UNIVERSIDAD NACIONAL AUTÓNOMA DE MÉXICO POSGRADO EN CIENCIAS FÍSICAS FROM QUANTUM RADIAL ANHARMONIC OSCILLATOR TO ZEEMAN EFFECT IN HYDROGEN ATOM TESIS QUE PARA OPTAR POR EL GRADO DE: DOCTOR EN CIENCIAS (FÍSICA) PRESENTA: M. EN C. JUAN CARLOS DEL VALLE ROSALES TUTOR PRINCIPAL: DR. ALEXANDER TURBINER ROSENBAUM (ICN-UNAM) COMITÉ TUTOR: DR. JUAN CARLOS LÓPEZ VIEYRA (ICN-UNAM) DR. JORGE ALEJANDRO REYES ESQUEDA (IF-UNAM) CIUDAD DE MÉXICO, 2020 UNAM – Dirección General de Bibliotecas Tesis Digitales Restricciones de uso DERECHOS RESERVADOS © PROHIBIDA SU REPRODUCCIÓN TOTAL O PARCIAL Todo el material contenido en esta tesis esta protegido por la Ley Federal del Derecho de Autor (LFDA) de los Estados Unidos Mexicanos (México). El uso de imágenes, fragmentos de videos, y demás material que sea objeto de protección de los derechos de autor, será exclusivamente para fines educativos e informativos y deberá citar la fuente donde la obtuvo mencionando el autor o autores. Cualquier uso distinto como el lucro, reproducción, edición o modificación, será perseguido y sancionado por el respectivo titular de los Derechos de Autor. iii Abstract In the present thesis, two fundamental systems of quantum mechanics are studied: (i) the D- dimensional radial polynomial anharmonic potential, and (ii) a hydrogen atom subjected to a uniform constant magnetic field. For both systems, the corresponding Schrödinger equation is non-exactly solvable: closed analytical expressions (not in terms of infinite asymptotic series) for the energies and wave functions may be found only in approximate form. This thesis presents the general formalism to construct locally accurate approximations of the wave functions in the coordinate space, which leads to highly accurate energies and expectations values. The first part of this thesis is devoted to constructing a locally accurate approximation, called the Approximant, for the wave functions of the low-lying states of the D-dimensional radial polynomial anharmonic potential. Perturbation Theory in the weak and strong coupling regimes, asymptotic expansions at small and large distances, and the true semi-classical approx- imation are basic ingredients of the formalism to construct the Approximant. The key result of this study is a compact expression for the exponential phase of the ground state wave function, which is parameter-dependent. This expression is the building block for constructing the Ap- proximants of the excited states. To fix the value of some parameters, the exact reproduction of all growing terms of the exponential phase at large distances is imposed as constraint. This procedure effectively reduces the number of parameters. Taking the Approximant as trial func- tion, the remaining free parameters are optimized in the framework of the Variational Method. In this manner, the energy is calculated as an expectation value of the Hamiltonian by means of a locally accurate Approximant. Explicit calculations are carried out for the first four low-lying states of the cubic, quartic, and sextic anharmonic potentials of the form V (r) = r2+gm−2rm, withm = 3, 4, 6, respectively. For the cubic case, a one-parametric Approximant leads to unprecedented accurate energies with a relative accuracy of order . 10−5 with respect to the exact result. In turn, the relative deviation of the Approximant from the exact wave function is . 10−2 at any point of the r-space. For the case of the quartic/sextic anharmonicity, a 2/4-parametric Approximant leads to energies with a relative accuracy of order . 10−8, while the relative accuracy of the Approximant is of order 10−6 for any point of the r-space. Consistency of those results is checked inside the Non-Linearization Procedure and the Lagrange-Mesh method. In the second part, the formalism used for the radial anharmonic potential is employed to construct a locally accurate Approximant for the ground state (1s0) of the hydrogen atom in a uniform constant magnetic field. A non-relativistic consideration with a static proton is assumed. The natural extension of the true semi-classical expansion was obtained for this system. The main result of this part is a compact 8-parametric expression for the Approximant of the ground state. When it is used as a trial function in the Variational Method, it leads to highly accurate energy (relative accuracy . 10−6), and an accurate quadrupole moment in the whole domain of accessible magnetic fields in Nature. iv Resumen En la presente tesis se estudian dos sistemas fundamentales de mecánica cuántica: (i) el po- tencial anarmónico radial D-dimensional, y (ii) un átomo de hidrógeno en presencia de un campo magnético constante y uniforme. Para ambos sistemas, la ecuación de Schrödinger cor- respondiente no admite soluciones exactas: las enerǵıas y las funciones de onda solamente se pueden encontrar de forma aproximada. Esta tesis presenta el formalismo general para construir aproximaciones localmente precisas de las funciones de onda en el espacio de coordenadas, esto conduce a enerǵıas y valores de expectaćıon altamente precisos. La primera parte de la tesis está dedicada a construir una aproximación localmente pre- cisa (llamada Aproximante) para las funciones de onda de los primeros estados ligados del potencial anarmónico radial D-dimensional. La teoŕıa de perturbaciones en los reǵımenes de acoplamiento débil y fuerte, las expansiones asintóticas a pequeñas y grandes distancias, y la verdadera aproximación semiclásica son los ingredientes básicos del formalismo para construir el Aproximante. El resultado principal de este estudio es una expresión compacta para la fase exponencial de la función de onda del estado base, la cual depende de ciertos parámetros. Esta expresión es la pieza clave para construir Aproximantes de estados excitados. Para fijar el valor de algunos parámetros, se impone como constricción la reproducción exacta de todos los términos crecientes de la fase exponencial a grandes distancias. Este procedimiento reduce efectivamente el número de parámetros. Tomando el Aproximante como función de prueba, los parámetros restantes se optimizan usando el Método Variacional. De esta manera, la enerǵıa se calcula como el valor esperado del Hamiltoniano evaluado por medio de un Aproximante localmente preciso. Se realizan cálculos expĺıcitos para los primeros cuatro estados ligados de los potenciales anarmónicos cúbico, cuártico, y séxtico de la forma V (r) = r2 + gm−2rm, con m = 3, 4, 6, respectivamente. Para el caso cúbico, un Aproximante con un solo parámetro conduce a enerǵıas sin precedentes de una precisión relativa de orden 10−5 en cualquier punto del espacio r. A su vez, la desviación relativa del Aproximante con respecto a la función de onda exacta es de orden 10−2. Para los casos de anarmonicidad cuártica/sextica, un Aproximante de 2/4 parámetros conduce a enerǵıas de una precisión relativa de orden 10−8, mientras que la precisión relativa del Aproximante es de orden 10−6 para cualquier punto del espacio r. La consistencia de esos resultados se verifica usando el Procedimiento de No-Linealización y por medio del método de la malla de Lagrange. En la segunda parte, el mismo formalismo utilizado para el potencial anarmónico radial se emplea para construir un Aproximante localmente preciso para el estado base (1s0) del átomo de hidrógeno en un campo magnético constante y uniforme. Las consideraciones del estudio son no-relativistas con un protón estático. El resultado principal de esta parte es una expresión compacta para el Aproximante del estado base. Cuando el Aproximante se usa como una función de prueba en el Método Variacional, este conduce a una enerǵıa altamente precisa (de precisión relativa 10−6), y a un momento cuadrupolar preciso en todo el dominio de campos magnéticos accesibles en la Naturaleza ( . 2.35× 1013G). Contents Overview ix 1 The Basic Tool: Variational Method in Quantum Mechanics 1 1.1 Variational Principle and Upper Bounds for the Energy . . . . . . . . . . . . . . 1 1.2 Variational Energy and Perturbation Theory . . . . . . . . . . . . . . . . . . . . 2 I Radial Anharmonic Oscillator 5 2 Introduction 7 2.1 Anharmonic Oscillator and Diatomic Molecules . . . . . . . . . . . . . . . . . . 7 2.2 Renewed Interest . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8 2.3 Present Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 3 Generalities 11 3.1 Radial Potential in D-dimension . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 3.2 Radial Polynomial Anharmonic Oscillator . . . . . . . . . . . . . . . . . . . . . 13 3.3 The Ground State: Riccati Equation . . . . . . . . . . . . . . . . . . . . . . . . 14 3.3.1 Riccati-Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 3.3.2 Generalized Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . 17 4 The Weak Coupling Regime 19 4.1 Riccati-Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19 4.1.1 D-dimensional Radial Anharmonic Oscillator . . . . . . . . . . . . . . . . 21 4.2 Generalized Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23 4.3 Connection Between the Expansions for Y and Z . . . . . . . . . . . . . . . . . 25 4.4 Connection with WKB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 5 The Strong Coupling Regime 29 5.1 Riccati-Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 5.1.1 Physically Relevant Trial Function . . . . . . . . . . . . . . . . . . . . . 31 5.1.2 Interpolation: Locally Accurate Approximation . . . . . . . . . . . . . . 32 v vi Contents 5.2 Generalized Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 6 The Approximant 37 6.1 The Approximant for the Ground State . . . . . . . . . . . . . . . . . . . . . . 37 6.2 Approximants for Excited States . . . . . . . . . . . . . . . . . . . . . . . . . . 39 7 Results for Two-Term Potentials 41 7.1 Numerical Realization of the Variational Method . . . . . . . . . . . . . . . . . 41 7.2 Cubic Anharmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 7.2.1 Perturbation Theory in the Weak Coupling Regime . . . . . . . . . . . . 43 7.2.2 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 7.2.3 The Approximant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 7.2.4 Numerical Results: Variational Calculations . . . . . . . . . . . . . . . . 47 7.2.5 First Terms in the Strong Coupling Expansion . . . . . . . . . . . . . . . 52 7.3 Quartic Anharmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . 54 7.3.1 Perturbation Theory in the Weak Coupling Regime . . . . . . . . . . . . 54 7.3.2 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55 7.3.3 The Approximant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 7.3.4 Numerical Results: Variational Calculations . . . . . . . . . . . . . . . . 59 7.3.5 First Terms in the Strong Coupling Expansion . . . . . . . . . . . . . . . 62 7.4 Sextic Anharmonic Oscillator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63 7.4.1 Perturbation Theory in the Weak Coupling Regime . . . . . . . . . . . . 64 7.4.2 Generating Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65 7.4.3 The Approximant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67 7.4.4 Numerical Results: Variational Calculations . . . . . . . . . . . . . . . . 68 7.4.5 First Terms in the Strong Coupling Expansion . . . . . . . . . . . . . . . 72 8 Conclusions 75 II Hydrogen Atom in Constant Magnetic Field 81 9 Introduction 83 9.1 Motivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 9.2 Present Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 10 Generalities 87 10.1 The Ground State: Riccati Equation: . . . . . . . . . . . . . . . . . . . . . . . 89 10.2 Riccati-Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 10.3 Generalized Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91 Contents vii 11 Weak Coupling Regime 93 11.1 Riccati Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93 11.1.1 Behavior of the Phase Φ at Small Distances . . . . . . . . . . . . . . . . 95 11.2 Generalized Bloch Equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 11.2.1 Asymptotic Behavior of Φ at Large Distances . . . . . . . . . . . . . . . 97 11.3 Connection between RB and GB Equations . . . . . . . . . . . . . . . . . . . . 98 12 The Approximant 101 12.1 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 12.1.1 Energy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 102 12.1.2 Quadrupole Moment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105 13 Conclusions 109 General Conclusions 115 Appendices 119 A The WKB (Semi-Classical) Approximation 119 B First PT Corrections and Generating Functions 121 B.1 Cubic Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 B.2 Quartic Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 B.3 Sextic Case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123 C Plots of Variational Parameters: Anharmonic Oscillator 125 C.1 Cubic: (7.2.24) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 C.2 Quartic: (7.3.26) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 C.3 Sextic: (7.4.27) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 128 D General Aspects of the Two-Term Radial Anharmonic Oscillator 130 E System of Coordinates {ρ, r, ϕ} 131 E.1 Definition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 E.2 Differentiation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 F Perturbation Theory for Hydrogen Atom in Constant Magnetic Field 133 F.1 Computational Code: Non-Linearization Procedure . . . . . . . . . . . . . . . . 133 F.2 Corrections in Perturbation Theory . . . . . . . . . . . . . . . . . . . . . . . . . 134 G Computational Code: Hydrogen Atom in Magnetic Field 137 viii Contents H Lagrange-Mesh Method for Hydrogen Atom in Magnetic Field: Code 143 I Plots of Variational Parameters: Hydrogen Atom in Magnetic Field 145 Bibliography 153 Overview It is well-known that the time-independent Schrödinger equation can be solved exactly for a few physical systems. Therefore, we need to use different approximation methods to handle the overwhelming majority of quantum-mechanical problems. In fact, much of the success of quantum mechanics has been achieved thanks to approximation methods. Among them, the Variational Method is one of the most powerful tools which allow us to deal with many physical systems of different nature. In all books on quantum mechanics, e.g. [1–3], the Variational Method plays a central role among the approximation methods presented. In the variational approach, the trial function is key for leading to approximations of the spectrum of a given system. A natural question is: how to choose an adequate trial function that yields reliable approximate energies and wave functions? The Variational Method per se does not provide a prescription about the choice of a trial function. Thus, the construction of a trial function is handmade, usually based on physical intuition. Symmetries of the system, perturbation theory, asymptotic analysis, semi-classical approximations, features of nodal surfaces, etc., are sources of information of the exact function that may be coded in the trial one. Nevertheless, one could choose a trial function based on technical arguments. For example, one may choose a trial function in such a way that all integrals involved in the realization of the Variational Method are calculated in closed analytical form or reduced to one-dimensional integrals. In fact, most of the trial functions chosen in atomic physics follow this line. Any trial function represents an approximate solution of the Schrödinger equation. With an approximate solution, one can calculate the expectation values of relevant operators. It is known that a high accuracy reached in the variational energy does not guarantee high accuracy in such expectation values, see [4]. This reflects one fact: a trial function may be close to the exact one in the domain where the dominant contribution of the integrals (involved in the Variational Method) occurs, but it may differ in the domains required by the calculation of the expectation value. Let us discuss a simple example, see [5], that illustrates these phenomena described above. We consider a particular system where the exact solutions are known: a hydrogen atom. The exact ground state wave function is well-known1 ψexact(r) = e−r , 1Atomic units used: ~ = me = −e = 1, where me and e are the mass and the electric charge of the electron, respectively. ix x Overview as well as its energy2 E = −1Ry . Suppose we are not aware of the exact solution, and we want to use the Variational Method to estimate the ground state energy. In particular, we can consider a Gaussian trial function ψI(r) = e−α r2 where α is a positive variational parameter. With this trial function, the realization of the Variational Method is completely analytical. The optimal parameter α turns out to be α = 8 9π , and it leads to the following variational energy Evar = −0.849Ry . This variational estimate deviates from the exact value around 15%. This result is surprising due to the high accuracy of the variational estimate and the simplicity of the trial function. The deviation can be dramatically reduced to 3% by considering a two-term superposition of the previous trial function, namely ψII(r) = e−α r2 + γ e−β r2 , α, β > 0 . Here α, β, and γ are free parameters that are fixed by means of the Variational Method.3 With this new trial function, the variational estimate of the energy is Evar = −0.972Ry . One could continue to reduce the deviation taking as trial function a larger combination of Gaussian functions. However, no matter how many Gaussian functions we consider, it is clear that their sum will never converge to ψexact. This simple example evidences one important fact: although a trial function can provide an accurate estimate of the energy, the trial function with optimal parameters may not be very close to the exact one. In our particular example, it can be roughly seen if we plot the exact solution and compare it with the trial ones, see Fig. 1. The closeness of the trial functions to the exact one may be qualitatively studied by means of the relative deviation (Rd), which is defined as Rd(r) = ∣ ∣ ∣ ∣ ψexact(r)− ψi(r) ψi(r) ∣ ∣ ∣ ∣ , i = I, II . (1) For both ψI and ψII , the plot of Rd as a function of r is presented in Fig. 2. In the domain 0 ≤ r . 3.5, the dominant contribution of the integrals (required by the Variational Method) occurs for both trial functions. Inside this domain, the relative deviation reaches a local maximum: ∼ 0.6 and ∼ 0.3 for ψI and ψII , respectively. Therefore, in percentage terms, ψexact may deviate from ψI around 60%, and 30% from ψII . Outside this domain, the relative deviation is a monotonically increasing function for both trial functions. Due to the large relative deviation xi 0 1 2 3 4 0.5 0.25 0.75 1 Figure 1: Plot of ψexact, ψI , and ψII . As normalization, we set ψexact(0) = ψI(0) = ψII(0) = 1. 0 1 2 3 4 5 1 2 3 4 Figure 2: Relative deviation of the functions ψI and ψII (from ψexact) as a function of r throughout the domain, we may say that neither ψI nor ψII are locally accurate approximations of ψexact. Non-locally accurate approximation may lead to inaccurate estimates of some expectation values. For instance, let us now compute 〈r2−n〉 using ψexact, ψI , and ψII . The numerical results are presented in Table 1. One can note that the deviation of the variational results, compared with the exact one, is significant for n > 5: it exceeds 30%. This deviation contrasts with the fact that for the trial functions ψI and ψII , the deviation of the variational energy from the exact one is 15% and 3%, respectively. The calculation of expectation values of more complex functions than simple monomials could be catastrophic for reaching highly accurate results. For example, one relevant quantity which involves expectation values is the cusp parameter C, C ≡ 〈 ψ ∣ ∣ δ(~r) ∂ ∂r ∣ ∣ψ 〉 〈ψ | δ(~r) |ψ〉 , 2Ry stands for the Rydberg constant, 1 Ry ≈ 13.6 eV. 3The optimal choice of (rounded) parameters is α = 0.20, β = 1.33, and γ = 1.38. The integration required by the Variational Method can be done analytically, while the optimization of parameters is a numerical procedure. xii Overview Table 1: Expectation values 〈rn−2〉 for different values of n using ψexact, ψI , and ψII . Interest- ingly, at n = 3 the trial function ψI provides the exact result. Results rounded to first displayed s. d. Atomic units used. n ψexact ψI ψII 0 2 1.13 1.71 1 1 0.85 0.97 3 1.5 1.5 1.48 4 3 2.65 2.80 5 7.5 5.30 6.28 6 22.5 11.71 16.01 7 78.75 28.11 44.90 see [6]. In atomic physics, this parameter is usually employed to measure the local quality of the approximate wave function near the Coulomb singularities. Using the exact wave function of the hydrogen atom ψexact, the exact cusp parameter results in Cexact = −1 . This is an exact number that the ground state wave function of the hydrogen atom must provide. On the other hand, if we compute the cusp parameter associated with the approximate functions ψI and ψII , it is zero for both! This is nothing but a reflection of the fact that the logarithmic derivative of ψI and ψII do not have the correct asymptotic behavior at small r. In fact, the difference between ψexact, ψI , and ψII is more pronounced when we plot their corresponding logarithmic derivatives, see Fig. 3. 0 1 2 3 4 0.5 1 1.5 2 Figure 3: Logarithmic derivative of ψexact, ψI , and ψII . This simple and illustrative example exhibits some important niceties that should be taken xiii into account when designing trial functions. However, it has to be noted that the above dis- cussion is based on the knowledge of the exact energy and wave function, which is frequently unknown. So, the question is: how to estimate the accuracy of the approximate energy (wave function) that the Variational Method provides when the exact energy (wave function) is not available? The answer to this highly non-trivial question lies in the connection between the vari- ational energy and perturbation theory discovered in [7]. This connection plays a fundamental role in this thesis. The present work aims to construct locally accurate trial functions for one fundamental sys- tem: the quantum radial polynomial anharmonic oscillator in arbitrary dimension D. This study is strongly motivated by the work presented in [8]. Due to the separation of the angular variables, this D-dimensional system is reduced to a one-dimensional problem. We present a formalism to design locally accurate trial functions of any state for an arbitrary anharmonic potential. The formalism culminates in a simple expression that produces trial functions that lead to unprecedented results in the weak and strong coupling regime: highly accurate estimates of the exact energies, and locally accurate approximate wave functions. It is demonstrated that an extension of these ideas and the formalism may be applied to a different physical system: the hydrogen atom in a constant uniform magnetic field in its ground state. The formalism leads to an adequate trial function that yields highly accurate results for the energy and quadrupole moment for all magnetic fields available in Nature. For both systems, the obtained approxi- mate wave functions actually represent approximate solutions of the corresponding Schrödinger equation. In fact, they can be considered as exact and compact solutions inside their domain of applicability. This will be discussed in detail for both systems. The thesis is divided in two parts. Part I is devoted to the D-dimensional quantum radial anharmonic oscillator. The hydrogen atom in a constant magnetic field is studied in the Part II. It is worth mentioning that Part I and II can be read independently since they are self- contained. Each one begins with a short Introduction and ends with Conclusions. Regarding Part I, some of the results shown in this thesis were presented in [9, 10]. For Part II, they will appear in [11]. Chapter 1 The Basic Tool: Variational Method in Quantum Mechanics In this Chapter, we present a brief and general description of the Variational Method (VM) in the framework of quantum mechanics. We focus the discussion on the ground state. However, when additional information about the nodal surface is available, it may be extended to study excited states. 1.1 Variational Principle and Upper Bounds for the En- ergy Let us consider an arbitrary Hamiltonian Ĥ which describes a particular system of interest. We assume that the wave function of the ground state, as well as the corresponding energy, are unknown. If the Schrödinger equation of Ĥ is non-solvable, the natural question is: How to find the energy and wave function approximately? The variational principle gives us an answer in terms of the variational energy, which is defined as Evar[ψ] = ∫ ψ∗ t Ĥψt dV ∫ ψ∗ tψt dV . (1.1) Here, ψt is an arbitrary normalizable function, ∫ ψ∗ψ dV <∞ , (1.2) usually called trial function. The importance of the variational energy is established via the Variational Principle. Variational Principle. Consider a system described by the Hamiltonian Ĥ with exact ground state energy E0, then the variational energy satisfies the inequality Evar[ψt] ≥ E0 (1.3) 1 2 Chapter 1. The Basic Tool: Variational Method in Quantum Mechanics for any trial function ψt. Besides, the equality is fulfilled if and only if ψt is the exact ground state wave function of the system. Proof. The trial function ψt can be expanded in terms of the (unknown) exact basis of or- thonormal eigenfunctions {ψk} of Ĥ. Thus, ψt = ∑ k ckψk , Ĥψk = Ekψk , (1.4) where ck are some coefficients, and Ek is the eigenvalue associated to ψk. Without loss of generality, we assume that eigenvalues are sorted, E0 ≤ E1 ≤ E2 ≤ ... . Hence, the variational energy (1.1) reads Evar[ψt] = ∑ k |ck|2Ek ∑ k |ck|2 . (1.5) Since E0 is the lowest energy, the following inequality holds ∑ k |ck|2Ek ≥ E0 ∑ k |ck|2 . (1.6) Using it in (1.5), it follows the variational principle Evar[ψt] ≥ E0 . (1.7) From another side, it is clear that the equality is fulfilled if and only if ψt is the exact ground state wave function of Ĥ. In this case, all coefficient ck is zero except c0. As one could imagine, the variational principle is a useful tool to estimate upper bounds of the exact ground state energy. The lower upper bound, the closer Evar is to the exact energy. In practice, the trial function can depend on a finite number of free parameters. In this manner, the variational energy will also be parameter-dependent. A proper choice of parameters may lead to the lowest upper bound that a given trial function can provide. This choice of parameters is known as the optimal configuration. The efficient way to find it is through an optimization procedure, which is, for most of the adequate trial functions, a numerical procedure. The Variational Method covers from the construction of the trial function to the calculation of the lowest upper bound. 1.2 Variational Energy and Perturbation Theory We now discuss an interesting connection between the variational energy and Perturbation Theory (PT). For a given trial function ψt, it is straightforward to calculate the Hamiltonian Ĥt for which ψt is an exact ground state eigenfunction. We call this procedure the Inverse Problem. In fact, since the kinetic operator T̂ of the Hamiltonian is usually given, we can 1.2. Variational Energy and Perturbation Theory 3 focus only on the construction of the associated potential1 Vt of such Hamiltonian. Thus, let us assume that Ĥt is of the form Ĥt = T̂ + Vt . (1.8) Since ψt is the exact eigenfunction of Ĥt, it is clear that it obeys Ĥtψt = T̂ψt + Vtψt = E (t) 0 ψt , (1.9) where E (t) 0 is the ground state energy2 of Ĥt. From equation (1.9) we can trivially solve for the (unknown) Vt, Vt = E (t) 0 − T̂ψt ψt . (1.10) Since E (t) 0 can be absorbed in the definition of Vt, its value is unimportant. It is evident that when solving the inverse problem, we can generate zillions of potentials for which a single eigenstate - the lowest - is known exactly. Once Vt is known, the connection between the variational energy and PT can be established. Suppose that we are interested in the ground state of the operator Ĥ = T̂ + V . (1.11) However, it turns out that the corresponding Schrödinger equation Ĥψ = Eψ (1.12) is non-solvable. To estimate E, we use the VM with a trial function ψt as entry. Thus, we calculate the variational energy associated to ψt Evar[ψt] = ∫ ψ∗ t (T̂ψt + V ψt) dV ∫ ψ∗ tψt dV . (1.13) Using (1.10), we can write (1.13) as follows Evar[ψt] = E (t) 0 + ∫ ψ∗ t (V − Vt)ψt dV ∫ ψ∗ tψt dV . (1.14) If we define E (t) 1 = ∫ ψ∗ t (V − Vt)ψt dV ∫ ψ∗ tψt dV , (1.15) the expression shown in (1.14) acquires a more compact form, Evar[ψt] = E (t) 0 + E (t) 1 . (1.16) This result suggests reinterpreting the variational energy as the first two terms of an expansion in PT developed for the perturbed Hamiltonian Ĥ = T̂ + Vt + λ (V − Vt) . (1.17) 1In this work, potentials are always real functions. 2The meaning of the subscript will be clarified, for now, it is just notation. 4 Chapter 1. The Basic Tool: Variational Method in Quantum Mechanics Here, λ is a formal parameter that ultimately should be placed equal to one, λ = 1. The meaning of the superscripts placed in (1.16) is now clear: E (t) 0 plays the role of the zero-order correction of the energy, while E (t) 1 - the expectation value of perturbation potential (V − Vt) - is the first order correction. To summarize, The computation of the variational energy of a given Hamiltonian Ĥ = T̂ + V , taking as trial function ψt, is equivalent to calculate the first two terms in a formal PT. The trial function ψt is the exact ground state of the unperturbed Hamiltonian Ĥt = T̂ + Vt, while V − Vt plays the role of perturbation. If higher order corrections, E (t) 2 , E (t) 3 , ..., are known, they may be used to estimate the accuracy of the variational energy. In principle, the exact value of the lowest energy of Ĥ is the sum of those corrections, E = ∞∑ n=0 E(t) n . (1.18) In general, the question about the convergence of the expansion (1.18) is crucial but difficult to answer. We restrict ourselves by saying that if ψt is chosen with the appropriate asymptotic behavior at large distances, a convergent expansion will occur. In this situation, our initial variational estimate of the energy can be improved by adding next-order corrections. In practice, it is sufficient to calculate few terms in the expansion (1.18) to estimate rates of convergence as well as highly accurate estimates of the exact energy via partial sums. For all one dimensional physical systems, a considerable number of corrections E (t) n can be calculated. This situation changes dramatically when we deal with true multidimensional systems. The calculation of any energy correction, i.e. E (t) 2 , E (t) 3 , ..., requires the knowledge of the corrections to the wave function in PT. In this manner, using corrections for ψt, the deviation (between the trial function and exact one) can also be analyzed and, more important, estimated. Part I Radial Anharmonic Oscillator 5 Chapter 2 Introduction 2.1 Anharmonic Oscillator and Diatomic Molecules The first modern study of the one-dimensional quantum anharmonic oscillator dates back to 1925. This study was carried out by Heisenberg and presented to the scientific community in [12]. In this work, Heisenberg considered the cubic and quartic anharmonic oscillators described by the classical equations of motion ẍ + w2 0 x + λx2 = 0 , ẍ + w2 0 x + λx3 = 0 , respectively. We have denoted the coupling constant as λ, in turn, w0 is the natural frequency of the harmonic oscillator. In the framework of what would later be called Matrix (quantum) Mechanics, he presented not only the energy spectrum of the harmonic oscillator (λ = 0) for the first time, but also gave the first corrections to the energy in PT in powers of λ. For the cubic case, a first order (vanishing) correction was obtained. For the quartic case, the first two corrections were presented. Results at λ = 0 were confirmed in the framework of quantum mechanics exposed by Schrödinger just one year later [13]. Around those years, the main motivation to study anharmonic perturbations of the harmonic oscillator came from atomic and molecular physics. A concrete example is the discovery of the spectral band at 1.76µm of the diatomic molecule HCl (hydrogen chloride). This result, obtained by Brinsmade and Kemble in 1917 [14], was of importance since it established that the HCl molecule behaves as an anharmonic oscillator and showed how this could be interpreted on the basis of the early quantum theory introduced by Planck. Once the Schrödinger equation was presented to the world in 1926, the diatomic molecule described by a 3-dimensional anharmonic potential was studied immediately, see e.g. [15]. Due to the exactly solvable nature of the one-dimensional Morse potential, it was widely used to model the interaction between the two atoms of a diatomic molecule near their equilibrium position [16,17]. The Morse potential is given by V (r) = V0(e −2a(r−r0) − 2 e−a(r−r0)) , 7 8 Chapter 2. Introduction where V0 is the energy of dissociation, r the nuclear separation, r0 the equilibrium value of r, and a is a positive constant. From the Taylor expansion of this potential around r0, V (r) = V0 ( −1 + a2(r − r0) 2 − a3(r − r0) 3 + 7a4 12 (r − r0) 4 + . . . ) , (2.1.1) one can easily check that the Morse potential contains an infinite number of anharmonic terms. Following the vibrational spectroscopic information, one can determine parameters a, V0, and r0. For example, the diatomic molecule BeO (beryllium oxide, see [16]) in the state Σ1 is characterized by a = 2.12 a−1 0 , V0 = 5.245 eV , r0 = 1.27 a0 , where a0 = 5.29×10−9cm is the Bohr radius. As far as vibrational energy levels are concerned, the Morse potential provides accurate results. However, when the information given by the rotational energy levels is taken into account, certain discrepancies are found [15]. A natu- ral direction, to avoid such discrepancies, was to consider a more general radial anharmonic polynomial potential1 of the form V (r) = K(r − r0) 2 ( 1 + b (r − r0) + c (r − r0) 2 + d (r − r0) 3 + . . . ) , with K, b, c, d, ..., being real parameters called anharmonic constants. With this kind of potential, the corresponding Schrödinger equation is non-exactly solvable: energies and wave functions can be found only in approximate form. The first approaches to solve it (∼1926) were based on the Bohr theory [18], and second order PT [19]. Over the next years, several studies were focused on determining the first anharmonic constants of several diatomic molecules using those approaches. 2.2 Renewed Interest As we have seen, immediately after the foundation of quantum mechanics, the study of anhar- monic oscillators was strongly motivated by molecular physics. From a more general point of view, it was recognized that the spectral problem of a quantum anharmonic oscillator has very interesting properties. The publication of seminal papers by Bender and Wu [20] (from 1969) refreshed the interest in the quantum mechanical properties of anharmonic oscillators. First, they studied the one-dimensional quartic anharmonic oscillator characterized by the quantum Hamiltonian Ĥ = − d2 dx2 + 1 4 x2 + 1 4 λx4 , where λ is a positive coupling constant. In particular, this Hamiltonian is interesting because it allows us to demonstrate that the perturbation series of the energy in powers of λ diverges for any state. They investigated the reason for this divergence using an analytical continuation of the energy levels into the complex λ-plane as well as the WKB approximation. It turned out 1Series are truncated up to some order. 2.3. Present Work 9 that the presence of an infinite number of branch points with a limit point at λ = 0 was the reason for this divergence. A generalization of this study was presented a few months later by Bender [21]. This time he considered a more general case Ĥ = − d2 dx2 + 1 4 x2 + 1 2M λx2M , where M = 2, 3, ... . Analytical features of the energy levels were investigated in this work. He concluded that those properties are qualitatively similar to the case M = 2. Interestingly, it was shown that the exact solution exists in the limit M → ∞. In a following paper (1973), Bender and Wu [22] developed two independent mathematical techniques for determining the large-n behavior of PT for the quartic anharmonicity. In fact, asymptotic expressions at large n of perturbation coefficients were immediately called Bender-Wu formulas. After the publication of their results, thousands of physical systems were suddenly connected in one way or another. For example, the Hamiltonian of the one-dimensional quantum quartic anharmonic oscillator represents a φ4 scalar quantum field theory in (0 + 1) dimensions. Many properties that appear in (0 + 1) dimension, like divergence in PT, can also be present in a realistic field theory. The apparently simple problem of a quantum anharmonic oscillator revealed an extremely rich internal structure, which looks intrinsic for any non-trivial system in quantum mechanics. In particular, they formally evidenced the impossibility of finding the spectrum in PT in powers of λ. These facts encouraged the development of various numerical and analytical non-perturbative methods over the next years. Thus, the anharmonic oscillator became a basic system to test new methods. That explains why, from the early ’70s until today, one can find many papers devoted to the exploration of the quantum spectra of the anharmonic potential2. It is worth mentioning some of the methods used throughout the years: Variational Method [23, 24], Rayleigh-Ritz method [25], PT in both the weak and strong coupling regimes [20,22,26–30], Padé Approximants [31], Hill Determinant method [32–34], WKBmethod [35,36], Self-Similar Approximations [37], the method of Characteristic Functions [38], among many, many others. 2.3 Present Work From previous considerations, it is clear that the anharmonic oscillator has a fundamental interest in physics. The present study is devoted to studying the energy spectra of the D- dimensional radial polynomial anharmonic oscillator. One may ask: why to consider it in D-dimension? Of course, the physical spatial dimensions, D = 1, 2, 3, are evidently the most popular. However, as it will be shown, in our study the dimension D does not play an essential role in the formalism that we present below: it appears only as an external parameter. We may easily deal with an arbitrary D, and ultimately set its value to the dimension of interest. Furthermore, some non-trivial analytical properties of the energy as a function of D appear at both physical and not-physical dimensions, [39, 40]. 2Specially in one-dimension for quartic and sextic cases. 10 Chapter 2. Introduction This work aims primarily to study the ground state of the radial anharmonic oscillator in ar- bitrary D using as tools PT in the weak and strong coupling regimes, asymptotic series at small and large distances, and a new version of the semi-classical approximation. In general, here we present the formalism to study an arbitrary polynomial anharmonic oscillator. Nevertheless, as we will see in Part II, it can be applied to physical systems with different nature. Our ultimate goal is to construct a locally-accurate approximation of the ground state wave function in arbitrary (integer) dimension D > 0. We will call this approximation the Approx- imant, and we will denote it by Ψ (t) 0,0. Moreover, we will show how the Approximant Ψ (t) 0,0 can be modified to construct Approximants for the wave functions of excited states. It will be demonstrated that when an Approximant is used as a trial function in the Variational Method (VM), it leads to a highly accurate estimate of the energy with a very small deviation from the exact result for all coupling constant. In fact, some of our results will serve as benchmarks for future works. Furthermore, a fast-convergent PT occurs for both the energy and the wave function if the trial function is used as the zeroth order approximation in a PT scheme3. This is explicitly checked for three anharmonicities: cubic, quartic, and sextic. 3By means of the connection between PT and the variational energy discussed in the previous Chapter. Chapter 3 Generalities In the forthcoming Sections, we present some general results that will be useful for our study of the D-dimensional anharmonic oscillator. 3.1 Radial Potential in D-dimension The time-independent Schrödinger equation for a particle subjected to a spherical-symmetric potential V (r) in a D-dimensional space (D = 1, 2, ...) has the form [ − ~2 2M ∇2 D + V (r) ] ψ = E ψ . (3.1.1) Here ~ is the Planck constant, M is the mass of the particle, ∇2 D = D∑ k=1 ∂2xk (3.1.2) is the D-dimensional Laplace operator in Cartesian coordinates {x1, x2, ...xD}, r = √ x21 + x22 + ...+ x2D , xi ∈ (−∞,∞) , (3.1.3) i = 1, ..., D, is the radius in D-dimension1, while ψ and E represent wave function and its corresponding energy, respectively. Due to the spherical symmetry of the potential V (r), in D > 1 it is convenient to introduce D-dimensional hyperspherical coordinates2 {r,Ω}, with r as in (3.1.3) and Ω parameterized with (D− 1) Euler angles. In these coordinates, the Laplace operator ∇2 D takes the form ∇2 D = ∂2r + D − 1 r ∂r + ∆SD−1 r2 , ∂r ≡ ∂ ∂r , (3.1.4) 1In the present work, we assume that at D = 1 the radius becomes r = |x|, see (3.1.3), thus V = V (|x|). 2See [41,42]. 11 12 Chapter 3. Generalities where ∆SD−1 is the Laplacian on the SD−1 sphere. Naturally, if we choose in (3.1.4) a dimension D = 2, 3, we obtain the familiar expressions of the Laplacian in polar and spherical coordinates, respectively. For the purposes of the present work, the explicit form of the operator SD−1 is not needed. It is enough to know that the operator ∆SD−1 is closely related with the D-dimensional orbital angular momentum L̂,3 ∆SD−1 = −L̂2 . (3.1.5) This relation4 brings two remarkable consequences at D > 1: (i) For any spherically symmetric potential, the D-dimensional angular momentum L̂ is conserved. (ii) In hyperspherical coordinates, we can separate the hyperradial coordinate r from the angular ones (Ω). Hence, in D > 1 the wave function ψ and its energy E can be labeled by one radial quantum number nr, one angular quantum momentum ℓ, and (D − 2) magnetic quantum numbers. As a consequence, the wave function is represented as the product of two functions ψnr,ℓ,{mℓ}(r,Ω) = Ψnr,ℓ(r) ξℓ,{mℓ}(Ω) , (3.1.6) where nr = 0, 1, 2, . . . , ℓ = 0, 1, 2, . . . . The set of (D − 2) magnetic quantum numbers {mℓ}, where each of them takes values from −ℓ to +ℓ, contains N (D, ℓ) different configurations, where N (D, ℓ) =    (2ℓ+D − 2) (ℓ+D − 3)! ℓ! (D − 2)! , ℓ ≥ 1, 1 , ℓ = 0 . (3.1.7) For D > 1, the angular part of the wave function ξℓ,{mℓ}(Ω) corresponds to a D-dimensional spherical harmonic [41, 42] which satisfies the spectral problem −∆SD−1 ξℓ,{mℓ} = L̂2 ξℓ,{mℓ} = ℓ (ℓ+D − 2) ξℓ,{mℓ} . (3.1.8) Hence, for given ℓ this equation is satisfied by N (D, ℓ) different linear independent spherical harmonics, which correspond to the eigenvalue ℓ (ℓ+D−2) with degeneracy5 equals to N (D, ℓ). 3The D(D − 1)/2 independent components of the D-dimensional angular momentum tensor are Lij = xipj − xjpi , i < j , where xi and pi are the ith coordinate and the linear quantum momentum, respectively. 4The operator L̂2 is defined as L̂ 2 = N∑ j=2 j−1 ∑ i=1 L2 ij in terms of its components, see [43]. 5Note that for D = 3, the eigenvalue of L̂2 and its degeneracy are given by the familiar formulas ℓ(ℓ + 1) and N (3, ℓ) = 2ℓ+ 1, respectively. 3.2. Radial Polynomial Anharmonic Oscillator 13 Not surprisingly, two spherical harmonics with different quantum numbers and same ℓ are orthogonal, ∫ Ω ξℓ,{mℓ}ξℓ,{m′ ℓ} dΩ = δmℓ,m ′ ℓ (3.1.9) where δmℓ,m ′ ℓ is the Kronecker delta, and dΩ is the solid angle in D−1 dimensions. Furthermore, two spherical harmonics with different angular momentum, say ℓ and ℓ′, are orthogonal too. Further properties of the spherical harmonics can be found in [41,42]. However, those properties do not play an important role in the present work. Since the theory of the spherical harmonics is well established, we have to focus solely on determining the radial part of the wave function Ψnr,ℓ(r). After substituting (3.1.4), (3.1.6) and (3.1.8) into (3.1.1), we arrive to a radial Schrödinger equation6 [ − ~2 2M ( ∂2r + D − 1 r ∂r − ℓ (ℓ+D − 2) r2 ) + V (r) ] Ψnr,ℓ(r) = Enr,ℓ Ψnr,ℓ(r) , (3.1.10) which determines the radial wave function Ψnr,ℓ(r). For any radial potential, the energy Enr,ℓ of an excited state is always degenerate with respect to the quantum numbers {mℓ}. Specifically, for a given (nr, ℓ) and D > 1 we have N (D, ℓ) different wave functions with the same energy. Note that the magnetic quantum numbers do not play an essential role. For this reason, since now on, we get rid of the label {mℓ} in the energy and wave function. 3.2 Radial Polynomial Anharmonic Oscillator From now on, we consider the radial polynomial anharmonic potential7 V (r) = 1 g2 m∑ k=2 ak g k rk , m ≥ 2 , (3.2.1) where g is a non-negative coupling constant, ak are real parameters. In addition, we consider only potentials such that V (r) > 0 at r > 0, with a2 and am positive. Hence, the minimum at r = 0 is global. For m > 3 there may be some other (local) minimums at r > 0. For the sake of simplicity, we assume that none of them is degenerate. Evidently, under these considerations the Hamiltonian Ĥ = − ~2 2M ∇2 D + 1 g2 m∑ k=2 ak g krk, (3.2.2) 6It is known that by means of a gauge rotation, equation (3.1.10) can be transformed into a one-dimensional Schrödinger one with an effective potential. However, we do not follow this approach since it does not provide advantages in the present study. 7Note that any radial polynomial potential, with a leading quadratic term at small r, can be rewritten as in (3.2.1) employing scalings in coefficients ak. 14 Chapter 3. Generalities has an infinite and discrete spectrum: the potential (3.2.1) admits bound states. Let us note that the potential (3.2.1) can be written in a compact form as V (r) = 1 g2 V̂ (gr) , (3.2.3) where we have defined V̂ (gr) = m∑ k=2 ak (gr) k . (3.2.4) As we discussed in the previous Section, separation of variables occurs for the Schrödinger equation associated to Hamiltonian (3.2.2). Therefore, we only have to determine the radial part of the wave function using the Schrödinger equation (3.1.10). We are not trying to find the complete spectra of the Hamiltonian operator (3.2.2). Our study is devoted to the low-lying states. However, as we will see through these pages, the ground state plays the most important role between the low-lying states. 3.3 The Ground State: Riccati Equation The ground state of any spherical symmetricD-dimensional Hamiltonian is non-degenerate (and nodeless). It has zero angular momentum ℓ = 0 (S-state) and zero radial quantum number nr = 0. The angular part corresponds to the zeroth spherical harmonic (which is a constant) with zero eigenvalue in (3.1.8). From (3.1.10), it can be seen that the ground wave function depends only on the radial coordinate r, we will denote it without labels by Ψ(r). For the D- dimensional radial anharmonic oscillator (3.2.1), the ground state wave function corresponds to the lowest energy eigenfunction of the radial Schrödinger operator ĥr = − ~2 2M ( ∂2r + D − 1 r ∂r ) + 1 g2 V̂ (gr) . (3.3.1) The corresponding spectral problem, with explicit boundary conditions, read ĥr Ψ(r) = EΨ(r) , ∫ ∞ 0 Ψ2 rD−1dr < ∞ , Ψ(0) = 1, Ψ(∞) = 0 . (3.3.2) Here E denotes the ground state energy. It is interesting to note that D may be considered as a continuous parameter in (3.3.2). It will be shown that some non-trivial analytical properties of the ground state energy as a function of D appear at non-physical dimensions, see e.g. [39] and [40]. To solve the spectral problem (3.3.2), it is convenient to adopt the so-called exponential representation, Ψ(r) = e− 1 ~ Φ(r). (3.3.3) The function Φ(r) has dimension of action, and it is usually called the phase of the wave function. Thus, instead of looking for Ψ, we have to find the phase Φ. From (3.3.2), it can 3.3. The Ground State: Riccati Equation 15 be shown that Φ satisfies a non-linear ordinary differential equation of first order and second degree, ~ ∂ry(r) − y(r) ( y(r)− ~(D − 1) r ) = 2M [ E − 1 g2 V̂ (r) ] , (3.3.4) where the function y(r) is nothing but the derivative of Φ(r) y(r) = ∂rΦ(r) . (3.3.5) In fact, y(r) coincides with the logarithmic derivative of the wave function Ψ, y(r) = −~ ∂r (logΨ(r)) . (3.3.6) Equation (3.3.4) is a particular case of the familiar Riccati (differential) equation. It shows that the logarithmic derivative y(r) is more fundamental than the wave function Ψ, and even more than the phase. The Riccati equation is the key equation of this Chapter. 3.3.1 Riccati-Bloch Equation The Riccati equation (3.3.4) can be transformed into one without the explicit dependence on the constants ~ and M . This is achieved by introducing the quantum coordinate8 v = ( 2M ~2 ) 1 4 r , (3.3.7) and making the replacements y = (2M~ 2) 1 4 Y (v) , E = ~ (2M) 1 2 ε . (3.3.8) In this way, the equation (3.3.4) becomes ∂vY(v) − Y(v) ( Y(v)− D − 1 v ) = ε (λ) − 1 λ2 V̂ (λv) , ∂v ≡ d dv , (3.3.9) where λ = ( ~2 2M ) 1 4 g (3.3.10) plays the role of the effective coupling constant replacing g. Using previous definitions, it is evident that λ v = g r . (3.3.11) The boundary conditions which we impose to the equation (3.3.9) are ∂vY(v) ∣ ∣ ∣ v=0 = ε D , Y(+∞) = +∞ . (3.3.12) 8We call v the quantum coordinate because it depends explicitly on the Planck constant ~. 16 Chapter 3. Generalities These two boundary conditions guarantee that the corresponding wave function is normalizable, leading to a bound state. Evidently, if ~ = 1 and M = 1/2, the equations (3.3.9) and (3.3.4) coincide with the same coupling constant λ = g. We call the equation (3.3.9) the Riccati-Bloch (RB) equation. It governs the dynamics of the logarithmic derivative in the v-space. The RB equation (3.3.9) allows us to construct the asymptotic expansions of Y(v) at small and large v.9 For fixed λ 6= 0 and small v, we have the Taylor series10 Y(v) = ε D v+ (ε2 − a2D 2) D2(D + 2) v3− a3λ D + 3 v4−a4D 3(D + 2)λ2 − 2 ε (ε2 − a2D 2) D3(D + 2)(D + 4) v5 + . . . . (3.3.13) For large v, Y(v) = a1/2m λ(m−2)/2vm/2 + am−1λ (m−4)/2 2a 1/2 m v(m−2)/2 + (4amam−2 − a2m−1)λ (m−6)/2 8a 3/2 m v(m−4)/2 + . . . . (3.3.14) It is easy to check that for λ = 0, hence, for the harmonic oscillator potential V (r) = a2 r 2, the equation (3.3.9) has the exact solution Y(v) = ε D v , ε = a 1/2 2 D . (3.3.15) Not surprisingly, this solution leads to the well-known ground state Gaussian wave function of the harmonic oscillator. Four remarks in a row: (i) The expansion (3.3.13) at λ = 0 is terminated, and it consists of the first term alone, leading to Y(v) = (ε/D) v, which is in agreement with (3.3.15). (ii) In general, the majority of the coefficients in the expansion of Y(v) in powers of v, see (3.3.13), depends on the energy ε explicitly. (iii) The fourth and subsequent (unwritten) terms in (3.3.14) can acquire explicit dependence on ε. (iv) For M fixed, large v can be obtained in two situations: large r and fixed ~, or fixed r and small ~, see (3.3.7). Hence, the expansion (3.3.14) can describe the semi-classical limit ~ → 0. Integrating (3.3.13) and (3.3.14) with respect to the coordinate v, both expansions are con- verted into series for 1 ~ Φ, namely, 1 ~ Φ = ε 2D v2 + (ε2 − a2D 2) 4D2(D + 2) v4 − a3λ 5(D + 3) v5 − a4D 3(D + 2)λ2 − 2ε(ε2 − a2D 2) 6D3(D + 2)(D + 4) v6 + . . . , (3.3.16) 9Ultimately, it allows us to construct the expansion of y(r) at large distances once variable r is restored. 10This is the expansion around the global minimum. 3.3. The Ground State: Riccati Equation 17 and 1 ~ Φ = 2a 1/2 m m+ 2 λ(m−2)/2v(m+2)/2+ am−1λ (m−4)/2 ma 1/2 m vm/2 + (4amam−2 − a2m−1)λ (m−6)/2 4(m− 2)a 3/2 m v(m−2)/2 + . . . , (3.3.17) for small and large v, respectively. These two expansions contain the asymptotic information about the exact wave function that later will be used. 3.3.2 Generalized Bloch Equation There exists an alternative way to transform the original Riccati equation (3.3.4) into a new one in which the constants ~ and M do not appear explicitly. This is achieved by introducing the classical coordinate11 u = g r , (3.3.18) and defining a new unknown function Z (u) = ( g2 2M )1/2 y . (3.3.19) It is straightforward to show that the function Z(u) satisfies a non-linear differential equation λ2 ∂uZ(u) − Z(u) ( Z(u)− λ2(D − 1) u ) = λ2 ε(λ) − V̂ (u) , ∂u ≡ d du , (3.3.20) c.f. (3.3.9). Here ε and λ are the same as in (3.3.8) and (3.3.10). Therefore, these two quantities play the role of energy and effective coupling constant in both spaces: the v-space and the u- space. Boundary conditions, which guarantee the normalizability of the wave function, are similar to those imposed for the RB equation, ∂uZ(u) ∣ ∣ ∣ u=0 = ε D , Z(+∞) = +∞ . (3.3.21) At D = 1, the equation (3.3.20) was called in [44–46] the (one-dimensional) Generalized Bloch equation. Here, the equation (3.3.20) is a natural extension of the one-dimensional GB equation to the D-dimensional radial case. For this reason, we continue to call it as the Generalized (radial) Bloch equation. In contrast with the RB equation, by setting ~ = 1 and M = 1/2 the GB equation does not take the form of the Riccati one shown in (3.3.4). This is one of the principal features that differentiate the RB equation from the GB one. Naturally, (3.3.20) can also be used to construct the asymptotic expansions similar to (3.3.13) and (3.3.14) for small u, Z(u) = ε D u+ (ε2 − a2D 2) D2(D + 2)λ2 u3 − a3 (D + 3)λ2 u4 − a4D 3(D + 2)λ2 − 2 ε (ε2 − a2D 2) D3(D + 2)(D + 4)λ4 u5 + . . . , (3.3.22) 11We called it like this since it is ~-independent. In fact, in the classical counterpart of the anharmonic oscillator, the classical equations of motion written in variable u are g-independent, see [44, 45]. 18 Chapter 3. Generalities and large u, Z(u) = a1/2m um/2 + am−1 2a 1/2 m u(m−2)/2 + (4amam−2 − a2m−1) 8a 3/2 m u(m−4)/2 + . . . , (3.3.23) respectively. We note the following: (i) Small u can be obtained in two situations, small r and fixed g, or fixed r and small coupling constant g. (ii) Large u occurs when r is large for fixed g, or for fixed r and g large. Therefore, it is not a surprise that expansions in v-space (3.3.13) and (3.3.14) can be connected with (3.3.22) and (3.3.23) in u-space, respectively. This is achieved by using the remarkable relation u = λ v , (3.3.24) which connects the two spaces. Chapter 4 The Weak Coupling Regime Assuming a small effective coupling constant λ, we describe how to solve the RB and GB equations using PT in λ. In particular, we will show that the perturbative solution of the RB equation is an application of the Non-Linearization Procedure [7]. On the other hand, the perturbative approach developed for the GB equation leads to a purely algebraic procedure to calculate arbitrary corrections. Finally, we discuss the non-trivial connection between these two solutions and its relation with the WKB approximation. 4.1 Riccati-Bloch Equation Let us take the RB equation (3.3.9) with an arbitrary potential V (v;λ) in the right-hand side1, ∂vY(v) − Y(v) ( Y(v)− D − 1 v ) = ε (λ) − V (v;λ) , (4.1.1) We assume that potential V (v;λ) admits a Taylor expansion in powers of the effective coupling constant λ, thus V (v;λ) = ∞∑ n=0 Vn(v)λ n . (4.1.2) Here the coefficient functions Vn(v), n = 0, 1, ..., are real functions in v. Let us now assume that the unperturbed equation at λ = 0, ∂vY0(v) − Y0(v) ( Y0(v)− D − 1 v ) = ε0 − V0(v) , (4.1.3) 1We take (3.3.9) and make the replacement 1 λ2 V̂ (λv) → V (v;λ). 19 20 Chapter 4. The Weak Coupling Regime can be solved explicitly by any means.2 Evidently, once we know Y0(v), the wave function can be found since Ψ0 = exp ( − ∫ v Y0(s) ds ) . (4.1.4) The only constraint when finding Y0(s) is that the function Ψ0 should be normalizable. Once the solution of the unperturbed problem is known, we can develop PT in powers of λ, ε(λ) = ∞∑ n=0 εn λ n , (4.1.5) and Y(v) = ∞∑ n=0 Yn(v)λ n . (4.1.6) Now the task is to find the corrections εn and Yn(v). To do so, we substitute (4.1.5) and (4.1.6) into the RB equation (4.1.1); it is easy to see that the nth correction Yn(v) satisfies the first order linear ordinary differential equation3 ∂v ( vD−1Ψ2 0 Yn(v) ) = (εn −Qn(v)) v D−1Ψ2 0 , (4.1.7) where Q1(v) = V1(v) , Qn(v) = Vn(v) − n−1∑ k=1 Yk(v)Yn−k(v) , n = 2, 3, . . . . (4.1.8) Evidently, the exact solution for Yn(v) is written in integral form Yn(v) = 1 vD−1 Ψ2 0 (∫ v 0 (εn −Qn)Ψ 2 0 s D−1 ds ) . (4.1.9) Since we are interested in finding bound states, it is necessary to impose the condition of the absence of current of particles for both v → 0 and v → ∞ as boundary condition [8], Yn(v) v D−1 Ψ2 0 ∣ ∣ ∣ v={0,∞} → 0 . (4.1.10) It can be easily checked that if v → 0, the condition (4.1.10) is satisfied automatically, while at v → ∞ the correction εn has to be chosen accordingly εn = ∫∞ 0 Qn Ψ 2 0 v D−1 dv ∫∞ 0 Ψ2 0 v D−1 dv (4.1.11) to satisfy (4.1.10). In D = 1, PT developed to solve the Riccati equation (associated with a given Schrödinger equation) was called the Non-Linearization Procedure [8]. We have given its straightforward extension to radial potentials in arbitrary D. We will continue to call it the Non-Linearization Procedure. Some remarks: 2It can always be achieved via the Inverse Problem in the framework of the RB equation: we take some function Y0(v) and then calculate the left-hand side of (4.1.3). Finally, we solve for the potential V0(v) and ε0. 3The product vD−1Ψ2 0 is the integrating factor of the differential equation. 4.1. Riccati-Bloch Equation 21 (i) In contrast with the Rayleigh-Schrödinger PT, see e.g. [1], the knowledge of the entire spectrum of the unperturbed problem is not required to find perturbative corrections in (4.1.5) and (4.1.6). It is sufficient to know the unperturbed state wave function of the unperturbed problem to which we are looking for corrections. (ii) This approach gives the closed analytic expression for both corrections εn and Yn(v) in the form of nested integrals. Therefore, this procedure is an efficient method to calculate several orders in PT, see e.g. [8] and [47]. (iii) In this framework, the convergence of the perturbation series (4.1.5) and (4.1.6) is guar- anteed for any D if the first correction Y1(v) is bounded |Y1(v)| ≤ C (4.1.12) where C is a positive constant, for discussion and details see [8]. (iv) Using the Non-Linearization Procedure we can exploit the connection between the varia- tional energy and PT to calculate corrections of a given trial function and its variational estimate of the energy. We have presented a brief review of the Non-Linearization Procedure applied to the ground state. This approach can be modified to study excited states by admitting a finite number of simple poles with residues equal to one in the corrections Yn(v). The position of the poles is also found in PT in powers in λ. Explicit formulas can be found in [8] for the one-dimensional case. The generalization of the D-dimensional radial potential case is straightforward. 4.1.1 D-dimensional Radial Anharmonic Oscillator Let us now consider the D-dimensional radial anharmonic oscillator potential (3.2.1). We study the weak coupling regime assuming that the effective coupling constant λ = ( ~2 2M ) 1 4 g is small. To apply the Non-Linearization Procedure, we choose in (4.1.2) Vk = ak+2 v k+2 for k = 0, 1, . . . ,m− 2 , Vk = 0 for k > m− 2 . (4.1.13) Therefore, our potential is a polynomial of finite degree m − 2 and m, in λ and v, respec- tively. The corresponding unperturbed equation (4.1.3) describes the D−dimensional spherical harmonic oscillator: an exactly solvable problem in any dimension D > 0, see e.g. [48]. The unperturbed ground state wave function is given by Ψ0 = e − √ M 2~2 a 1/2 2 r2 = e− 1 2 a 1/2 2 v2 , (4.1.14) 22 Chapter 4. The Weak Coupling Regime and it does not depend on the dimension D. In general, the explicit calculation of the perturbative corrections Yn(v) and εn can be carried out analytically by using (4.1.9). However, the final expressions involve the incomplete gamma function and the Meijer G-function explicitly. Due to the form of those expressions, they are irrelevant for our study since the information that they may provide is hidden in cumbersome formulas. Hence, we take a different but more relevant approach: asymptotic series in particular domains. Let us find the asymptotic expansions of the nth correction Yn(v) in two limits: v → 0, and v → ∞. For small v, the correction Yn(v) is given by the Taylor expansion Yn(v) = v ∞∑ k=0 b (n) k vk , (4.1.15) where b (n) k are some coefficients. It can be easily shown that the first coefficient b (n) 0 is related to the energy correction εn, b (n) 0 = εn D . (4.1.16) The next coefficient always vanishes, b (n) 1 = 0 , (4.1.17) due to the absence of the linear term in the radial anharmonic potential (3.2.1). On the other hand, for large v we have a Laurent-like series, Yn(v) = v ∞∑ k=0 c (n) k vn−k , n > 0 , (4.1.18) where c (n) k are some coefficients. Interestingly, the leading coefficient c (n) 0 does not depend on D while the next-to-leading one always vanishes, c (n) 1 = 0 . (4.1.19) Using expansions (4.1.15) and (4.1.18), it can be shown that in the non-physical dimension D = 0 all corrections εn vanish, εn = 0 . (4.1.20) From the integral form of correction εn shown in (4.1.11), one can show that if D → 0, the numerator of (4.1.11) is bounded, ∫ ∞ 0 Qn Ψ 2 0 v D−1 dv < ∞ , (4.1.21) while the denominator behaves like ∫ ∞ 0 Ψ2 0 v D−1 dv = 1 D a −D 4 2 + O(D0) (4.1.22) 4.2. Generalized Bloch Equation 23 for D → 0. Consequently, the correction εn vanishes linearly in D when D → 0. Note that the coefficient b (n) 0 remains finite in this limit, see (4.1.16). If any correction εn vanishes for arbitrary n, then their formal sum (4.1.5) results in ε = 0, and the energy vanishes, E = 0. In general, for D 6= 0 it can be shown that expansion for ε is asymptotic: εn grows at least factorially as n → ∞, see e.g. [20, 22, 28, 49]. Therefore, series (4.1.5) and (4.1.6) are divergent. This is not surprising since a Dyson [50] instability is present for most of the radial anharmonic oscillators of the form (3.2.1). The divergent nature of the series could discourage to calculate corrections in PT. However, for some particular cases of (3.2.1) it has been proved that perturbation series for the energy can be summed to the exact one by using different summation techniques : calculating the Borel sum [27], taking Padé approximants [31] and via renormalization [51]. These regularization procedures can lead to highly accurate estimates of the energy for some systems. In particular, it is successful for some cases of the anharmonic oscillator. An interesting situation occurs when all odd monomial terms in r in potential (3.2.1) are absent, i.e. the potential is (formally) an even function in r, V (r) = V (−r). In this case, all odd corrections Y2n+1(v) and ε2n+1 vanish. Likewise, any even correction Y2n(v) has the form of a polynomial of finite degree Y2n(v) = v n∑ k=0 c (2n) 2k v2(n−k) , Y2n(−v) = −Y2n(v) , (4.1.23) where c (2n) 2n = ε2n D . (4.1.24) This implies that ε2n and Y2n(v) can be calculated by linear algebra means4. In turn, the energy correction ε2n is a finite-degree polynomial in D, ε2n = n+1∑ k=1 d (2n) k Dk , (4.1.25) where d (2n) k are real coefficients. However, it is enough to have a single odd monomial term ∼ v2q+1 in the potential (3.2.1) (with q positive and integer) to break the features (4.1.23) and (4.1.25). Hence, all coefficients in front of the singular terms in the expansion (4.1.18), as well as all higher order terms in the expansion (4.1.15) at k > n, are proportional to a2q+1. 4.2 Generalized Bloch Equation We can develop PT in powers of λ for the GB equation (3.3.20). Of course, the expansion of ε in powers of λ is the same as the one used in the RB equation, see (4.1.5), ε(λ) = ∞∑ n=0 εn λ n . 4In this situation, a large number of corrections can be calculated by using a computational code, see e.g. [47]. 24 Chapter 4. The Weak Coupling Regime In turn, the expansion for the function Z(u) reads Z(u) = ∞∑ n=0 Zn(u)λ n . (4.2.1) We assume that the energy corrections εn are already known: they can be found using the Non- Linearization Procedure, the standard Rayleigh-Schrödinger PT, or any other suitable method. It is immediate to see that the correction Zn(u) is calculated by algebraic means. Its calculation depends on corrections of a smaller order, just like it happens in any perturbative approach. The first 3 corrections are Z0(u) = ± √ V̂ (u) , Z1(u) = 0 , (4.2.2) Z2(u) = u ∂uZ0(u) + (D − 1)Z0(u) − u ε0 2 uZ0(u) . In general, the nth correction is given by Zn(u) = u ∂uZn−2(u) + (D − 1)Zn−2(u) − u2 ∑n−2 i=2 Zi(u)Zn−i(u) − u εn−2 2 uZ0(u) , n > 2 . (4.2.3) Note that the boundary condition (3.3.21), Z(∞) = +∞, implies that the positive sign of the square root must be chosen in the expression for Z0(u). Analyzing the Zn(u) correction (4.2.3), one can see that the boundary condition at u = 0, see (3.3.21), can not be fulfilled in the case of an arbitrary potential (3.2.1) without defined parity with respect to the change r → −r. In this situation, any Zn(u) (and its derivative) at n > 1 diverges at small u. For instance, for the cubic potential where a3 6= 0 the nth correction Zn(u) with n > 1 behaves like Zn(u) ∼ u−n+2 (4.2.4) when u → 0. In turn, for the quintic potential with a3 = 0 but a5 > 0 the nth correction at n > 3 behaves like Zn(u) ∼ u−n+4 (4.2.5) when u tends to zero. For the polynomial potential of degree (2k + 1) in u, if all other odd terms are absent a3 = a5 = . . . = a2k−1 = 0 but a2k+1 > 0, the function Zn(u) behaves like Zn(u) ∼ u−n+2k (4.2.6) for small u as long as n > 2k − 1. It might be an indication that the radius of convergence of the expansion of Z(u) (see (4.2.1)) in 1/u is finite. However, if the anharmonic potential V (r) is formally even, V (r) = V (−r) the boundary condition at u = 0 can always be satisfied. This allows us to determine the correction εn by imposing ∂uZn(u) ∣ ∣ ∣ u=0 = εn D . (4.2.7) 4.3. Connection Between the Expansions for Y and Z 25 At this point, it is worth mentioning that the expansion in powers of λ (4.2.1) is divergent for any anharmonic oscillator due to the Dyson instability mentioned below. This fact can be seen in the partial sums of (4.2.1), see Figs. 4.1 and 4.2. Let us emphasize that the presence of a single odd degree monomial term in u in the potential V (u) breaks the feature (4.2.7): the expansion for Z(u) in powers of λ only satisfies the single boundary condition at u = ∞. It is necessary to discuss the connection between the expansions (4.1.6) and (4.2.1) to explain why this happens. Figure 4.1: Partial sums of the expansion for Z(u) as functions of u for a general anharmonic potential with monomials of even degrees. The function Z(u) (solid blue line), represents the exact solution, see text. Near u = 0, the deviation of the partial sums from the exact solution is a consequence of the divergent nature of series (4.1.6). 