Keating [31] to cut the sum in ZBK, eq. (1.12), at nB ≥ n0 . Moreover, we show in section 7.3 that the probability amplitudes in the Berry state |φBinB , defined by eq. (7.4), are almost the same as for the initial truncated Riemann state |σ, 0inB , eq. (5.4), due to the fact that E = 1 for small values of n. The largest deviations between |φBinB and |σ, 0inB occur at the discontinuities τ = 2π n20 of the Riemann-Siegel formula, eq. (1.9), and fade between the discontinuities for increasing τ . This behavior becomes clear when we consider that between the jumps the dots marking the integers of n are shifted to the left on the curves shown in fig. F.2. Figure F.2.: Sketch of the function E, defined by eq. (F.2), for a fixed time τ in dependence on a continuous variable n with p p dots marking integers. The dashed vertical orange line indicates n = τ /(2π) − 1, where E( τ /(2π) − 1, τ ) = 1/2. The absolute value |E|, depicted on the left, is for almost all n < n0 equal to one and drops after a slight increase in the vicinity of n0 − 1 to zero. In contrast, the argument of E, pictured on the right, is at first zero, becomes slightly positive around n0 − 1 and then rotates clockwise in the complex plane for n on the right of the orange line. 122 F.3. Berry state F.3. Berry state The norm of the Berry state |φB(σ, τ )i, eq. (7.4), reads en (σ, τ ) ≡ N B "n −1 #−1/2 B X |E(n, τ )|2 n=0 (n + 1)σ . (F.3) Due to the τ -dependence of the function E, eq. (F.2), we get different normalization factors for different values of τ even if the summation limit nB ≥ n0 (τ ) is the same. The en2 for nB = n0 (τ ) chosen for different values of τ . Like left picture in fig. F.3 shows N B en2 increases to the limit one. However, it starts the curves of Nν2 depicted in fig. F.1, N B at "n −1 #−1 B X 2 2 en (0, τ ) = N |E(n, τ )| B n=0 e2 − N2 which is larger than NnB (0) = 1/nB if nB > 1. Moreover, the difference N nB nB between the normalizations of |φB(σ, τ )inB and |σ, τ inB are distinct at the jumps τ = 2πn20 of the Riemann-Siegel formula, eq. (1.9), and tend to zero for increasing τ until the next jump occurs, as the right pictures in fig. F.3 confirm. Of course, the difference fades also for increasing summation limit nB . e2 N n0 en2 − Nn2 N 0 0 1.0 0.3 0.2 0.8 0.1 0.6 0.003 0.4 0.002 0.2 0.001 2 4 6 8 10 2 4 6 8 10 σ en2 , eq. (F.3), for 1 < nB = n0 (τ ) at Figure F.3.: The curves depicting the normalization factor N 0 different values of τ approach the limit one for large σ, just like the norm Nν2 shown in fig. F.1. However, due to the τ -dependence of E, the curves are different for different values of τ and they start at larger values compared with the ones of Nn20 . The right pictures show the difference e 2 − N 2 between the normalizations. It is largest at the jumps τ = 2πn2 (solid lines) of the N n0 n0 0 Riemann-Siegel formula and tends to zero before the next jump at 2π(n0 + 1)2 . As expected, the difference also becomes smaller for increasing summation limit. The dashed and the dotted gray line mark σ = 1/2 and σ = 1, respectively. The parameters are chosen as follows: n0 = 2 (blue): τ = 8π ∼ = 25.1 (solid), τ = 30 (dashed), τ = 56 (dotted); n0 = 3 (purple): τ = 18π ∼ = 56.5 (solid), τ = 75 (dashed), τ = 100 (dotted); n0 = 20 (red): τ = 800π ∼ = 2513.3 (solid), τ = 2600 (dashed), τ = 2770 (dotted). 123 F. Normalization and proportionality factors F.4. Truncated alternating sum In section 5.3.1, the proportionality factor, eq. (5.10), ) M(ν,N (σ, τ ) ≡ NN (σ) Nν (σ) (21−σ−iτ − 1) a (F.4) appears, connecting the overlap C, eq. (5.8), with the truncated alternating sum ζN , eq. (5.7). For the sake of simplicity, we have chosen there ν = N in all pictures. Hence, (N,N ) (N ) we investigate now only the behavior of Ma ≡ Ma . (N ) The factor Ma displays a sinusoidal dependence on τ , as fig. F.4 shows. Moreover, its absolute value as well as the absolute value of its real and imaginary part decrease (N ) for increasing truncation parameter N or decreasing value σ. Hence, Ma yields very small values of the overlap C if we improve the approximation ζN of the Riemann zeta function by increasing the truncation parameter N . (N ) |Ma (σ, τ )| 0.3 0.2 0.1 2 4 6 8 10 12 n o (N ) Re Ma (σ, τ ) 2 4 τ n o (N ) Im Ma (σ, τ ) 0.2 6 8 10 12 τ 0.1 -0.1 2 4 6 8 10 12 τ -0.2 -0.1 -0.3 -0.2 (N ) Figure F.4.: The dependence on τ of the proportionality factor Ma (σ, τ ), defined by eq. (F.4) for ν = N , displays a sinusoidal shape for the absolute value as well as for the real and imaginary part. As expected, the values for N = 100, indicated by the red curves, are smaller than for (N ) N = 50 (black) and N = 30 (blue). The solid and dashed lines indicate Ma for σ = 2/3 and σ = 1/2, respectively, to emphasize that the curves of smaller σ are closer to the τ -axis. 124 F.5. Riemann-Siegel reference state F.5. Riemann-Siegel reference state Like Nν2 (see section F.1), the square of the normalization, eq. (6.8), 1 Nn0 (σ, τ ) ≡ q 1 + |γ(σ, τ )|2 with γ(σ, τ ) ≡ Nn0 (σ) χ(σ + iτ ) Nn0 (2 − 3σ) (F.5) of the Riemann-Siegel reference state |ΦRSi increases with increasing σ from values smaller than 1/2 to the limiting value 1, as the upper pictures of fig. F.5 shows for three different τ for each value n0 . However, the curves of Nn20 cross each other on the critical line where Nn20 = 1/2. In the Wigner matrix Ŵ|ΦRSi , eq. (6.15), the function γ appears as additional factor which inverts the behavior of Nn20 . Indeed, the absolute value |γ| Nn20 in the off-diagonal matrix elements (middle row of fig. F.5) increases for σ < 1/2 before it tends to zero. In contrast, the factor |γ|2 Nn20 in the diagonal matrix element connected to |gi (bottom row) starts at values larger than 1/2 before dropping to zero. Due to the fact that |γ| = 1 holds on the critical line independently of τ , the absolute values of the factors in all matrix elements are the same for σ = 1/2. F.6. Berry-Keating reference state For the Berry-Keating reference state |ΦBKi the normalization, eq. (7.6), reads 1 en (σ, τ ) ≡ q N B 1 + |γB(σ, τ )|2 with γB(σ, τ ) ≡ en (σ, τ ) N B en (2 − 3σ, τ ) N B χ(σ + iτ ) . (F.6) The structure of the formula is the same as for the norm Nn0 , eq. (F.5), of the Riemanneν of the Berry states, (F.3), are involved. Siegel reference state. But now the norms N However, for σ = √ 1/2 these norms cancel each other and we get the same value en (1/2, τ ) = 1/ 2 as for Nn (1/2). Moreover, N e1 (σ, τ ) = N1 (σ, τ ) since the factor N 0 B −1 e N1 = |E(0, τ )| is independent of σ and therefore γB = γ. en2 for nB = n0 (τ ) shown in the first row of fig. F.6 look like Although the curves of N B the ones for Nn0 depicted in fig. F.5, they are slightly different. The inset confirms that en − Nn is largest at the at the jumps τ = 2πn2 and fades in the interval the difference N 0 0 0 2 2πn0 ≤ τ < 2π(n0 + 1)2 and, of course, with increasing summation limits. The bottom pictures in fig. F.6 reveal that the factor |γB| is more restricting than |γ|. e 2 in the middle row do not exceed the value 1/2 and the ones in the The curves of |γB| N n0 en2 ≤ 1 for σ < 1/2. Again, all factors are the bottom row are limited to 1/2 ≤ |γB|2 N 0 same on the critical line. 125 F. Normalization and proportionality factors Figure F.5.: The factors appearing in the Wigner matrix elements of Ŵ|ΦRSi , eq. (6.11), differ in their behavior. The left pictures show the dependence on σ up to 10 while the right ones are magnifications of the critical strip. The first row depicts the factor Nn20 , eq. (F.5) in the diagonal matrix element connected to the excited state, which increases for increasing σ from values smaller than 1/2 to 1. However, the factor |γ| in the absolute value of the prefactor of the off-diagonal matrix elements (second row) causes a moderate increase for σ < 1/2 before the curves drop to zero. In contrast to that, |γ|2 in the prefactor of the diagonal matrix element connected to |gi generates a decreasing curve. The dashed gray line indicates the critical line σ = 1/2, where the prefactors are identical since |γ| = 1. The dotted gray line marks σ = 1. The curves are depicted for the following values: n0 = 1 (orange): τ = 2π ∼ = 6.3 (solid), τ = 10 (dashed), τ = 25 (dotted); n0 = 2 (blue): τ = 8π ∼ 25.1 (solid), τ = 30 (dashed), τ = 56 (dotted); = n0 = 20 (red): τ = 800π ∼ 2513.3 (solid), τ = 2600 (dashed), τ = 2770 (dotted). = 126 F.6. Berry-Keating reference state e2 N n0 1.0 0.5 0 10 0 05 2 4 6 8 10 σ -0 05 1.0 02 2 4 6 8 10 σ 2 4 6 8 10 σ -0 2 e2 |γB| N n0 -0 4 0.5 1.0 02 en2 |γB|2 N 0 -0 -0 -0 -0 -1 2 4 6 8 0 0.5 2 4 6 8 10 0.2 0.4 0.6 0.8 1.0 σ Figure F.6.: The factors in the Wigner matrix elements of Ŵ|ΦBKi , eq. (7.10), resemble in their en2 , eq. (F.6), connected to behavior the factors in Ŵ|ΦRSi , shown in fig. F.5: (i) The factor N 0 |ei (first row) increases to the limit 1. (ii) The absolute value of the one in the off-diagonal matrix element (second row) first increases and then drops to zero. (iii) The one connected to |gi (bottom row) only decreases and (iv) all curves cross each other on the critical line (dashed gray line) at the value 1/2. However, the higher resolution of the critical strip shown on the right reveals that |γB| changes the behavior of the factors especially for σ < 1/2: The factor in the off-diagonal matrix elements (middle picture) does not exceed the value 1/2 and (bottom) e 2 ≤ 1 for σ = 0, in contrast to the behavior of |γ| N 2 and |γ|2 N 2 depicted in fig. F.5. |γB|2 N n0 n0 n0 en2 − Nn2 , |γ | N en2 − |γ| Nn2 and |γ |2 N en2 − |γ|2 Nn2 In the left column, the inset pictures of N B B 0 0 0 0 0 0 emphasize the difference between the factors in Ŵ|ΦBKi and Ŵ|ΦRSi . The dotted gray line indicates the border of the critical strip σ = 1 and the colors represent the following values: n0 = 1 (orange): τ = 2π ∼ = 6.3 (solid), τ = 10 (dashed), τ = 25 (dotted); n0 = 2 (blue): τ = 8π ∼ = 25.1 (solid), τ = 30 (dashed), τ = 56 (dotted); n0 = 20 (red): τ = 800π ∼ = 2513.3 (solid), τ = 2600 (dashed), τ = 2770 (dotted). 