4.3 Connection Between the Expansions for Y and Z Thus far, we have constructed two different representations for the logarithmic derivative y(r), see (3.3.6). Due to the uniqueness of the solutions of the Schrödinger equation, there must be a connection between those representations. According to the RB equation, we have the y(r) = (2M~ 2) 1 4 Y(v(r)) , Y(v) = ∞∑ n=0 Yn(v)λ n , v(r) = ( 2M ~2 ) 1 4 r . (4.3.1) 26 Chapter 4. The Weak Coupling Regime Figure 4.2: Partial sums of the expansion for Z(u) as functions of u for a generic anharmonic potential which includes odd monomials. The function Z(u) (solid blue line) represents the exact solution, see text. The divergence in the vicinity of u = 0 of the partial sums is the result of the impossibility of satisfying the boundary condition at u = 0. From the other side, the GB equation leads to y(r) = ( 2M g2 ) 1 2 Z(u(r)) , Z(u) = ∞∑ n=0 Zn(u)λ n , u(r) = g r . (4.3.2) The coupling constant λ is defined in (3.3.10). The connection can be established if we use explicitly the expansion of Yn(v) at large v, see (4.1.18). In this case, y(r) is given formally by y = (2M ~ 2) 1 4 v ∞∑ n=0 λn ∞∑ k=0 c (n) k vn−k . (4.3.3) By using the relation between the v-space and u-space (3.3.24), u = λ v , and changing the order of summation, it follows that (4.3.3) is equivalent to y = ( 2M g2 ) 1 2 ∞∑ n=0 λn(gr) ∞∑ k=1 c(k)n (gr)k−n . (4.3.4) Now, the connection is clear. If we compare the expansions (4.3.3) and (4.3.4), we can conclude that Zn(u) = u ∞∑ k=1 c(k)n uk−n (4.3.5) 4.4. Connection with WKB 27 due to the uniqueness of the Taylor series. Therefore, Zn(u) is a generating function of the coefficients c (n) k , k = 0, 1, ... , see Fig. 4.3 for a graphical illustration. Equation (4.3.5) exhibits y(r) = (2M) 1 2 g ( c (0) 0 (gr)1 + c(1)0 (gr)2 + λ2 c(1)2 (gr)0 + λ3 c(1)3 (gr)−1 + λ4 c(1)4 (gr)−2 + . . . + c(2)0 (gr)3 + λ2 c(2)2 (gr)1 + λ3 c(2)3 (gr)0 + λ4 c(2)4 (gr)−1 + . . . + c(3)0 (gr)4 + λ2 c(3)2 (gr)2 + λ3 c(3)3 (gr)1 + λ4 c(3)4 (gr)0 + . . . + c(4)0 (gr)5 + λ2 c(4)2 (gr)3 + λ3 c(4)3 (gr)2 + λ4c(4)4 (gr)1 + . . . + ... + ... + ... + ... + . . . ) ︸ ︷︷ ︸ ︸ ︷︷ ︸ ︸ ︷︷ ︸ ︸ ︷︷ ︸ ≡ (2M) 1 2 g ( Z0(gr) + λ2 Z2(gr) + λ3 Z3(gr) + λ4 Z4(gr) + . . . ) Figure 4.3: Representation of the generating functions Zn constructed from the perturbative series in powers of λ for the function y(r). the reason why we cannot satisfy, in general, the boundary condition at u = 0: Zn(u) is constructed from the expansion (4.1.18), which is valid in the limit r → ∞ (u → ∞ for fixed g). It also clarifies the reason why this boundary condition can be satisfied only for even potentials: when formally V (r) = V (−r), the (finite) series of any Yn(v) at small distances is the same than for large ones. Some interesting properties of the corrections Zn(u) occur when considering r → ∞. It will be discussed in detail for the cubic, quartic, and sextic anharmonic oscillators, see below. 4.4 Connection with WKB Let us conclude this Chapter by showing the relation between the perturbative approach im- plemented for the GB equation and the semi-classical WKB approximation. By integrating in r, the expansion (4.3.2) is converted into an expansion in terms of generating functions for the phase Φ(r) = ∞∑ n=0 λnGn(r) , G1(r) = 0 , (4.4.1) where Gn(r) is equal to Gn(r) = ( 2M g2 )1/2 ∫ r Zn(gr) dr . (4.4.2) 28 Chapter 4. The Weak Coupling Regime Note that keeping g and M fixed, (4.4.1) can be regarded as a semi-classical expansion of the phase in powers of ~ 1 2 , see (3.3.10). The leading order of this expansion is given by G0(r) = ∫ r√ 2M V (r) dr . (4.4.3) This leading order term is the classical action for vanishing energy5 E = 0, and it is related to the zeroth order correction of the standard D-dimensional (radial) WKB method6 (developed in the classically forbidden region), see below. This is the only term in the expansion (4.4.1) which contains no dependence on D. The second order term in (4.4.1) is λ2G2(r) = ~ ( D − 1 2 log[r] + 1 4 log [2M V (r)] − ε0 2 ∫ r dr √ V (r) ) . (4.4.4) Except for the integral - the third term - it looks like the first order correction to WKB at E = 0 and arbitrary D, see Appendix A. Thus, it defines the functional determinant of a path integral formalism [45]. The appearance of the extra term in the form of an integral can be explained as follows. Let us take the zeroth order term in the standard WKB method (in the forbidden region) and expand it in powers of ~1/2 using the fact that the energy is ~-dependent, see (3.3.8) and (3.3.10), thus ∫ r√ 2M (V (r)− E) dr = ∫ r√ 2M V (r) dr − ~ ε0 2 ∫ r dr √ V (r) + O(~3/2) . (4.4.5) Note that the first term is nothing but G0(r) while the second term is exactly the integral which appears in (4.4.4). Consequently, one can see that (4.4.1) is the standard semi-classical WKB expansion in the classically-forbidden region, re-expanded in powers of ~1/2. Higher order generating functions G3(r), G4(r), . . ., are related with higher order corrections of the WKB expansion in a similar way. This is an alternative consideration to the one-dimensional case presented in [44–46]. In these papers, it was described how to obtain expansion (4.4.1) using a path integral formulation (in Euclidean time) considering a special classical solution of zero energy (E = 0) called flucton. In a quite straightforward way, this consideration can be extended to the D-dimensional radial case, where the flucton path appears in the radial direction. Hence, the perturbative approach developed for the GB Equation can be called the true semi-classical approximation: both ε(λ) and Φ(r;λ) are expanded in powers ~1/2. 5A vanishing energy corresponds to the classical energy of a particle located at the global minimum. 6We present a brief review of the WKB method in Appendix A. Chapter 5 The Strong Coupling Regime In Chapter 4, assuming a small effective constant λ = ( ~2 2M ) 1 4 g, we studied the ground state energy E and the phase in the form of a Taylor expansion around λ = 0. As a result, the harmonic oscillator was taken as the unperturbed problem. We now consider the opposite limit: large λ in both RB and GB equations. In this case, we take as unperturbed potential the leading monomial term amg m−2rm in (3.2.1). It implies that all other terms in the potential (3.2.1), except for the leading one, are taken as perturbations. This situation can be achieved if we consider the expansion of the energy in fractional powers 1/λ. 5.1 Riccati-Bloch Equation To develop the strong coupling expansion in the framework of the RB equation, let us first introduce the new quantum coordinate1 w = λ m−2 m+2 v , (5.1.1) and define the new function Ỹ(w) = λ 2−m m+2Y(v) . (5.1.2) When we introduce these two expressions in the RB equation (3.3.9), we have ∂wỸ(w) − Ỹ(w) ( Ỹ(w)− D − 1 w ) = ε̃(λ̃) − 1 λ̃m V̂ ( λ̃ w ) , ∂w ≡ d dw , (5.1.3) 1The insertion of this coordinate in the RB equation is equivalent to perform a Symanzik scaling transfor- mation to the Schrödinger equation, see [9]. 29 30 Chapter 5. The Strong Coupling Regime where ε̃ = λ 4−2m m+2 ε , λ̃ = λ 4 m+2 , (5.1.4) cf. (3.3.9). This equation does not contain explicit dependence on ~ and M , while ε̃ plays the role of energy. Furthermore, (5.1.3) is the RB equation that is appropriate to construct the strong coupling expansion. This can be easily seen if we note that the potential is now given by 1 λ̃m V̂ ( λ̃ w ) = 1 λ̃m m∑ k=2 ak (λ̃ w) k = amw m + am−1 λ̃ wm−1 + . . .+ a2 λ̃m−2 w2 . (5.1.5) Thus, the parameter λ̃−1 is now the coupling constant. Therefore, when large λ̃ is considered, the potential becomes a single monomial in the variable w in power m. This monomial now represents the unperturbed potential. The latter implies that Ỹ(w) and ε̃(λ̃) should eventually be expanded in series in the following form Ỹ(w) = ∞∑ n=0 Ỹn(w) λ̃ −n . (5.1.6) and ε̃(λ̃) = ∞∑ n=0 ε̃n λ̃ −n , (5.1.7) respectively. Here ε̃n, n = 0, 1, ... , are the coefficients of the strong coupling expansion. A simi- lar expansion to (5.1.7) occurs for the excited states. Contrary to the weak coupling expansion, it has been demonstrated that, for some particular cases of the anharmonic potential (3.2.1), PT developed in the strong coupling regime (5.1.7) has a finite radius of convergence, see e.g. [52] and references therein. It is related to the fact that the dominant term in the potential (3.2.1) is rm at r → ∞ while other monomials are subdominant. Therefore, the phenomenon of the Dyson instability does not occur. It has to be emphasized that the equation (5.1.3) has a similar form than the RB equation with arbitrary potential (4.1.1). This implies that all the formulas, derived in Section 4.1 in the framework of the Non-Linearization Procedure, are suitable to calculate Ỹn(w) and ε̃n: we only have to make the replacement v → w in (4.1.1). Thus, let us choose the potential that appears in (4.1.2) in the following form Vk = am−k w m−k for k = 0, 1, ..., m− 2 , Vk = 0 for k > m− 2 . (5.1.8) In this case, the unperturbed equation (4.1.3) is ∂vỸ0(w) − Ỹ0(w) ( Ỹ0(w)− D − 1 w ) = ε̃0 − amw m . (5.1.9) If the solution of this equation were known, the corresponding square-integrable unperturbed wave function Ψ0 would be obtained using (4.1.4). Then, we would use the formulas (4.1.9) and 5.1. Riccati-Bloch Equation 31 (4.1.11) to construct PT. As a result, we would obtain the coefficients ε̃k of the strong coupling expansion. However, since the exact solution of (5.1.9) for m > 2 is unknown, we need to find Ỹ0(w) in approximate form. In the spirit of the present work, there are two possible ways to find Ỹ0(w): (i) One can choose the simplest physically relevant trial function for the potential W0(w) = a2pw 2p , m = 2p (5.1.10) which defines the unperturbed dimensionless radial Schrödinger equation − 1 2 ( ∂2w + D − 1 w ∂w ) Ψ0,0 + W0(w)Ψ0,0 = ε̃0,0Ψ0,0 . (5.1.11) Then, we use the Non-Linearization Procedure employing the connection between PT and the variational energy, see (1.16). (ii) We can construct a more sophisticated locally-accurate approximation of Ỹ0(w) for w ≥ 0. Let us describe both options. 5.1.1 Physically Relevant Trial Function Probably, the simplest physically relevant trial function, see [8] for discussion, was proposed for the first time in [7, 53] handling the one-dimensional case. Undoubtedly, the similar trial function is also appropriate for the D-dimensional case of the radial potentials, Ψ0,0(w) = e−a 1/2 2p wp+1 p+1 , Ỹ0,0(w) = a 1/2 2p wp , (5.1.12) where the second subscript marks the order of the correction to Ψ0. By solving the inverse problem, one can find the potential W0,0(w) for which Ψ0,0 is the exact solution, namely W0,0(w) = a2pw 2p − a 1/2 2p (p+D − 1)wp−1 , ε̃0,0 = 0 . (5.1.13) Taking the difference between the potentials (5.1.10) and (5.1.13), W0,1(w) = W0(w) − W0,0(w) = a 1/2 2p (p+D − 1)wp−1 , (5.1.14) one can expand PT as in (4.1.5) and (4.1.6) with W0,1(w) as the perturbative potential for W0,0(w), see Section 1.2 for details. In this manner, it allows us to calculate the zeroth order coefficient ε̃0 in the strong coupling expansion (5.1.7) iteratively, which appears in the form of an expansion ε̃0 = ε̃0,0 + ε̃0,1 + ε̃0,2 + . . . . (5.1.15) 32 Chapter 5. The Strong Coupling Regime The first two coefficients can be found analytically, ε̃0,0 = 0 , (5.1.16) ε̃0,1 = a 1 p+1 2p (p+D − 1) ( p+ 1 2 ) p−1 p+1 Γ(D+p−1 p+1 ) Γ( D p+1 ) . (5.1.17) The next ones can be obtained numerically2. The function Ỹ0(w) is expanded in a similar way as in (5.1.15), Ỹ0(w) = Ỹ0,0(w) + Ỹ0,1(w) + Ỹ0,2(w) + . . . , (5.1.18) where Ỹ0,0(w) is given by (5.1.12). Evidently, this expansion can be converted into the expansion of the phase, Φ0 = Φ0,0 + Φ0,1 + Φ0,2 + . . . (5.1.19) where Φ0,0 = a 1/2 2p ~ wp+1 p+ 1 . Using Ψ0,0 as entry one can also calculate explicitly the first ap- proximation to ε̃1, the subdominant coefficient in the strong coupling expansion (5.1.7), which results in ε̃1,0 = am−1 ( p+ 1 2a 1/2 2p ) 2p−1 p+1 Γ(D+2p−1 p+1 ) Γ( D p+1 ) . (5.1.20) The next terms, e.g. ε̃1,1, ε̃1,2, ..., can be only written in the form of nested integrals, similar to those as in expression (4.1.11). Their computation is numerical. Of course, this procedure can be extended to calculate higher coefficients in the expansion (5.1.7). In general terms, it can be shown that in this approach series (5.1.15) and (5.1.19) are convergent. However, the rate of convergence is very slow. It implies that several corrections are needed to obtain highly accurate results from their partial sums. Since corrections are calculated numerically, the accumulation of error in the computations is a drawback. This inconveniences can be avoided if we accelerate the convergence using a more appropriate zeroth order approximation than (5.1.12). 5.1.2 Interpolation: Locally Accurate Approximation One can construct another physically relevant trial function for the RB Equation (5.1.9) that leads to a fast convergent series (5.1.15) and (5.1.19). To do so, we employ the asymptotic behavior of Ỹ0(w) at small w, Ỹ0(w) = ( ε̃0 D w + ε̃20 D2 (D + 2) w3 − a2p δp,3/2 D + 3 w4 + 2ε̃30 − a2pD 3(D + 2)δp,2 D3(D + 2)(D + 4) w5 + . . . ) , (5.1.21) 2 In Section 7.2, which is devoted to the cubic anharmonic oscillator, the first coefficients will be presented explicitly. 5.2. Generalized Bloch Equation 33 and at large w Ỹ0(w) = ( a 1/2 2p wp + 1 2 (D + p− 1)w−1 − 1 2 ε̃0w −p + . . . ) . (5.1.22) The dominant term in (5.1.22) corresponds to (5.1.12). A more sophisticated approximation of Ỹ0,0(w) is the result of the interpolation between expansions (5.1.21) and (5.1.22). The con- struction of such interpolation will be presented in Chapter 6. We show how this interpolation leads to a locally accurate approximation for the cubic, quartic, and sextic anharmonic oscil- lators. Furthermore, it leads to highly accurate estimates of the first two coefficients of the strong coupling expansion. 5.2 Generalized Bloch Equation There is an interesting connection between the strong coupling expansion developed for the GB equation (3.3.20) and the expansion of y(r) at small r, see (3.3.13). To construct the GB equation for the strong coupling regime, it is necessary to note that u and w are related in a very simple way, u = λ̃ w . (5.2.1) This can be checked straightforwardly just by noting that λ̃ w = λ 4 m+2 ( λ m−2 m+2 v ) = λ v = u , (5.2.2) see (5.1.1) and (5.1.4). Interestingly, the equation (5.2.2) looks very similar to (3.3.24). This remark suggests introducing into the RB equation (5.1.3) the classical coordinate u = λ̃ w , (5.2.3) and a new function Z̃(u) = λ̃ Ỹ (w) . (5.2.4) In this way, we obtain λ̃2 ∂uZ̃(u) − Z̃(u) ( Z̃(u) − λ̃2 (D − 1) u ) = λ̃2 ε̃ − λ̃−m+2 V̂ (u) . (5.2.5) This equation is the GB Equation for the strong coupling regime: the left-hand side is the same as in (3.3.20) - the GB equation for weak coupling regime - while the right-hand side is different3. Now we can develop PT in inverse powers of λ̃−1 in the form ε̃(λ̃) = ∞∑ n=0 ε̃n λ̃ −n. 3Except when m = 2. In this irrelevant case - the harmonic oscillator - there is no coupling constant, see (3.2.1). 34 Chapter 5. The Strong Coupling Regime see (5.1.7), and Z̃(u) = ∞∑ n=0 Z̃n(u) λ̃ −n , (5.2.6) This is nothing but the expansions at λ̃ = ∞. We assume the coefficients ε̃n are known a priori. In particular, they may be computed from PT developed for the RB equation in the strong coupling regime. Inserting the series (5.1.7) and (5.2.6) in (5.2.5), one can determine corrections Z̃n(u) iteratively by solving the same first order differential equation with a different right-hand side ∂uZ̃i(u) + (D − 1) u Z̃i(u) = ε̃i , i = 0, 1 , (5.2.7) and ∂uZ̃n(u) + (D − 1) u Z̃n(u) = ε̃n + n−2∑ i=0 Z̃i(u) Z̃n−i−2(u) − δm,n V̂ (u) , n = 2, 3, . . . . (5.2.8) Imposing the boundary condition Z̃(0) = ε̃/D to the original equation (5.2.5), we can find the boundary condition for the nth correction, ∂uZ̃(u) ∣ ∣ ∣ u=0 = ε̃n D . (5.2.9) The equations (5.2.7) and (5.2.8) are solved in an elementary way. In particular, Z̃0(u) = ε̃0 D u , Z̃1(u) = ε̃1 D u , Z̃2(u) = ε̃2 D u + ε̃20 D2(D + 2) u3 , Z̃3(u) = ε̃3 D u + 2 ε̃0 ε̃1 D2(D + 2) u3 − δm,3 m∑ k=2 ak D + k uk+1 , Z̃4(u) = ε̃4 D u + 2 ε̃0 ε̃2 + ε̃21 D2(D + 2) u3 + 2 ε̃30 D3(D + 2)(D + 4) u5 − δm,4 m∑ k=2 ak D + k uk+1 . The general nth correction is given by Z̃n(u) = ε̃n D u + u1−D n−2∑ k=0 ∫ u 0 Z̃k(s) Z̃n−k−2(s) s D−1 ds − δm,n m∑ k=2 ak u k+1 D + k (5.2.10) at n ≥ 2. Note that it is not guaranteed that the boundary condition Z̃(∞) = +∞ is fulfilled for all these corrections. It implies that PT in inverse powers of λ̃ for Z̃(u) has a finite radius of convergence in u. Thus, the corrections Z̃n(u) make sense in a bounded domain in u only, see 5.2. Generalized Bloch Equation 35 discussion below. From (5.2.10), one can see that any correction is a finite-degree polynomial in u, Z̃n(u) =    [n 2 ] ∑ k=0 α (n) 2k u 2k+1 − δm,n m∑ k=2 ak u k+1 D + k for n ≤ m+ 1 , n∑ k=0 α (n) k uk+1 for n > m+ 1 , (5.2.11) where α (n) k are real coefficients, and the symbol [ ] denotes the integer part function. It is easy to see that α (n) 0 = ε̃n D (5.2.12) following (5.2.9). The next coefficient always vanishes, α (n) 1 = 0 , (5.2.13) for any n. The coefficient α (n) 2 can be found in the form of finite sum α (n) 2 = 1 D2(D + 2) n−2∑ k=0 ε̃kε̃n−k−2 , (5.2.14) while the next coefficient again vanishes, α (n) 3 = 0 , (5.2.15) for any n, c.f. (5.2.13). From (5.2.11), one can see that Z̃n(u) for n < m and n = m + 1 is always an odd polynomial in u: Z̃n(−u) = −Z̃n(u) , while for n = m and n > m + 1 this property may be lost. In the case of an even anharmonic potential, V (r) = V (−r), all odd corrections Z̃2n+1(u) and ε̃2n+1 vanish. In turn, all even corrections Z̃2n(u) are odd polynomials in u of degree (2n+ 1) if 2n > m+ 1. The perturbative approach, which was used to calculate the strong coupling expansion of Z̃(u), is related to the asymptotic behavior of Z̃(u) at small u. Once we established the poly- nomial structure of the corrections Z̃n(u), one can sum up some subseries of the perturbative expansion (5.2.6), keeping the degree of u fixed. For example, the sum ( ∞∑ n=0 α (n) 0 λ̃−n ) u = ε̃ D u , (5.2.16) corresponds to summing up all terms O(u) in (5.2.6). The next sum is carried out with all terms O(u3). Using (5.2.14) one can show that it corresponds to ∞∑ n=2 α (n) 2 λ̃−nu3 − a2 λ̃ −m D + 2 u3 = λ̃−2 ε̃2 − a2 λ̃ −mD2 D2(D + 2) u3 . (5.2.17) 36 Chapter 5. The Strong Coupling Regime The next sum contains coefficients in front of u4 terms. Following (5.2.15), this sum has a single term, − a3 λ̃ −m D + 3 u4 . (5.2.18) This summation procedure can be extended to higher order terms un, n > 4. After performing such summations, we obtain the coefficients in front of all terms in the Taylor expansion in powers of u for the function Z̃(u), and, eventually, for y(r). From another side, one can construct the expansion in powers of u directly from (5.2.5), Z̃(u) = ε̃ D u + λ̃−2 ε̃2 − a2 λ̃ −mD2 D2(D + 2) u3 − a3 λ̃ −m D + 3 u4 − a4 λ̃ −mD3(D + 2)− 2 ε̃ λ̃−2(ε̃2λ̃−2 − a2 λ̃ −mD2) D3(D + 2)(D + 4) u5 + . . . . (5.2.19) One can note the exact correspondence between the first three terms of the expansion (5.2.19) and the subseries previously defined in (5.2.16), (5.2.17) and (5.2.18). In this way, we can see the connection between the strong coupling expansion, developed on the basis of the GB equation, and the small u one: both expansions lead to the same representation - the Taylor series - of Z̃(u) and, eventually, of y(r). Finally, this connection clarifies why the boundary condition Z̃(∞) = +∞ can not always be fulfilled. Chapter 6 The Approximant Now we formulate a prescription for the construction of a locally-accurate, uniform approxima- tion of the wave function of a given state. Such approximate function is called The Approximant. 6.1 The Approximant for the Ground State We denote this approximation by Ψ (t) 0,0 and call it the Approximant. Here, the subscripts indicate explicitly the quantum numbers of the ground state (nr = 0, l = 0) at any D.1 It is convenient to assume the exponential representation of the wave function, Ψ (t) 0,0(r) = e− 1 ~ Φt(r) , (6.1.1) since the phase Φ of the exponential representation is not that sharp as the wave function Ψ.2 Hence, we focus on the construction of the approximate phase Φt. We follow the prescription that the phase Φt has to interpolate the expansions for small and large distances, (3.3.16) and (3.3.17), respectively. Besides, it should interpolate the same expansions constructed in the strong coupling regime (5.1.21) and (5.1.22). One of the simplest interpolations with these characteristics is of the following form 1 ~ Φt(r) = ã0 + ã1 g r + 1 g2 V (r ; ã2, . . . , ãm) √ 1 g2 r2 V (r ; b̃2, . . . , b̃m) + Logarithmic Terms (r ; {c̃}) . (6.1.2) 1 At D = 1, the angular quantum number is not defined. However, we may fix it (formally) to zero, ℓ = 0. 2 This was explicitly seen in a particular example described in the Overview. 37 38 Chapter 6. The Approximant Without loss of generality, we put b̃2 = 1 as normalization. Here, V (r; {ã}) and V (r; {b̃}) are modified potentials of the form of the original one (3.2.1): instead of the external given parameters {a} some free parameters {ã} and {b̃} are taken, respectively. The logarithmic terms can depend on free parameters {c̃} as well. These terms will eventually insert a prefactor in (6.1.1). Parameters {c̃} are chosen in such a way that this prefactor has no real nodes. All those remaining free parameters will be fixed by making the variational calculation with (6.1.2) taken as the phase of the trial function. At g = 0, the phase (6.1.2) becomes ã0 + ã2r 2, which corresponds to the radial harmonic oscillator phase. Formula (6.1.2) is the key result of this Chapter. The phase Φt has the outstanding property that, with an appropriate choice of parameters (ã0, ã1; ã2, b̃2, . . . ãk, b̃k; {c̃}), the generating function G0(r) is partially (or exactly) reproduced. In particular, to reproduce exactly the dominant term in the expansion (3.3.17), the condition ãm = 2 a 1/2 m m+ 2 b̃m−2 (6.1.3) is imposed3. Further constraints to parameters can be imposed in order to reproduce all growing terms in the expansion of the exact function Y(v), see (3.3.14). If such constraints are imposed, all the corrections in PT to Y0 = ∂rΦt will be bounded, and it will guarantee the convergence of the perturbation series (1.18). To mimic the possible appearance of logarithmic terms in the expansion (3.3.17), in particular, those which may occur in G0(r) and G2(r), some logarithmic terms to phase Φt should be added, see (4.4.4). For example, for the two-term anharmonic oscillator potential V (r) = r2 + gm−2 rm , (6.1.4) the generating function G2(r) always contains logarithmic terms, g2G2(r) (2M)1/2 = 1 4 log [ 1 + (gr)m−2 ] + D m+ 2 log [ 1 + √ 1 + (gr)m−2 ] , (6.1.5) see (4.4.2). In this case, the logarithmic terms added in (6.1.2) should be a certain minimal modification of the logarithmic terms that appear in G0(r) and G2(r). For instance, a minimal modification of g2G2(r)/(2M)1/2 is of the form 1 4 log [ 1 + c̃ (gr)m−2 ] + D m+ 2 log [ 1 + √ 1 + c̃ (gr)m−2 ] , (6.1.6) where c̃ is a free parameter. If c̃ = 1, it coincides with (6.1.5). In general, some constraints on parameters have to be imposed to guarantee that Φt has the structure of (3.3.16) and that Ψ (t) 0,0 is a normalizable function. Other constraints come from 3Without loss of generality, this constraint is written for the case ~ = 1 and M = 1/2. 6.2. Approximants for Excited States 39 the symmetry of the phase. For example, if the potential is formally even, V (r) = V (−r), the function Φt has to be even too. After imposing all the constraints described above, the remaining free parameters can be fixed by giving them the values of optimal parameters obtained in variational calculations. To realize it, we use the VM for the radial Hamiltonian (3.3.2) taking the Approximant4 Ψ (t) 0,0(r) = Prefactor(r ; {c̃})× exp  − ã0 + ã1 g r + 1 g2 V (r ; ã2, . . . , ãm) √ 1 g2 r2 V (r ; b̃2, . . . , b̃m)   , (6.1.7) as a trial function, see (6.1.1) and (6.1.2). The accuracy of variational calculations can be estimated using the connection between the variational energy and PT5. This kind of estimates will be presented in the forthcoming Chapter for particular anharmonicities. 6.2 Approximants for Excited States Case D = 1 In the one-dimensional situation r = |x|, where x ∈ (−∞,∞), see (3.1.3). Therefore, the potential is an even function V = V (|x|). Due to the symmetry of the potential under the reflection x → −x the quantum number which indicates the parity ν = (−1)p, p = 0, 1, is well defined. It implies that we can use it to label any wave function6, namely, Ψ(t) nr,p(x) = xpP (p) nr (x2)e− 1 ~ Φt(x) , nr = 0, 1, ..., p = 0, 1 , (6.2.1) where P (p) nr (x 2) is a polynomial of degree nr with real coefficients and all real roots, and with P (p) nr (0) = 1 chosen for normalization. The polynomial in (6.2.1) defines the nodal surface of the approximate wave function. With vanishing coupling constant (g = 0) the polynomial xpP (p) nr (x 2) is well-known, it is written in terms of the Hermite polynomials xpP (p) nr (x2) = H2nr+p(x) . (6.2.2) The phase Φt has exactly the same form as in (6.1.2) but with a different set of free parameters {ã}, {b̃}, {c̃}. The same constraints imposed on the parameters of the ground state should be 4It is beyond our purpose to study the analytical properties of the Approximant in the complex r-plane. However, except for the real axis, it is clear that the Approximant develops nodes in the complex plane. 5See Section 1.2 6The energies (which now are labeled as Enr,p) obey the following inequality Enr,0 < Enr,1 < Enr+1,0 for any coupling constant g. 40 Chapter 6. The Approximant imposed for the excited ones. To fix the value of the coefficients of P (p) nr (x 2), some orthogonality constraints are imposed. For fixed (nr, p), the (nr − 1) free parameters of P (p) nr (x 2) are found by demanding7 〈Ψ(t) nr,p|Ψ (t) kr,p 〉 = 0 , kr = 0, . . . , (nr − 1) . (6.2.3) These constraints will ensure that the function Ψ (t) nr,p(r) has the correct number of nodes (n = 2nr + p).8 The remaining free parameters are fixed through the VM. Case D > 1 Due to the separation of the Schrödinger equation (3.1.1) in spherical coordinates, it is sufficient to approximate the radial part of the wave function. One can label the eigenstates by the pair of quantum numbers (nr, ℓ), omitting the magnetic quantum number. In this case, the Approximant is of the form Ψ (t) nr,ℓ (r) = rℓ P (ℓ) nr (r2) e−Φt , nr = 0, 1, ..., ℓ = 0, 1, ..., (6.2.4) where P (ℓ) nr (r 2) is a polynomial of degree nr with real roots. We set P (ℓ) nr (0) = 1 as normalization. Note that at g = 0, the polynomial P (ℓ) nr (r 2) is given in terms of the Laguerre associated polynomials [48] P (ℓ) nr (r2) = Lℓ−D/2−1 nr (r2) . (6.2.5) Analogously to the one-dimensional case, in (6.2.4) Φt has the same form of the phase con- structed for the ground state (6.1.2), but with a different set of parameters {ã}, {b̃}, {c̃}. The same set of constraints imposed on the parameters of the ground state should be imposed for the excited ones. For a given (nr, ℓ), the coefficients of the polynomial P (ℓ) nr (r) are fixed imposing the orthogonality constraints 〈Ψ(t) nr,ℓ |Ψ(t) kr,ℓ 〉 = 0 , kr = 0, . . . , (nr − 1) . (6.2.6) These constraints will make sure that the function Ψ (t) nr,p(r) has the correct number of nodes (n = 2nr + l). Note that two states with different angular momentum, say ℓ and ℓ′ are already orthogonal, independently of the value of their radial quantum numbers, see Section 3.1. The remaining free parameters are fixed by means of the VM. 7 We used the familiar bra-ket notation to represent the scalar product. 8 In practice, n can be used as the only quantum number to label the states. Chapter 7 Results for Two-Term Potentials The formalism, described in the previous Chapters, is now applied to 3 particular cases of two-term radial anharmonic potentials: cubic, quartic, and sextic. Hence, the forthcoming discussion is focused on potentials of the form V (r) = r2 + gm−2rm , (7.0.1) with m = 3, 4, 6. For each anharmonicity, we present results of analytical and numerical nature in the following order. We begin by presenting some general features of the PT coming from the RB and GB equations developed in the weak coupling regime.1 Then, we construct the Approximants of the first 4 low-lying states for which we estimate their corresponding energy by means of the VM at gm−2 = 0.1, 1, 10 in D = 1, 2, 3, 6. For some of those states, we use the connection between PT and the variational energy to estimate its accuracy, as well as the accuracy of the Approximant. Finally, employing the Approximant, we compute in VM the first two coefficients of the strong coupling expansion of the ground state energy. As we describe below, the realization of the VM and the estimation of its accuracy is a numerical procedure. In the next Section, we briefly describe the computational codes used in numerical computations. Full details can be found in [54], [55], and [56]. Finally, let us mention that the Section for each anharmonicity is self-contained, written in such a way that they can be read independently. 7.1 Numerical Realization of the Variational Method It is not a surprise that the calculation of the variational energy Evar[ψt] with trial function ψt = Ψ (t) 0,0, shown in (6.1.7), requires to perform a one-dimensional numerical integration, see (1.13). In addition, a numerical minimization is needed in order to obtain the optimal configuration of parameters. A computational code was written in FORTRAN 77. It uses the integration routine D01FCF 1 Some general properties of the two-term potential are summarized in Appendix D. 41 42 Chapter 7. Results for Two-Term Potentials from the NAG-LIB, which was built using the algorithm described in [57]. The routine of optimization that we used to find the optimal variational parameters was MINUIT of the CERN-LIB. Needless to say, the use of Approximants for excited states as trial functions in VM of the form (7.2.25) and (7.2.26) also require numerical integration and minimization. Without loss of generality in (numerical) variational calculations we set ~ = 1 and M = 1/2, thus v = r, ε = E, λ = g, and Y = y see (3.3.7), (3.3.8) and (3.3.10). With this setting, it is clear that Evar corresponds to a variational estimate of the exact energy ε. For each anharmonicity in D = 1, we performed variational calculations for the first three low-lying states which are characterized by the quantum numbers (0, 0), (0, 1), and (1, 0), see (6.2.1). In turn, at D = 2, 3, 6 we consider the first four states with quantum numbers (0,0), (1,0), (0,1), and (0,2), see (6.2.4). We calculate the variational energy Evar = E (t) 0 + E (t) 1 (7.1.1) for those states. Using the connection between the variational energy and PT2, we take y0 = ∂rΦt as zeroth order correction. Then, the accuracy of the variational estimate Evar is found by calculating the second correction3 E (t) 2 (7.1.2) in the expansion (1.18). Furthermore, we can define the corrected variational energy Evar + E (t) 2 (7.1.3) which, in principle, should provide a more accurate estimate of the exact energy. Correction E (t) 2 is found numerically using the formalism of the Non-Linearization Procedure presented in Section 4.1. The calculation of E (t) 2 requires the knowledge of y1, which is already known in integral form, see (4.1.9). A computational code in Mathematica package was designed; it includes all relevant formulas. Note that once y1 is known, the accuracy of the Approximant itself can be estimated via the relative deviation from the exact wave function ∣ ∣ ∣ ∣ ∣ Ψ0,0(r)−Ψ (t) 0,0(r) Ψ (t) 0,0(r) ∣ ∣ ∣ ∣ ∣ . 10−δ . (7.1.4) Here δ is a positive constant that is obtained numerically. In (7.1.4), the function Ψ0,0 denotes the (unknown) exact one. If δ is sufficiently large, we will say that Ψ (t) 0,0 is a locally accurate approximation of the exact wave function Ψ0,0. We can compare the corrected variational energies Evar + E (t) 2 with those coming from an independent numerical method. We choose the Lagrange-Mesh method [58]: an approximate variational method simplified by a Gauss quadrature associated with the mesh. It is known that this method, in the form proposed and developed by D. Baye and co-authors, provides easily 12 - 13 s. d. in energies for various problems of atomic and molecular physics. A detailed and complete description can be found in [58]. At D = 1, we use the Hermite mesh, while for D > 1, the Laguerre one. 2Presented in Section 1.2 3 It is the second correction in formal PT, but the first one to the variational energy. 7.2. Cubic Anharmonic Oscillator 43 7.2 Cubic Anharmonic Oscillator The simplest radial anharmonic oscillator is characterized by a cubic anharmonicity, V (r) = r2 + g r3 , (7.2.1) see (7.0.1) at m = 3. It is worth mentioning that many features that the cubic anharmonic potential exhibits are also present in the general potential without symmetry of reflection r → −r, i.e. V (r) 6= V (−r). Thus, the cubic anharmonic oscillator covers the basis of understanding of more complex potentials. Let us note that at D = 1, r → |x|. Therefore, the one-dimensional version of the potential (7.2.1) is V (x) = x2 + g|x|3 , −∞ < x <∞ , (7.2.2) which clearly holds bound states. 7.2.1 Perturbation Theory in the Weak Coupling Regime For the cubic anharmonic oscillator, the perturbative expansions of ε and Y(v) can be derived for the RB equation (3.3.9), ∂vY(v) − Y(v) ( Y(v)− D − 1 v ) = ε (λ) − v2 − λ v3 , ∂v ≡ d dv , (7.2.3) with v and λ defined in (3.3.7) and (3.3.10), respectively. Those expansions remain functionally the same as for the general oscillator, ε = ε0 + ε1 λ + ε2 λ 2 + . . . , ε0 = D , and Y(v) = Y0(v) + Y1(v)λ + Y2(v)λ 2 + . . . , Y0(v) = v . Since potential (7.2.1) contains an odd degree monomial r3, the correction Yn(v) is characterized by the infinite series (4.1.15) and (4.1.18) at small and large v, respectively. The first three corrections Yn(v), at both v → 0 and v → ∞, are presented in Appendix B. The first corrections ε1 and Y1(v) can be calculated in closed analytic form using (4.1.9) and (4.1.11), ε1 = Γ(D+3 2 ) Γ(D 2 ) , Y1(v) = ev 2 2 vD−1 { ε1 γ ( D 2 , v2 ) − γ ( D + 3 2 , v2 )} , (7.2.4) where Γ(a) and γ(a, b) denote the complete and incomplete gamma functions, respectively, see [59]. Higher energy corrections εn can be computed only numerically.4 For example, the first six corrections5 at D = 1 are ε1 = 0.564 190 , ε2 = −0.373 405 , ε3 = 0.512 464 , ε4 = −0.941 502 , ε5 = 2.051 176 , ε6 = −5.022 987 . (7.2.5) 4Interestingly, the behavior of εn at large D can be obtained analytically in the 1/D-expansion, see [29]. 5Rounded to the first displayed s. d. 44 Chapter 7. Results for Two-Term Potentials 7.2.2 Generating Functions We can determine the coefficients of the expansion of Yn(v) at large v, c (n) k in (4.1.18), by algebraic means. In general, they are written in terms of the corrections ε0, ε1, . . ., εn, see Appendix B for explicit formulas for n = 1, 2, 3. It was pointed out in [8,39,60,61] that the general expression for the coefficients c (n) 0 , c (n) 2 , . . . can be obtained by solving recurrence relations. For example, c (n) 0 satisfies non-linear recurrence relation c (n) 0 = −1 2 n−1∑ k=1 c (k) 0 c (n−k) 0 , c (1) 0 = 1 2 . (7.2.6) In turn, the recurrence relation for c (n) 2 is linear c (n) 2 = 1 2 (2n+D) c (n) 0 − n−1∑ k=1 c (k) 0 c (n−k) 2 , c (1) 2 = 1 4 (D + 2) . (7.2.7) The solution of (7.2.7) requires the knowledge of the coefficients c (n) 0 . It is evident that in order to find c (n) 3 , the coefficients c (n) 0 and c (n) 2 are needed, and so on. The simplest way to solve these recurrence relations is by means of generating functions. Surprisingly, we can calculate these generating functions straightforwardly via the algebraic and iterative procedure derived from the GB equation in the weak coupling regime. For the cubic anharmonic oscillator, the GB equation reads λ2 ∂uZ(u) − Z(u) ( Z(u)− λ2(D − 1) u ) = λ2 ε(λ) − u2 − u3 , ∂u ≡ d du . (7.2.8) For the particular cubic anharmonicity, the function Z(u) admits an expansion of the form Z(u) = Z0(u) + Z2(u)λ 2 + Z3(u)λ 3 + . . . , (7.2.9) with Zk(u) = u ∞∑ n=1 c (n) k un−k . (7.2.10) Evidently, the generating function of the coefficient c (n) k is nothing but Zk(u). Generating functions can be easily calculated using formulas presented in (4.2.2) and (4.2.3). For example, the first two are Z0(u) = u √ 1 + u , (7.2.11) Z2(u) = u+ 2D ( 1 + u− √ 1 + u ) 4u(1 + u) . (7.2.12) From these expressions, the explicit solutions of the equations (7.2.6) and (7.2.7) may be easily derived, c (n) 0 = (−1)n+1 Γ(2n+ 1) 22n−1Γ(n) Γ(n+ 1) , c (n) 2 = (−1)n+1 2 ( 1 +D Γ(2n+ 1) 22n Γ(n+ 1)2 ) . (7.2.13) 7.2. Cubic Anharmonic Oscillator 45 Note that the coefficients c (n) 0 are D-independent, while c (n) 2 depends on D linearly. As men- tioned before, an important property of the generating functions Z0(u), Z2(u), Z3(u), ... is related to the asymptotic expansion of the function y(r) at large r, keeping fixed the effective coupling constant λ. This expansion is conveniently written in variable v, y = (2M~ 2) 1 4 ( λ1/2v3/2 + 1 2λ1/2 v1/2 − 1 8λ3/2 v−1/2 + 2D + 1 4 v−1 − (8λ2ε− 1) 16λ5/2 v−3/2 + . . . ) (7.2.14) see (3.3.7). Note that the first four terms of this expansion are ε-independent, but only first three are D-independent. From another side, the expansion of (2M g2 )1/2Z0(u) at large u is ( 2M g2 )1/2 Z0(u) = ( 2M g2 )1/2( u3/2 + 1 2 u1/2 − 1 8 u−1/2 + 1 16 u−3/2 + . . . ) . (7.2.15) To compare the expansions (7.2.14) and (7.2.15), let us replace the classical coordinate u by the quantum one v, taking into account that u = λv, see (3.3.24). Then, the expansion (7.2.15) becomes ( 2M g2 )1/2 Z0(λv) = (2M~ 2) 1 4 ( λ1/2v3/2 + 1 2λ1/2 v1/2 − 1 8λ3/2 v−1/2 + 1 16λ5/2 v−3/2 + . . . ) . (7.2.16) One can intermediately see that it reproduces exactly the expansion (7.2.14) up to O(v−1/2). With a similar approach, we can demonstrate that at large r the function ( 2M g2 )1/2 λ2Z2(λv) contributes to the expansion (7.2.14) from O(v−1), since ( 2M g2 )1/2 λ2Z2(λv) = (2M~ 2) 1 4 ( 2D + 1 4 v−1 − D 2λ1/2 v−3/2 − 1 4λ v−2 + . . . ) . (7.2.17) Note that the expansion of (2M g2 )1/2(Z0(gr) + λ2Z2(gr)) at large v reproduces exactly the first four terms in (7.2.14). The expansion of the next generating functions (2M g2 )1/2λ3Zn(λv), with n > 2, starts from O(v−3/2). Explicitly, ( 2M g2 )1/2 λnZn(λv) = (2M~ 2) 1 4 ( −εn−2 λ n−5/2 2 v−3/2 + . . . ) , v → ∞ , n > 2 , (7.2.18) where εn−2 is the energy PT correction of the order (n− 2). As a consequence, no matter how many generating functions we consider in the expansion (2M g2 )1/2(Z0(λv) + λ2Z2(λv) + ...), the term of order O(v−3/2) in (7.3.18) can not be reproduced exactly. 46 Chapter 7. Results for Two-Term Potentials 7.2.3 The Approximant By means of (4.4.2), we can calculate the first two6 generating functions, G0(r) and G2(r) in the expansion (4.4.1). They are given by G0(r) = (2M)1/2 g2 ( 2(−2 + 3gr)(1 + gr)3/2 15 ) , (7.2.19) G2(r) = (2M)1/2 g2 ( 1 4 log[1 + gr] +D log [ 1 + √ 1 + gr ]) . (7.2.20) In contrast with G2(r), the function G0(r) contains no logarithmic terms. Theses generating functions serve as the building blocks for the construction of the Approximant. Following the general prescription (6.1.2), we write the Approximant of the ground state in the exponential representation Ψ (t) 0,0 = e− 1 ~ Φt with 1 ~ Φt(r) = ã0 + ã1 gr + ã2 r 2 + ã3 g r 3 √ 1 + b̃1 g r + 1 4 log[1 + b̃1 g r] + D log [ 1 + √ 1 + b̃1 g r ] . (7.2.21) We set ã3 = 2 5 b̃ 1/2 1 , (7.2.22) in order to reproduce exactly the leading term in the asymptotic behavior of the phase at r → ∞. Here, the logarithmic terms in Φt represent a minimal modification of those coming G2(r), cf. (7.2.20). Additionally, we impose the constraint ã1 = b̃1 4 (2 ã0 −D − 1) (7.2.23) in order to guarantee a vanishing Φ′ t at r = 0. Thus, there is no linear term in the expansion of (7.2.21) at small distances, just as it is indicated in (3.3.16). Eventually, the Approximant in its final form depends on three free parameters {ã0, ã2, b̃1} only, and it reads Ψ (t) 0,0(r) = 1 ( 1 + b̃1 g r )1/4 ( 1 + √ 1 + b̃1 g r )D exp  − ã0 + ã1 g r + ã2 r 2 + ã3 g r 3 √ 1 + b̃1 g r   , (7.2.24) 6The next two functions G3(r) and G4(r) are presented explicitly in Appendix B. 7.2. Cubic Anharmonic Oscillator 47 where ã3,1 are defined by constraints (7.2.22), (7.2.23). This is the eventual expression of the 3-parametric approximate ground state wave function. To construct 3-parametric Approximants for excited states, we follow the prescription pre- sented in Section 6.2. Hence, at D = 1 we have Ψ(t) nr,p(x) = xpP (p) nr (x 2) ( 1 + b̃1g|x| )1/4 ( 1 + √ 1 + b̃1g|x| )D exp  − ã0 + ã1g|x|+ ã2x 2 + ã3g|x|3 √ 1 + b̃1g|x|   , (7.2.25) while at D > 1 Ψ (t) nr,ℓ (r) = rℓ P (ℓ) nr (r 2) ( 1 + b̃1 g r )1/4 ( 1 + √ 1 + b̃1 g r )D exp  − ã0 + ã1 g r + ã2 r 2 + ã3 g r 3 √ 1 + b̃1 g r   . (7.2.26) Definitions of P (p) nr (x 2) and P (ℓ) nr (r 2) are presented in (6.2.1) and (6.2.4), explicitly. 7.2.4 Numerical Results: Variational Calculations The calculations of the variational energy for the first low-lying states with quantum numbers (0, 0), (0, 1), (0, 2), (1, 0) are presented in Tables 7.1 - 7.4 for different values of D and g. For some states, the variational energy Evar = E (t) 0 + E (t) 1 , the first correction E (t) 2 to it as well as its corrected value Evar + E (t) 2 are shown. All states are studied in dimension D = 1, 2, 3, 6 and coupling constant g = 0.1, 1, 10. For all cases, the variational energy Evar is obtained with absolute deviation from the exact result of the order 10−7−10−8 (7−8 s.d.7). This deviation is found by calculating the second correction E (t) 2 (the first correction to the variational energy). To check the energies obtained in Non-Linearization Procedure, we calculated them numeri- cally in the Lagrange-Mesh method. These numerical calculations show that all digits presented for Evar +E (t) 2 - the variational energy with the first (second order in PT) correction E (t) 2 taken into account - are exact. We obtain not less than 9 d.d.8 correctly in Non-Linearization Proce- dure. This accuracy is confirmed independently by calculating the second correction E (t) 3 to the 7From now on, s.d. stands for significant digits. 8Decimal digits is abbreviated as d.d. 48 Chapter 7. Results for Two-Term Potentials variational energy: this correction is . 10−9 for any D and g that we have studied. It indicates a very fast rate of convergence in Non-Linearization Procedure with trial functions (7.2.24), (7.2.25), and (7.2.26) taken as zeroth order approximations of the wave function. In general, the energies grow with the increase of D and/or g. At D > 1, we can classify the states by radial quantum number nr and angular momentum ℓ as (nr, ℓ); we mention the hierarchy of eigenstates which holds for any fixed integer D > 1 and g > 0: (0,0), (0,1), (0,2), (1,0). It has to be noted that to the best of our knowledge, these results represent the first accu- rate calculations of the energy of the low-lying states for the D-dimensional cubic anharmonic oscillator. Hence, our results will serve as the benchmark for future studies. The Devi- ation of the Approximant from the exact (unknown) eigenfunction can be estimated via the Non-Linearization Procedure. It can be shown that for the ground state function the relative deviation is extremely small and bounded. ∣ ∣ ∣ ∣ ∣ Ψ0,0(r)−Ψ (t) 0,0(r) Ψ (t) 0,0 ∣ ∣ ∣ ∣ ∣ . 10−4 , (7.2.27) This small deviation occurs in the whole range of r ∈ [0,∞) at any dimension D that we explored and at any coupling constant g > 0 considered. Therefore, the Approximant leads to a locally accurate approximation of the exact wave function Ψ0,0(r) once the optimal parameters for Ψ (t) 0,0 are chosen. In general, optimal variational parameters are smooth and slow changing functions of g without a strong D dependence. As an example, for the ground state (0,0) the plots of the parameters ã0,2,3 vs g for D = 2, 3, 6 are shown in Appendix C, see Fig. C.1. A similar situation appears for the excited states for different D and g. Analyzing those plots, one can see the (accidental) appearance of another constraint, ã0 = 125 ã23 + 12 150 ã3 . (7.2.28) It is fulfilled by the optimal parameters with high accuracy. This constraint guarantees an approximate reproduction of the subdominant term in the expansion (7.2.14). One can demand the Approximant to fulfill exactly this constraint. When doing so, no deterioration in the accuracy of the wave function and energy occurs. In all cases studied, the first correction y1 to the logarithmic derivative y0 = ∂rΦt is un- bounded. The reason is clear: we have only imposed the constraint (7.2.22) to Φt, the exact reproduction of the dominant term in the expansion (7.2.14).9 If in addition to constraints (7.2.22), (7.2.23), and (7.2.28), we impose − g2ã1 = 9375 a43 − 1000 a23 + 48 15000 a33 , (7.2.29) Φt reproduces (exactly) all growing terms of the exact phase at r → ∞: r5/2, r3/2, r1/2, and log r see (7.2.14). Thus, y1 will be bounded. Surprisingly, the accuracy of the variational estimate 9The approximate reproduction of the subdominant term was imposed by the VM. 7.2. Cubic Anharmonic Oscillator 49 Table 7.1: Ground state energy for the cubic potential V (r) = r2 + g r3 for D = 1, 2, 3, 6 and g = 0.1, 1, 10, labeled by quantum numbers (0,0) at any D. Variational energy Evar = Evar, the first correction E (t) 2 (rounded to three s.d.) found with use of Ψ (t) 0,0, see text, and the corrected energy Evar +E (t) 2 = Evar +E (t) 2 shown. Evar +E (t) 2 coincides with Lagrange Mesh results (see text) in 9 displayed d. d., hence all printed digits are exact. g D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 1.053 120 300 5.39× 10−7 1.053 119 761 2.124 027 648 4.40× 10−7 2.124 027 208 1.0 1.387 428 891 4.00× 10−8 1.387 428 851 2.877 490 906 3.76× 10−8 2.877 490 868 10.0 2.729 533 139 6.56× 10−7 2.729 532 483 5.794 213 459 5.58× 10−7 5.794 212 901 g D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.208 922 743 4.00× 10−7 3.208 922 343 6.528 432 540 2.02× 10−7 6.528432338 1.0 4.442 965 260 3.15× 10−8 4.442 965 229 9.465 319 951 1.85× 10−8 9.465319933 10.0 9.094 985 589 4.23× 10−7 9.094 985 166 19.981 458 504 1.96× 10−7 19.981 458 308 Table 7.2: The first excited state energy for the cubic potential V (r) = r2 + g r3 for different D and g. Variational energy Evar, the first correction E (t) 2 found with use of Ψ (t) 0,1 for D = 1 and Ψ (t) 0,1 for D > 1, see (7.2.25) and (7.2.26). The corrected energy Evar + E (t) 2 = Evar + E (t) 2 shown, the correction E (t) 2 rounded to 3 s.d. All printed digits for Evar + E (t) 2 are exact. g D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.208 922 765 4.21× 10−7 3.208 922 343 4.305 557 665 3.55× 10−7 4.305 557 309 1.0 4.442 965 265 3.59× 10−8 4.442 965 229 6.068 723 537 2.92× 10−8 6.068 723 507 10.0 9.094 985 630 4.64× 10−7 9.094 985 166 12.579 594 377 3.48× 10−7 12.579 594 029 g D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 5.412 425 220 2.86× 10−7 5.412 424 933 8.784 695 351 1.21× 10−7 8.784 695 230 1.0 7.745 092 165 2.41× 10−8 7.745 092 141 13.018 486 318 1.49× 10−8 13.018 486 303 10.0 16.215 748 127 2.66× 10−7 16.215 747 861 27.841 430 199 1.37× 10−7 27.841 430 061 50 Chapter 7. Results for Two-Term Potentials Table 7.3: The second excited state energy for the cubic potential V (r) = r2 + g r3 for different D and g. Variational energy Evar found with use of Ψ (t) 1,0 for D = 1 and Ψ (t) 0,2 for D > 1, see (7.2.25) and (7.2.26). The first correction E (t) 2 and the corrected energy Evar+E (t) 2 = Evar+E (t) 2 shown. Correction E (t) 2 rounded to 3 s.d. All 9 printed d. d. in Evar + E (t) 2 are exact. g D = 1 D = 2 Evar Evar −E(t) 2 Evar + E (t) 2 0.1 5.436 849 553 6.528 432 582 2.43× 10−7 6.528 432 338 34 1.0 7.879 141 644 9.465 319 955 2.21× 10−8 9.465 319 932 56 10.0 16.641 305 904 19.981 458 531 2.23× 10−7 19.981 458 308 14 g D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 7.652 743 974 1.87× 10−7 7.652 743 787 11.069 434 802 6.73× 10−8 11.069 434 735 1.0 11.224 406 591 1.87× 10−8 11.224 406 573 16.699 837 135 1.22× 10−8 16.699 837 123 10.0 23.860 743 313 1.78× 10−7 23.860 743 135 36.070 426 676 1.01× 10−7 36.070 426 576 Table 7.4: The third excited state energy for the cubic potential V = r2 + g r3 for different D and g. The radial node r0 calculated in Lagrange-Mesh method with 9 correct d.d. Variational energy Evar and its radial node r (0) 0 given by underlined digits, both found with Ψ (t) 1,0, see (7.2.26). g D = 2 D = 3 D = 6 Evar r (0) 0 Evar r (0) 0 Evar r (0) 0 0.1 6.570 941 476 0.953 355 772 7.709 696 057 1.162 462 764 11.158 149 471 1.626 259 519 1.0 9.690 376 045 0.780 270 156 11.517 370 455 0.942 382 083 17.128 462 904 1.289 488 891 10.0 20.681 622 942 0.532 038 126 24.758 598 047 0.638 712 790 37.346 045 150 0.865 039 246 will not be dramatically deteriorated for the ground state: it can provide not less than 5 s.d for the energy in the whole domain studied. Furthermore, the Approximant results in a locally accurate approximation of the exact ground state wave function. The relative deviation does not exceed ∼ 0.02. This is surprising, taking into account that the Approximant will be one parametric. Interestingly, the dependence of the single parameter as function of D and g is very weak, see Fig. 7.1. Results of similar accuracy occur for excited states. In our present calculations10, the ratio |y1/y0| is bounded and small. Thus, y1 is a small function in comparison with y0. In Fig. 7.2, it is presented y0 and y1 vs r for g = 1 for dimension D = 1. Similar plots appear for D = 2, 3, 6. Making an analysis of these plots, one can see that in the domain 0 ≤ r . 1 , which gives the dominant contribution to the integrals 10In which we have only imposed (7.3.27) and (7.2.23) as constraints. 7.2. Cubic Anharmonic Oscillator 51 0 5 10 15 20 0 0.2 0.4 0.6 0.8 1 g a 3 Figure 7.1: Variational parameter ã3 for the one-parametric ground state (0, 0) Approximant as function of g for D = 1, 2, 3, 6. From g ≈ 2.3 the single parameter ã3 becomes approximately D-independent. 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0 1 2 3 4 5 6 0.0 0.5 1.0 1.5 2.0 2.5 3.0 0.000 0.002 0.004 0.006 0.008 0.010 0.012 Figure 7.2: Function y0 = ∂rΦt (on left) and its first order correction y1 (on right) at g = 1 as a function of r at D = 1. which define the energy corrections, |y1| is extremely small compared to |y0|, being close to zero. It explains why the energy correction E (t) 2 is small being of order ∼ 10−7, or ∼ 10−8. In a similar way, one can show numerically that higher corrections y2, y3, . . . drop down to zero in the domain 0 ≤ r . 1 even faster. This is a clear indication of the convergence of the expansion (1.18) as n→ ∞. At D > 1, we checked that the Approximant Ψ (t) 1,0 also provides an accurate estimate of the position of the radial node of the exact wave function, see (7.2.26). The orthogonality condition described in Section 6.2 provides a simple analytic expression for the zeroth order approximation r (0) 0 to r0. Making a comparison of r (0) 0 with numerical estimates which come from the Lagrange-Mesh method with ∼ 50 mesh points, we can see the coincidence of r0 and r (0) 0 for at least 5 d.d. at integer dimension D at any coupling constant g ≥ 0. The results are presented in Table 7.4. 52 Chapter 7. Results for Two-Term Potentials 7.2.5 First Terms in the Strong Coupling Expansion We present the estimates of the first two terms in the strong coupling expansion (5.1.7) for the ground state energy of the cubic oscillator. Assuming for simplicity 2M = ~ = 1, this expansion has the form E = g2/5 ( ε̃0 + ε̃1g −4/5 + ε̃2g −8/5 + . . . ) . (7.2.30) It is worth mentioning that similar expansion holds for any excited state. This expansion corresponds to PT in powers of λ̂ for the potential V (w) = w3 + λ̂ w2 , λ̂ = g−4/5 , (7.2.31) defined in w ∈ [0,∞). Evidently, it has a finite radius of convergence. As the first step, we focus on the calculation of ε̃0 in (7.2.30), which is the ground state energy in the pure cubic potential V = w3 (the ultra-strong coupling regime). To do it, we use the simplest physically relevant trial function (5.1.12), Ψ (t) 0,0 = e− 2 5 w5 . (7.2.32) Taking Ψ (t) 0,0 as zero approximation, we calculate variational energy analytically and then develop the PT procedure described in Section 5.1. In this manner we obtain different PT corrections11 ε̃0,k , k = 0, 1, 2, 3, . . . to the exact value of ε̃0 and then form the partial sums (5.1.15). The first non-trivial partial sum (ε̃0,k + ε̃0,1) is equal to variational energy. Explicit results, including partial sums up to sixth order, are presented in Table 7.5 for the one-dimensional case. We have to note a slow convergence of the partial sums to the exact result with respect to the increase of the number of perturbative terms in the expansion (5.1.15). For example, the partial sum ∑6 k=0 ε̃0,k with seven terms included allows us to reproduce 7 s.d. of the exact result12 It is natural to accelerate convergence taking more advanced trial functions as entry. Table 7.5: Different partial sums (5.1.15) for the leading term ε̃0 of the strong coupling ex- pansion (7.2.30) for the one-dimensional cubic anharmonic oscillator in PT with trial function (7.2.32) as zero approximation. ε̃0,0 ∑1 k=0 ε̃0,k ∑2 k=0 ε̃0,k ∑3 k=0 ε̃0,k ∑4 k=0 ε̃0,k ∑5 k=0 ε̃0,k ∑6 k=0 ε̃0,k Exact ε̃0 0 1.053 006 976 1.021 174 929 1.022 989 568 1.022 956 899 1.022 946 414 1.022 947 763 1.022 947 875 As an alternative to the trial function (7.2.32), let us use the Approximant13 (7.2.21) in order to calculate the first two terms in the strong coupling expansion (7.2.30), see also (5.1.7). In Table 7.6, we present for different D the leading coefficient of the strong coupling expansion 11The calculation of corrections is a numerical process. 12By saying exact, we mean that the (truncated) displayed d.d. in ε̃0 are exact digits. 13We fix g = 1 as normalization of the free parameters. 7.2. Cubic Anharmonic Oscillator 53 ε̃0 found in VM, denoted by ε̃ (1) 0 , and the second PT correction to it ε̂2 calculated via the Non-Linearization Procedure. We introduce the partial sum ε̃ (2) 0 = ε̃ (1) 0 + ε̂2. The final results are verified using the Lagrange-Mesh method with 200 mesh points. One can see that system- atically, the correction ε̂2 is of order ∼ 10−7. Hence, the first six d. d. in variational energy are correct. It defines the accuracy of variational calculations of ε̃0 with the Approximant (7.2.21) as the trial function. One can estimate the order of the third PT correction: for all D studied ε̂3 ∼ 10−2 ε̂2. It indicates an extremely high rate of convergence ∼ 10−2 in PT. In Table 7.7 we present the first two approximations of the coefficient ε̃1 - the expectation value of w2 - , see (7.2.30). Table 7.6: Ground state energy ε̃0, see (7.2.30), for the potential W (w) = w3 for D = 1, 2, 3, 6 found in PT with the Approximant Ψ (t) 0,0 as the entry: ε̃ (1) 0 is the variational energy, ε̂2 is the second PT correction and ε̃ (2) 0 = ε̃ (1) 0 + ε̂2 is corrected variational energy. Eight d.d. in ε̃ (2) 0 confirmed in Lagrange mesh calculation. D = 1 D = 2 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 1.022948250 3.75× 10−7 1.022 947 875 2.187 461 809 3.09× 10−7 2.187 461 499 D = 3 D = 6 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 3.450 562 918 2.29× 10−7 3.450 562 689 7.647 118 254 1.01× 10−7 7.647 118 153 Table 7.7: Subdominant (next-to-leading) term ε̃1 in the strong coupling expansion (7.2.30) of the ground state energy for the cubic anharmonic radial potential for D = 1, 2, 3, 6. The first correction to it ε̃1,1 as well as the corrected value ε̃ (2) 1 = ε̃ (1) 1 + ε̃1,1 displayed. D = 1 D = 2 ε̃ (1) 1 ε̃1,1 ε̃ (2) 1 ε̃ (1) 1 ε̃1,1 ε̃ (2) 1 0.410 598 524 6.78× 10−7 0.410 599 202 0.766 573 847 5.24× 10−7 0.766 574 371 D = 3 D = 6 ε̃ (1) 1 ε̃1,1 ε̃ (2) 1 ε̃ (1) 1 ε̃1,1 ε̃ (2) 1 1.092 125 224 2.67× 10−7 1.092 125 491 1.967 599 668 1.42× 10−7 1.967 599 810 54 Chapter 7. Results for Two-Term Potentials 7.3 Quartic Anharmonic Oscillator The simplest and most popular formally even radial anharmonic oscillator potential is charac- terized by a quartic anharmonicity V (r) = r2 + g2 r4 , (7.3.1) cf. (7.0.1) at m = 4. Many properties that the quartic anharmonic radial oscillator exhibits are typical for any even anharmonic potential, V (r) = V (−r). In particular, the polynomial nature of corrections εn and Yn(v), see (4.1.11) and (4.1.9) is one of such common features. 7.3.1 Perturbation Theory in the Weak Coupling Regime For the quartic anharmonic oscillator, the perturbative expansions of ε and Y(v) may be derived from the RB equation (3.3.9), ∂vY(v) − Y(v) ( Y(v)− D − 1 v ) = ε (λ) − v2 − λ2 v4 , ∂v ≡ d dv , (7.3.2) where v and λ are defined in (3.3.7) and (3.3.10), respectively. These expansions are of the form ε = ε0 + ε2 λ 2 + ε4 λ 4 + . . . , ε0 = D , (7.3.3) and Y(v) = Y0(v) + Y2(v)λ 2 + Y4(v)λ 4 + . . . , Y0(v) = v , (7.3.4) respectively. All odd terms in λ vanish in both expansions, see (4.1.5) and (4.1.6). In general, a large number of corrections can be calculated by linear algebra means. In particular, the first non-vanishing corrections are ε2 = 1 4 D (D + 2) , Y2(v) = 1 2 v3 + 1 4 (D + 2) v . (7.3.5) The next two corrections ε4,6 and Y4,6(v) are presented in Appendix B. The algebraic procedure of finding PT corrections holds for all even anharmonic potential, V (r) = V (−r). However, it is enough to have one single odd monomial term in the potential to break this property down. In this situation, the calculation of the correction εn becomes a numerical procedure, just like it happened for the cubic case. In general, all corrections Y2n(v) are odd-degree polynomials in v of the form, Y2n(v) = v n∑ k=0 c (2n) 2k v2(n−k) , (7.3.6) where any coefficient c (2n) 2k is a polynomial in D of degree k, c (2n) 2k = P (2n) k (D) , c (2n) 2n = ε2n D . (7.3.7) 7.3. Quartic Anharmonic Oscillator 55 The energy corrections ε2n are of the form [60] ε2n(D) = D (D + 2)Rn−1(D) , (7.3.8) where Rn−1(D) is a polynomial of degree (n− 1) in D. In particular, R0(D) = 1 4 . From (7.3.8), one can see that any energy correction ε2n vanishes when D = 0. In this situation, the formal sum of corrections results in ε = 0, and ultimately in E = 0. Thus, the radial Schrödinger equation is reduced to − ~2 2M ( d2Ψ(r) dr2 − 1 r dΨ(r) dr ) + (r2 + g2 r4)Ψ(r) = 0 (7.3.9) at D = 0. This equation defines the zero mode of the radial Schrödinger operator at D = 0. It can be solved exactly in terms of Airy functions [39], Ψ = C1 Ai ( 1 + (λv)2 λ4/3 ) + C2 Bi ( 1 + (λv)2 λ4/3 ) , (7.3.10) see (3.3.7) and (3.3.10) for the definition of λ and v, respectively. However, this linear com- bination can not be made normalizable14 by any choice of constants C1 and C2. Therefore, it is an indication of the absence of a normalizable zero mode at D = 0. Interestingly, at the non-physical dimension D = −2, ε2n with n > 1 also vanish. Thus, the formal sum of the corrections results is ε = −2, see (7.3.8). In this case, no exact solution for the corresponding radial Schrödinger equation was found. 7.3.2 Generating Functions For the quartic anharmonic oscillator, the function Z(u) satisfies the GB equation λ2 ∂uZ(u) − Z(u) ( Z(u)− λ2(D − 1) u ) = λ2 ε(λ) − u2 − u4 , ∂u ≡ d du , (7.3.11) cf. (3.3.20). The solution of this equation can be written as an expansion in terms of generating functions, namely, Z(u) = Z0(u) + Z2(u)λ 2 + Z4(u)λ 4 + . . . , (7.3.12) where each coefficient is of the form Z2k(u) = u ∞∑ n=k c (2n) 2k u2(n−k) , k = 0, 1, . . . . (7.3.13) Here, the expansion of ε in powers of λ is given by (7.3.3). Note that all generating functions Z2k+1(u), k = 1, 2, ..., of odd order λ2k+1 are absent in expansion (4.2.1). Interestingly, from the 14At D = 0, the normalizability of the wave function is guaranteed if ∫ ∞ 0 ψ2 r−1dr <∞. 56 Chapter 7. Results for Two-Term Potentials polynomial form of the coefficient c (2n) 2k in D, see (7.3.8), one can find the structure of generating function in the variable D, namely Z2k(u) = u k∑ n=0 f (k) n (u2)Dn , (7.3.14) where f (k) n (u2), n = 0, 1, ..., k, are real functions. Hence, Z2k(u) is a polynomial in D of degree k. For instance, this feature can be explicitly noted in the first two terms of the expansion (7.3.12), Z0(u) = u √ 1 + u2 , (7.3.15) Z2(u) = u2 +D ( 1 + u2 − √ 1 + u2 ) 2u(1 + u2) . (7.3.16) Thus, f (0) 0 (u2) = √ 1 + u2 , f (2) 0 (u2) = 1 2(1 + u2) , f (2) 1 (u2) = 1 + u2 − √ 1 + u2 2u2(1 + u2) . (7.3.17) The asymptotic behavior of the generating functions Z2k(u), k = 0, 1, 2, . . ., in the expansion (7.3.12) at large u is related to the asymptotic behavior of the function y(r) at large r in a quite interesting manner. It can be easily found that for fixed effective coupling constant λ, the asymptotic expansion of y at large r, written in variable v, see (3.3.7), has the form y = (2M~ 2) 1 4 ( λv2 + 1 2λ + D + 1 2 v−1 − 4λ2ε+ 1 8λ3 v−2 + . . . ) , v → ∞ . (7.3.18) Note that the first three terms of the expansion are ε-independent, but only the first two are D-independent. On the other hand, the first three terms in the expansion of lowest generating function (2M g2 )1/2Z0(u) at large u are ( 2M g2 )1/2 Z0(u) = ( 2M g2 )1/2( u2 + 1 2 − 1 8 u−2 + . . . ) , u→ ∞ , (7.3.19) see (7.3.15). To compare the expansions (7.3.18) and (7.3.19), let us replace the classical coordinate u by the quantum one v via u = λ v, see (3.3.24). Evidently, large v implies large u and vice versa (as long as λ is fixed). Then the expansion (7.3.19) becomes ( 2M g2 )1/2 Z0(λv) = (2M~ 2) 1 4 ( λv2 + 1 2λ − 1 8λ3 v−2 + . . . ) , v → ∞ . (7.3.20) It reproduces exactly the first two terms in (7.3.18) but fails to reproduce the term of order O(v−1), which is absent. However, the next generating function (2M g2 )1/2λ2Z2(λv) at large v- expansion reproduces the term O(v−1) exactly in the original expansion (7.3.18), ( 2M g2 )1/2 λ2Z2(λv) = (2M~ 2) 1 4 ( D + 1 2 v−1 − D 2λ v−2 + . . . ) , v → ∞ . (7.3.21) 7.3. Quartic Anharmonic Oscillator 57 In turn, it fails to reproduce the term O(v−2) correctly. Thus, the expansion of the sum (2M g2 )1/2(Z0(λv) + λ2Z2(λv)) at large v reproduces exactly the first three ε-independent terms in the expansion (7.3.18). These three terms are responsible of the normalizability of the wave function at large v. All higher generating functions Z4(λv), Z6(λv) . . . contribute at large v to the same term O(v−2) as follows ( 2M g2 )1/2 λ2nZ2n(λv) = (2M~ 2) 1 4 ( −ε2n−2 λ 2n−3 2 v−2 + . . . ) , v → ∞ , n > 2 , (7.3.22) where ε2n−2 is the energy PT correction of the order (2n − 2). As a consequence, no matter how many generating functions we consider in the expansion (2M g2 )1/2(Z0(λv) + λ2Z2(λv) + ...), the term of order O(v−2) of (7.3.18) can not be reproduced exactly. 7.3.3 The Approximant The first two15 generating functions of the phase in the expansion (4.4.1), G0(r) and G2(r), are given by G0(r) = (2M)1/2 g2 ( (1 + g2r2) 3/2 3 ) , (7.3.23) G2(r) = (2M)1/2 g2 ( 1 4 log[1 + g2r2] + D 2 log [ 1 + √ 1 + g2r2 ]) . (7.3.24) These two functions serve as building blocks for the construction of the Approximant. Following the general prescription (6.1.2), we write the Approximant of the ground state in the exponential representation Ψ (t) 0,0 = e− 1 ~ Φt where 1 ~ Φt(r) = ã0 + ã2 r 2 + ã4 g 2 r4 √ 1 + b̃4 g2 r2 + 1 4 log [ 1 + b̃4 g 2 r2 ] + D 2 log [ 1 + √ 1 + b̃4 g2 r2 ] , (7.3.25) here ã0,2,4, and b̃4 are free parameters. The logarithmic terms, added in (7.3.25), generate a prefactor to the exponential function. These terms are just a certain minimal modification of those which occur in the second generating function (7.3.24). As a result, the Approximant of the ground state for arbitrary D = 1, 2, 3, . . . is given by 15The generating functions G4(r) and G6(r) are presented explicitly in Appendix B. 58 Chapter 7. Results for Two-Term Potentials Ψ (t) 0,0(r) = 1 ( 1 + b̃4 g2 r2 )1/4 ( 1 + √ 1 + b̃4 g2 r2 )D/2 exp  − ã0 + ã2 r 2 + ã4 g 2 r4 √ 1 + b̃4 g2 r2   . (7.3.26) This is the central formula of this Section devoted to the quartic anharmonicity. We will show that it provides a highly accurate uniform approximation of the exact ground state wave function. We impose the constraint b̃4 = 9 ã24 , (7.3.27) which allows us to reproduce the dominant term in the expansion (7.3.18) exactly16. Hence, it reproduces the exact asymptotic behavior of the phase at large distances. Therefore, the final form of the Approximant contains only 3 free parameters, say {ã0, ã2, ã4}. Note that by choosing ã0 = (2M)1/2 3g2 , ã2 = (8M)1/2 3 , ã4 = (2M)1/2 3 , (7.3.28) the phase Φt reproduces exactly the first two terms in the expansion in generating functions (4.4.1). However, parameters ã0, ã2, ã4 in (7.3.28) are not optimal from the point of view of the VM. To construct 3-parametric Approximants for excited states, we follow the prescription presented in the previous Section 6.2. Hence, at D = 1 we have Ψ(t) nr,p(x) = xpP (p) nx (x 2) ( 1 + b̃4 g2 x2 )1/4 ( 1 + √ 1 + b̃4 g2 x2 )D/2 exp  − ã0 + ã2 x 2 + ã4 g 2 x4 √ 1 + b̃4 g2 x2   , (7.3.29) while at D > 1 Ψ (t) nr,ℓ (r) = rℓ P (ℓ) nr (r 2) ( 1 + b̃4 g2 r2 )1/4 ( 1 + √ 1 + b̃4 g2 r2 )D/2 exp  − ã0 + ã2 r 2 + ã4 g 2 r4 √ 1 + b̃4 g2 r2   . (7.3.30) Definitions of P (p) nr (x 2) and P (ℓ) nr (r 2) are presented in (6.2.1) and (6.2.4), explicitly. 16In numerical computations we set ~ = 1 and M = 1/2. 7.3. Quartic Anharmonic Oscillator 59 7.3.4 Numerical Results: Variational Calculations The calculations of the variational energy for the first low-lying states with quantum numbers (0,0), (0,1), (0,2), (1,0) are presented in Tables 7.8 - 7.11 for different values of D and g2. For some of those states, the variational energy Evar = E (t) 0 +E (t) 1 , the first correction E (t) 2 to it, as well as its corrected value Evar+E (t) 2 are shown. All states are studied in dimensionD = 1, 2, 3, 6 and coupling constant g2 = 0.1, 1, 10. The variational energy Evar is obtained with absolute deviation (from the exact result) in the order 10−8−10−14 (8−14 s.d.). This deviation is found by calculating the second correction E (t) 2 (the first correction to the variational energy). To check the energies obtained in Non-Linearization Procedure, we calculated them numeri- cally in the Lagrange-Mesh method. These numerical calculations show that all digits presented for Evar + E (t) 2 are exact: we obtain not less than 12 exact d.d. in the Non-Linearization Pro- cedure. This accuracy is confirmed independently by calculating the second correction E (t) 3 to variational energy: this correction is . 