127 G. Effective Hamiltonian We give here a short derivation of eq. (8.7) for the effective Hamiltonian. G.1. Time-dependent Hamiltonian Since our calculations are carried out in the interaction picture, we have to deal with a time-dependent Hamiltonian Ĥ(t). The solution of the time-dependent Schrödinger equation d i~ |ψ(t)i = Ĥ(t)|ψ(t)i dt for this Hamiltonian is formally given by |ψ(t)i = Û (t)|ψ(0)i , where the time-evolution operator i Zt dt0 Ĥ(t0 ) Û (t) ≡ T̂ exp − ~ 0 contains the time ordering operator h i T̂ Ĥ(t1 ) Ĥ(t2 ) = ( Ĥ(t1 ) Ĥ(t2 ) , t2 ≤ t1 Ĥ(t2 ) Ĥ(t1 ) , t1 < t2 . One can show [28, 45] that t tZν−1 Z Zt Zt 1 T̂ dt1 . . . dtν Ĥ(t1 ) · . . . · Ĥ(tν ) = dt1 . . . dtν Ĥ(t1 ) · . . . · Ĥ(tν ) . ν! 0 0 0 0 Hence, the time-evolution operator is approximately Û (t) ∼ = 1̂ − i ~ Zt 0 dt0 Ĥ(t0 ) − 1 ~2 Zt 0 dt0 Zt0 dt00 Ĥ(t0 )Ĥ(t00 ) (G.1) 0 up to second order in time. We will now use a special form of the Hamiltonian, following from the transformation into the interaction picture, to calculate Û (t). 129 G. Effective Hamiltonian G.2. Interaction picture and effective Hamiltonian The transformation of a Hamiltonian Ĥ ≡ Ĥ0 + Ĥint with the free Hamiltonian Ĥ0 and the interaction Hamiltonian Ĥint into the interaction picture is given by i i I Ĥint ≡ e ~ Ĥ0 t Ĥint e − ~ Ĥ0 t and yields a Hamiltonian of the form I Ĥint (t) ≡ X Ĥj σ̂ e −iνj t + Ĥj† σ̂ † e iνj t j in the case of the Jaynes-Cummings-Paul model. Here, σ̂ ≡ |gihe| is the Pauli operator. The first order term of the time-evolution operator, eq. (G.1), then contains the integral # " Zt † X Ĥj σ̂ Ĥj σ̂ † iν t −iνj t 0 0 e −1 − e j −1 I1 ≡ dt Ĥ(t ) = i νj νj 0 j and the second order contribution is proportional to I2 ≡ Zt 0 =i dt0 Zt0 dt00 Ĥ(t0 )Ĥ(t00 ) 0 t XZ k,j 0 dt0 " # Ĥk Ĥ † σ̂σ̂ † Ĥk† Ĥj σ̂ † σ̂ i(νk −νj )t0 0 0 0 j e − e iνk t − e −i(νk −νj )t − e −iνk t . νj νj Here, we have used the fact that σ̂ 2 = 0. When we now carry out the integration over t0 , we arrive at i X i h † Ĥj Ĥj σ̂ † σ̂ − Ĥj Ĥj† σ̂σ̂ † t νj j X 1 h † i † † iνj t † −iνj t − Ĥ Ĥ σ̂ σ̂ e − 1 + Ĥ Ĥ σ̂σ̂ e − 1 j j j j νj2 j " ! X e i(νk −νj )t − 1 e iνk t − 1 † † + Ĥk Ĥj σ̂ σ̂ − νj (νk − νj ) νk νj j6=k !# e −i(νk −νj )t − 1 e −iνk t − 1 † † − . + Ĥk Ĥj σ̂σ̂ νj (νk − νj ) νk νj I2 = The second and third contribution contain oscillatory terms and can be neglected since they are of the order O(ν −2 ) and therefore much smaller than the first term (O(ν −1 )). Thus, we only keep the contributions of I2 which are linear in time. 130 G.2. Interaction picture and effective Hamiltonian Moreover, the first order term I1 can be neglected in comparison to I2 since its contributions are constant, creating an overall phase, or oscillate much faster in time than the linear term of I2 . Hence, the time-evolution operator, eq. (G.1), is approximately given by † † † † it X Ĥj Ĥj σ̂ σ̂ − Ĥj Ĥj σ̂σ̂ . (G.2) Û (t) ∼ 1̂ − = ~ ~νj j Up to first order in time t, eq. (G.2) is equivalent to the time evolution produced by an effective Hamiltonian X Ĥj† Ĥj σ̂ † σ̂ − Ĥj Ĥj† σ̂σ̂ † Ĥeff ≡ ~νj j which transforms into i X 1 h † Ĥj Ĥj − Ĥj Ĥj† 1̂ + Ĥj† Ĥj + Ĥj Ĥj† σ̂z Ĥeff ≡ 2~νj j when we use 1̂ ± σ̂z = 2 ( |eihe| = σ̂ † σ̂ |gihg| = σ̂σ̂ † . This formula is employed to calculate the effective Hamiltonian of the Jaynes-CummingsPaul model and to approximate the Riemann Hamiltonian ĤR in chapter 8. 131 H. Matrix elements In chapter 8, we need the matrix elements hn + j|fˆ|ni to find an effective Hamiltonian which approximates the Riemann Hamiltonian ĤR , eq. (3.2). We will now give the calculation of these matrix elements for a polynomial operator fˆ ≡ f (x̂). H.1. Calculation with â and ↠When we recall that the position operator κ x̂ ≡ √ â + ↠2 for the harmonic oscillator is the sum of the annihilation operator â and the creation operator ↠, we immediately see that |hn ± j|x̂µ |ni|2 is a polynomial in n of degree µ (see tab. H.1 for µ = 0, . . . , 4). Therefore, we find for a polynomial operator fˆ ≡ f (x̂) = that µX max fµ x̂µ µ=0 2 max 2 µX µ ˆ fµ hn ± j|x̂ |ni hn ± j|f |ni = µ=0 is a polynomial of degree nµmax . A closer examination of |hn ± j|x̂µ |ni|2 shows that only matrix elements with j ≤ µ survive, that is 2 2 (H.1) hn ± j|fˆ|ni = |f0 + f1 g1 + f2 g2 + · · · + fµmax gµmax | , where gν ≡ ν 2 P pk δj,2k , k=0 ν+1 2 P pek δj,2k−1 , ν even