10−13 for any D and g2 that we have studied. It indi- cates a very fast rate of convergence in Non-Linearization Procedure with trial function either (7.3.29) or (7.3.30) taken as zero approximation. In general, the energies grow with an increase of D and/or g2. At D > 1 we can classify the states by radial quantum number nr and angular momentum ℓ as (nr, ℓ), we mention the hierarchy of eigenstates which holds for any fixed integer D > 1 and g2: (0,0), (0,1), (0,2), (1,0). There is a considerable number of calculations devoted to estimating the energy of the first low-lying states. Our results, see Tables 7.8 - 7.11, are in complete agreement with [23] and [24] for D = 1, and superior considerably of those obtained for different D > 1 and g2, see e.g. [25], [26] and [32]. At D = 6 the calculations are carried out for the first time. The Deviation of the Approximant from the exact (unknown) eigenfunction can be estimated via the Non-Linearization Procedure. For the ground state function, this deviation is extremely small and bounded, ∣ ∣ ∣ ∣ ∣ Ψ0,0(r)−Ψ (t) 0,0(r) Ψ (t) 0,0(r) ∣ ∣ ∣ ∣ ∣ . 10−6 (7.3.31) in the whole range of r ∈ [0,∞) at any D and g2. Therefore, the Approximant of the ground state is a locally accurate approximation of the exact wave function Ψ0,0(r) once the optimal pa- rameters are chosen. In general, the optimal variational parameters are smooth, slow changing as functions of g2 without a strong D dependence. For the ground state (0,0), the plots of the parameters ã0,2,3 vs g for D = 1, 2, 3, 6 are shown in Appendix C, Fig. C.2. In turn, Fig. C.3 shows the plots of the parameters as functions of D for fixed g2. Similar plots appear for the excited states for different D and g2. Making an analysis of the parameters ã0,2,4 vs g2 for different D, one can see the (accidental) appearance of another D-independent constraint on the parameters, namely ã2 = 1 + 27 ã24 18ã4 . (7.3.32) It corresponds to the fact that the coefficient in front of r - another growing term at r → ∞ in 60 Chapter 7. Results for Two-Term Potentials Table 7.8: Ground state energy for the quartic potential V (r) = r2 + g2 r4 for D = 1, 2, 3, 6 and g2 = 0.1, 1, 10, labeled by quantum numbers (0,0) at any D. Variational energy Evar = Evar, the first correction E (t) 2 (rounded to 3 s.d.) found with use of Ψ (t) 0,0, see text, and the corrected energy Evar +E (t) 2 = Evar +E (t) 2 shown. Evar +E (t) 2 coincides with Lagrange Mesh results (see text) in 12 displayed d. d., hence all printed digits are exact. g2 D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 1.065 285 509 544 3.00× 10−14 1.065 285 509 544 2.168 597 211 269 5.28× 10−14 2.168 597 211 269 1.0 1.392 351 641 563 3.37× 10−11 1.392 351 641 530 2.952 050 091 995 3.17× 10−11 2.952 050 091 962 10.0 2.449 174 072 588 4.69× 10−10 2.449 174 072 118 5.349 352 819 751 3.44× 10−10 5.349 352 819 462 g2 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.306 872 013 152 2.20× 10−13 3.306 872 013 152 6.908 332 111 232 9.80× 10−14 6.908 332 111 232 1.0 4.648 812 704 237 2.69× 10−11 4.648 812 704 210 10.390 627 295 514 9.68× 10−12 10.390 627 295 504 10.0 8.599 003 455 030 2.22× 10−10 8.599 003 454 807 19.936 900 374 076 6.48× 10−11 19.936 900 374 011 Table 7.9: The first excited state energy for the quartic potential V (r) = r2 + g2 r4 for different D and g2. Variational energy Evar, the first correction E (t) 2 found with use of Ψ (t) 0,1 for D = 1 and Ψ (t) 0,1 for D > 1, see (7.3.29) and (7.3.30). The corrected energy Evar + E (t) 2 = Evar + E (t) 2 shown, the correction E (t) 2 rounded to 3 s.d. All printed digits for Evar + E (t) 2 are exact. g2 D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.306 872 013 236 8.33× 10−11 3.306 872 013 153 4.477 600 360 878 1.10× 10−10 4.477 600 360 768 1.0 4.648 812 707 206 2.99× 10−9 4.648 812 704 212 6.462 906 003 251 3.39× 10−9 6.462 905 999 864 10.0 8.599 003 467 556 1.27× 10−8 8.599 003 454 810 12.138 224 752 729 1.38× 10−8 12.138 224 738 901 g2 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 5.678 682 663 377 1.33× 10−10 5.678 682 663 243 9.447 358 518 278 1.80× 10−10 9.447 358 518 099 1.0 8.380 342 533 658 3.56× 10−9 8.380 342 530 101 14.658 513 816 952 3.39× 10−9 14.658 513 813 563 10.0 15.927 096 988 667 1.40× 10−8 15.927 096 974 709 28.536 810 849 436 1.21× 10−8 28.536 810 837 360 7.3. Quartic Anharmonic Oscillator 61 Table 7.10: The second excited state energy for the quartic potential V (r) = r2 + g2 r4 for different D and g2. Variational energy Evar found with use of Ψ (t) 1,0 for D = 1 and Ψ (t) 0,2 for D > 1, see (7.3.29) and (7.3.30). The first correction E (t) 2 and the corrected energy Evar+E (t) 2 = Evar + E (t) 2 shown. Correction E (t) 2 rounded to 3 s.d. All 12 printed d. d. in Evar + E (t) 2 are exact. g2 D = 1 D = 2 Evar Evar −E(t) 2 Evar + E (t) 2 0.1 5.747 959 269 942 6.908 332 112 167 9.35× 10−10 6.908 332 111 232 1.0 8.655 049 995 062 10.390 627 321 799 2.63× 10−8 10.390 627 295 506 10 16.635 921 650 401 19.936 900 479 247 1.05× 10−7 19.936 900 374 040 g2 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 8.165 006 438 494 1.00× 10−9 8.165 006 437 493 12.084 471 853 886 1.11× 10−9 12.084 471 852 776 1.0 12.485 556 075 670 2.47× 10−8 12.485 556 051 000 19.217 523 515 555 1.97× 10−8 19.217 523 495 879 10.0 24.145 857 689 623 9.48× 10−8 24.145 857 594 824 37.811 402 320 699 6.90× 10−8 37.811 402 251 702 Table 7.11: The third excited state energy for the quartic potential V = r2+g2 r4 for different D and g2. The radial node r0 calculated in Lagrange-Mesh method with 9 correct d.d. Variational energy Evar and its radial node r (0) 0 given by underlined digits, both found with Ψ (t) 1,0, see (7.3.30). g2 D = 2 D = 3 D = 6 E (1) 0 r (0) 0 E (1) 0 r (0) 0 E (1) 0 r (0) 0 0.1 7.039 707 584 0.918 783 458 8.352 677 825 1.111 521 078 12.415 256 177 1.522 966 591 1.0 10.882 435 576 0.733 724 778 13.156 803 922 0.875 567 486 20.293 829 707 1.166 753 149 10.0 21.175 135 370 0.524 083 057 25.806 276 215 0.621 795 290 40.388 142 970 0.820 068 428 the trial phase (7.3.25) - is reproduced almost exactly in accordance to (7.3.18). Not surprisingly, if we demand to fulfill exactly constraint (7.3.32), no deterioration in the accuracy of the energy and wave function occurs. Thus, it can be concluded that the approximate phase (7.3.25), at large r, reproduces (almost) exactly all three growing with r terms: r3, r, and log r. Eventually, if we require to reproduce all those terms exactly by the Approximant in its final form, it will contain two free parameters {ã0, ã4} only. In this case, the parameters {ã2, b̃4} obey constraints (7.3.32), (7.3.27), respectively. The first order correction y1 to the logarithmic derivative of the ground state y0 = ∂rΦt is a bounded function at different D and g2. For example, for g2 = 1 the first correction y1 has the bound |y1|max ∼    0.0101 , D = 1 0.0087 , D = 2 0.0077 , D = 3 0.0060 , D = 6 (7.3.33) It is the consequence of the fact that the derivative of Φt reproduces exactly the growing terms at large r in the expansion (7.3.18). “Boundness” of y1, together with the small value of the 62 Chapter 7. Results for Two-Term Potentials maximum, implies that we deal with smartly designed zeroth order-approximation Ψ (t) 0,0; it leads, in the framework of the Non-Linearization Procedure, to a fast convergent series of the energy and wave function. In Fig. 7.3 y0 and y1 vs r are presented for g2 = 1 in D = 1. It is not a surprise that a similar plot should appear for D = 2, 3, 6 (not shown). An analysis of these plots indicates that |y1| is an extremely small function in comparison with |y0| in the domain 0 ≤ r . 1 where the dominant contribution to variational integrals occurs, see (1.1). It explains why the energy correction E (t) 2 is small being the order of ∼ 10−8, or ∼ 10−11. In a similar way one can show numerically that the higher corrections y2, y3, . . ., drop down to zero in the domain 0 ≤ r . 1 even faster, indicating the convergence of the expansion (1.18) as n→ ∞. 0.5 1 1.5 2 2.5 3 0 2 4 6 8 10 5 10 15 20 25 0 0.002 0.004 0.006 0.008 0.01 0.012 Figure 7.3: Quartic oscillator at D = 1: function y0 = (Φt) ′ (on left) and its first correction y1 (on right) vs r for g2 = 1. At D > 1 we checked that the Approximant Ψ (t) 1,0 also provides an accurate estimate of the position of the radial node of the exact wave function, see (7.2.26). The orthogonality condition described in Section 6.2 provides a simple analytic expression for the zeroth order approximation r (0) 0 to r0. Making a comparison of r (0) 0 with numerical estimates which come from the Lagrange-Mesh method with 200 mesh points, we can see the coincidence of r0 and r (0) 0 for at least 6 d.d. at integer dimension D at any coupling constant g2 ≥ 0. The results are presented in Table 7.11. 7.3.5 First Terms in the Strong Coupling Expansion In this Section, we will explicitly find the first two terms of the strong coupling expansion (5.1.7) of the ground state energy for the quartic anharmonic oscillator (7.3.1), E = g2/3(ε̃0 + ε̃2g −4/3 + ε̃4g −8/3 + . . .) , (7.3.34) assuming 2M = ~ = 1. In contrast with the weak coupling expansion, the expansion (7.3.34) has a finite radius of convergence. This expansion corresponds to PT in powers of λ̂ for the potential V (w) = w4 + λ̂ w2 , λ̂ = g−2/3 , (7.3.35) 7.4. Sextic Anharmonic Oscillator 63 in the radial Schrödinger equation defined in w ∈ [0,∞). In order to calculate the first two terms ε̃0 and ε̃2 of the strong coupling expansion (7.3.34), we use the Approximant (7.3.26). In Table 7.12, the leading coefficient ε̃0 and the second perturbative correction to it ε̂2, as well as the corrected value ε̃ (2) 0 = ε̃ (1) 0 + ε̂2, calculated via the Non-Linearization Procedure, are presented for different D. Numerical results for ε̃0, based on the Lagrange-Mesh method and obtained with 12 d.d., indicate that ε̃ (2) 0 = ε̃ (1) 0 + ε̂2 reproduce not less than 10 d.d. This accuracy is verified independently by calculating the next correction ε̂3, which results in the order of ε̂3 ∼ 10−2ε̂2. In turn, Table 7.13 contains the results of the first two approximations for the coefficient ε̃2 in (7.3.34). It should be mentioned that our final results for the coefficient ε̃0 reproduce and sometimes exceed the best results available in the literature, see [62], [63], [64]. Hence, we have established some benchmarks. At D = 6 the calculations are carried out for the first time. Table 7.12: Ground state energy ε̃0 for the potential W = w4 (see (7.3.35)) for D = 1, 2, 3, 6 found in PT based on the Approximant Ψ (t) 0,0: ε̃ (1) 0 corresponds to the variational energy, ε̂2 is the second PT correction, ε̃ (2) 0 = ε̃ (1) 0 + ε̂2 is the corrected variational energy. 10 d.d. in ε̃ (2) 0 confirmed independently in Lagrange-Mesh method. D = 1 D = 2 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 1.060 362 090 491 7.02× 10−12 1.060 362 090 484 2.344 829 072 753 9.27× 10−12 2.344 829 072 744 D = 3 D = 6 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 3.799 673 029 810 9.27× 10−12 3.799 673 029 801 8.928 082 199 890 4.07× 10−11 8.928 082 199 850 Table 7.13: Subdominant coefficient ε̃2 in the strong coupling expansion (7.3.34) for the ground state energy for the quartic radial anharmonic potential (7.3.35) for different D = 1, 2, 3, 6. The first correction to it ε̃2,1 as well as the corrected value ε̃ (2) 2 = ε̃ (1) 2 + ε̃2,1 displayed D = 1 D = 2 ε̃ (1) 2 ε̃2,1 ε̃ (2) 2 ε̃ (1) 2 ε̃2,1 ε̃ (2) 2 0.362 022 648 388 3.96× 10−10 0.362 022 648 784 0.651 477 773 845 4.38× 10−10 0.651 477 774 283 D = 3 D = 6 ε̃ (1) 2 ε̃2,1 ε̃ (2) 2 ε̃ (1) 2 ε̃2,1 ε̃ (2) 2 0.901 605 894 682 2.03× 10−9 0.901 605 896 709 1.526 804 282 772 −3.06× 10−8 1.526 804 252 175 7.4 Sextic Anharmonic Oscillator In this Section, it will be considered the two-term sextic anharmonic radial oscillator potential V (r) = r2 + g4 r6 , (7.4.1) see (7.0.1) at m = 6. 64 Chapter 7. Results for Two-Term Potentials 7.4.1 Perturbation Theory in the Weak Coupling Regime In the Weak Coupling Regime, the perturbative expansion for ε and Y(v) can be developed from the RB equation (3.3.9), ∂vY(v) − Y(v) ( Y(v)− D − 1 v ) = ε (λ) − v2 − λ4 v6 , ∂v ≡ d dv . (7.4.2) The definition of v and λ is shown in (3.3.7) and (3.3.10), respectively. The perturbation series for ε and Y(v) are of the form ε = ε0 + ε4 λ 4 + ε8 λ 8 + . . . , ε0 = D , (7.4.3) and Y(v) = Y0(v) + Y4(v)λ 4 + Y8(v)λ 8 + . . . , Y0(v) = v , (7.4.4) respectively. The terms of order λ4n+1 and λ4n+2 are absent in both expansions. Since the potential (7.4.1) is even, PT can be constructed by algebraic means due to the polynomial nature of correction Yn(v). The first non-vanishing correction is ε4 = 1 8 D (D+2) (D+4) , Y4(v) = 1 2 v5 + 1 4 (D+4) v3 + 1 8 (D+2) (D+4) v . (7.4.5) The next two corrections, ε8,12 and Y8,12(v), are presented in Appendix B. It can be shown that the correction Y4n(v) has the form of an odd polynomial in v, namely Y4n(v) = v 2n∑ k=0 c (4n) 2k v2(2n−k) , (7.4.6) with coefficients c (4n) 2k being polynomials in D of degree k, c (4n) 2k = P (4n) k (D) , c (4n) 4n = ε4n D , (7.4.7) cf. (7.3.6), (7.3.7). The correction ε4n has the factorization property ε4n(D) = D (D + 2) (D + 4)R2n−2(D) , (7.4.8) where R2n−2(D) is a polynomial in D of degree (2n − 2), cf. (7.3.8). In particular, R0 = 1 8 . From (7.4.7), one can see that all corrections ε4n vanish at D = 0,−2,−4. Hence, their formal sum results in ε = 0,−2,−4, respectively. In the case D = 0, the radial Schrödinger equation takes the form − ~2 2M ( d2Ψ(r) dr2 − 1 r dΨ(r) dr ) + (r2 + g4 r6)Ψ(r) = 0 , (7.4.9) cf.(7.3.9). Its formal solution is given in terms of the parabolic cylinder functions [65] (also known as Weber functions), Ψ = C1Dν−(λv 2) + C2Dν− ( i λv2 ) , ν± = −1 2 ± 1 4λ2 , (7.4.10) 7.4. Sextic Anharmonic Oscillator 65 where v and λ are defined in (3.3.7) and (3.3.10), respectively. This formal solution has the meaning of the zero mode of the Schrödinger operator at D = 0. The function shown in (7.4.10) cannot be made normalizable by any choice of constants C1 and C2. Hence, the Schrödinger operator at D = 0 for the sextic potential (7.4.1) has no zero mode in the Hilbert space. It complements a similar statement made for quartic potential (7.3.1). One can guess that the zero mode in the Hilbert space at D = 0 is absent for the Schrödinger operator with anharmonicity r2m. At D = −2,−4, all corrections ε4n vanish. As a result, we formally have ε = −2 and ε = −4, respectively. For both cases, no exact solution has been found for their corresponding radial Schrödinger equations. 7.4.2 Generating Functions For the sextic anharmonic oscillator, the GB equation is given by λ2 ∂uZ − Z ( Z − λ2(D − 1) u ) = λ2 ε(λ) − u2 − u6 , ∂u ≡ d du , (7.4.11) see (3.3.20). The expansion of Z(u) in terms of generating functions has the form Z(u) = Z0(u) + Z2(u)λ 2 + Z4(u)λ 4 + . . . . (7.4.12) In turn, the expansion of the energy reads ε = ε0 + ε4 λ 4 + ε8 λ 8 + . . . , ε0 = D , (7.4.13) see (7.4.3). Interestingly, (7.4.12) has the same structure as the expansion for the quartic anharmonic case: all generating functions Z2k+1(u), k = 1, 2, ... of odd orders λ2k+1 are absent in expansion, cf. (7.3.12). It contrasts with the expansion of Y(v) and ε(λ) in which the powers λ4n are present only. In fact, for any even radial anharmonic potential, V (r) = V (−r), the function Z(u) is written in terms of generating functions as an expansion in powers of λ2. Further properties of the generating function for the sextic anharmonic oscillator can be established. For example, there are two different families of generating functions, Z4k(u) = u ∞∑ n=k c (4n) 4k u4(n−k) , (7.4.14) and Z4k+2(u) = u ∞∑ n=k+1 c (4n) 4k+2u 4(n−k)−2 . (7.4.15) Those families occur in correspondence to ε4k+2 = 0 and ε4k 6= 0, respectively. Following (7.4.6) and (7.4.7), it is easy to see that both families of generating functions are of the form Z2k(u), Z2k(u) = u k∑ n=0 f (k) n (u2)Dn , (7.4.16) 66 Chapter 7. Results for Two-Term Potentials where f (p) n (u2) are some real functions. It implies that Z2k(u) is a polynomial in D of degree k. For example, the first two terms in the expansion (7.4.12) are Z0(u) = u √ 1 + u4 , (7.4.17) Z2(u) = 2u4 +D ( 1 + u4 − √ 1 + u4 ) 2u (1 + u4) . (7.4.18) From these expressions, we can conclude immediately that f (0) 0 (u2) = √ 1 + u4 , f (2) 0 (u2) = u2 (1 + u4) , f (2) 1 (u2) = 1 + u4 − √ 1 + u4 2u2(1 + u4) . (7.4.19) The asymptotic behavior at large u of Z2p(u) is related to the expansion of y at large r. For the sextic anharmonic oscillator, the expansion of y at large r and fixed effective coupling λ is rewritten conveniently in variable v, see (3.3.7), y = (2M~ 2) 1 4 ( λ2v3 + (D + 2)λ2 + 1 2λ2 v−1 − ε 2λ2 v−3 + . . . ) , v → ∞ . (7.4.20) The first two terms of this expansion are ε-independent, but only the first one is D-independent. Following an analogous procedure to that used for the cubic and quartic oscillator, we can transform the expansion of Z2k(u) at large u into an expansion at large v via the connection between the classical and quantum coordinate shown in (3.3.24). The first two terms in (7.4.12) expanded at large v read ( 2M g2 )1/2 Z0(λv) = (2M~ 2) 1 4 ( λ2v3 + 1 2λ2 v−1 − 1 8λ6 v−5 + . . . ) , v → ∞ , (7.4.21) and ( 2M g2 )1/2 λ2Z2(λv) = (2M~ 2) 1 4 ( D + 2 2 v−1 − D 2λ2 v−3 − 1 λ4 v−5 + . . . ) , v → ∞ , (7.4.22) see (7.4.17) and(7.4.18). The sum (2M g2 )1/2(Z0(λv) + λ2Z2(λv)) at large v reproduces exactly the first two terms in the expansion (7.4.20). However, higher generating functions, Z4(λv), Z6(λv), ... contribute to the same order O(v−3) at large v, ( 2M g2 )1/2 λ2kZ2k(λv) = (2M~ 2) 1 4 ( −ε4k−4λ 4k−6 2 v−3 + . . . ) , v → ∞ . (7.4.23) Here ε4k−4 denotes the energy correction of order λ4k−4. Therefore, it does not matter how many generating functions are considered in the expansion (2M g2 )1/2(Z0(λv) + λ2Z2(λv) + ...) re-expanded at large v, the ε-dependent coefficient in front of the term of order O(v−3) will never be reproduced exactly. A similar situation occurred for the cubic and quartic cases. 7.4. Sextic Anharmonic Oscillator 67 7.4.3 The Approximant The first two17 generating functions of the phase, G0(r) and G2(r), are given by G0(r) = (2M)1/2 g2 ( r2 4 √ 1 + g4r4 + log [ g2r2 + √ 1 + g4r4 ]) , (7.4.24) and G2(r) = (2M)1/2 g2 ( 1 4 log [ 1 + g4r4 ] + D 4 log [ 1 + √ 1 + g4r4 ]) . (7.4.25) These two functions serve as building blocks for the construction of the Approximant. Following the general prescription (6.1.2), we write the Approximant of the ground state in the exponential representation Ψ (t) 0,0 = e− 1 ~ Φt where 1 ~ Φt(r) = ã0 + ã2 r 2 + ã4 g 2 r4 + ã6 g 4 r6 √ 1 + b̃4 g2 r2 + b̃6 g4 r4 + 1 4g2 log [ c̃2 g 2 r2 + √ 1 + b̃4 g2 r2 + b̃6 g4 r4 ] + 1 4 log [ 1 + b̃4 g 2 r2 + b̃6 g 4r4 ] + D 4 log [ 1 + √ 1 + b̃4 g2 r2 + b̃6g4r4 ] . (7.4.26) At this stage ã0,2,4,6, b̃4,6, and c̃2 are 7 free parameters. The logarithmic terms added in (7.4.26) generate prefactor to the exponential function, Those terms are a certain minimal modification of those that occur in the first two generating functions (7.3.24) and (7.4.25). As a result, the Approximant of the ground state for arbitrary D = 1, 2, 3, . . . is given by Ψ (t) 0,0(r) = 1 ( 1 + b̃4 g2 r2 + b̃6 g4r4 )1/4 ( 1 + √ 1 + b̃4 g2 r2 + b̃6g4 r4 )D/4 × 1 ( c̃2g2r2 + √ 1 + b̃4 g2 r2 + b̃6 g4 r4 )1/4g2 exp  − ã0 + ã2 r 2 + ã4 g 2 r4 + ã6 g 4 r6 √ 1 + b̃4 g2 r2 + b̃6 g4 r4   . (7.4.27) This is the central formula of this Section devoted to the sextic anharmonicity. Once the optimal parameters coming from the VM are chosen, we will show that it provides a highly accurate and uniform approximation for the exact ground state eigenfunction. We impose the constraint b̃6 = 16 ã26 (7.4.28) 17The generating functions G4(r) and G6(r) are presented explicitly in Appendix B. 68 Chapter 7. Results for Two-Term Potentials which allows us to reproduce the dominant term in the expansion (7.4.20) exactly18. Hence, it reproduces the exact dominant asymptotic behavior of the phase at large distances. To construct 5-parametric Approximants for excited states, we follow the prescription presented in Section 6.2. Hence, at D = 1 we have Ψ(t) nr,p(x) = xpP (p) nx (x 2) ( 1 + b̃4 g2 x2 + b̃6 g4x4 )1/4 ( 1 + √ 1 + b̃4 g2 x2 + b̃6g4 x4 )D/4 × 1 ( c̃2g2x2 + √ 1 + b̃4 g2 x2 + b̃6 g4 x4 )1/4g2 exp  − ã0 + ã2 x 2 + ã4 g 2 x4 + ã6 g 4 x6 √ 1 + b̃4 g2 x2 + b̃6 g4 x4   . (7.4.29) while at D > 1 Ψ (t) nr,ℓ (r) = rℓ P (ℓ) nr (r 2) ( 1 + b̃4 g2 r2 + b̃6 g4r4 )1/4 ( 1 + √ 1 + b̃4 g2 r2 + b̃6g4 r4 )D/4 × 1 ( c̃2g2r2 + √ 1 + b̃4 g2 r2 + b̃6 g4 r4 )1/4g2 exp  − ã0 + ã2 r 2 + ã4 g 2 r4 + ã6 g 4 r6 √ 1 + b̃4 g2 r2 + b̃6 g4 r4   . (7.4.30) Definitions of P (p) nr (x 2) and P (ℓ) nr (r 2) are presented in (6.2.1) and (6.2.4), explicitly. 7.4.4 Numerical Results: Variational Calculations The calculations of the variational energy for the first low-lying states with quantum numbers (0,0), (0,1), (0,2), (1,0) are presented in Tables 7.14 - 7.17 for different values of D and g4. For some states, the variational energy Evar = E (t) 0 +E (t) 1 , the first correction E (t) 2 , and the corrected value Evar + E (t) 2 are shown. All states are studied in dimension D = 1, 2, 3, 6 and coupling constant g4 = 0.1, 1, 10. For all cases, the variational energy Evar is obtained with absolute 18In numerical computations we set ~ = 1 and M = 1/2. 7.4. Sextic Anharmonic Oscillator 69 deviation (from the exact result) in the order 10−9−10−12 (9−12 s.d.). This deviation is found by calculating the second correction E (t) 2 (the first correction to the variational energy). To check the energies obtained in Non-Linearization Procedure, we calculated them numeri- cally in the Lagrange-Mesh method. These numerical calculations show that all digits presented for Evar + E (t) 2 are exact: we obtain not less than 13 d.d. correctly in the Non-Linearization Procedure. In general, the energies grow with an increase of D and/or g4. At D > 1, we classify the states by radial quantum number nr and angular momentum ℓ as (nr, ℓ); we mention the hierarchy of eigenstates which holds for any fixed integer D > 1 and g: (0,0), (0,1), (0,2), (1,0). This hierarchy is the same that we obtained for the cubic and quartic anharmonic oscillators. In contrast with the quartic oscillator case, the available literature with estimates of the energy of the low-lying states for the sextic radial anharmonic oscillator is very limited. Most of the calculations are made for the one-dimensional case. In general, our results reproduce all known numerical ones that can be found for D = 1, while for D = 2, 3, we sometimes exceeded their accuracy, see e.g. [25], [26], [66], and [67]. At D = 6, the calculations are carried out for the first time. The relative deviation of Ψ (t) 0,0 from the exact (unknown) ground state eigenfunction Ψ0,0 is estimated with the Non-Linearization Procedure. The deviation is bounded and very small, ∣ ∣ ∣ ∣ ∣ Ψ0,0(r)−Ψ (t) 0,0(r) Ψ (t) 0,0(r) ∣ ∣ ∣ ∣ ∣ . 10−6 , (7.4.31) in the whole range of r ∈ [0,∞) at any integer D and at any coupling constant g4 that we considered. Thus, our Approximant leads to a locally accurate approximation of the exact wave function once optimal parameters are chosen. In general, optimal variational parameters are smooth, slow changing as functions of g4 without strong D dependence. The plots of the optimal parameters for the ground state (0,0) a0,2,3 vs g for D = 1, 2, 3, 6 are shown in Fig. C.4 in Appendix C. In turn, Fig. C.5 presents the plots of the parameters as functions of D for fixed g4. Similar plots occur for excited states for different D and g4. Making an analysis of those plots, one can check that they obey (with high accuracy) another constraint, b̃4 ≈ 32 ã6 ã4 . (7.4.32) This constraint guarantees the exact reproduction of the first two terms in the expansion (7.4.20). If we demand to fulfill exactly constraint (7.4.32), no deterioration in the accuracy of the energy and wave function occurs. Thus, the Approximants either (7.4.29) or (7.4.30) contains ultimately 5 free parameters, namely {ã0, ã2, ã4, ã6, c̃2}. Now we focus on the ground state. For all D and g4, the correction |y1| to y0 = ∂rΦt is a very small and bounded function in 0 . r . 1.7. In this domain, the dominant contribution of integrals - required by the variational method - occurs. For example, for g4 = 1 andD = 1, 2, 3, 6 70 Chapter 7. Results for Two-Term Potentials Table 7.14: Ground state energy for the sextic potential V (r) = r2 + g4 r6 for D = 1, 2, 3, 6 and g4 = 0.1, 1, 10, labeled by quantum numbers (0,0) at any D. Variational energy Evar = Evar, the first correction E (t) 2 (rounded to three s.d.) found with use of Ψ (t) 0,0, see text, the corrected energy Evar +E (t) 2 = Evar +E (t) 2 shown. Evar +E (t) 2 coincides with Lagrange Mesh results (see text) in 12 displayed d. d., hence all printed digits are exact. g4 D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 1.109 087 078 465 1.20× 10−13 1.109 087 078 465 2.307 218 600 932 7.04× 10−13 2.307 218 600 931 1.0 1.435 624 619 003 3.22× 10−13 1.435 624 619 003 3.121 935 474 246 9.81× 10−13 3.121 935 474 246 10.0 2.205 723 269 598 3.22× 10−12 2.205 723 269 595 4.936 774 524 584 1.72× 10−12 4.936 774 524 582 g4 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.596 036 921 222 1.76× 10−12 3.596 036 921 220 7.987 905 269 800 7.11× 10−13 7.987 905 269 799 1.0 5.033 395 937 721 5.21× 10−12 5.033 395 937 720 11.937 202 695 862 9.62× 10−13 11.937 202 695 862 10.0 8.114 843 118 826 7.60× 10−12 8.114 843 118 819 19.880 256 604 739 3.12× 10−12 19.880 256 604 736 Table 7.15: The first excited state energy for the sextic potential V (r) = r2 + g4 r6 for different D and g4. Variational energy Evar, the first correction E (t) 2 found with use of Ψ (t) 0,1 for D = 1 and Ψ (t) 0,1 for D > 1, see (7.4.29) and (7.4.30). The corrected energy Evar + E (t) 2 = Evar + E (t) 2 shown, the correction E (t) 2 rounded to 3 s.d. All printed digits for Evar + E (t) 2 are exact. g4 D = 1 D = 2 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 3.596 036 921 295 7.50× 10−11 3.596 036 921 220 4.974 197 493 807 9.01× 10−11 4.974 197 493 717 1.0 5.033 395 937 795 7.52× 10−11 5.033 395 937 720 7.149 928 601 496 5.84× 10−11 7.149 928 601 438 10.0 8.114 843 118 966 1.48× 10−10 8.114 843 118 818 11.688 236 0345 77 1.81× 10−10 11.688 236 034 396 g4 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 6.439 143 322 388 6.64× 10−11 6.439 143 322 321 11.324 899 788 818 8.15× 10−10 11.324 899 788 004 1.0 9.455 535 276 950 1.09× 10−10 9.455 535 276 841 17.387 207 808 723 1.26× 10−9 17.387 207 807 460 10.0 15.619 579 279 334 5.05× 10−10 15.619 579 278 830 29.302 506 554 618 1.22× 10−9 29.302 506 553 402 7.4. Sextic Anharmonic Oscillator 71 Table 7.16: The second excited state energy for the sextic potential V (r) = r2 + g4 r6 for different D and g4. Variational energy Evar found with use of Ψ (t) 1,0 for D = 1 and Ψ (t) 0,2 for D > 1, see (7.4.29) and (7.4.30). The first correction E (t) 2 and the corrected energy Evar+E (t) 2 = Evar + E (t) 2 shown. Correction E (t) 2 rounded to 3 s.d. All 12 printed d. d. in Evar + E (t) 2 are exact. g4 D = 1 D = 2 Evar Evar −E(t) 2 Evar + E (t) 2 0.1 6.644 391 710 782 7.987 905 270 111 3.12× 10−10 7.987 905 269 799 1.0 9.966 622 004 356 11.937 202 696 127 2.66× 10−10 11.937 202 695 862 10.0 16.641 218 168 076 19.880 256 605 756 1.02× 10−9 19.880 256 604 742 g4 D = 3 D = 6 Evar −E(t) 2 Evar + E (t) 2 Evar −E(t) 2 Evar + E (t) 2 0.1 9.617 462 285 440 1.50× 10−10 9.617 462 285 290 14.962 630 328 506 1.60× 10−10 14.962 630 328 346 1.0 14.584 132 948 883 3.15× 10−9 14.584 132 945 729 23.431 551 835 405 2.19× 10−9 23.431 551 833 215 10.0 24.447 468 037 325 5.42× 10−9 24.447 468 031 906 39.815 551 142 800 7.05× 10−9 39.815 551 135 750 Table 7.17: The third excited state energy for the sextic potential V = r2+g4 r6 for different D and g4. The radial node r0 calculated in Lagrange-Mesh method with 9 correct d.d. Variational energy Evar and its radial node r (0) 0 given by underlined digits, both found with Ψ (t) 1,0, see (7.4.30). g4 D = 2 D = 3 D = 6 Evar r0 Evar r0 Evar r0 0.1 8.402 580 462 0.837 310 052 10.237 873 721 0.995 787 872 16.154 260 610 1.308 543 484 1.0 12.914 938 793 0.671 821 606 15.989 440 787 0.790 364 964 25.938 441 037 1.019 166 200 10.0 21.792 578 251 0.515 914 526 27.155 085 604 0.604 322 682 44.521 781 513 0.773 860 964 the first correction y1 has the upper bound |y1|max ∼    0.0078 , D = 1 0.0065 , D = 2 0.0048 , D = 3 0.0031 , D = 6 (7.4.33) It is the consequence of the fact that, by construction, the derivative y0 = ∂rΦt reproduces the growing terms in the expansion (7.4.20). It implies that we deal with a smartly designed zeroth order-approximation Ψ (t) 0,0 which leads (in the framework of the Non-Linearization Procedure) to a fast convergent series for the energy and wave function. In Fig. 7.4, y0 and y1 vs r are presented for g4 = 1 in D = 1. We emphasize that all curves in these figures are slow-changing vs D. Similar plots appear for D = 2, 3, 6 (not shown) as well as for other values of g4 > 0. An analysis of these plots indicates that (−y1)2 is an extremely small function in comparison with y0 in the domain 0 ≤ r . 1.7, thus, in the domain which provides the dominant contribution 72 Chapter 7. Results for Two-Term Potentials 0.5 1 1.5 2 0 2 4 6 8 10 5 10 15 0 0.002 0.004 0.006 0.008 Figure 7.4: Sextic oscillator at D = 1: function y0 = (Φt) ′ (on left) and its first correction y1 (on right) vs r for g4 = 1. in variational integrals. It is the real reason why the energy correction E (t) 2 is extremely small being of order ∼ 10−8, or sometimes even smaller, ∼ 10−10. A similar situation occurs for the phase (and its derivative) of the Approximants for the excited states. At D > 1, it was checked that the Approximant provides an estimate of the radial node position of the wave function Ψ (t) 1,0 by imposing the orthogonality constraint to the Approximant of the ground state. For sextic anharmonic potential, the zeroth order approximation r (0) 0 gives not less than 7 d.d. of accuracy. The first order correction r (1) 0 gives a contribution to the 6 d.d. Variational results are presented in Table 7.17. They are compared with ones obtained in Lagrange-Mesh method with 200 mesh points. It can be noted that the radial node r0 grows with an increase of D at fixed g4, but decreases with the increase of g4 at fixed D. 7.4.5 First Terms in the Strong Coupling Expansion In this Section, we calculate the first two terms in the strong coupling expansion (5.1.7). For the sextic anharmonic oscillator potential (7.4.1) this (convergent) expansion has the form ε = g(ε̃0 + ε̃6 g −3 + ε̃12 g −6 + . . .) . (7.4.34) We have assumed that 2M = ~ = 1. Evidently, this expansion corresponds to PT in powers of λ̂ for the potential V (w) = w6 + λ̂ w2 , λ̂ = g−3 (7.4.35) in the Schrödinger equation defined in w ∈ [0,∞). We use the Approximant (7.4.27) and develop the perturbation procedure already described in Chapter 5 to calculate ε̃0 and ε̃6. In Table 7.18, we present for different D the variational estimate for ε̃0 (denoted by ε̃ (1) 0 ), the first correction to it ε̂2, and the corrected value ε̃ (2) 0 = ε̃ (1) 0 + ε̂2 calculated via the Non-Linearization Procedure. Using the Lagrange-Mesh method it was verified that ε̃ (2) 0 provides not less than 12 exact d.d. at every D considered. Hence, 7.4. Sextic Anharmonic Oscillator 73 the first 10 d.d. in variational energy ε̃ (1) 0 displayed in Table 7.18 are exact. In this case, the ε̂3 ∼ 10−2ε̂2, it indicates a high rate of convergence in PT. Finally, in Table 7.19, we present the first two approximations of the coefficient ε̃6 - the expectation value of w2 - , see (7.4.34). Comparing our results for ε̃0 and ε̃6 with those available in the literature, see e.g. [25] and [66], one can see that we usually reproduce, although often exceed them. Table 7.18: Ground state (0, 0) energy ε̃0 for the potential W = r6 at D = 1, 2, 3, 6 found in PT based on the Approximant Ψ (t) (0,0) : ε̃ (1) 0 corresponds to the variational energy, ε̂2 is the second PT correction, ε̃ (2) 0 = ε̃ (1) 0 + ε̂2 is the corrected variational energy. 