ν odd . k=1 Here, pν denotes a polynomial of order ν and peν represents the square root of a polynomial of order 2ν − 1. The Kronecker deltas δj,ν cause a splitting of the absolute square in eq. (H.1) into 2 2 2 (H.2) hn ± j|fˆ|ni = |f0 + f2 g2 + . . . | + |f1 g1 + f3 g3 + . . . | 133 H. Matrix elements µ hn − j| (â + ↠)µ |ni hn + j| (â + ↠)µ |ni 0 1 2 3 4 δj,0 √ n δj,1 p (2n + 1) δj,0 + n(n − 1) δj,2 p 3n3/2 δj,1 + n(n − 1)(n − 2) δj,3 (6n2 + 6n + 3) δj,0 p +(4n − 2) n(n − 1) δj,2 p + n(n − 1)(n − 2)(n − 3) δj,4 √ n + 1 δj,1 p (2n + 1) δj,0 + (n + 1)(n + 2) δj,2 p 3(n + 1)3/2 δj,1 + (n + 1)(n + 2)(n + 3) δj,3 (6n2 + 6n + 3) δj,0 p +(4n + 6) (n + 1)(n + 2) δj,2 p + (n + 1)(n + 2)(n + 3)(n + 4) δj,4 Table H.1.: Matrix elements hn ± j| (â + ↠)µ |ni for n = 0, . . . , 4. the absolute square of the terms with µ odd and µ even, respectively. Furthermore, this splitting causes the first part of eq. (H.2) to be a polynomial of degree 2ν in n whereas the second part, containing the odd coefficients, is a polynomial of order 2ν − 1. H.2. General expression Here, we derive a general formula for the matrix elements which can be used to numerically calculate them. Therefore, we use again the series expansion f (x̂) ≡ ∞ X fµ x̂µ µ=0 of the operator fˆ. The matrix elements are then given by hn + j|f (x̂)|ni = ∞ X µ=0 µ fµ hn + j|x̂ |ni = 2n p κ π n! 2j (n + j)! ∞ X µ=0 fµ Ien,j,µ . (H.3) In the last step, we made use of the position representation of the Fock states, eq. (D.15), to express the integral by Ien,j,µ ≡ Z∞ 2 dx xµ Hn+j (κx) Hn (κx) e −(κx) . −∞ When we now substitute y = κx and use the familiar relation Hn (−x) = (−1)n Hn (x) of the Hermite polynomials to rewrite the negative part of the integral, we arrive at Ien,j,µ = 134 1 κµ−1 1 + (−1)µ+j In,j,µ . 2 H.2. General expression Therefore, Ien,j,µ vanishes if j is even and µ is odd or vice versa. The remaining integral In,j,µ Z∞ 2 ≡ 2 dy y µ Hn+j (y) Hn (y) e −y (H.4) 0 can be evaluated with the help of the Taylor series Hn (x) ≡ n X an,α xα α=0 of the Hermite polynomials. The expansion coefficients (α) an,α ≡ Hn (0) α! (H.5) contain the αth derivative of Hn . Integration over y in eq. (H.4) yields In,j,µ = n X α=0 an,α n+j X an+j,β Γ β=0 µ+α+β+1 2 (H.6) which vanishes for µ = 0 and j 6= 0 since the Hermite polynomials are orthogonal to each other [50], that is √ In,j,0 = π 2n−1 n! δj,0 . Finally, we can express the matrix elements, eq. (H.3), by ∞ X fµ 1 + (−1)µ+j p hn + j|f (x̂)|ni ≡ κ f0 δj,0 + In,j,µ 2 2n 2j π n! (n + j)! µ=1 κµ ∞ P f2µ In,j,2µ for j even κ2µ κ2 µ=0 = p (H.7) ∞ P f2µ+1 2n 2j π n! (n + j)! I for j odd κ2µ+1 n,j,2µ+1 2 κ2 µ=1 in terms of In,j,µ , eq. (H.6). Since the Hermite polynomials Hn are either even or odd, half of the coefficients an,α , eq. (H.5), are zero. Nevertheless, eq. (H.7) is quite useful to calculate the matrix elements numerically. 135 H. Matrix elements H.3. Explicit expression up to fourth order This section contains the explicit expressions of the matrix elements calculated with a polynomial operator fˆ of fourth order, used in section 8.2.2 to approximate the Riemann Hamiltonian. For µmax = 4, eq. (8.16) and eq. (8.17) read Fn(4) ≡ ε0 f˜0 + and Cn(4) ≡ respectively. The coefficients 4 X j=1 4 X j=1 h i ε(j+) Θ(n − j) f˜−j + f˜j h i ε(j−) Θ(n − j) f˜−j − f˜j , 2 f˜±j ≡ hn ± j|fˆ|ni denote the matrix elements. With the help of tab. H.1, we get for the even coefficients f˜0 = f˜−2 = f˜2 = f˜−4 = f˜4 = 2 3κ4 1 1 2 f2 + n +n+ f4 f0 + κ n + 2 2 2 2 κ4 n(n − 1) f2 + κ2 (2n − 1) f4 4 2 κ4 (n + 1)(n + 2) f2 + κ2 (2n + 3) f4 4 κ8 n(n − 1)(n − 2)(n − 3) f42 16 κ8 (n + 1)(n + 2)(n + 3)(n + 4) f42 , 16 2 that is polynomials of order O(n4 ), and for the odd coefficients 2 3κ2 κ2 ˜ n f1 + n f3 f−1 = 2 2 2 κ2 3κ2 ˜ f1 = (n + 1) f1 + (n + 1) f3 2 2 κ6 n(n − 1)(n − 2) f32 f˜−3 = 8 κ6 f˜3 = (n + 1)(n + 2)(n + 3) f32 8 polynomials of order O(n3 ). 136 Bibliography [1] M. Planat, P. Solé, and S. Omar Riemann hypothesis and quantum mechanics J. Phys. A: Math. Theor. 44, 145203 (2011) [2] E. Landau Handbuch der Lehre von der Verteilung der Primzahlen (Chelsea Publishing Company, New York, 1974) [3] B. Riemann Ueber die Anzahl der Primzahlen unter einer gegebenen Grösse Monatsberichte der Berliner Akademie (Nov. 1859) Transcribed German version and English translation by D. Wilkins see http://www.claymath.org/millennium/Riemann Hypothesis/1859 manuscript/ [4] E. C. Titchmarsh The theory of the Riemann zeta-function (Clarendon Press, Oxford, 1967) [5] A. A. Karatsuba and S. M. Voronin The Riemann zeta-function (De Gruyter, Berlin, 1992) [6] D. Schumayer and D. A. W. Hutchinson Colloquium: Physics of the Riemann hypothesis Rev. Mod. Phys. 83, 307 (2011) [7] M. V. Berry and J. P. Keating The Riemann zeros and eigenvalue asymptotics SIAM Review 41, 236 (1999) [8] G. Sierra and J. Rodrı́guez-Laguna H = xp model revisited and the Riemann zeros Phys. Rev. Lett. 106, 200201 (2011) [9] R. K. Bhaduri, A. Khare, and J. Law Phase of the Riemann ζ function and the inverted harmonic oscillator Phys. Rev. E 52, 486 (1995) [10] J.-B. Bost and A. Connes Hecke algebras, type III factors and phase transitions with spontaneous symmetry breaking in number theory Selecta Mathematica 1, 411 (1995) 137 Bibliography [11] F. Gleisberg, R. Mack, K. Vogel, and W. P. Schleich Factorization with a logarithmic energy spectrum New J. Phys. 15, 023037 (2013) [12] C. Weiss, S. Page, and M. Holthaus Factorising numbers with a Bose-Einstein condensate Physica A 341, 586 (2004) [13] R. Mack, J. P. Dahl, H. Moya-Cessa, W. T. Strunz, R. Walser, and W. P. Schleich Riemann ζ function from wave-packet dynamics Phys. Rev. A 82, 032119 (2010) [14] J. Twamley and G. J. Milburn The quantum Mellin transform New J. Phys. 8, 328 (2006) [15] C. Feiler and W. P. Schleich Entanglement and analytical continuation: an intimate relation told by the Riemann zeta function New J. Phys. 15, 063009 (2013) [16] E. T. Jaynes and F. W. Cummings Comparison of quantum and semiclassical radiation theories with applications to the beam maser Proc IEEE 51, 89 (1963) [17] H. Paul Induzierte Emission bei starker Einstrahlung Ann. Phys. 466, 411 (1963) [18] E. P. Wigner On the quantum correction for thermodynamic equilibrium Phys. Rev. 40, 749 (1932) [19] S. Wallentowitz, R. L. de Matos Filho, and W. Vogel Determination of entengled quantum states of trapped atom Phys. Rev. A 56, 1205 (1997) [20] J. E. Moyal Quantum mechanics as a statistical theory Math. Proc. Camb. Phil. Soc. 45, 99 (1949) [21] H. J. Groenewold On the principles of elementary quantum mechanics Physica 12, 405 (1946) [22] C. Feiler Zahlentheorie und Quantenmechanik - Die ζ-Funktion im Phasenraum Diploma Thesis (2007) 138 Bibliography [23] E. Schrödinger Die gegenwärtige Situation in der Quantenmechanik Die Naturwissenschaften 48, 807, ibid. 49, 823, ibid. 50, 844 (1935) English translation in: J. A. Wheeler and W. H. Zurek Quantum theory and measurement (Princeton University Press, Princeton, 1983) [24] W. Schleich, M. Pernigo, and Fam Le Kien Nonclassical state from two pseudoclassical states Phys. Rev. A 44, 2172 (1991) [25] M. Brune, E. Hagley, J. Dreyer, X. Maı̂tre, A. Maali, C. Wunderlich, J. M. Raimond, and S. Haroche Observing the progressive decoherence of the ‘meter’ in a quantum measurement Phys. Rev. Lett. 77, 4887 (1996) [26] S. Deléglise, I. Dotsenko, C. Sayrin, J. Bernu, M. Brune, J.-M. Raimond, and S. Haroche Reconstruction of non-classical cavity field states with snapshots of their decoherence Nature 455, 510 (2008) [27] C. Monroe, D. M. Meekhof, B. E. King, and D. J. Wineland A “Schrödinger cat” superposition state of an atom Science 272, 1131 (1996) [28] W. P. Schleich Quantum optics in phase space (Wiley-VCH, Berlin, 2001) [29] C. L. Siegel Über Riemanns Nachlaß zur analytischen Zahlentheorie Quellen und Studien zur Geschichte der Mathematik, Astronomie und Physik, Abt. B: Studien 2, 45 (1932) [30] Clay Mathematics Institute Millennium prize problems: Riemann hypothesis http://www.claymath.org/millennium/Riemann Hypothesis/ [31] M. V. Berry and J. P. Keating A new asymptotic representation for ζ( 12 + it) and quantum spectral determinants Proc. R. Soc. Lond. A 437, 151 (1992) [32] G. H. Hardy and J. E. Littlewood The zeros of Riemann’s zeta-function on the critical line Mathematische Zeitschrift 10, 283 (1921) [33] E. Jahnke and F. Emde Funktionentafeln mit Formeln und Kurven (Teubner, Leipzig, 1933) 139 Bibliography [34] A. A. Utzinger Die reellen Züge der Riemann’schen Zetafunktion (Dissertation, Zürich, 1934) [35] A. Speiser Geometrisches zur Riemannschen Zetafunktion Math. Ann. 110, 514 (1935) [36] J. Arias-de-Reyna X-ray of Riemann zeta function arXiv:math/0309433 (2003) [37] K. A. Broughan and A. R. Barnett The holomorphic flow of the Riemann zeta function Math. Comp. 73, 987 (2003) [38] J. W. Neuberger Steepest descent and differential equations J. Math. Soc. Japan 37, 187 (1985) [39] J. W. Neuberger Sobolev gradients and differential equations (Springer, Heidelberg, 1997) [40] J. W. Neuberger, C. Feiler, and W. P. Schleich Continuous Newton method, lines of constant phase and the Riemann zeta function To be published. [41] J. W. Neuberger Continuous Newton’s method for polynomials Math. Intell. 21, 18 (1999) [42] N. Levinson and H. L. Montgomery Zeros of the derivatives of the Riemann zeta-function Acta mathematica 133, 49 (1974) [43] S. Haroche and J.