10 d.d. in ε̃ (2) 0 confirmed independently in Lagrange-Mesh method, see text. D = 1 D = 2 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 1.144 802 453 800 3.21× 10−12 1.144 802 453 797 2.609 388 463 259 5.72× 10−12 2.609 388 463 253 D = 3 D = 6 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 ε̃ (1) 0 −ε̂2 ε̃ (2) 0 4.338 598 711 518 4.73× 10−12 4.338 598 711 513 10.821 985 609 895 7.21× 10−12 10.821 985 609 888 Table 7.19: Subdominant coefficient ε̃6 in the strong coupling expansion (7.4.34) of the ground state (0, 0) energy for the sextic radial anharmonic potential (7.4.1) for D = 1, 2, 3, 6. The first correction to it ε̃6,1 as well as the corrected value ε̃ (2) 6 = ε̃ (1) 6 + ε̃6,1 displayed D = 1 D = 2 ε̃ (1) 6 −ε̃6,1 ε̃ (2) 6 ε̃ (1) 6 −ε̃6,1 ε̃ (2) 6 0.307 920 304 114 3.83× 10−10 0.307 920 303 731 0.534 591 069 789 2.85× 10−10 0.534 591 069 504 D = 3 D = 6 ε̃ (1) 6 −ε̃6,1 ε̃ (2) 6 ε̃ (1) 6 -ε̃6,1 ε̃ (2) 6 0.718 220 134 970 1.55× 10−9 0.718 220 133 425 1.137 762 108 070 2.68× 10−10 1.137 762 107 802 Chapter 8 Conclusions It has been developed a formalism to study the bound states of the quantum radial polynomial anharmonic potential V (r) = 1 g2 m∑ k=2 ak g k rk , m ≥ 2 . For this potential, our primary goal was to construct a locally accurate approximate solution of the D-dimensional radial Schrödinger equation. In this context, the key question was: how to design an appropriate approximation leading to locally accurate solutions and also to accurate estimates for the energies? Our answer is based on the available analytical information about the (unknown) exact ground state wave function. General Overview There are two parameterizations of the configuration space to study the radial anharmonic oscillator: (i) Quantum coordinate v-space, v ∝ r/ √ ~. In this case, the anharmonic potential is a finite-degree perturbation in variable v of the harmonic oscillator through the effective coupling constant λ ∝ √ ~ g. The dynamics of the wave function, taken in the exponential representation Ψ = e−Φ, is defined by the derivative of the phase y(r) = ∂rΦ. Written in the quantum variable v, the unknown function y(r) satisfies a non-linear ~-independent differential equation: the so-called Riccati-Bloch (RB) equation. Relevant analytical information can be extracted by studying this equation. One can construct the asymptotic series of y(r) at small and large r. To solve the RB equation, Perturbation Theory (PT) for the energy and the wave function can be developed in the weak and strong coupling regimes. For both cases, the construction of PT is nothing but an application of the Non- Linearization Procedure developed a long time ago. For the general anharmonic potential, 75 76 Chapter 8. Conclusions the construction of PT is a non-algebraic procedure. However, for even anharmonic potentials, V (r) = V (−r), the Non-Linearization Procedure in the weak coupling regime is purely algebraic at any dimension D. (ii) Classical coordinate u-space, with u = gr. Here, the potential is a finite degree polynomial in u. The coupling constant appears in front of the potential as the factor 1/g2. In this space, the dynamics of y(r) = ∂rΦ is governed by the Generalized Bloch (GB) equation: a new non-linear ~-independent differential equation different from the RB one. Just as it occurred with the RB equation, we can obtain analytical information from the GB one as well. In particular, the same asymptotic series for small and large r described in (i) can be constructed from the GB equation. When PT is developed for the GB equation in the weak coupling regime, we checked that it generates the true type of semi-classical expansion (in the non-classical domain, beyond the turning point) for the phase. The main difference between the standard WKB expansion and the present new one is based on the fact that in the true expansion the energy does have an explicit dependence on the Planck constant ~. On the other hand, when PT is developed for the strong coupling regime, we found an interesting connection between the expansion for large g and that for small r: both expansions describe the same domain in r. Both spaces are related through a simple expression: u = λ v. Furthermore, it turned out that there is a remarkable connection between them in the weak coupling regime. This non- trivial connection can be summarized as follows: perturbative corrections calculated for the GB equation are generating functions of those corrections calculated on the basis of the RB equation. All the information collected from (i) and (ii), such as asymptotic series and PT in weak and strong regimes, was used to design a general prescription to construct a locally accurate approximation of the ground state wave function at any D: the Approximant Ψ (t) 0,0(r). When the exponential representation is assumed for the Approximant, Ψ (t) 0,0(r) = e− 1 ~ Φt(r) , we may focus on finding approximately Φt. One key result of this Thesis is a compact expression for the phase 1 ~ Φt(r) = ã0 + ã1 g r + 1 g2 V (r ; ã2, . . . , ãm) √ 1 g2 r2 V (r ; b̃2, . . . , b̃m) + Logarithmic Terms (r ; {c̃}) . It contains a finite number of parameters, {ã, b̃, c̃}. At the weak and strong coupling regimes, it interpolates the asymptotic series of the phase at small and large distances. Furthermore, 77 we showed that this formula contains the analytical information obtained from (i) and (ii) if appropriate constraints are imposed on free parameters. Interestingly, those constraints came from the asymptotic series at large r. Let us emphasize that the potential V (r) appeared as a building block of the approximate phase. The recipe to insert logarithmic terms in Φt(r) was also established: they should mimic the logarithmic terms of the first two generating functions for the phase: G0(r) and G2(r). Imposing orthogonality constraints, the prescription may be used to design Approximants of excited states. Some of these parameters were fixed requiring Φt to reproduce the coefficients of all growing terms in the expansion of the exact phase at large distances. The remaining parameters are considered as variational. In this way, the Approximant is taken as a trial function in the VM. The Approximant is not only suitable to estimate (with high accuracy) the energy of the first low-lying states for finite values of g, but also, due to its features, it can be used to construct the strong coupling expansion. Through the connection of PT with the variational energy, the (relative) deviation between the Approximant and the exact (unknown) wave function may be estimated. A similar situation occurs with the variational energy: we can determine its deviation from the exact result. Results for three particular anharmonic oscillators are summarized below. Two-Term Potentials We applied the formalism summarized above for the anharmonic potential V (r) = r2 + gm−2rm , where m = 3, 4, 6 (cubic, quartic, and sextic). In this Section, we present the conclusions associated with each anharmonicity. Cubic The key expression for the approximate phase of the Approximant is 1 ~ Φt(r) = ã0 + ã1 gr + ã2 r 2 + ã3 g r 3 √ 1 + b̃1 g r + 1 4 log[1 + b̃1 g r] + D log [ 1 + √ 1 + b̃1 g r ] , where {ã0, ã1, ã2, ã3, b̃1} are parameters. If the constraints b̃1 = 25 4 ã23 , ã2 = 125 ã23 + 12 150ã3 , −g2ã1 = 9375 a43 − 1000 a23 + 48 15000 a33 , 78 Chapter 8. Conclusions are imposed, Φt reproduces all growing terms of the exact phase at large distances. Furthermore, the constraint ã0 = 2ã1 b̃3 + D + 1 2 is imposed to have vanishing ∂rΦt at r = 0. As a result, Φt depends only on 1 free parameter. Taking Ψ (t) 0,0 as trial function, the single parameter can be chosen as optimal in the VM. Thus, the variational energy of the ground state is obtained with absolute accuracy ∼ 10−5 (4- 5 s.d.) for D = 1, 2, 3, 6 and g = 0.1, 1.0, 10.0. This is confirmed by the calculation of the first correction E (t) 2 to the variational energy Evar, which is always of the order ∼ 10−6, while the rate of convergence of the energy in Non-Linearization Procedure is ∣ ∣ ∣ ∣ E (t) 3 E (t) 2 ∣ ∣ ∣ ∣ ∼ 10−4. The relative deviation of Ψ (t) 0,0(r) from the exact wave function is bounded and small, it does not exceed ∼ 10−2 for any value of r. Hence, the Approximant leads to a locally accurate approximation to the exact ground state wave function. Results of this accuracy were obtained for the low-lying states: (0, 0), (0, 1), (0, 2), (1, 0). Undoubtedly, a similar accuracy would be reached for other dimensions D and coupling constants g. For those states, the optimal parameters are smooth, slow-changing functions of g andD. These results represent the first accurate calculations of the energy of the low-lying states for the D-dimensional cubic anharmonic oscillator. Hence, these results may serve as a benchmark for other calculations. To the best of the knowledge of the author, there are not many examples in the literature in which a one-parametric Approximant (trial function) leads to such accuracies for the energy and wave functions. Quartic For this anharmonicity, the prescription generates an approximate phase of the form 1 ~ Φt(r) = ã0 + ã2 r 2 + ã4 g 2 r4 √ 1 + b̃4 g2 r2 + 1 4 log [ 1 + b̃4 g 2 r2 ] + D 2 log [ 1 + √ 1 + b̃4 g2 r2 ] , where {ã0, ã2, ã4, b̃4} are parameters, which are constrained b̃4 = 9 ã24 , ã2 = 1 + 27 ã24 18ã4 . These constraints guarantee that Φt reproduces exactly all two growing terms of the exact phase at large distances. The remaining 2 free parameters are chosen to be optimal in the VM with Ψ (t) 0,0(r) as trial function. The variational energy of the ground state is obtained with absolute accuracy ∼ 10−8 (8 - 9 s.d.) for D = 1, 2, 3, 6 and g = 0.1, 1.0, 10.0. This is confirmed by the first correction E (t) 2 , which is always of the order ∼ 10−8. The rate of convergence of the 79 energy in the Non-Linearization Procedure is ∣ ∣ ∣ ∣ E (t) 3 E (t) 2 ∣ ∣ ∣ ∣ ∼ 10−4. The relative deviation of Ψ (t) 0,0(r) from the exact wave function is bounded and very small, it does not exceed ∼ 10−6. Hence, the Approximant led to a locally accurate approximation to the exact wave function. Results of similar accuracy were obtained for the low-lying states: (0, 0), (0, 1), (0, 2), (1, 0). Clearly, the same accuracy would be reached for other dimensions and coupling constants. For those states, the optimal parameters are smooth and slow-changing functions of g and D. The variational estimates energy have confirmed, and some times exceeded, the results found in the literature. Sextic According to the prescription, the approximate phase of the sextic case is given by 1 ~ Φt(r) = ã0 + ã2 r 2 + ã4 g 2 r4 + ã6 g 4 r6 √ 1 + b̃4 g2 r2 + b̃6 g4 r4 + 1 4g2 log [ c̃2 g 2 r2 + √ 1 + b̃4 g2 r2 + b̃6 g4 r4 ] + 1 4 log [ 1 + b̃4 g 2 r2 + b̃6 g 4r4 ] + D 4 log [ 1 + √ 1 + b̃4 g2 r2 + b̃6g4r4 ] , where {ã0, ã2, ã4, ã6, b̃2, b̃4} are parameters. which are constrained b̃6 = 16 ã26 , b̃4 = 32 ã6 ã4 . These constraints reduced the number of parameters to 4; they guarantee that Φt reproduces exactly all growing terms of the exact phase at large distances. The remaining 4 free parameters are chosen to be optimal in the VM with Ψ (t) 0,0(r) as trial function. The variational energy coincides with the exact one with not less than 8 s.d. for the states (0, 0), (0, 1), (0, 2), (1, 0) for D = 1, 2, 3, 6 and g4 = 0.1, 1.0, 10.0. However, there are no doubts that it will be the same for any integer D and g4 > 0. For those states, the optimal parameters are smooth slow-changing as functions of g and D. The rate of convergence of energy in the Non-Linearization Procedure remains ∣ ∣ ∣ ∣ E (t) 3 E (t) 2 ∣ ∣ ∣ ∣ ∼ 10−4. The relative deviation of Ψ (t) 0,0(r) from the exact eigenfunction is a bounded function and very small, it does not exceed ∼ 10−6. Hence, the Approximant led to a locally accurate approximation of the exact wave function. Results of similar accuracy were obtained for the first low-lying states. The variational estimates of the energy have confirmed, and some times exceeded, the results found in the literature. 80 Chapter 8. Conclusions Final Remarks The reproduction of the growing terms at large distances of the phase reduces the number of parameters that the Approximant contains. Hence, the dimensionality of the space of param- eters is reduced. If the remaining free parameters are chosen to be optimal in the VM, highly accurate estimates of the energy will occur. In addition, the Approximant will become a locally accurate approximation of the exact wave function. We noticed that the hierarchy of the energies of the first four low-lying states of the cubic, quartic, and sextic anharmonic oscillators is the same: (0, 0), (0, 1), (0, 2), (1, 0) in arbitrary D > 1 and non-vanishing coupling constant g. See Fig 8.1. E n er gy (0, 0) (1, 0) (0, 1) (0, 2) E E E Figure 8.1: Schematic energy diagram for the first four bound states (nr, ℓ) for the potential V (r) = r2 + gm−2rm, m = 3, 4, 6, at g > 0 in arbitrary dimension D. The first two terms in the strong coupling expansion were estimated with not less than 9 d.d. for the cubic, quartic, and sextic anharmonic oscillator. For the cubic potential, the first two terms were calculated for the first time. For quartic and sextic, our estimates reproduced the benchmarks known in the literature at D = 1, 2, 3. In D = 6, we established the benchmark. It is clear that the formalism presented in this Part can be used to study some other systems in quantum mechanics. For example, a similar formalism was used in the past to study the low-lying states of the Yukawa potential leading to highly accurate results, see [47]. Part II Hydrogen Atom in Constant Magnetic Field 81 Chapter 9 Introduction The hydrogen atom is the simplest atomic system in Nature. It contains a positively charged proton and a negatively charged electron bound to the nucleus via the Coulomb force. Due to the separation of variables in spherical coordinates (and in some other systems of coordinates), the Schrödinger equation associated with this system was solved exactly in the seminal paper of modern quantum mechanics [13]. It is not an exaggeration to say that it laid the groundwork for subsequent studies of more complex atoms and molecules. In a non-relativistic consideration, when the hydrogen atom is placed in a constant uniform magnetic field, separation of variables no longer occurs for the Schrödinger equation, and it becomes non-solvable. The impossibility of obtaining exact solutions makes necessary the use of approximate methods to find the spectrum of the system. This feature makes the hydrogen atom in a constant uniform magnetic field an appropriate system to study using the formalism presented in Part I. 9.1 Motivation The study of the atomic properties in the presence of a magnetic field dates back to a series of experiments carried out by Zeeman around 1896. In those experiments, he noticed the splitting of the spectral lines of sodium when their source is placed under the influence of a magnetic field.1 Early atomic models were incapable of describing adequately such phenomena. It was not until the theory of quantum mechanics was established (1926) that a satisfactory theoretical description occurred. Due to the success of this new theory, it was intermediately clear that a study in this quantum mechanical framework could explain in full detail the splitting observed by Zeeman. It was Schrödinger who presented the first results of quantum nature for the 1It is necessary to mention that around that time, the notions of proton and electron were absent. The electron would be experimentally detected one year later, while the proton 15 years later. 83 84 Chapter 9. Introduction splitting of the spectral lines of atoms with one electron of valence2 [68]. The splitting of a given spectral line into components in the presence of a small magnetic field became known as (linear) Zeeman effect. In stronger fields, the resulting splitting was called Paschen-Back effect [69]. At ultra-strong fields, the splitting was known as Quadratic Zeeman effect. Each regime was treated mainly in divergent perturbation theory (PT) and some other approximate methods such as the adiabatic approximation and the Variational Method. Early attempts to explore atomic systems in a magnetic field are summarized in the review paper by Garstang [70]. Some of those early attempts were motivated by astrophysics discoveries. This motivation dates from the discovery by Hale in 1908 of magnetic fields in sunspots from the Zeeman splitting of their spectral lines. Some other works were motivated by plasma and semiconductor physics, where the effects of a large magnetic field can be present. In recent decades, the interest has been renewed due to the discovery of white dwarfs and neutron stars. Their atmospheres, rich in hydrogen, experience the strongest magnetic fields γ that can be found in Nature until today. For example, a neutron star may have magnetic fields in the range 1012 − 1013 G, while magnetars in the range 1013 − 1015 G. Due to its simplicity, in comparison with other atoms, the hydrogen one has played a fun- damental role in the understanding of the behavior of more complex atoms in magnetic fields. In fact, the hydrogen atom is the most studied Coulombic system in weak, intermediate, and strong magnetic fields. In the non-relativist consideration, there is a large number of works de- voted to finding its spectrum approximately as a function of the magnetic field. Most of them are devoted to study the energy levels, but no other relevant quantities like the quadrupole moment or even the wave function. 9.2 Present Work The purpose of Part II is to extend and apply the formalism presented in Part I to one of the most relevant atomic systems in the presence of a constant uniform magnetic field: a hy- drogen atom. Our ultimate goal is to design an adequate approximation of the ground state wave function following the spirit of Part I. However, it should be noted that the Schrödinger equation associated with the hydrogen atom in a magnetic field is effectively a two-dimensional partial differential equation. It contrasts with the Schrödinger equation that corresponds to the D-dimensional radial anharmonic oscillator, which was effectively one-dimensional due to the separation of angular variables from the radial one. Despite this significant difference, as well as of the obvious ones (potentials with different physical and analytical natures), we will show that the formalism can be applied with success. In particular, the Riccati-Bloch and Gener- alized Bloch equations can be derived to calculate PT in the weak coupling regime and study asymptotic series. All the analytical information is collected to construct The Approximant : a locally accurate and uniform approximation of the wave function of the ground state. This Part may be read independently on the previous one, it is self-contained. However, 2 Just like the sodium atoms considered experimentally by Zeeman. 9.2. Present Work 85 sometimes we will make some analogies between the results presented here and those of Part I. Chapter 10 Generalities The Hamiltonian Ĥ = 1 2me ( p̂ + e c A )2 − e2 r , r = √ x2 + y2 + z2 , (10.0.1) may describe a hydrogen atom in the presence of a constant magnetic field B = γ ẑ. In (10.0.1), the mass of the proton is assumed to be infinite, mp = ∞ . Here me and −e are the mass and charge of the electron, respectively, e corresponds to the charge of the infinitely massive proton located at the origin of coordinates, and c is the speed of light. Besides, p̂ = −i~∇ is the momentum operator of the electron, r is its distance to the origin. See Fig. 10.1 for the geometrical setting. It is straightforward to show that (10.0.1) is equivalent to Ĥ = 1 2me p̂2 + 1 mec p̂ ·A + 1 mec A · p̂+ e2 2mec2 A2 − e2 r (10.0.2) From now on, the symmetric gauge A = 1 2 B× r , B = γ ẑ , (10.0.3) is assumed1, it implies that the Hamiltonian (10.0.2) takes the form Ĥ = − ~2 2me ∆ + |e|γ 2mec L̂z + V , ∆ = ∂2x + ∂2y + ∂2z . (10.0.4) Here, we defined the potential V = −e 2 r + e2γ2 8mec2 ρ2 , ρ = √ x2 + y2 , (10.0.5) 1This choice of the vector potential leads to a constant magnetic field directed along the z-axis. The magni- tude of this magnetic field is γ, measured in Gauss/Teslas. 87 88 Chapter 10. Generalities depends on r and ρ only, L̂z is the projection of the momentum operator on the direction of magnetic field, L̂z = −i~ (x ∂y − y ∂x) , (10.0.6) which is conserved. The Schrödinger equation associated to (10.0.4) is Ĥψ = Eψ , (10.0.7) with boundary conditions imposed in such a way that the wave function is normalizable ∫ ψ2 dV <∞ . (10.0.8) The dependence of the potential on variables ρ and r hints that it is convenient to write the Schrödinger equation (10.0.7) in the non-orthogonal system of coordinates (ρ, r, ϕ), see Fig. 10.1. In these coordinates it reads2 x y z e r ϕ ρ p Figure 10.1: System of coordinates (ρ, r, ϕ) at half-space z ≥ 0. The infinitely heavy proton (p) is located at the origin. − ~2 2me [ ∂2ρ + 2ρ r ∂ρr + ∂2r + 1 ρ ∂ρ + 2 r ∂r + 1 ρ2 ∂2ϕ ] ψ − i|e|~γ 2mec ∂ϕψ + [ −e 2 r + γ2e2 8mec2 ρ2 ] ψ = E ψ . (10.0.9) In Appendix E we present some relevant expressions involving differential operators written in this system of coordinates. Needless to say, equation (10.0.9) is non-solvable for γ 6= 0: energy and wave function can be found only in approximate form. 2Note that in this coordinates L̂z = −i~ ∂ϕ. 10.1. The Ground State: Riccati Equation: 89 Due to cylindrical symmetry of the system, any wave function is characterized by two quan- tum numbers: the magnetic quantum number m (eigenvalue of L̂z) and the parity ν = ±1 with respect to a reflection z → −z. It suggests considering the representation ψ(ρ, r, ϕ) = zpρ|m|Ψ(ρ, r) eimϕ , m = 0,±1,±2, ... , p = ±1 (10.0.10) for the states with definite parity, hence ν = (−1)p. Here z = √ r2 − ρ2, see Fig. 10.1. The function Ψ(ρ, r) satisfies the equation − ~2 2me [ ∂2ρ + 2ρ r ∂ρr + ∂2r + 2|m|+ 1 ρ ∂ρ + 2(|m|+ p+ 1) r ∂r ] Ψ + [ −e 2 r + γ2e2 8mec2 ρ2 ] Ψ = Em,p Ψ , (10.0.11) where Em,p = Em,p − |e|~γ m 2mec (10.0.12) is the energy with the linear Zeeman term subtracted. Note that the quantum number m now plays the role of a parameter. From equation (10.0.11) one can explicitly see that Em,p should be an even function with respect to the magnetic quantum number m, Em,p = E−m,p , (10.0.13) for any possible value of p. Therefore Em,p = E−m,p + |e|~γ m mec . (10.0.14) Some general properties of the spectrum of the Schrödinger equation, such as the appearance quasi-crossings and crossings of energy levels, its structure as well as analytical information, can be found in the literature, see [70]. 10.1 The Ground State: Riccati Equation: The present study focuses on the ground state, which corresponds to quantum numbers m = 0 and ν = +. Thus, its wave function depends only on variables (ρ, r). At γ = 0, it corresponds to the 1s0 state of the hydrogen atom. From now on, we present the (nodeless) ground wave function and its energy dropping the labels of quantum numbers, presenting them as Ψ and E , respectively. It is usual to denote these states as 1s0 even for γ 6= 0. Following (10.0.11), the equation that determines Ψ and E reads − ~2 2me [ ∂2ρ + 2ρ r ∂ρr + ∂2r + 1 ρ ∂ρ + 2 r ∂r ] Ψ + [ −e 2 r + γ2e2 8mec2 ρ2 ] Ψ = E Ψ . (10.1.1) We now study the function Ψ(ρ, r) when it is written in the exponential representation, namely Ψ(ρ, r) = e−Φ(ρ,r) . (10.1.2) 90 Chapter 10. Generalities One can show that the phase Φ(ρ, r) satisfies a non-linear partial differential equation of second order, ∂2ρΦ + 2ρ r ∂ρrΦ + ∂2rΦ + 1 ρ ∂ρΦ + 2 r ∂rΦ − (∂ρΦ) 2 − 2ρ r (∂ρΦ)(∂rΦ) − (∂rΦ) 2 = 2me ~2 [ E + e2 r − γ2e2 8mec2 ρ2 ] . (10.1.3) This equation is defined in the domain 0 ≤ ρ ≤ r and 0 ≤ r <∞, see Fig. 10.2. ρ = r r ρ Figure 10.2: Infinite wedge domain (shaded in gray) of the equation (10.1.3) in (ρ, r) space. Note that (10.1.3) can be regarded as a generalization for two dimensions of the well-known one-dimensional Riccati equation, cf. (3.3.4). It is worth mentioning that (10.1.3) is the key equation of the present Chapter. Naturally, this equation can be solved exactly at γ = 0; it will generate the 1s0 orbital. For γ 6= 0 it becomes non-solvable: its solution has to be found in approximate form. We have already met in the past this situation for the D-dimensional radial anharmonic oscillator. In the next Sections, we will show that the Riccati-Bloch and Generalized Bloch equations may be found from equation (10.1.3). 10.2 Riccati-Bloch Equation Let us define the dimensionless variables s = ρ r0 , t = r r0 , (10.2.1) where r0 = ~2 mee2 ≃ 5.29× 10−9cm (10.2.2) 10.3. Generalized Bloch Equation 91 is the Bohr radius. In these new variables (10.2.1), the Riccati equation (10.1.3) appears without explicit dependence on c, e, ~, and me, namely ∂2sΦ + 2s t ∂stΦ + ∂2tΦ + 1 s ∂sΦ + 2 t ∂tΦ − (∂sΦ) 2 − 2s t (∂sΦ)(∂tΦ) − (∂tΦ) 2 = ε + 2 t − λ2s2 4 , (10.2.3) where ε = E E0 , E0 = mee 4 2~2 (10.2.4) and λ = γ γ0 , γ0 = c|e|3m2 e ~3 . (10.2.5) Note that E0 is the Rydberg constant while γ0 is the atomic unit of magnetic field, respectively; E0 ≈ 2.18× 10−18J = 13.6 eV , (10.2.6) γ0 ≈ 2.35× 105 T = 2.35× 109 G . (10.2.7) Expressions (10.2.6) and (10.2.7) suggest that λ is the effective magnetic field measured in atomic units γ0, which occurs instead of γ, while ε plays the role of energy measured in Rydbergs (Ry). Equation (10.2.3) is nothing but the dimensionless version of the Riccati equation, and we call it Riccati-Bloch (RB) equation. Both equations coincide when we set ~ = 1, me = 1 and e = −1. The RB equation governs the dynamics via the phase Φ in the (s, t)-space. Note that (10.2.3) is analogous to equation (3.3.9). 10.3 Generalized Bloch Equation Let us introduce in (10.1.3) different dimensionless variables u = ρ ρ0 , v = r r0 , (10.3.1) where ρ0 = mec|e| ~γ . (10.3.2) Note that ρ0 has explicit dependence on γ, while v coincides with t, see (10.2.1) and (10.3.1). Introducing (10.3.1) into the Riccati equation (10.1.3), we obtain a two-dimensional Generalized Bloch (GB) equation λ2 ∂2uΦ + 2u v ∂uvΦ + ∂2vΦ + λ2 u ∂uΦ + 2 v ∂vΦ − λ2(∂uΦ) 2 − 2u v (∂uΦ)(∂vΦ) − (∂vΦ) 2 = ε + 2 v − u2 4 . (10.3.3) 92 Chapter 10. Generalities cf. (3.3.20), see [44,45]. Note that the potential in the right hand side does not have any explicit dependence on the parameters of the problem. The definitions of ε and λ are given in (10.2.4) and (10.2.5), respectively. Just as it occurred for the RB equation, all variables/quantities involved in (10.3.3) are dimensionless. Let us note that the variables u and s are remarkable easy related u = λ s , (10.3.4) see (10.2.1), (10.2.5) and (10.3.1). This relation looks very similar to the one which occurred for the anharmonic oscillator between the classical coordinate and the quantum one, see (3.3.24). Chapter 11 Weak Coupling Regime 11.1 Riccati Bloch Equation Just like we did for the D-dimensional radial anharmonic oscillator, we can use PT assuming a weak coupling regime to solve the RB equation (10.2.3), Φ(s, t) = ∞∑ n=0 λ2nΦn(s, t) , ε = ∞∑ n=0 λ2nεn . (11.1.1) The zeroth order correction is Φ0(s, t) = t , ε0 = −1 , (11.1.2) it corresponds to the exact phase and energy of ground state (1s0) of the hydrogen atom at γ = 0, respectively. It must be emphasized that if we set c = 1, me = 1 and e = 1, then the effective coupling constant results in λ = γ~3. It implies that the PT developed for the ε in powers of γ coincides with the semi-classical expansion in powers of ~3.1 For n > 0, one can show that corrections Φn and εn are determined by a linear partial differential equation, ∂ssΦn + 2s t ∂stΦn + ∂ttΦn + ( 1 s − 2s t ) ∂sΦn + ( 2 t − 1 ) ∂tΦn = εn − Qn , (11.1.3) where Qn = − n−1∑ k=1 [ ∂sΦk∂sΦn−k + ∂tΦk∂tΦn−k + s t (∂sΦk∂tΦn−k + ∂tΦk∂sΦn−k) ] (11.1.4) 1We already showed in the previous Part that a similar situation occurs for the radial anharmonic potential. In that case, the semi-classical expansion of the energy is in powers of ~1/2. 93 94 Chapter 11. Weak Coupling Regime is constructed from the previous (n − 1) corrections. The explicit analytical solution can be found for equation (11.1.3), for example, the first order correction is Φ1(s, t) = 1 24 s2t + 1 16 s2 + 1 24 t2 , ε1 = 1 2 . (11.1.5) In general, the nth correction Φn is a polynomial in variables (s, t) with the following structure Φn(s, t) = n−1∑ j=0 n∑ k=j ( a (n) j,k t + b (n) j,k ) s2(n−k)t2(k−j) , a (n) 0,n = 0 . (11.1.6) For all order n, the coefficients a (n) j,k and b (n) j,k are rational real numbers. Interestingly, the equation εn = 4 b (n) n−1,n−1 + 6 b (n) n−1,n (11.1.7) holds for any n > 1. Hence the correction εn is also a rational number. Several corrections can be easily constructed in this framework as a consequence of the polynomial nature of the corrections Φn. They are determined by solving algebraic linear equations via elementary algebra. Hence, the realization of PT is ultimately an algebraic procedure. As a consequence, many terms in expansions (11.1.1) can be calculated explicitly. In Appendix F, we present the first corrections of the Φ for n = 2, 3, 4. In turn, we present some corrections up to 100th order for the energy ε; see Tables F.1 and F.2. It must be emphasized that it is the first time that the coefficient ε100 has been calculated exactly. The corrections were computed using Mathematica with absolute accuracy (in the form of rational numbers). The code designed for this purpose is very compact: only 15 lines of commands are needed. The code is presented in explicit form in Appendix F. The calculation of the first 50 corrections does not take more than one hour of computations in a common laptop. To arrive up to order 100, it was necessary to leave the laptop working for about two weeks. Despite the simplicity of the calculation of PT corrections, it should be mentioned that, according to the Dyson instability argument [50], both series (11.1.1) should be divergent. This can be seen from the asymptotic behavior of εn for large n [71], εn = 64 (−1)n+1 π 5 2 +2n Γ ( 2n+ 3 2 ) ( (1 + O(n−1) ) , n→ ∞ . (11.1.8) Following this formula, it is clear that the PT for ε is Borel-summable2. However, the Padé- Borel procedure does not provide accurate results, even taking into account the first hundred coefficients εn, see below. A similar situation occurs with conformal mapping [73]. From equations (11.1.8) and (11.1.7), we note that the coefficients b (n) n−1,n−1 and b (n) n−1,n also grow factorially at large n. It is worth mentioning that the perturbative approach used to solve the Riccati-Bloch equation is nothing but an example of the application of the so-called Non-Linearization Procedure [8] (sometimes referred as logarithmic PT). In the previous Part, it was used to solve the one-dimensional version of the RB equation. Here we have shown that it can also be applied successfully to two-dimensional problems. A general description of this method can be found in [8]. 2See [72] for a detailed description of this method. 11.2. Generalized Bloch Equation 95 11.1.1 Behavior of the Phase Φ at Small Distances For one-dimensional spectral problems, the exponential representation is convenient to study the asymptotic behavior of the phase at some particular limits. For the D-dimensional radial anharmonic oscillator, we showed that both RB and GB equations are suitable to construct asymptotic series of the phase. This situation changes dramatically for the hydrogen atom in a constant magnetic field. In this case, the phase is determined by a non-linear partial (RB or GB) differential equation in two variables; it complexifies the construction of asymptotic series. Interestingly, from the structure of corrections Φn, we can overcome this difficulty. The structure of the Taylor series of the phase Φ at small s and t can be obtained from the polynomial form of the correction Φn, see (11.1.6). Collecting the same powers in s and t, coming from different corrections Φn, their formal sums will result in the asymptotic expansion at small (s, t), namely Φ(s, t) = t + σ1(λ 2) s2 + σ2(λ 2) t2 + σ3(λ 2) s2t + . . . , (s, t) → 0 . (11.1.9) The functions σ1, σ2 and σ3 are given (formally) by σ1(λ 2) = ∞∑ n=1 b (n) n−1,n−1λ 2n , σ2(λ 2) = ∞∑ n=1 b (n) n−1,nλ 2n , σ3 = ∞∑ n=1 a (n) n,n−1λ 2n . (11.1.10) From equation (11.1.7), it is clear that ε(λ2) = −1 + 4 σ1(λ 2) + 6 σ2(λ 2) . (11.1.11) From Taylor series (11.1.9), once variables ρ and r are restored, one can immediately conclude that the presence of a magnetic field does not break the cusp condition for the exact wave function C ≡ 〈 ψ ∣ ∣ δ(~r) ∂ ∂r ∣ ∣ψ 〉 〈ψ | δ(~r) |ψ〉 = − 1 r0 . (11.1.12) In atomic physics, the parameter C is known as the cusp parameter [6]. For non-solvable systems which consist of electrons and (infinitely massive) positive charges, it is usually used to measure the local quality of the approximate wave function near the Coulomb singularities. 11.2 Generalized Bloch Equation To solve the GB equation, we can develop PT in powers of λ2, Φ(u, v) = ∞∑ n=0 λ2nφn(u, v) , ε = ∞∑ n=0 λ2nεn . (11.2.1) The expansion for ε coincides with the one presented in (11.2.1). The zeroth order correction φ0(u, v) is determined by the non-linear partial differential equation 2u v ∂uvφ0 + ∂2vφ0 + 2 v ∂vφ0 − 2u v (∂uφ0)(∂vφ0) − (∂vφ0) 2 = ε0 + 2 v − u2 4 . (11.2.2) 96 Chapter 11. Weak Coupling Regime Its explicit solution in closed analytical form is given by φ0(u, v) = A (0) 0 (u) v + B (0) 0 (u) (11.2.3) where A (0) 0 (u) = √ 1 + u2 12 , B (0) 0 (u) = 1 2 log ( 1 + u2 12 ) + log ( 1 + √ 1 + u2 12 ) . (11.2.4) The correction φn(u, v) with n ≥ 1 obeys a linear partial differential equation, namely 2u v ∂u,vφn + ∂2vφn + 2 v ∂vφn − 2u v (∂uφ0 ∂vφn + ∂uφn ∂vφ0) − 2(∂vφn ∂vφ0) = εn − qn (11.2.5) where qn = ∂2uφn−1 + ( 1 u − ∂uφ0 ) ∂uφn−1 − n−1∑ k=1 { ∂uφn−k−1∂uφk + 2u v ∂uφn−k−1∂vφk + ∂vφn−k−1∂vφk } . (11.2.6) It can be shown that all correction φn(u, v) is a polynomial in v with u-dependent coefficients, φn(u, v) = n∑ k=0 { A (n) k (u) v + B (n) k (u) } v2(n−k) . (11.2.7) In general terms, the functions A (n) k (u) and B (n) k (u) are determined by solving (ordinary) linear differential equations of first degree. To write the first order correction (and next ones), it is convenient to introduce the variable w(u) = √ 1 + u2 12 . (11.2.8) Thus, φ1(u, v) = A (1) 0 (u) v3 + B (1) 0 (u) v2 + A (1) 1 (u) v + B (1) 1 (u) (11.2.9) where A (1) 0 = −(w − 1)(w + 1) 120w3 , (11.2.10) B (1) 0 = −6w3 − w2 − 9w − 6 120 (w + 1)w4 , (11.2.11) A (1) 1 = −(w − 1) (30w4 + 52w3 + 54w2 + 42w + 15) 120 (w + 1)w5 , (11.2.12) B (1) 1 = −(w − 1) (9w6 + 18w5 + 38w4 + 46w3 + 42w2 + 30w + 10) 80 (w + 1)w6 . (11.2.13) 11.2. Generalized Bloch Equation 97 The next corrections φ2, φ3, ..., can also be calculated explicitly. However, we omit to present them since they involve cumbersome expressions. As we discuss in the next Section, some interesting properties of functions A (n) k (u) and B (n) k (u) are related to their asymptotic behavior3 at large u and v. 11.2.1 Asymptotic Behavior of Φ at Large Distances Using the corrections in PT obtained from the generalized Bloch equation, the asymptotic series can be calculated in some particular directions in the domain (s, t). Let us consider the line s = α t (11.2.14) where α ∈ (0, 1] is a parameter. One can see that along this line (keeping the value of α fixed), the dominant asymptotic behavior of the nth correction to the phase φn(u, v) = φn(λs, t)s=αt (11.2.15) at large t comes from the term A (n) 0 (λ(αt))t2n+1, see (11.2.7). Concisely, φn(λ s, t)s=αt ∼ An (αλ)2n−1 t2 + O ( t0 ) , t→ ∞ , n = 0, 1, 2, ... , (11.2.16) where An is a coefficient. The first three coefficients are A0 = 1 2 √ 3 , A1 = − 1 20 √ 3 , A2 = − 23 2800 √ 3 . (11.2.17) From (11.2.16), one can note that the behavior of φ(λ s, t)s=αt is given by φ(λ s, t)s=αt = λ ( α ∞∑ n=0 Anα −2n ) t2 + 2 log(t) + . . . , t → ∞ . (11.2.18) The logarithmic (subdominant) term comes from the function B (0) 0 (u) displayed in (11.2.4). In order to determine the sum in (11.2.18), we can define the generating function A(α) = α ∞∑ n=0 Akα −2n . (11.2.19) One can check that it satisfies a non-linear differential equation, namely (1− α2) (A′)2 + 4A2 − 1 4 α2 = 0 , A(1) = 1 4 . (11.2.20) 3At this stage, one can suspect that corrections φn play the role of generating functions, just like Gn(r) did in the case of the radial anharmonic oscillator, see (4.4.2). Let us remind that, based on the functions Gn(r), one can construct the asymptotic expansion of the phase at large distances. 