-M. Raimond Exploring the quantum: atoms, cavities and photons (Oxford University Press, Oxford, 2006) [44] D. Leibfried, R. Blatt, C. Monroe, and D. Wineland Quantum dynamics of single trapped ions Rev. Mod. Phys. 75, 281 (2003) [45] M. O. Scully and M. S. Zubairy Quantum optics (Cambridge University Press, Cambridge, 1997) [46] J. P. Dowling A quantum state of ultra-low phase noise Opt. Comm. 86, 119 (1991) 140 Bibliography [47] Wolfram Research, Inc. Mathematica Version 8.0 (Champaign, Illinois, 2010) [48] K. Vogel, V. M. Akulin, and W. P. Schleich Quantum state engeneering of the radiation field Phys. Rev. Lett. 71, 1816 (1993) [49] A. S. Parkins, P. Marte, P. Zoller, and H. J. Kimble Synthesis of arbitrary quantum states via adiabatic transfer of Zeeman coherence Phys. Rev. Lett. 71, 3095 (1993) [50] I. N. Bronstein, K. A. Semendjajew, G. Musiol, and H. Mühlig Taschenbuch der Mathematik (Harri Deutsch, Frankfurt am Main, 2000), 5 edn. [51] Z.-H. Li, D.-W. Wang, H. Zheng, S.-Y. Zhu, and M. S. Zubairy Effect of the counterrotating-wave terms on the spontaneous emission from a multilevel atom Phys. Rev. A 80, 023801 (2009) [52] R. L. de Matos Filho and W. Vogel Engineering the Hamiltonian of a trapped atom Phys. Rev. A 58, R1661 (1998) [53] W. Vogel and R. L. de Matos Filho Nonlinear Jaynes-Cummings dynamics of a trapped ion Phys. Rev. A 52, 4214 (1995) [54] P. H. Fuss (editor) Correspondance mathématique et physique de quelques célèbres géomètres du XVIIIème siècle (Saint Petersburg, 1843) Volume 1: Letter from L. Euler to C. Goldbach (1729) Volume 2: Letter from D. Bernoulli to C. Goldbach (1729) [55] E. T. Whittaker and G. N. Watson A course of modern analysis (University Press, Cambridge, 1969) [56] M. Abramowitz and I. A. Stegun Handbook of mathematical functions (Dover Publications, New York, 1972), 9 edn. [57] A. Sommerfeld Elektrodynamik (Verlag Harri Deutsch, Thun-Frankfurt, 1988) [58] A. M. Odlyzko The 1022 -nd zero of the Riemann zeta function Contemp. Math. 290, 139 (2001) Tables of zeros of the Riemann zeta function see http://www.dtc.umn.edu/∼odlyzko/zeta tables/ 141 Bibliography [59] X. Gourdon and P. Demichel The 1013 first zeros of the Riemann Zeta function, and zeros computation at very large height (2004) http://numbers.computation.free.fr/Constants/Miscellaneous/zetazeros1e131e24.pdf [60] K. Husimi Some formal properties of the density matrix Proc. Phys.-Math. Soc. Japan 22, 264 (1940) [61] Y. Kano A new phase-space distribution function in statistical theory of the electromagnetic field J. Math. Phys. 6, 1913 (1965) [62] R. J. Glauber Photon correlations Phys. Rev. Lett. 10, 84 (1963) [63] E. C. G. Sudarshan Equivalence of semiclassical and quantum mechanical descriptions of statistical light beams Phys. Rev. Lett. 10, 277 (1963) [64] D. Leibfried, D. M. Meekhof, B. E. King, C. Monroe, W. M. Itano, and D. J. Wineland Experimental determination of the motional quantum state of a trapped atom Phys. Rev. Lett. 77, 4281 (1996) [65] C. Kurtsiefer, T. Pfau, and J. Mlynek Measurement of the Wigner function of an ensemble of helium atoms Nature 386, 150 (1997) [66] M. Freyberger, P. J. Bardroff, C. Leichtle, G. Schrade, and W. P. Schleich The art of measuring quantum states Phys. World 10, 41 (Nov. 1997) [67] M. B. Plenio and S. Virmani An introduction to entanglement measures Quant. Inf. Comput. 7, 1 (2007) [68] M. A. Nielsen and I. L. Chuang Quantum computation and quantum information (Cambridge University Press, Cambridge, 2000) 142 Zusammenfassung Die Riemannsche Zetafunktion ζ bildet durch ihre enge Verknüpfung mit der Zahlentheorie nicht nur eine wesentliche Grundlage der Mathematik, sondern beeinflusst auch die Physik. Deshalb ist es nicht verwunderlich, dass neben mathematischen Ansätzen zum Beweis der Riemannschen Hypothese, welche die Lage der sogenannten nicht-trivialen Nullstellen von ζ beschreibt, auch physikalische Methoden herangezogen werden [6]. Ziel dieser Arbeit ist es, durch verschiedene mathematische wie physikalische Darstellungen der Zetafunktion Einblicke in ihr Verhalten zu erlangen. Dazu werden im ersten Kapitel die benötigten mathematischen Formeln zusammengetragen, die ζ exakt oder näherungsweise beschreiben. Das zweite Kapitel widmet sich der Darstellung der Zetafunktion mit Hilfe der kontinuierlichen Newton-Methode. Die daraus erhaltenen Newtonschen Flüsse veranschaulichen die Topologie der Zetafunktion durch Kurven konstanter Phase, welche die Lage des einzigen Pols sowie der Nullstellen der Funktion und ihrer ersten Ableitung in der komplexen Ebene besonders hervorheben. Deshalb wäre auch eine Verletzung der Riemann-Hypothese im Newtonschen Fluss sichtbar. In den folgenden Kapiteln wird ein physikalischer Zugang, basierend auf dem Überlapp hΦ|Ψ(t)i zweier quantenmechanischer Zustände, vorgestellt. Dabei werden die Phasen in den Summanden der verschiedenen Darstellungen von ζ mit Hilfe der Zeitentwicklung des verschränkten Zustands |Ψ(t)i realisiert, welcher analog den Zuständen des Jaynes-Cummings-Paul-Modells gewählt wurde: Das Feld eines Hohlraumresonators, beschrieben durch eine Superposition von Fock-Zuständen, wird mit einem Zwei-NiveauAtom gekoppelt. Die Wechselwirkung wird allerdings durch den Riemann-Hamiltonian ĤR ≡ ~ω ln(n̂ + 1) σ̂z gesteuert, der neben dem Pauli-Operator σ̂z , welcher das Vorzeichen der auftretenden Phasen bestimmt, eine logarithmische Abhängigkeit vom Photonenzahloperator n̂ aufweist. Es stellt sich im vierten Kapitel heraus, dass für Argumente s ≡ σ + iτ von ζ mit Realteil σ größer als eins auf Verschränkung mit dem Zwei-Niveau-Atom verzichtet werden kann. Die Riemann-Zustände |σ, τ i, die lediglich das Hohlraumfeld beschreiben, genügen, um die Dirichlet-Summe zu erzeugen, da diese nur die Phasen −iτ ln(n + 1) aufweist. Allerdings sind die Riemann-Zustände aufgrund ihrer Normierung ebenfalls auf den Bereich σ > 1 beschränkt. Mathematisch lässt sich der Definitionsbereich mittels einer alternierenden Summe auf alle positiven σ erweitern. Die daraus resultierenden Zustände sind aber weiterhin auf den Bereich σ > 1 beschränkt, solange man die exakte, also unendliche Summe darstellen möchte. Selbst Verschränkung hilft in diesem Fall nicht. Erst die Beschränkung auf eine endliche Anzahl an Fock-Zuständen erzielt das gewünschte Ergebnis. Die abgeschnittenen Riemann-Zuständen |σ, τ iν sind auch im Bereich 0 < σ < 1 definiert. Jedoch ist die 143 Zusammenfassung Näherung der Zetafunktion durch eine abgeschnittene alternierende Summe für kleine Abschneideparameter nicht besonders gut. Große Abschneideparameter führen hingegen zur Verkleinerung der Wahrscheinlichkeitsamplituden der einzelnen Fock-Anteile und somit zu sehr kleinen Überlappwerten. In den Kapiteln 6 und 7 verwenden wir deshalb die Riemann-Siegel sowie die BerryKeating-Formel als Näherung für ζ. Beide Darstellungen bestehen aus zwei endlichen Summen mit relativ kleiner Summationsgrenze, in deren Summanden sowohl die Phasen −iτ ln(n+1) als auch deren komplex Konjugierte +iτ ln(n+1) auftreten. Dies macht die Verwendung von verschränkten Zuständen zur physikalischen Darstellung von ζ nötig. Die Beträge der einzelnen Summanden unterscheiden sich in den beiden Näherungen nur minimal. Trotzdem führt dieser kleine Unterschied zu einer erheblichen Verbesserung der Näherung durch die Berry-Keating-Formel, die bereits den ersten Korrekturterm zur Riemann-Siegel-Formel enthält. Des Weiteren unterscheiden sichhdie Summationsgreni p zen: In der Riemann-Siegel-Formel wird bis zu einem Wert n0 ≡ τ /(2π) summiert, wobei mit [. . . ] die Gaußklammer gemeint ist. Deshalb springen die Funktionswerte sobald τ = 2π n20 erreicht ist. Diese Diskontinuitäten werden durch die Berry-KeatingFormel vermieden, deren Summationsgrenze lediglich nB ≥ n0 erfüllen muss. Bei der physikalischen Darstellung der beiden Formeln machen wir uns ihre Ähnlichkeit zu Nutze. Zur Realisierung der Phasenfaktoren genügt die Zeitentwicklung des Riemann-Siegel-Zustands |ΨRS(t)i, der an die Darstellung von Schrödingerkatzen erinnert. Wir koppeln allerdings zwei entgegengesetzt rotierende abgeschnittene RiemannZustände mit den beiden Atomzuständen, da diese eine logarithmische anstelle einer linearen Abhängigkeit der Phase von der Photonenzahl n aufweisen. Durch die Zeitentwicklung wird der anfangs unverschränkte Zustand |ΨRS(0)i verschränkt. Im Phasenraum wirkt sich die Zeitentwicklung auf unterschiedliche Weise aus: Die Diagonalelemente der Wignermatrix Ŵ|ΨRSi rotieren ebenfalls in entgegengesetzte Richtungen, wohingegen die Nichtdiagonalelemente die anfängliche Symmetrie in p beibehalten und nur ihre interne Struktur verändern. Die Unterschiede der beiden Näherungen werden im physikalischen Bild durch Verwendung unterschiedlicher Referenzzustände erreicht. Der Riemann-Siegel-Referenzzustand |ΦRSi besteht aus den abgeschnittenen Riemann-Zuständen |σ, 0in0 und |2 − 3σ, 0in0 gekoppelt an das Atom. Dadurch ist der Zustand für σ 6= 1/2 verschränkt und wird nur auf der kritischen Achse, also für σ = 1/2, unverschränkt. Im Phasenraum sind deshalb zwar die Diagonalelemente der Wignermatrix Ŵ|ΦRSi für alle σ symmetrisch in p, die Nichtdiagonalelemente hingegen nur für σ = 1/2. Im Gegensatz dazu setzt sich der Berry-Keating-Referenzzustand |ΦBKi aus den beiden Berry-Zuständen |φB(σ, τ )inB und |φB(2−3σ, τ )inB zusammen, die sich nur sehr wenig von den abgeschnittenen Riemann-Zuständen zur Zeit τ = 0 unterscheiden. Dennoch führt der kleine Unterschied dazu, dass |ΦBKi auch auf der kritischen Achse verschränkt ist und lediglich die Nichtdiagonalelemente der Wignermatrix Ŵ|ΦBKi für σ = 1/2 symmetrisch in p sind. Nichtsdestotrotz zeigt die kritische Achse in den Phasenraumbildern beider Näherungen ihre