98 Chapter 11. Weak Coupling Regime The exact solution to this equation is given by A(α) = 1 4 α2 . (11.2.21) Hence, the behavior of the phase (11.2.15) at t→ ∞ results in Φ(λ s, t)s=αt = 1 4 α2 λ t2 + 2 log(t) + O(t0) , t→ ∞ . (11.2.22) Some other relevant limits in different directions can also be studied. For example, if s is fixed, say s = s0, the asymptotic expansion at large t is of the form Φ(s0, t) = C0(s0, λ) t + C1(s0, λ) log(t) + O(t0) , t→ ∞ . (11.2.23) Here C1 and C2 are some unknown functions, except at λ = 0. As we have seen, the perturbative solution of the GB equation in powers of λ2 can provide information about the behavior of the exact phase at large distances. In contrast with the solution of the GB equation of the radial anharmonic oscillator, further manipulations4 in corrections φn are needed to obtain the asymptotic behavior of the phase. 11.3 Connection between RB and GB Equations We have constructed two different representations for the phase Φ of the ground state wave function. From one side Φ = ∞∑ n=0 λ2nΦn(s, t) , s = ρ r0 , t = r r0 , (11.3.1) and from the other Φ = ∞∑ n=0 λ2nφn(u, v) , u = ρ ρ0 , v = r r0 . (11.3.2) The definitions of r0 and ρ0 appear in (10.2.2) and (10.3.2), respectively. Due to uniqueness of the phase Φ, it is clear that there must be a connection between corrections Φn(s, t) and φn(u, v). In order to establish it, we use the polynomial structure of corrections (11.1.6) in (11.3.1), thus Φ = ∞∑ n=0 λ2n n∑ j=0 n∑ k=j ( a (n) j,k t + b (n) j,k ) s2(n−k)t2(k−j) . (11.3.3) 4In the case of the D-dimensional anharmonic oscillator, it was enough to expand the first functions Gn(r) at large distances to obtain the asymptotic behavior of the phase. 11.3. Connection between RB and GB Equations 99 Changing the order of summation and using the relation (10.3.4) between variables s and u, it can be shown that (11.3.3) is equivalent to Φ = ∞∑ n=0 λ2n n∑ k=0 ∞∑ j=n ( a (j) k,n v + b (j) k,n ) u2(j−n)v2(n−k) . (11.3.4) Comparing expansions (11.3.2) and (11.3.4), we conclude that φn(u, v) = n∑ k=0 ∞∑ j=n ( a (j) k,n v + b (j) k,n ) u2(j−n)v2(n−k) (11.3.5) due to the uniqueness of the Taylor series. After further manipulations, (11.3.5) is expressed as follows φn(u, v) = n∑ k=0 v2(n−k) ∞∑ j=0 ( a (n+j) k,n v + b (n+j) k,n ) u2j . (11.3.6) Comparing (11.2.7) and (11.3.6), we explicitly find that A (n) k (u) = ∞∑ j=0 a (n+j) k,n u2j , (11.3.7) and B (n) k (u) = ∞∑ j=0 b (n+j) k,n u2j . (11.3.8) Therefore, the meaning of the connection between the GB and RB equations is the following: the coefficient functions A (n) k (u) and B (n) k (u) are nothing but generating functions of the coefficients a (n) j,k and b (n) j,k , respectively. 5 5A similar statement was formulated for the D-dimensional radial anharmonic oscillator, see Chapter 2, Section 4.3. Chapter 12 The Approximant The analytical information about the phase, obtained from the RB and GB equations, Taylor and asymptotic series, is now used to design the Approximant : an approximation of the exact (unknown) ground state wave function denoted by ψ(t). To do so, we follow (as long as pos- sible) the prescription proposed and applied in the previous Part, where it was successful in constructing the Approximant of some particular cases of the D-dimensional radial polynomial anharmonic oscillator. First, we assume the exponential representation for the Approximant in coordinates (ρ, r), ψ(t)(ρ, r) = e−Φt(ρ,r) . (12.0.1) Therefore, we can focus only on the construction of Φt(ρ, r). According to the prescription, the approximate phase has to interpolate the expansions of small and large distances, i.e., the expansions (11.1.9), (11.2.22), and (11.2.23). Following the reflection symmetry of the system ρ → −ρ for fixed r, Φt(ρ, r) should be a function of ρ2. One of the simplest interpolations, which accomplishes the prescription given above, is of the form Φt(ρ, r) = α0 + α1 r + α3 r 2 + a3 γ ρ 2 + α4 γ ρ 2 r √ 1 + β1 r + β2 r2 + β3 ρ2 + log(1 + β1 r + β3 r 2 + β4 ρ 2) (12.0.2) where {α0, α1, α2, α3, α4, β1, β3, β4} are 8 free real parameters that later will be fixed via the VM. Note that this formula is analogous to (6.1.2). The appearance of a square root in the denominator is motivated by the generating function φ0, see (11.2.3). The logarithmic term is included to mimic the appearance of logarithmic terms in the exact wave function, see (11.2.4). It will generate a prefactor in the approximate wave function, which ultimately is given by 101 102 Chapter 12. The Approximant ψ (t) 1s0 (ρ, r) = 1 1 + β1 r + β2 r2 + β3 ρ2 exp ( −α0 + α1 r + α2 r 2 + α3 γ ρ 2 + α4 γ ρ 2 r √ 1 + β1 r + β2 r2 + β3 ρ2 ) . (12.0.3) This is the expression of the trial function that is going to be used in numerical calculations. We have labeled this Approximant with 1s0, indicating the usual notation for the ground state hydrogen atom in the absence of a magnetic field. As was discussed in Chapter 1, the variational method allows us to fix the value of the free parameters. We take the Approximant (12.0.3) and compute its associated variational energy Evar: the parameter-dependent expectation value of the Hamiltonian (10.0.4). The variational principle guarantees that the variational energy is an upper bound of the exact energy, Evar ≥ Eexact. Then, we minimize with respect to the free parameters to obtain their optimal configuration that provides the lowest upper bound. 12.1 Results 12.1.1 Energy From now on we set c = ~ = e = me = 1 in numerical computations, see (10.2.4). It is not a surprise that a numerical integration and minimization are needed to calculate and minimize the variational energy associated with (12.0.3). A code in FORTRAN 77 was written for this purpose. It uses the integration routine D01FCF from NAG-LIB as well as the minimization routine MINUIT of CERN-LIB. Further details of the computational code can be found in [54], [55], and [56]. Variational estimates of the energy are presented in Table 12.1 for some representative magnetic fields in the range γ ∈ [0.01, 10 000] a.u. Plots of optimal variational parameters as functions of log(1 + γ2) are shown in Appendix I, see Fig. C.1. In general terms, they are smooth changing parameters in the whole domain of magnetic fields studied. Our variational calculations are compared with some benchmarks coming from literature. However, for some magnetic fields, no estimate of the energy was found. Hence, to analyze the accuracy of our results, we carried out some numerical calculations via the Lagrange- Mesh method in the formulation presented in [58] using spherical coordinates {r, θ, ϕ}. The complete description and application of this method to the present system - hydrogen atom in a constant magnetic field - was already presented, see Section 2.5 in [75]. We implemented the Lagrange-Mesh method in a notebook of Mathematica. For the whole range of magnetic fields studied, the mesh was kept unchanged, it consisted of Nr = 80 radial functions and 12.1. Results 103 Table 12.1: Energy E and quadrupole moment Qzz of the ground state 1s0 of the hydrogen atom in constant magnetic field calculated in Variational Method (results without superscript) for γ ∈ [0.01, 10 000] with trial function (12.0.3). Benchmarks are included. Energy presented in Ry. γ E −Qzz γ E −Qzz 0.01 -0.999 950 005 51 0.000 248 5.0 2.239 209 0.506 493 -0.999 950 005 52a,e 0.000 249e 2.239 202a,e 0.506 331e 0.1 -0.995 052 960 5 0.023 270 10.0 6.504 427 0.445 22 -0.995 052 960 8a,e 0.023 2712e 6.504 405a,d 0.445 09d 0.5 -0.894 421 065 0.256 143 100.0 92.420 7 0.217 5 -0.894 421 075a,e 0.256 156 21e 92.420 3a,d 0.216 8d 1.0 -0.662 337 66 0.417 618 500.0 487.487 0.125 1 -0.662 337 79a,d 0.417 654d 487.485a,e 0.123 8e 2.0 -0.044 426 7 0.511 354 1 000.0 984.678 0.099 4 -0.044 427 8a,e 0.511 432e 984.675a,d 0.098 2d γ0 0.000 001 0.513 561 10 000.0 9 971.74 0.049 3 0.000 000e 0.513 537e 9 971.72e,f 0.047 9e γ0 = 2.065 211 858, see (12.1.2). a Power series [74], d Lagrange Mesh [75], e Lagrange Mesh (present work), f Basis of Splines [76]. Nu = 200 angular ones1. Hence the approximate ground state function is an expansion in terms of 80 × 200 = 16 000 functions. With this enormous mesh, we reproduced the results obtained for all magnetic fields studied in [75]: 1, 10, 100, and 1 000 a.u. It seems that the maximum accuracy in energy - 19 exact s. d. - is obtained for magnetic fields γ . 1 a.u. In general terms, for γ . 1 000 a.u. the method reaches in energy an accuracy of not less than 10 exact s. d. The accuracy decays dramatically once we study fields beyond γ ≈ 1 000 a.u. For example, it reaches only 6 exact digits at γ = 10 000 a.u. Using this method, we have confirmed and extended the results shown in [75]. Making a comparison between our results with benchmarks, we concluded the relative de- viation of the variational estimate of the energy to the exact one is very small in the whole domain of considered magnetic fields, ∣ ∣ ∣ ∣ Evar − Eexact Eexact ∣ ∣ ∣ ∣ . 10−6 , γ ∈ [0.01, 10 000] . (12.1.1) To the best of the author’s knowledge, there is no other variational calculation with a compact 1 u = cos θ. 104 Chapter 12. The Approximant trial function that provides such an accurate estimate of the energy. This may be seen from Table 12.2, where we show a collection of estimates of the ground state energy at γ = 1 a.u. In this Table, we have introduced the result obtained from the Padé-Borel summation applied to the first 100 coefficients in the perturbation expansion (11.1.1) of ε. This result provides 11 exact s.d. for the energy at γ ≤ 1a.u. However, the accuracy of the energy calculated by this procedure decays dramatically for γ & 2 a.u. Table 12.2: The ground state energy E1s0 in Ry of the hydrogen atom in constant magnetic field γ = 1. The results ranked by accuracy. Infinite massive proton assumed. Reference E1s0 Method [77] Wallis & Bowlden, 1958 -0.346 Variational Method [78] Yafet et al., 1956 -0.523 Variational Method [79] Galindo & Pascual, 1976 -0.690 Rational Approximants [80] Turbiner, 1984 -0.61 Variational Method [81] Potekhin & Turbiner, 2001 -0.63 Adiabatic Approximation [82] Kobori & Ohyama, 1997 -0.654 Variational Method [83] Pokatilov & Rusanov, 1969 -0.659 1 Variational Method [84] Larsen, 1968 -0.661 Variational Method [85] Lee et al., 1973 -0.661 6 Variational Method [86] Makado & McGill, 1986 -0.662 2 Variational Method [87] Cabib et al., 1971 -0.662 4 Variational Method [88] Arteca et al., 1983 -0.662 7 Regularization of PT [89] Chen & Goldman, 1992 -0.662 35 Expansion of the Wavefunction [90] Praddaude, 1972 -0.662 33 Power Series [81] Potekhin & Turbiner, 2001 -0.662 332 Variational Method [91] Vieyra & Olivares, 2008 -0.662 332 Variational Method [92] Rösner et al., 1984 -0.662 338 Multi-Configuration Hartree-Fock [93] Ivanov, 1988 -0.662 337 4 Finite Difference Method Present Work, 2020 -0.662 337 7 Variational Method [94] Xi et al., 1992 -0.662 337 78 Spline-basis [76] Wang & Hsue, 1995 -0.662 337 79 Spline-basis [73] Le Guillou & Zinn-Justin, 1983 -0.662 337 79 Borel-Leroy Summation [95] Wunner & Ruder, 1982 -0.662 337 79 Expansion of the Wavefunction [96] Fonte et al., 1990 -0.662 337 793 Variational Rayleigh-Ritz Present Work, 2020 -0.662 337 793 46 Padé-Borel Summation [97] Stubbins et al., 2004 -0.662 337 793 466 Variational Rayleigh-Ritz [74] Kravchenko et al., 1996 -0.662 337 793 466 Method of Moments [75] Baye et al., 2008 -0.662 337 793 466 315 9 Lagrange Mesh Present Work, 2020 -0.662 337 793 466 316 071 2 Lagrange Mesh [97] Stubbins et al., 2004 -0.662 337 793 466 316 071 220 596 ... Multiconfiguration [98] Nakashima & Nakatsuji, 2010 -0.662 337 793 466 316 071 220 596 ... Free Complement We also paid attention to the critical magnetic field γ0, for which the ground state energy 12.1. Results 105 vanishes2, E(γ0) = 0 . (12.1.2) The value of γ0 in the Lagrange-Mesh method, results in γ0 = 2.065 211 858 a.u. , (12.1.3) with E ∼ 10−10Ry. For this magnetic field, the variational trial function (12.0.3) provides E ∼ 10−6Ry. The local deviation between the Approximant and the exact wave function can be estimated by studying the vicinity of the Coulomb singularity3 via the cusp parameter (11.1.12). A straightforward calculation shows that the cusp parameter C(t), derived from to the Approxi- mant (12.0.3), is given by a simple formula C(t) = α1 + ( 1− α0 2 ) β1 , (12.1.4) see (11.1.12). The results for C(t) are presented in Table 12.3. It can be seen that C(t), calculated with optimal parameters as entry, satisfies the cusp condition accurately for γ . 500 a.u. with an error of . 16%. For the largest magnetic field studied, γ = 10 000 a.u. the deviation increases up to 70%. The cusp parameter C was also estimated in the Lagrange Mesh Method. In general, this calculation provides a parameter C with not less than 6 exact s. d. in the whole range of magnetic fields considered. Table 12.3: Estimate of the cusp parameter for the ground state 1s0 coming from the Approx- imant (12.0.3). Only s. d. presented. Atomic units used. γ C(t) γ C(t) 0.01 1.000 002 5.0 0.997 0.1 0.999 97 10.0 1.002 0.5 0.999 7 100.0 1.065 1.0 0.999 3 500.0 1.159 2.0 0.996 5 1 000.0 1.23 γ0 0.996 7 10 000.0 1.7 12.1.2 Quadrupole Moment It is well-known that the hydrogen atom in its ground state acquires a quadrupole moment at γ > 0, see [81]. Based on the cylindrical symmetry of the system, it can be easily shown that 2Note that γ0 defines a normalizable zero mode of the Hamiltonian (10.0.4). 3Located at r = 0, see (10.0.5). 106 Chapter 12. The Approximant 0 3 6 9 12 15 18 0.1 0.2 0.3 0.4 0.5 0.6 0 3 6 9 12 15 18 0.5 1 1.5 2 Figure 12.1: Left: Plot of the absolute value of the quadrupole moment (−Qzz) in (a.u)2 as a function log(1+ γ2). Variational estimate (see 12.0.3) of the quadrupole moment as well as the one coming from Padé-Borel summation presented. In the Lagrange-Mesh method, the mini- mum of the quadrupole moment −Qzz = 0.524 52 (a.u)2 is reached at γ ≅ 2.968 69a.u. Right: Expectation values 〈z2〉 and 〈ρ2〉 calculated by means of Approximant (12.0.3) as functions of log(1 + γ2). The intersection of the curves occurs at γ ≈ 2.806a.u.; at this magnetic field 〈z2〉 = 〈ρ2〉 ≈ 0.524. For γ & 700a.u. the value of 〈z2〉 is at least two orders of magnitude larger than 〈ρ2〉. Atomic units used. the quadrupole moment tensor is diagonal and only contains one single independent element, say Qzz = 〈ρ2〉 − 2 〈z2〉 . (12.1.5) Once optimal parameters are found for the Approximant (12.0.3), the Qzz component can be estimated for different values of γ using the square of the Approximant as a compact analytical expression for the electronic charge distribution. Plots of Qzz and expectations values, 〈z2〉 and 〈ρ2〉, as functions of γ are presented in Fig. 12.1. The numerical results are shown in Table 12.1. In the literature, there are not many accurate benchmarks for Qzz. Hence, in order to estimate the deviation of Qzz (calculated via the Approximant (12.0.3)) with respect to the exact value, we use the Lagrange-mesh method to obtain highly accurate estimates of Qzz that can be regarded as the exact values, see Table 12.1. These estimates indicate that for small magnetic fields γ . 2, we reproduce not less than 4 exact s.d. In turn, for γ & 2 we reproduce exactly 2-3 s.d. In addition to those calculations, the quadrupole moment is estimated using the PT calculated for Φn, see (11.1.1). In this manner, we can construct the expansion of the quadrupole moment in powers of γ2, Qzz(γ) = ∞∑ n=0 Q(n) zz γ 2n . (12.1.6) 12.1. Results 107 Up to O(γ8) this expansion4 looks like Qzz(γ) = −5 2 γ2 + 615 32 γ4 − 2564987 11520 γ6 + 19550772149 5529600 γ8 + . . . . (12.1.7) Due to the polynomial nature of corrections Φn with rational coefficients, see (11.1.6), correc- tions Q (n) zz in the expansion (12.1.6) will be rational numbers as well. The calculation of the first Q (n) zz indicates that series (12.1.6) are alternating with very fast growing coefficients as n grows. It is an indication that (12.1.6) could be a divergent series. The largest coefficient that we calculated was Q (26) zz whose (rounded) value is Q(26) zz = −8.188× 1047 . (12.1.8) Making a comparison of the first 26 coefficients Q (n) zz in the expansion (12.1.6) with those εn in (11.1.1), one can notice that the series for the quadrupole moment diverges faster than for the energy. With a motivation coming from the results obtained for the energy, we tested the applicability of the Padé-Borel summation to the first 26 terms of the series (12.1.6), assuming the same growth of Q (n) zz as in (11.1.8) for εn. In general, the Padé-Borel summation provides accurate results for γ . 2 a.u: it confirms not less than 3 correct s. d. of Qzz calculated either by means of the Approximant or the Lagrange-Mesh method. The accuracy of the Padé-Borel summation decays dramatically for magnetic fields larger than γ = 2 a.u. The plot of the Qzz as a function of γ, calculated in this scheme, is presented in Fig. 12.1. 4First ten coefficients are presented explicitly in Appendix, see F.3. Chapter 13 Conclusions Following the formalism presented for the D-dimensional radial anharmonic oscillator in Part I, a compact wave function Approximant for the ground state of a hydrogen atom in a constant magnetic field was constructed. PT, asymptotic expansions at small and large distances, the choice of coordinates (ρ, r) to work with, as well as the RB and GB1 equations, were basic ingredients to construct the Approximant. The main result of the study is summarized in the following expression for the Approximant ψ (t) 1s0 (ρ, r) = 1 1 + β1 r + β2 r2 + β3 ρ2 exp ( −α0 + α1 r + α2 r 2 + α3 γ ρ 2 + α4 γ ρ 2 r √ 1 + β1 r + β2 r2 + β3 ρ2 ) , where {α0, α1, α2, α3, α4, β1, β3, β4} are 8 free parameters. To fix their value, the Variational Method using the Approximant ψ (t) 1s0 as trial function was employed. It led to a variational estimate of the ground state energy with a very small relative deviation, of order ∼ 10−6 in the whole studied range of magnetic fields γ ∈ [0.01, 10 000] a.u. These results reproduced with high accuracy the benchmark energies (some of them are established by this work in the Lagrange-Mesh method). As an additional calculation, the quadrupole moment was computed for the same range using the square of the Approximant as a compact expression of the electronic density. Results coincide with benchmarks (coming from the Lagrange-Mesh method) in not less than 3 s. d. for γ . 100 a.u. It was explicitly demonstrated that the cusp condition of the exact hydrogen atom holds even in the presence of a constant magnetic field. Our Approximant provides a cusp parameter with a very small deviation from the exact value. For the weak and intermediate magnetic fields γ . 500 a.u., its relative deviation is about 15%. Due to the success of the Approximant when describing the ground state, it is clear that it can be modified to study excited states. Multiplying the Approximant by a suitable polynomial, 1Which leads to the true semi-classical expansion. 109 110 Chapter 13. Conclusions some particular families of excited states can be studied, just like it was done in [91]. For example, z ψ (t) 1s0 (ρ, r) where z = √ r2 − ρ2, is a suitable Approximant to describe the excited state 2p0. However, it should be mentioned that the general structure of the nodal surfaces, in the form of a prefactor, is still unknown. Due to the algebraic nature of PT in the framework of the Non-Linearization Procedure, the first 100 energy corrections (in powers of the effective magnetic field) with absolute accuracy were obtained: corrections are presented for the first time as rational numbers. Using these corrections, it was summed up the (divergent) PT for the energy via the Padé-Borel summation technique. It led to highly accurate results (not less than 11 s.d.) for small magnetic field γ . 1 a.u. Further Studies • PT can be constructed in the framework of the Non-Linearization Procedure to estimate the accuracy of the variational energy and wave function. In this context, the Approx- imant plays the role of the unperturbed wave function. This approach was successfully applied to the D-dimensional radial anharmonic oscillator: corrections in PT of the vari- ational energy were computed. As a result, its deviation with respect to the exact value was estimated, see Eq. (7.2.27). One could try to follow the same procedure for the hydrogen atom in a magnetic field by considering corrections to the approximate phase and the energy, Φexact = Φ (0) t + Φ (1) t + Φ (2) t + . . . , Eexact = Evar + E (t) 2 + E (t) 3 + . . . where Φ (0) t ≡ Φt, see Eq. (12.0.2). In the Non-Linearization procedure, it can be eas- ily shown that the first order correction to the phase, denoted by Φ (1) t , obeys a two- dimensional linear partial differential equation (PDE), namely ∆Φ (1) t − 2∇Φ (0) t · ∇Φ (1) t = Evar − V (1) . Here we have denoted V (1) = V − V0, where V = −1 r + γ2 8 ρ2 , V0 ≡ ∆Ψ(t) Ψ(t) . If the solution for ∇Φ (1) t is found, one could calculate the first correction to Evar using that E (t) 2 = − ∫ ( ∇Φ (1) t )2 ψ(t)ψ(t) dV ∫ ψ(t)ψ(t) dV , see [8]. If V (1) is subdominant in comparison to V , i.e. the ratio of V (1) to V is bounded at large distances, then PT developed in the Non-Linearization Procedure will be convergent. 111 • We did not take into account the information coming from the strong coupling regime (γ → ∞) in the design of the Approximant. For the ground state energy, the first two terms of the strong coupling expansion are well-known, E = γ − log2 γ + . . . , see [99]. It would be interesting to use this expansion in the RB and GB equations and study the strong coupling regime for the phase. This piece of missing information could lead to an improvement of the Approximant.2 • In this work, an infinitely massive and static proton was assumed. In the Schrödinger equation, finite mass effects can be incorporated making a pseudoseparation of variables of the center of mass and the relative coordinates, see [100] for details. Naturally, taking into account the finite mass of the proton, the spectrum of the hydrogen atom in a constant magnetic field will be modified. It turns out that when the (finitely) massive proton is fixed in the space (hydrogen atom at rest), the Schrödinger equation that describes this situation looks very similar to Eq. (10.0.1). The only difference is that the reduced mass of the hydrogen atom replaces the electron mass me in Eq. (10.0.1), see [101]. As a consequence, the Approximant that we have presented in this Part is also appropriate to describe the ground state of the finite massive static hydrogen atom in a constant magnetic field, see [11]. The study of finite mass effects was beyond of the purpose of the present Thesis, it will be done elsewhere. 2For example, an extra parameter, say q, may be inserted in the form of a degree of the prefactor of the Approximant. In this manner ψ (t) 1s0 (ρ, r) = 1 (1 + β1 r + β2 r2 + β3 ρ2)q exp ( −α0 + α1 r + α2 r 2 + α3 γ ρ 2 + α4 γ ρ 2 r √ 1 + β1 r + β2 r2 + β3 ρ2 ) . The insertion of this extra parameter surely will lead to higher accuracy for both the energy and the cusp parameter. General Conclusions 113 115 • Part I For the cubic, quartic, and sextic radial anharmonic oscillator potentials, whose cor- responding Schrödinger equations do not allow finding exact solutions, highly accurate approximate eigenfunctions of the low-lying states are written for arbitrary coupling con- stant and space dimension. Associated energies lead to unprecedented relative accuracies of (at least) order . 10−5 for any coupling constant in any dimension. These energies are in complete agreement with the highly accurate results of the Lagrange-Mesh calcu- lations, performed in the present thesis. Those results can be generalized for the radial anharmonic oscillator with arbitrary anharmonicity. It manifests a solution of the spectral problem of the radial anharmonic oscillator. • Part II A locally accurate ground state wave function (the Approximant) for the hydrogen atom subjected to a constant uniform magnetic field is found. This approximate function leads to high accuracy for the ground state energy (with relative accuracy . 10−5), as well as for the quadrupole moment. In the whole domain of accessible magnetic fields in Nature, the obtained energies are in complete agreement with the highly accurate ones provided by the Lagrange-Mesh method and the Padé-Borel summation3. In a non-relativistic consideration with a static proton, the Approximant can be regarded as the solution of the Schrödinger equation, which does not admit exact solutions. 3The domain of applicability of the Padé-Borel sum occurs for weak and intermediate fields. Appendices 117 Appendix A The WKB (Semi-Classical) Approximation We present an extension of the WKBmethod that can be found in elementary books on quantum mechanics, e.g. [1,2]. In particular, we focus the discussion on the radial Schrödinger equation in arbitrary dimension D = 1, 2, ..., which describes an S-state1 of an arbitrary radial potential − ~2 2M ( ∂2r + D − 1 r ∂r ) Ψ + V (r)Ψ = EΨ . (A.0.1) If we assume that the potential fulfills (i) V (r) ≥ 0 (ii) V (0) = 0 (iii) V (r → ∞) → +∞ we can guarantee the existence of bound states. We adopt the exponential representation for the wave function to find them, Ψ(r) = eiS(r)/~ . (A.0.2) The reason why we include ~ explicitly will be clear at the end of this Appendix. For now, it is clear that the physical units of S are those of ~, i.e. units of action. In this representation (A.0.1) becomes a Riccati-type equation i~S ′′(r)− S ′(r) ( S ′(r)− i~ (D − 1) r ) = 2M(V (r)− E) (A.0.3) cf. (3.3.4). For our purposes, it is enough to study this equation in the so-called non-classical domain in r (which is characterized by the inequality V (r) > E) beyond the largest turning 1Characterized by zero angular momentum ℓ = 0. 119 120 Appendix A. The WKB (Semi-Classical) Approximation point. To solve (A.0.3) in PT we consider the expansion S(r) = ∞∑ n=0 Sn(r)~ n (A.0.4) keeping E fixed. It is straight forward to show that equations S ′ 0 = ±i √ 2M(V − E) , (A.0.5) and S ′ 1 = i(D − 1) 2r + i 2 S ′′ 0 S0 (A.0.6) determine the first 2 corrections. In general, S ′ n is given by the expression S ′ n = iS ′′ n−1 + i(D−1) 2r S ′ n−1 − ∑n−1 k=1 S ′ kS ′ n−k 2S ′ 0 , (A.0.7) as we can see, the calculation of S ′ n is an algebraic procedure. To guarantee the normalizability of the wave function (A.0.2) we need to take the positive sign in (A.0.5). Thus, it will decay exponentially at large distances. Taking this into account, it is clear that the wave function is of the form Ψ(r) = 1 r D−1 2 (2M(V (r)− E)1/4) exp ( −1 ~ ∫ r√ 2M(V (r)− E) dr + O(~) ) , (A.0.8) as long as r → ∞. Some remarks should be pointed out: (i) Contrary to the function S(r), the energy E is considered fixed as a function of ~. Of course, this is not always true. For most of the potentials, the energy E should be ~-dependent. (ii) The integrand that occurs in (A.0.8) has a physical meaning, it is (almost) the classical action presented in the form ∼ ∫ p dq. Further properties and results can be obtained from the WKB approximation. To do so, it is necessary to study the behavior of the wave function in the classical domain(s) and then match the solutions for each region (the non-classical and classical one). Some of those results are well known, for example, the celebrated Bohr-Sommerfeld from early quantum mechanics. A review of those properties is found in [1]. Appendix B First PT Corrections and Generating Functions B.1 Cubic Case The expansions in powers of v for the first three corrections Yn(v), n = 1, 2, 3 for the cubic anharmonic oscillator potential (7.2.1), see (4.1.15) and (4.1.18) are presented. • At v → 0 Y1(v) = ǫ1v + 2 ǫ1 D + 2 v3 − 1 D + 3 v4 + ǫ1 (D + 2)(D + 4) v5 + . . . , Y2(v) = ǫ2 v + ǫ21 + 2 ǫ2 D + 2 v3 + 6 ǫ21 + 4 ǫ2 (D + 2)(D + 4) v5 − 2 ǫ1 (D + 3)(D + 5) v6 + . . . , Y3(v) = ǫ3 v + 2(ǫ1 ǫ2 + ǫ3) D + 2 v3 + 2(ǫ31 + 6 ǫ1 ǫ2 + 2 ǫ3) (D + 2)(D + 4) v5 − 2 ǫ2 (D + 3)(D + 5) v6 + . . . . (B.1.1) In these formulas we have denoted ǫn = εn D . • At v → ∞ Y1(v) = 1 2 v2 + 1 4 (D + 1) − ε1 2 v−1 + 1 8 (D2 − 1)v−2 + . . . , Y2(v) = − 1 8 v3 − 1 16 (3D + 4)v + ε1 4 − 1 32 (6D2 + 6D + 16 ε2 − 1)v−1 + 1 8 (3D − 2) ε1v −2 + . . . , Y3(v) = 1 16 v4 + 1 32 (5D + 8)v2 − 3 ε1 16 v + 1 64 (15D2 + 26D + 16 ε2 + 10) − 1 32 (15D ε1 + 16 ε3)v −1 + . . . . (B.1.2) A straightforward application of the formulas (4.2.3) and (4.4.2) allows us to calculate the 121 122 Appendix B. First PT Corrections and Generating Functions generating functions. In particular, G3 and G4 are given by G3 = − (2M)1/2 g2 ( ε1 2 log [ w − 1 w + 1 ]) , G4 = − (2M)1/2 g2 ( 5 + (5 + 12D)w + (1− 6D(D + 1))(w + 1)w2 (3w2 − 2) 48(w − 1)(w + 1)2w3 (1− 16 ε2 − 6D(D + 1)) 32 log [ w − 1 w + 1 ]) , (B.1.3) where w = (1 + gr)1/2. Explicit formulas for the first corrections εn in PT at D = 1 have been presented in (7.2.5). B.2 Quartic Case We present the corrections ε4,6 and Y4,6 in explicit form for the quartic anharmonic potential (7.3.1) in the expansion (7.3.6) and (7.3.7), ǫ2 = 1 4 (D + 2) , ǫ4 = − 1 16 (D + 2)(2D + 5) , ǫ6 = 1 64 (D + 2)(8D2 + 43D + 60) , (B.2.1) Y2(v) = 1 2 v3 + ǫ2v , Y4(v) = −1 8 v5 − 1 16 (3D + 8)v3 + ǫ4v , Y6(v) = 1 16 v7 + 1 32 (5D + 16)v5 + 1 16 (3D2 + 17D + 25)v3 + ǫ6v , (B.2.2) with ǫ2n = ε2n D . We also present two generating functions for the phase, see expansion (4.4.1), G4 = 5 24g2w3 + D (1 + w + w2) 4g2(w + 1)w2 + D2 8g2w , G6 = − 5 16g2w6 − D (15 + 30w + 20w2 + 16w3 + 20w4 + 30w5 + 15w6) 48g2(w + 1)2w5 − D2 (4 + 8w + 8w2 + 12w3 + 18w4 + 9w5) 32g2(w + 1)2w4 − D3 (1 + 3w2) 48g2w3 , (B.2.3) B.3. Sextic Case 123 where w = √ 1 + g2v2. Two remarks in a row: (i) in the variable w all generating functions are rational functions, (ii) the polynomial structure in variable D of generating functions becomes evident. B.3 Sextic Case We present explicitly the first corrections ε8,12 and Y8,12 for the sextic anharmonic oscillator (7.4.1). See (7.4.3) and (7.4.4). ǫ4 = 1 8 (D + 2)(D + 4) , ǫ8 = − 1 128 (D + 2)(D + 4)(9D2 + 72D + 152) , ǫ12 = 1 1024 (D + 2)(D + 4)(81D4 + 1404D3 + 9624D2 + 31152D + 40384) . (B.3.1) Y4(v) = 1 2 v5 + 1 4 (D + 4)v3 + ǫ4v , Y8(v) = − 1 8 v9 − 1 16 (3D + 16)v7 − 1 16 (3D2 + 27D + 64)v5 − 1 32 (4D3 + 49D2 + 204D + 288)v3 + ǫ8v , Y12(v) = 1 16 v13 + 1 32 (5D + 32)v11 + 5 64 (3D2 + 34D + 104)v9 + 1 32 (8D3 + 125D2 + 688D + 1344)v7 + 1 256 (55D4 + 1038D3 + 7708D2 + 26784D + 36800)v5 + 1 256 (36D5 + 783D4 + 7040D3 + 32768D2 + 78912D + 78336)v3 + ǫ12v, (B.3.2) where it has been defined ǫ4n = ε4n D . Now we present explicitly two generating functions G4, G6 in the expansion (4.4.1), G4 = r2 (5 + w2) 12w3 + Dr2 (1 + w + w2) 4w2(w + 1) + D2r2 16w , G6 = 5− 3w2 4g2w6 + D (15 + 15w − 4w2 + 2w3 + 6w4 + 6w5) 24g2(w + 1)w5 + D2 (2 + 2w + w2 + 3w3 + 3w4) 16g2(w + 1)w4 + D3 (1 + 3w2) 96g2w3 , (B.3.3) 124 Appendix B. First PT Corrections and Generating Functions where w = √ 1 + g4r4. Two remarks: i) in the variable w, generating functions are rewritten as rational ones; ii) The polynomial structure in variable D of these generating functions is evident. Appendix C Plots of Variational Parameters: Anharmonic Oscillator In this Appendix, we present the plots of the optimal parameters in the Variational Method for the Approximants of the ground state of the cubic, quartic, and sextic anharmonic oscillator. The number of equation of each Approximant is indicated in the title of each Section. C.1 Cubic: (7.2.24) 2 4 6 8 10 g -10 -5 0 5 10 0 0.4 0.8 0.45 0.55 0.65 2 4 6 8 10 g 0.45 0.5 0.55 0.6 0.65 0 2 4 6 8 10 g 0.2 0.4 0.6 0.8 Figure C.1: Variational parameters ã0, ã2 and ã3 of the ground state as functions of g for D = 2, 3, 6. 125 126 Appendix C. Plots of Variational Parameters: Anharmonic Oscillator C.2 Quartic: (7.3.26) ✷ ✹ ✻ ✽ ✶ ❣ ✁ -✶ ✶ ✷ ✸ ❛ ✂ ✵ (a) ✵ ✵✁ ✵✂ ✵✄ ✵☎ ✵✆ ✵✆✆ ✵✝ ✞ ✷ ✹ ✻ ✽ ✶✞ ❣ ✟ ✞✠✡ ✞✠✻ ✞✠☛ ✞✠✽ ✞✠☞ ✶✠ ❛ ✌ ✟ (b) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣ ✵✁✵✂ ✵✁✶ ✵✁✶✂ ✵✁✷ ✵✁✷✂ ✵✁✄ ✵✁✄✂ ❛ ☎ ✆ (c) Figure C.2: Ground state: Variational parameters ã0 (a), ã2 (b) and ã4 (c) vs the coupling constant g for D = 1, 2, 3, 6. Parameters (7.3.28), which allow us to reproduce the first two terms G0, G2 in the expansion (4.4.1) shown by solid (black) line which is horizontal for ã2 (b) and ã4 (c). C.2. Quartic: (7.3.26) 127 ✷ ✸ ✹ ✺ ✻ ❉ -✶ ✵ ✶ ✷ ✸ ❛ ✁ (a) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✺ ✵✻ ✵✁ ✵✂ ✵✄ ❛ ☎ ✆ (b) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✵✺ ✵✶ ✵✶✺ ✵✷ ✵✷✺ ✵✸ ✵✸✺ ✵✹ ❛ ✁ ✂ (c) Figure C.3: Variational parameters ã0(a), ã2(b), ã4(c) vs D for fixed g2 = 0.1, 1, 10. D- independent parameters (7.3.28), which allow us to reproduce the first two terms G0, G2 in the expansion (4.4.2), shown by solid (black) horizontal lines, see (7.3.28). In (a) the horizontal lines correspond to ã0 = 1 3g2 at g2 = 0.1(i), 1(ii), 10(iii). 