besondere Rolle: nur für σ = 1/2 ist der Integralkern der Wignerdarstellung K 144 sowie der Betrag des Integralkerns in Moyaldarstellung K symmetrisch in p. Im letzten Kapitel wird gezeigt, dass die Ähnlichkeit zwischen dem RiemannHamiltonian und dem effektiven Hamiltonian des Jaynes-Cummings-Paul-Modells zur Berechnung eines genäherten Hamiltonian verwendet werden kann. Eine grobe Abschätzung mit Hilfe eines Polynoms vierter Ordnung führt zu einer guten Simulation des allgemeinen Verhaltens der Riemannschen Zetafunktion obwohl der genäherte Hamiltonian vom Riemann-Hamiltonian abweicht. Abschließend bleibt zu erwähnen, dass unsere physikalische Beschreibung der Riemannschen Zetafunktion auch auf komplexe Funktionen f (s) mit Argument s ≡ σ + iτ angewendet werden kann, wenn sich diese als Reihe f (σ + iτ ) ≡ ∞ X an (σ, τ ) bn (σ, τ ) + (n + 1) iτ (n + 1)−iτ n=0 darstellen lassen. Wie bei ζ können auch hier die Faktoren (n + 1)±iτ als Zeitentwicklung eines verschränkten Quantenzustands |Ψ(t)i interpretiert werden, wobei die Wechselwirkung wieder durch ĤR beschrieben wird, also logarithmisch vom Photonenzahloperator n̂ abhängt. τ = ωt ist hier die skalierte Zeit. Die Faktoren an und bn werden auf die Wahrscheinlichkeitsamplituden des Zustands |Ψ(t)i und des entsprechend gewählten Referenzzustands |Φi aufgeteilt. Hierbei ist darauf zu achten, dass die Wahrscheinlichkeitsamplituden des Anfangszustands |Ψ(0)i zeitunabhängig sind, da sonst die Phasen nicht durch die Zeitentwicklung mit ĤR entstehen. Folglich müssen alle τ -Abhängigkeiten der Faktoren an und bn in den Referenzzustand einfließen, da dort τ = Im s lediglich den Imaginärteil des Punktes s im Phasenraum darstellt. Auf diese Weise können die Eigenschaften der Funktion auf die Zustände transferiert und für Messungen zugänglich gemacht werden. 145 Publikationen 1. E. Kajari, M. Buser, C. Feiler und W. P. Schleich Rotation in Relativity and the Propagation of Light in Atom Optics and Space Physics, Proceedings of the International School of Physics “Enrico Fermi”, Course CLXVIII, edited by E. Arimondo, W. Ertmer, W. P. Schleich and E. M. Rasel (Elsevier, Amsterdam, 2009) ebenfalls erschienen in: Rivista del nuovo cimento 32, 339 – 437 (2009) 2. C. Feiler, M. Buser, E. Kajari, W. P. Schleich, E. M. Rasel and R. F. O’Connell New Frontiers at the Interface of General Relativity and Quantum Optics Space Sci Rev 148, 123 – 147 (2009) ebenfalls erschienen in: Probing the Nature of Gravity – Confronting Theory and Experiment in Space, Editors: C. W. F. Everitt, M. C. E. Hulrt, R. Kallenbach, G. Schäfer, B. F. Schultz und R. A. Treumann (Springer, New York, 2010) 3. S. Wölk, C. Feiler and W. P. Schleich Factorization of numbers with truncated Gauss sums at rational arguments J. Mod. Opt. 56, 2118 – 2124 (2009) 4. S. Wölk, C. Feiler and W. P. Schleich Quantum Mechanics Meets Number Theory Conference Paper zu International Conference on Quantum Information, Ottawa, Kanada (Optical Society of America, 2011) 5. C. Feiler and W. P. Schleich Entanglement and analytical continuation: an intimate relation told by the Riemann zeta function New J. Phys. 15, 063009 (2013) 6. J. W. Neuberger, C. Feiler, H. Maier and W. P. Schleich Continuous Newton Method, Lines of Constant Phase and the Riemann Zeta Function To be published. 147 Poster und Vorträge 1. C. Feiler, R. Mack and W. P. Schleich Wignerfunktion zur Beschreibung der Riemann-Zeta-Funktion“ ” Poster bei der DPG-Tagung in Düsseldorf (März 2007) 2. C. Feiler, R. Mack and W. P. Schleich Riemanns Zeta Function in Phase-Space“ ” Poster bei der DPG-Tagung in Darmstadt (März 2008) 3. C. Feiler, R. Mack and W. P. Schleich The Riemann ζ-Function in Phase Space“ ” Poster bei der DPG-Tagung in Hamburg (März 2009) 4. C. Feiler, R. Mack and W. P. Schleich Zeta-States in Phase Space“ ” Poster bei der DPG-Tagung in Hannover (März 2010) 5. C. Feiler Riemann zeta function in phase space“ ” Vortrag bei der Netzwerktagung der Alexander von Humboldt-Stiftung in Ulm (Oktober 2010) 6. C. Feiler, R. Mack and W. P. Schleich Zeta-States in Phase Space“ ” Poster bei der Conferece on Quantum Simulations“ in Benasque (Spanien, ” Februar/März 2011) und der DPG-Tagung in Dresden (März 2011) 7. C. Feiler Newton flow of the Riemann zeta function“ ” Vortrag beim Ulm-Augsburg Meeting 2013 in Augsburg (April 2013) 149 Curiculum vitae Der Inhalt dieser Seite wurde aus Datenschutzgründen entfernt. 151 Danksagung Der Inhalt dieser Seite wurde aus Datenschutzgründen entfernt. 153 Erklärung Hiermit versichere ich, dass ich die vorliegende Dissertation selbständig verfasst und keine anderen als die angegebenen Quellen und Hilfsmittel verwendet habe, sowie die wörtlich oder inhaltlich übernommenen Stellen als solche kenntlich gemacht habe. Ulm, den Cornelia Feiler