128 Appendix C. Plots of Variational Parameters: Anharmonic Oscillator C.3 Sextic: (7.4.27) ✷ ✹ ✻ ✽ ✶ ❣✁ -✸ -✷ -✶ ❛ ✂ ✵ (a) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣✵✁✷ ✵✁✷✂ ✵✁✄ ✵✁✄✂ ✵✁✹ ❛ ☎ ✆ (b) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣✵✁✶ ✵✁✶✷ ✵✁✶✹ ✵✁✶✻ ✵✁✶✽ ❛ ✂ (c) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣✵✁✷✂ ✵✁✄ ✵✁✄✂ ✵✁✹ ✵✁✹✂ ✵✁✂ ❜ ☎ (d) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣ ✵✁✵✂ ✵✁✶ ✵✁✶✂ ✵✁✷ ✵✁✷✂ ❜ ✄ ☎ (e) ✵ ✷ ✹ ✻ ✽ ✶✵ ❣✵✁✶✂ ✵✁✷ ✵✁✷✂ ✵✁✄ ✵✁✄✂ ✵✁✹ ❝ ☎ ✆ (f) Figure C.4: Ground state (0, 0): Variational parameters ã0 (a), ã2 (b), ã4 (c), b̃4 (d), b̃6 (e), c̃2 (f) vs the coupling constant g4 in domain g4 ∈ [0, 10] for D = 1, 2, 3, 6. C.3. Sextic: (7.4.27) 129 ✷ ✸ ✹ ✺ ✻ ❉✵ -✶ -✷ -✸ ❛ ✁ (a) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✷ ✵✷✺ ✵✸ ✵✸✺ ✵✹ ❛ ✁ ✂ (b) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✶ ✵✶✷ ✵✶✹ ✵✶✻ ✵✶✁ ❛ ✂ ✄ (c) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✷✺ ✵✸ ✵✸✺ ✵✹ ✵✹✺ ✵✺ ❜ ✁ ✂ (d) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵ ✵✵✺ ✵✶ ✵✶✺ ✵✷ ✵✷✺ ❜ ✁ ✂ (e) ✶ ✷ ✸ ✹ ✺ ✻ ❉✵✶ ✵✷ ✵✸ ✵✹ ❝ ✁ ✂ (f) Figure C.5: Ground state (0, 0): Variational parameters ã0 (a), ã2 (b), ã4 (c), b̃4 (d), b̃6 (e), c̃2 (f) vs dimension D for g4 = 0.1, 1.0, 10.0. Appendix D General Aspects of the Two-Term Radial Anharmonic Oscillator Let us consider the general two-term radial anharmonic oscillator potential, V (r) = r2 + gm−2rm , m > 2 . (D.0.1) cf. (3.2.1), where a2 = am = 1 and a3, a4, ..., am−1 = 0. Expansion of ε in powers of λ has the form ε(λ) = ε0 + λm−2εm−2 + λ2(m−2)ε2(m−2) + ... , (D.0.2) see (4.1.5). Not surprisingly, for even potentials (m = 2p, p = 1, 2, ...) the PT coefficients ε2n(p−1), n = 1, 2, ..., are polynomials in D factorized as follows ε2n(p−1)(D) = D(D + 2)(D + 4) . . . (D + 2p− 2)R(n−1)(p−1)(D) (D.0.3) where R(n−1)(p−1)(D) is a polynomial inD of degree (n−1)(p−1), see [29]. We emphasize that, in the framework of the Non-Linearization Procedure, any PT coefficient ε2n(p−1)(D) is calculated by linear algebra means. Series (D.0.2) is divergent due to a Dyson instability argument. It is evident that for larger D and m, the index of divergence is larger, see e.g. [93]. For potential (D.0.1) generating functions G0(r) and G2(r) can be written in closed form, G0(r) (2M)1/2 = 2 m+ 2 r2 { √ 1 + (gr)m−2 + m− 2 4 2F1 ( 1 2 , 2 m− 2 ; m m− 2 ;−(gr)m−2 )} , (D.0.4) g2G2(r) (2M)1/2 = 1 4 log [ 1 + (gr)m−2 ] + D m+ 2 log [ 1 + √ 1 + (gr)m−2 ] , (D.0.5) where 2F1 is the hypergeometric function. 130 Appendix E System of Coordinates {ρ, r, ϕ} A general description of the non-orthogonal system of coordinates {ρ, r, ϕ} is presented in this Appendix. Here we show some relevant expressions for the differentiation of a given function. E.1 Definition The system of coordinates {ρ, r, ϕ}, shown schematically in Fig. 10.1, is defined by the change of variables ρ = √ x2 + y2 (E.1.1) r = √ x2 + y2 + z2 (E.1.2) ϕ =    arctan (y x ) x > 0 , y > 0 2π + arctan (y x ) x > 0, y < 0 π 2 sgn(y) x = 0 π + arctan (y x ) x < 0 (E.1.3) where {x, y, z} denote Cartesian coordinates. The new system of coordinates is defined in the domain 0 ≤ ρ ≤ r , 0 ≤ r <∞ , 0 ≤ ϕ ≤ 2π . (E.1.4) The change of variables presented in (E.1.1), (E.1.2) and (E.1.3) is unambiguously defined for the half-space z ≥ 0. 131 132 Appendix E. System of Coordinates {ρ, r, ϕ} Conversely, the Cartesian coordinates may be retrieved from (E.1.1), (E.1.2) and (E.1.3), x = ρ cosϕ , (E.1.5) y = ρ sinϕ , (E.1.6) z = √ r2 − ρ2 . (E.1.7) E.2 Differentiation The Non-Linearization Procedure requires some expressions that involve differential operators; we present them below. The basic expressions come from the partial derivatives, for which a straightforward calculation via the chain rule indicates that ∂x = cosϕ∂ρ + ρ cosϕ r ∂r − sinϕ ρ ∂ϕ , (E.2.1) ∂y = sinϕ∂ρ + ρ sinϕ r ∂r + cosϕ ρ ∂ϕ , (E.2.2) ∂z = √ r2 − ρ2 r ∂r . (E.2.3) Using these expressions, one can easily check that the Laplacian operator ∇2 takes the form ∇2 = ∂2ρ + 2ρ r ∂ρr + ∂2r + 1 ρ ∂ρ + 2 r ∂r + 1 ρ2 ∂2φ . (E.2.4) Another relevant quantity comes from the dot product of two gradients. Let us consider func- tions f and g written in variables (ρ, r, ϕ), explicitly f = f(ρ, r, ϕ) , g = g(ρ, r, ϕ) . (E.2.5) A direct calculation using (E.2.1), (E.2.2), and (E.2.3), leads to ∇f · ∇g = ∂ρf ∂ρg + ∂r f∂rg + ρ r (∂rf ∂ρg + ∂ρf ∂rg) + 1 ρ2 ∂φf ∂φg . (E.2.6) Appendix F Perturbation Theory for Hydrogen Atom in Constant Magnetic Field F.1 Computational Code: Non-Linearization Procedure Below it is presented in explicit form the Mathematica Notebook which realizes the Non- Linearization procedure for the hydrogen atom in presence of a uniform and constant magnetic field. The program is optimized to run it directly from the terminal/console. (*This program calculates PT series of the enery and phase \ of hydrogen atom in magnetic field in its the ground state \ using the Non -Linerization procedure in variables (rho=s,r) *) (* Load Module: module load Mathematica /11.2 *) (* Usage: math -noprompt -run < file.txt *) (*----------------------------------------------------*) (* Autor: Juan Carlos de Valle 2019( revisited) *) (*----------------------------------------------------*) (*----------------------------------------------------*) (* Order (Input Data) *) m = 100; (*----------------------------------------------------*) (* Initial Data *) f[0] = r; (*----------------------------------------------------*) (* Constraint *) Do[ Subscript[Overscript[a, n], 0, 2*n + 1] = 0, {n, 1, m} ]; (*----------------------------------------------------*) (* ******************************************** *) (* MAIN PROGRAM *) (* ******************************************** *) (* Definitions *) f[n_] := Expand[Sum[( Subscript[Overscript[A, n], k, j]*r \ + Subscript[Overscript[B, n], k, j])*s^(2*(n - k))*r^(2*\ (k - j)), {j, 0, n - 1}, {k, j, n}]]; Subscript[E, 0] = -1^2; Q[1] := (1/4)*s^2; Q[n_] := -Expand[Sum[D[f[k], s]*D[f[n - k], s] + D[f[k], r]*D[f[n - k], r] + (s/r)*(D[f[k], s]*D[\ f[n - k], r] + D[f[k], r]*D[f[n - k], s]), {k, 1, n - 1}]]; eq[n_] := Expand[D[f[n], s, s] + ((2*s)/r)*D[f[n], s, r] \ + D[f[n], r, r] + (1/s - (2*s)/r)*D[f[n], s] + 2*(1/r - 1)*D[f[n], r] - Subscript[E, n] + Q[n]\ ]; (*----------------------------------------------------*) (* OUTPUT *) (*----------------------------------------------------*) Energy ={}; (* Solving Algebraic Equations *) Do[ sol = SolveAlways[eq[i] == 0, {r, s}]; Apply[Set , sol[[1]], {1}]; PrintTemporary[i]; AppendTo[Energy ,{i,Subscript[E, i]}]; Export["Coefficients.dat", Energy], {i, 1, m} ]; 133 134 Appendix F. Perturbation Theory for Hydrogen Atom in Constant Magnetic Field F.2 Corrections in Perturbation Theory We present explicitly the first corrections φn, n = 2, 3, 4 in the expansion (11.1.1). Additionally, in Tables F.1 - F.2 high order corrections εn are shown. Finally, the first 10 coefficients in the expansion of the quadrupole moment, see (12.1.6), are displayed in Table F.3. −φ2(s, t) = 1 1152 s4t+ 1 1440 s2t3 + 11 4608 s4 + 13 1440 s2t2 + 1 2880 t4 + 193 5760 s2t+ 1 120 t3 + 193 3840 s2 + 337 5760 t2 (F.2.1) φ3(s, t) = 1 27648 s6t+ 1 11520 s4t3 + 1 60480 s2t5 + 7 55296 s6 + 163 138240 s4t2 + 131 241920 s2t4 + 1 181440 t6 + 61 11520 s4t+ 8063 1209600 s2t3 + 53 201600 t5 + 803 92160 s4 + 33311 806400 s2t2 + 2927 604800 t4 + 90877 691200 s2t+ 2027 43200 t3 + 90877 460800 s2 + 188173 691200 t2 (F.2.2) −φ4(s, t) = 5 2654208 s8t+ 110592 s6t3 + 163 29030400 s4t5 + 1 2419200 s2t7 + 163 21233664 s8 + 293 2211840 s6t2 + 9833 58060800 s4t4 + 727 29030400 s2t6 + 1 9676800 t8 + 8819 13271040 s6t+ 1663979 812851200 s4t3 + 24733 40642560 s2t5 + 167 20321280 t7 + 10577 8847360 s6 + 13945163 1083801600 s4t2 + 22721 2822400 s2t4 + 5989 22579200 t6 + 27927329 650280960 s4t+ 29335139 451584000 s2t3 + 4828099 1016064000 t5 + 816005783 13005619200 s4 + 1349713153 4064256000 s2t2 + 146213807 2709504000 t4 + 16222576613s2t 16257024000 + 141801871t3 338688000 + 16222576613s2 10838016000 + 36642046037t2 16257024000 (F.2.3) F.2. Corrections in Perturbation Theory 135 Table F.1: First ten coefficients εn (in rational form) of the series expansion for ε calculated in the Non-Linearization Procedure, see (11.1.1). n (−1)n+1εn 0 1 1 1/2 2 53/96 3 5581/2304 4 21577397/1105920 5 31283298283/132710400 6 13867513160861/3538944000 7 5337333446078164463/62426972160000 8 995860667291594211123017/419509252915200000 9 86629463423865975592742047423/1057163317346304000000 10 6127873544613551793091647103033033/1776034373141790720000000 Table F.2: Coefficients ε10n, n = 1, 2, ..., 10, of the series expansion for ε calculated in Non- Linearization Procedure, see (11.1.1). Coefficients rounded to 3 displayed d. d. n −ε10n n −ε10n 1 3.450× 109 6 5.655× 10140 2 2.160× 1029 7 1.410× 10173 3 3.215× 1053 8 6.046× 10206 4 3.720× 1080 9 3.127× 10241 5 6.263× 10109 10 1.479× 10277 136 Appendix F. Perturbation Theory for Hydrogen Atom in Constant Magnetic Field Table F.3: First ten coefficients Q (n) zz of the series expansion in powers of γ2 for Qzz, see (12.1.6). n (−1)n+1Q (n) zz 0 0 1 5/2 2 615/32 3 2564987/11520 4 19550772149/5529600 5 16195465214963/221184000 6 1664517172621593361/867041280000 7 3895122983853951241069/62426972160000 8 12095392486876132744589230813/4894274617344000000 9 4348730023196338444920601655427467/37000716107120640000000 10 646955302702442553065579733086802864719/97681890522798489600000000 Appendix G Computational Code: Hydrogen Atom in Magnetic Field In this Appendix, the computational Hinb.f which performs the variational method using (12.0.3) as the trial function is explicitly shown. It uses the integration routine D01FCF from NAG-LIB as well as the minimization routine MINUIT of CERN-LIB. The input is data (such as magnetic field, domain of integrations, etc.) is written in data sheet as shown below. Magnetic field (in units of B_0): H (change of vars in rho) 2.35D0 Number of regions in rho (nrho) nrho+1 limits (NUM / DEN) 2 2 Limits for rho NUM (according to Number of Regions) 0.0 1.24 7.86 Limits for rho DEN 0.0 1.24 7.86 Number of regions in z (nz) nz+1 limits (NUM / DEN) 3 3 Limits for z (NUM) (according to Number of Regions) 0.0D0 0.78 2.85 15.4 Limits for z (DEN) 0.0D0 0.78 2.85 15.4 Default Acc. NUM/DEN , Factor , Flag for corr., acc. for corr , screen =1/ file=0 1.D-10 1.D-10 1.0D0 1 1.0D-02 1 SET TITLE ’Variational Energy for H in a magnetic field ’ PARAMETERS 1 ’a0=’ 2.2425240 0.1 2.0 3.99 2 ’a1=’ 1.0104802 0.1 0.9 1.2 3 ’a2=’ -0.2722381 0.01 -0.43 0.51 4 ’a3=’ 0.0411529 0.01 0.0 0.4 5 ’a4=’ 0.0545618 0.01 -0.1 0.2 6 ’a5=’ 0.0369813 0.01 0.0 0.057 7 ’b1=’ 0.0268046 0.01 0.0 0.03 8 ’b2=’ 0.0466748 0.01 0.0 0.1 9 ’b3=’ 0.0000000 0.01 0.00 0.06 10 ’b4=’ 0.0223345 0.01 0.0 0.8 11 ’c =’ 0.0833333 0.00 0.0 0.24 12 ’d =’ 2.0000000 0.0 0.9 2.8 set strategy 2 minimize minimize 2000 stop The main body of the program is presented in explicit form in the following pages. 137 138 Appendix G. Computational Code: Hydrogen Atom in Magnetic Field c P r o g r a m H i n B . f c c C o m p i l e a s : g f o r t r a n - o H i n B . e H i n B . f c c o r u s e g 7 7 i n s t e a d o f g f o r t r a n c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c H i n a m a g n e t i c f i e l d c J u a n C a r l o s d e l V a l l e c N e w A n s a t z w i t h w - v a r i a b l e c c B o = 2 . 3 5 x 1 0 ^ 9 G ( A T O M I C U N I T O F M A G N E T I C F I E L D ) c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c U S A G E : . / H i n B . e < f i l e . d a t > f i l e . o u t & c c c - - - U s i n g M I N U I T f r o m C E R N - l i b a n d c c - - - D 0 1 F C F I N T E G R A T I O N S U B R O U T I N E N A G - l i b c c P R O G R A M M A I N I M P L I C I T D O U B L E P R E C I S I O N ( A - H , O - Z ) C O M M O N / N I N P U T 3 / n r h o N , n r h o D , n z N , n z D , i c o r r C O M M O N / X I N P U T 2 / x B , E P S i n , E P S i d , F C T , E P S C C O M M O N / X I N P U T 1 / R L N ( 5 ) , R L D ( 5 ) , Z L N ( 1 0 ) , Z L D ( 1 0 ) , E P N U M ( 2 0 ) $ , E P D E N ( 2 0 ) C O M M O N / M O N I T O R / e b e s t , i c a l l , i c a l l b E X T E R N A L F C N , f u t i l i c a l l = 0 i c a l l b = 0 e b e s t = 1 . D 1 0 c r e a d ( 5 , * ) ! M a g n e t i c F i e l d s t r e n g h t r e a d ( 5 , * ) B x ! - > x B ( i n p r o p e r u n i t s ) r e a d ( 5 , * ) r e a d ( 5 , * ) n r h o N , n r h o D ! N u m b e r o f r e g i o n s i n r h o ( N U M / D E N ) r e a d ( 5 , * ) r e a d ( 5 , * ) ( R L N ( i r ) , i r = 1 , n r h o N + 1 ) ! r h o N u m l i m i t s r e a d ( 5 , * ) r e a d ( 5 , * ) ( R L D ( i r ) , i r = 1 , n r h o D + 1 ) ! r h o D e n l i m i t s r e a d ( 5 , * ) r e a d ( 5 , * ) n z N , n z D ! N u m b e r o f r e g i o n s i n z ( N U M / D E N ) r e a d ( 5 , * ) r e a d ( 5 , * ) ( Z L N ( i r ) , i r = 1 , n z N + 1 ) ! z N u m l i m i t s r e a d ( 5 , * ) r e a d ( 5 , * ) ( Z L D ( i r ) , i r = 1 , n z D + 1 ) ! z D e n l i m i t s r e a d ( 5 , * ) r e a d ( 5 , * ) E P S i n , E P S i d , F C T , i c o r r , e p s c , n s c r e e n c u n i t s = 2 . 3 5 D 0 ! O u r u n i t s ( P r o p e r d e f i n i t i o n o f R y ) x B = B x / u n i t s ! U n i t s c o n v e r s i o n f o r B c ! B i n u n i t s o f B o = 1 / 2 . 3 5 c ! S e e c o m m o n D A T O S c i f ( n s c r e e n . e q . 0 ) t h e n o p e n ( u n i t = 6 , f i l e = ’ m i n u i t . o u t ’ , s t a t u s = ’ u n k n o w n ’ ) e n d i f c w r i t e ( 6 , * ) w r i t e ( 6 , * ) ’ * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * ’ w r i t e ( 6 , * ) ’ T h i s p r o g r a m m i n i m i z e s t h e v a r i a t i o n a l E n e r g y f o r ’ w r i t e ( 6 , * ) ’ H a t o m i n a m a g n e t i c f i e l d ’ w r i t e ( 6 , * ) w r i t e ( 6 , * ) ’ * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * ’ w r i t e ( 6 , * ) w r i t e ( 6 , * ) ’ M a g n e t i c F i e l d S t r e n g t h : ’ , x B , ’ a u ( ’ , B x , ’ B _ 9 ) ’ w r i t e ( 6 , * ) w r i t e ( 6 , * ) c w r i t e ( * , * ) ’ L I M I T S : ’ w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ w r i t e ( * , * ) w r i t e ( * , * ) ( R L N ( i r ) , i r = 1 , n r h o N + 1 ) ! r h o N u m l i m i t s w r i t e ( * , * ) w r i t e ( * , * ) ( R L D ( i r ) , i r = 1 , n r h o D + 1 ) ! r h o D e n l i m i t s w r i t e ( * , * ) w r i t e ( * , * ) ( Z L N ( i r ) , i r = 1 , n z N + 1 ) ! z N u m l i m i t s w r i t e ( * , * ) w r i t e ( * , * ) ( Z L D ( i r ) , i r = 1 , n z D + 1 ) ! z D e n l i m i t s w r i t e ( * , * ) w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ w r i t e ( * , * ) w r i t e ( * , * ) ’ E P S : ’ , E P S i n , E P S i d w r i t e ( * , * ) ’ F A C T O R : ’ , F C T w r i t e ( * , * ) c c " F L A T " D I S T R I B U T I O N ( S A M E A C C U R A C Y F O R A L L S U B I N T E R V A L S ) D O i n = 1 , 2 0 ! m a x r e g i o n s = 2 0 E P N U M ( i n ) = E P S I n E P D E N ( i n ) = E P S i d E N D D O c c 5 0 0 1 f o r m a t ( 1 x , 5 ( F 1 0 . 5 , 2 x ) ) 5 0 0 2 f o r m a t ( 1 x , 4 ( F 1 0 . 5 , 2 x ) ) 5 0 0 3 f o r m a t ( 1 x , 2 ( F 1 0 . 5 , 2 x ) ) c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c I n i t i a l i z i n g M i n u i t c E s t a b l i s h i n g u n i t n u m b e r s f o r d a t a i n p u t / o u t p u t c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - i r d = 5 ! F o r t r a n U n i t n u m b e r f o r r e a d i n g i w r = 6 ! F o r t r a n U n i t n u m b e r f o r w r i t i n g i s a v = 7 ! F o r t r a n U n i t n u m b e r f o r s a v i n g c a l l m n i n i t ( i r d , i w r , i s a v ) c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + c C a l l i n g t o M I N U I T C A L L M I N U I T ( F C N , f u t i l ) c s t o p e n d ! E N D M A I N P R O G R A M c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c s u b r o u t i n e f u t i l ( ) e n d c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c S U B R O U T I N E F C N c S U B R O U T I N E F C N ( N P A R , G R , F V A L , X V , I F L A G , F U T I L ) I M P L I C I T D O U B L E P R E C I S I O N ( A - H , O - Z ) c M I N U I T p a r a m e t e r s a n d c o m m o n d e f i n i t i o n s 139 D I M E N S I O N G R ( * ) , X V ( * ) c . . P a r a m e t e r s f o r t h e i n t e g r a t i o n r o u t i n e D 0 1 F C F . . . I N T E G E R N D I M , M A X P T S , L E N W R K P A R A M E T E R ( N D I M = 2 , M A X P T S = 1 0 0 0 0 0 0 0 0 , L E N W R K = 5 0 0 0 0 0 0 ) c . . L O C A L A R R A Y S . . . D O U B L E P R E C I S I O N A N U M ( N D I M ) , B N U M ( N D I M ) , A D E N ( N D I M ) , B D E N ( N D I M ) , $ W R K S T R ( L E N W R K ) , X N U ( 2 0 ) , X D E ( 2 0 ) c . . C O M M O N B L O C K F O R I N P U T D A T A . . . C O M M O N / N I N P U T 3 / n r h o N , n r h o D , n z N , n z D , i c o r r C O M M O N / X I N P U T 2 / x B , E P S i n , E P S i d , F C T , E P S C C O M M O N / X I N P U T 1 / R L N ( 5 ) , R L D ( 5 ) , Z L N ( 1 0 ) , Z L D ( 1 0 ) , E P N U M ( 2 0 ) $ , E P D E N ( 2 0 ) C O M M O N / M O N I T O R / e b e s t , i c a l l , i c a l l b c . . C O M M O N B L O C K F O R W A V E F U N C T I O N S P A R A M E T E R S . . . C O M M O N / P A R M / B , a 0 , a 1 , a 2 , a 3 , a 4 , a 5 , b 1 , b 2 , b 3 , b 4 , c , d C O M M O N / F L A G S / i q f l a g c . . C O M M O N B L O C K F O R A B B R E V I A T I O N S . . . C O M M O N / A B B R / B q 2 , S Q R B , F C T R c . . I N T R I N S I C F U N C T I O N S . . . I N T R I N S I C A B S c . . E X T E R N A L F U N C T I O N S . . . E X T E R N A L D 0 1 F C F , E V A R N , E V A R D c . . E X E C U T A B L E S T A T E M E N T S . . . c i c a l l = i c a l l + 1 F V A L = 0 . 0 D 0 N C E N = 1 c c U S E F U L A B B R E V I A T I O N S : ( C O M M O N ) c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - B = x B B q 2 = B * B S Q R B = D S Q R T ( B ) F C T R = F C T c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c R S = 1 . 0 D 0 ! N a t u r a l S i z e o f t h e S y s t e m s e t t o 1 ( H - a t o m ) c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c A S S I G N I N G M I N U I T V A R I A B L E S ( V A R I A T I O N A L P A R A M E T E R S ) c a 0 = X V ( 1 ) a 1 = X V ( 2 ) a 2 = X V ( 3 ) a 3 = X V ( 4 ) / B / B a 4 = X V ( 5 ) / B a 5 = X V ( 6 ) / B b 1 = X V ( 7 ) b 2 = X V ( 8 ) b 3 = X V ( 9 ) / B / B b 4 = X V ( 1 0 ) / B / B c = X V ( 1 1 ) d = X V ( 1 2 ) c I N S E R T C O S T R A I N T S H E R E c o n s t r a i n t = a l p h a + b e t a * B I F ( c o n s t r a i n t . L T . 0 . D 0 ) t h e n F V A L = X B r e t u r n E N D I F c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c V A R I A T I O N A L E N E R G Y N U M E R A T O R c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c i f ( i c o r r . e q . 1 ) t h e n w r i t e ( * , * ) w r i t e ( * , * ) w r i t e ( * , * ) w r i t e ( * , * ) ’ I R E G , X N U ( I R E G ) , E P S , A C C , I F A I L ’ w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ e n d i f c e n e r n u = 0 . 0 D 0 q u a d m = 0 . 0 D 0 I R E G = 0 D O I r h o = 1 , n r h o N D O I z = 1 , n z N I R E G = I R E G + 1 E P S = E P N U M ( I R E G ) c c N U M E R A T O R L I M I T S ( 1 ) = r h o , ( 2 ) = z A N U M ( 1 ) = R L N ( I r h o ) B N U M ( 1 ) = R L N ( I r h o + 1 ) A N U M ( 2 ) = Z L N ( I z ) * R S B N U M ( 2 ) = Z L N ( I z + 1 ) * R S c M I N P T S = 0 I F A I L = 1 F I N V N = 0 . 0 D 0 X N U ( I R E G ) = 0 . 0 D 0 c C A L L D 0 1 F C F ( N D I M , A N U M , B N U M , M I N P T S , M A X P T S , E V A R N , E P S , A C C , L E N W R K , + W R K S T R , F I N V N , I F A I L ) c X N U ( I R E G ) = F I N V N c i f ( i c o r r . e q . 1 ) t h e n w r i t e ( * , 5 0 0 4 ) I R E G , X N U ( I R E G ) , E P S , A C C , I F A I L e n d i f 5 0 0 4 f o r m a t ( 1 x , I 4 , 2 x , E 1 2 . 6 , 2 x , 2 ( E 8 . 3 , 2 x ) , 1 x , I 2 ) c e n e r n u = e n e r n u + X N U ( I R E G ) c _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ c I n t h i s s e c t i o n w e c a l c u l a t e t h e q u a d r u p o l e m o m e n t : i q f l a g = 1 ! f l a g t o c a l c u l a t e Q u a d r u p o l e M o m e n t M I N P T S = 0 I F A I L = 1 F I N V Q = 0 . 0 D 0 c C A L L D 0 1 F C F ( N D I M , A N U M , B N U M , M I N P T S , M A X P T S , E V A R N , E P S , A C C , L E N W R K , + W R K S T R , F I N V Q , I F A I L ) c q u a d m = q u a d m + F I N V Q c i q f l a g = 0 ! b a c k t o n u m e r c _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ _ c E N D D O w r i t e ( * , * ) E N D D O c c E N D N U M E R A T O R C A L C U L A T I O N c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * 140 Appendix G. Computational Code: Hydrogen Atom in Magnetic Field c V A R I A T I O N A L E N E R G Y D E N O M I N A T O R c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c i f ( i c o r r . e q . 1 ) t h e n w r i t e ( * , * ) w r i t e ( * , * ) ’ I R E G , X D E ( I R E G ) , E P S , A C C , I F A I L ’ w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ e n d i f c e n e r d e = 0 . 0 D 0 I R E G = 0 D O I r h o = 1 , n r h o D D O I z = 1 , n z D I R E G = I R E G + 1 E P S = E P D E N ( I R E G ) c L I M I T S F O R T H E D E N O M I N A T O R A D E N ( 1 ) = R L D ( I r h o ) B D E N ( 1 ) = R L D ( I r h o + 1 ) A D E N ( 2 ) = Z L D ( I z ) * R S B D E N ( 2 ) = Z L D ( I z + 1 ) * R S c M I N P T S = 0 I F A I L = 1 F I N V D = 0 . 0 D 0 c C A L L D 0 1 F C F ( N D I M , A D E N , B D E N , M I N P T S , M A X P T S , E V A R D , E P S , A C C , L E N W R K , + W R K S T R , F I N V D , I F A I L ) c X D E ( I R E G ) = F I N V D i f ( i c o r r . e q . 1 ) t h e n w r i t e ( * , 5 0 0 4 ) I R E G , X D E ( I R E G ) , E P S , A C C , I F A I L e n d i f c e n e r d e = e n e r d e + X D E ( I R E G ) c E N D D O w r i t e ( * , * ) E N D D O c c E N D D E N O M I N A T O R C A L C U L A T I O N c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c * * * * * * * * * * * * * * * * * * * * * * * * * F V A L = e n e r n u / e n e r d e c * * * * * * * * * * * * * * * * * * * * * * * * * c * * * * * * * * * * * * * * * * * * * * * * * * * i f ( F V A L . L E . e b e s t ) t h e n i c a l l b = i c a l l e b e s t = F V A L e n d i f c c q u a d m o = q u a d m / e n e r d e c w r i t e ( * , * ) w r i t e ( * , * ) w r i t e ( * , * ) ’ H - a t o m N e w T r i a l ( 3 1 / 0 7 / 1 9 ) c a l l : ’ , i c a l l w r i t e ( * , * ) ’ * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * ’ w r i t e ( * , * ) ’ V a r i a t i o n a l E n e r g y = ’ , F V A L , ’ R y ’ w r i t e ( * , * ) ’ B i n d i n g E n e r g y = ’ , ( x B - F V A L ) , ’ R y ’ w r i t e ( * , * ) ’ M a g n e t i c F i e l d S t r e n g t h = ’ , x B , ’ a u ’ w r i t e ( * , * ) ’ ( ’ , x B * 2 . 3 5 D 0 , ’ B _ 9 ) ’ w r i t e ( * , * ) ’ Q u a d r u p o l e M o m e n t = ’ , q u a d m o w r i t e ( * , * ) w r i t e ( * , 1 0 1 0 ) ’ a 0 = ’ , a 0 w r i t e ( * , 1 0 1 0 ) ’ a 1 = ’ , a 1 w r i t e ( * , 1 0 1 0 ) ’ a 2 = ’ , a 2 w r i t e ( * , 1 0 1 0 ) ’ a 3 = ’ , a 3 * B * B w r i t e ( * , 1 0 1 0 ) ’ a 4 = ’ , a 4 * B w r i t e ( * , 1 0 1 0 ) ’ a 5 = ’ , a 5 * B w r i t e ( * , 1 0 1 0 ) ’ b 1 = ’ , b 1 w r i t e ( * , 1 0 1 0 ) ’ b 2 = ’ , b 2 w r i t e ( * , 1 0 1 0 ) ’ b 3 = ’ , b 3 * B * B w r i t e ( * , 1 0 1 0 ) ’ b 4 = ’ , b 4 * B * B w r i t e ( * , 1 0 1 0 ) ’ c = ’ , c w r i t e ( * , 1 0 1 0 ) ’ d = ’ , d w r i t e ( * , * ) w r i t e ( * , * ) ’ * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * ’ w r i t e ( * , * ) ’ E b e s t = ’ , e b e s t , ’ c a l l : ’ , i c a l l b w r i t e ( * , * ) c 1 0 1 0 f o r m a t ( 1 x , 1 A 1 2 , 1 F 1 8 . 1 4 ) c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c C O R R E C T I O N S c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c i f ( i c o r r . e q . 1 ) t h e n c w r i t e ( * , * ) ’ C a l c u l a t i n g R e l a t i v e C o r r e c t i o n s . . . ’ w r i t e ( * , * ) ’ - - - - - - - - - - ’ w r i t e ( * , * ) c 1 ) D e l t a R H O N U M E R A N U M ( 1 ) = R L N ( n r h o N + 1 ) B N U M ( 1 ) = 2 . 0 D 0 * R L N ( n r h o N + 1 ) A N U M ( 2 ) = Z L N ( 1 ) * R S B N U M ( 2 ) = Z L N ( n z N + 1 ) * R S E P S = E P S C M I N P T S = 0 I F A I L = 1 F I N V = 0 . 0 D 0 C A L L D 0 1 F C F ( N D I M , A N U M , B N U M , M I N P T S , M A X P T S , E V A R N , E P S , A C C , L E N W R K , + W R K S T R , F I N V , I F A I L ) w r i t e ( * , * ) w r i t e ( * , * ) ’ C o r r e c t i o n s i n r h o E P S A C C ( N U M E R ) ’ w r i t e ( * , 5 0 0 5 ) A B S ( F I N V / e n e r n u ) , E P S , A C C c w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ c c 2 ) D e l t a R H O D E N O M A D E N ( 1 ) = R L D ( n r h o D + 1 ) B D E N ( 1 ) = 2 . 0 D 0 * R L D ( n r h o D + 1 ) A D E N ( 2 ) = Z L D ( 1 ) * R S B D E N ( 2 ) = Z L D ( n Z d + 1 ) * R S E P S = E P S C M I N P T S = 0 I F A I L = 1 F I N V = 0 . 0 D 0 C A L L D 0 1 F C F ( N D I M , A D E N , B D E N , M I N P T S , M A X P T S , E V A R D , E P S , A C C , L E N W R K , + W R K S T R , F I N V , I F A I L ) w r i t e ( * , * ) w r i t e ( * , * ) ’ C o r r e c t i o n s i n r h o E P S A C C ( D E N O M ) ’ w r i t e ( * , 5 0 0 5 ) A B S ( F I N V / e n e r d e ) , E P S , A C C c w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ c c 3 ) D e l t a Z + N U M E R 141 A N U M ( 1 ) = R L N ( 1 ) B N U M ( 1 ) = R L N ( n r h o N + 1 ) A N U M ( 2 ) = Z L N ( n z N + 1 ) * R S B N U M ( 2 ) = 2 . 0 D 0 * Z L N ( n z N + 1 ) * R S E P S = E P S C M I N P T S = 0 I F A I L = 1 F I N V = 0 . 0 D 0 C A L L D 0 1 F C F ( N D I M , A N U M , B N U M , M I N P T S , M A X P T S , E V A R N , E P S , A C C , L E N W R K , + W R K S T R , F I N V , I F A I L ) w r i t e ( * , * ) w r i t e ( * , * ) ’ C o r r e c t i o n s i n z + E P S A C C ( N U M E R ) ’ w r i t e ( * , 5 0 0 5 ) A B S ( F I N V / e n e r n u ) , E P S , A C C c w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ c c 4 ) D e l t a Z + D E N O M A D E N ( 1 ) = R L D ( 1 ) B D E N ( 1 ) = R L D ( n r h o D + 1 ) A D E N ( 2 ) = Z L D ( n Z D + 1 ) * R S B D E N ( 2 ) = 2 . 0 D 0 * Z L D ( n Z D + 1 ) * R S E P S = E P S C M I N P T S = 0 I F A I L = 1 F I N V = 0 . 0 D 0 C A L L D 0 1 F C F ( N D I M , A D E N , B D E N , M I N P T S , M A X P T S , E V A R D , E P S , A C C , L E N W R K , + W R K S T R , F I N V , I F A I L ) w r i t e ( * , * ) w r i t e ( * , * ) ’ C o r r e c t i o n s i n z + E P S A C C ( D E N O M ) ’ w r i t e ( * , 5 0 0 5 ) A B S ( F I N V / e n e r d e ) , E P S , A C C c w r i t e ( * , * ) ’ - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - ’ c e n d i f c c 5 0 0 5 f o r m a t ( 4 x , 3 ( E 1 2 . 6 , 2 x ) ) c c * * * * * * * * * * * * * * * * * * E N D C O R R E C T I O N S * * * * * * * * * * * * * * * * * * * * * * * c R E T U R N E N D c + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + + c E N D F C N c c * * * c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c N U M E R A T O R B E G I N c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c H - a t o m i n a M a g n e t i c F i e l d ( N = 1 c e n t e r s ) c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c T h i s S U B R O U T I N E c a l c u l a t e s t h e I N T E G R A N D o f t h e E N E R G Y N U M E R A T O R c c N O T E S c c - - I n t e g r a t i o n c y l i n d r i c a l c o o r d i n a t e s : r h o _ e a n d z _ e c - - S u b r o u t i n e t o b e u s e d b y t h e N A G i n t e g r a t i o n r o u t i n e c D 0 1 F C F N A G - l i b c D O U B L E P R E C I S I O N F U N C T I O N E V A R N ( N D I M , X N ) I M P L I C I T D O U B L E P R E C I S I O N ( A - H , O - Z ) c . . . E L E C T R O N C O O R D I N A T E S ( I N T E G R A T I O N V A R I A B L E S ) . . c X N ( 1 ) = r h o e X N ( 2 ) = z _ e D I M E N S I O N X N ( N D I M ) c . . . D I S T A N C E S F R O M E L E C T R O N T O E A C H C E N T E R A N D O T H E R V A R I A B L E S c . . . C O M M O N B L O C K S W A V E F U N C T I O N S P A R A M E T E R S C O M M O N / P A R M / B , a 0 , a 1 , a 2 , a 3 , a 4 , a 5 , b 1 , b 2 , b 3 , b 4 , c , d c C O M M O N / F L A G S / i q f l a g c c . . . C O M M O N B L O C K F O R A B B R E V I A T I O N S C O M M O N / A B B R / B q 2 , S Q R B , F C T R c . . I n t r i n s i c F u n c t i o n s . . I N T R I N S I C D S Q R T , D E X P , D C O S c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c E V A R N = 0 . 0 D 0 Z E R O = 0 . 1 D - 1 0 N C E N = 1 c c E L E C T R O N C O O R D I N A T E S r h o = X N ( 1 ) z = X N ( 2 ) c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c D i s t a n c e s f o r m t h e e l e c t r o n t o t h e c e n t e r : c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c r = D S Q R T ( r h o * * 2 + z * * 2 ) c c D e f i n i t i o n s : c - - - - - - - - - - - - - - - - - - - - - - - - - - - - g a m m a = B ! m a g n e t i c f i e l d i n a . u . c c D e f i n i t i o n s : w = d s q r t ( 1 . d 0 + c * g a m m a * * 2 * r h o * * 2 ) c p h i = ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + a 4 * g a m m a * * 2 * r h o * * 2 + & a 5 * g a m m a * * 2 * r * r h o * * 2 + a 2 * w ) / d s q r t ( 1 + b 1 * r + b 3 * g a m m a * * 2 * r * * 2 + & b 4 * g a m m a * * 2 * r h o * * 2 + b 2 * w ) c p f = ( 1 . d 0 + b 1 * r + b 3 * g a m m a * * 2 * r * * 2 + b 4 * g a m m a * * 2 * r h o * * 2 + b 2 * w ) * * ( - d / 2 . d 0 ) P s i 0 = p f * d e x p ( - p h i ) c c c c Q u a d r u p o l e M o m e n t i f ( i q f l a g . E Q . 1 ) t h e n E V A R N = - P s i 0 * P s i 0 * ( r h o * * 2 - 2 . 0 D 0 * z * * 2 ) * r h o R E T U R N e n d i f c c * * * * * * * * * * * * * * * * B L O C K F O R G E O M E T R I C A L T E R M S * * * * * * * * * * * * * * * * c c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c A U X I L I A R Y F U N C T I O N S c w = d s q r t ( 1 . d 0 + c * g a m m a * * 2 * r h o * * 2 ) c f 1 = 1 . d 0 + b 1 * r + b 3 * g a m m a * * 2 * r * * 2 + b 4 * g a m m a * * 2 * r h o * * 2 + b 2 * s q r t ( 1 . d 0 + & c * g a m m a * * 2 * r h o * * 2 ) c c c 142 Appendix G. Computational Code: Hydrogen Atom in Magnetic Field c P A R T I A L D E R I V A T I V E S c p a 2 0 = ( g a m m a * * 2 * ( 4 . d 0 * f 1 * * 2 * ( 2 * a 4 + 2 * a 5 * r + & a 2 * c * ( - ( ( c * g a m m a * * 2 * r h o * * 2 ) / ( 1 + c * g a m m a * * 2 * r h o * * 2 ) * * 1 . 5 d 0 ) + & 1 / w ) ) - 2 . d 0 * f 1 * ( 2 * b 4 + b 2 * c * ( - ( ( c * g a m m a * * 2 * r h o * * 2 ) / ( 1 + & c * g a m m a * * 2 * r h o * * 2 ) * * 1 . 5 d 0 ) + 1 / w ) ) * ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + & a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + a 2 * w ) + & ( 3 . d 0 * g a m m a * * 2 * r h o * * 2 * ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + & a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + a 2 * w ) * ( b 2 * c + & 2 * b 4 * w ) * * 2 ) / w * * 2 - ( 4 . d 0 * f 1 * g a m m a * * 2 * r h o * * 2 * ( b 2 * c + & 2 * b 4 * w ) * ( a 2 * c + 2 * ( a 4 + a 5 * r ) * w ) ) / w * * 2 ) ) / ( 4 . d 0 * f 1 * * 2 . 5 d 0 ) c p a 1 0 = ( g a m m a * * 2 * r h o * ( 2 . d 0 * a 2 * c * f 1 + 4 . d 0 * f 1 * ( a 4 + a 5 * r ) * w - 1 . d 0 * ( a 0 + & a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + a 4 * g a m m a * * 2 * r h o * * 2 + & a 5 * g a m m a * * 2 * r * r h o * * 2 + a 2 * w ) * ( b 2 * c + & 2 * b 4 * w ) ) ) / ( 2 . d 0 * f 1 * * 1 . 5 d 0 * w ) c p a 0 2 = ( 8 . d 0 * a 3 * f 1 * * 2 * g a m m a * * 2 - 4 . d 0 * f 1 * ( b 1 + 2 * b 3 * g a m m a * * 2 * r ) * ( a 1 + & g a m m a * * 2 * ( 2 * a 3 * r + a 5 * r h o * * 2 ) ) - 4 . d 0 * b 3 * f 1 * g a m m a * * 2 * ( a 0 + a 1 * r + & a 3 * g a m m a * * 2 * r * * 2 + a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + & a 2 * w ) + 3 . d 0 * ( b 1 + 2 * b 3 * g a m m a * * 2 * r ) * * 2 * ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 & + a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + & a 2 * w ) ) / ( 4 . d 0 * f 1 * * 2 . 5 d 0 ) c p a 0 1 = ( 2 . d 0 * f 1 * ( a 1 + g a m m a * * 2 * ( 2 * a 3 * r + a 5 * r h o * * 2 ) ) - 1 . d 0 * ( b 1 + & 2 * b 3 * g a m m a * * 2 * r ) * ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + & a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + & a 2 * w ) ) / ( 2 . d 0 * f 1 * * 1 . 5 d 0 ) c p a 1 1 = ( g a m m a * * 2 * r h o * ( 8 . d 0 * a 5 * f 1 * * 2 * w - 2 . d 0 * f 1 * ( a 1 + g a m m a * * 2 * ( 2 * a 3 * r + & a 5 * r h o * * 2 ) ) * ( b 2 * c + 2 * b 4 * w ) + 3 . d 0 * ( b 1 + 2 * b 3 * g a m m a * * 2 * r ) * ( a 0 + a 1 * r & + a 3 * g a m m a * * 2 * r * * 2 + a 4 * g a m m a * * 2 * r h o * * 2 + a 5 * g a m m a * * 2 * r * r h o * * 2 + & a 2 * w ) * ( b 2 * c + 2 * b 4 * w ) - 2 . d 0 * f 1 * ( b 1 + 2 * b 3 * g a m m a * * 2 * r ) * ( a 2 * c + 2 * ( a 4 & + a 5 * r ) * w ) ) ) / ( 4 . d 0 * f 1 * * 2 . 5 d 0 * w ) c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c H A M I L T O N I A N c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c V V i = ( ( - 1 . d 0 * b 1 * * 2 * d * ( 2 + d ) ) / f 1 * * 2 + ( 8 . d 0 * b 4 * d * g a m m a * * 2 ) / f 1 - & 4 . d 0 * p a 0 1 * * 2 + 4 . d 0 * p a 0 2 - 4 . d 0 * p a 1 0 * * 2 + 4 . d 0 * p a 2 0 - 8 . d 0 / r + & ( 8 . d 0 * p a 0 1 ) / r - ( 4 . d 0 * b 3 * * 2 * d * ( 2 + d ) * g a m m a * * 4 * r * * 2 ) / f 1 * * 2 + & ( 4 . d 0 * p a 1 0 ) / r h o - ( 8 . d 0 * b 4 * d * g a m m a * * 2 * p a 1 0 * r h o ) / f 1 - & ( 8 . d 0 * p a 0 1 * p a 1 0 * r h o ) / r + ( 8 . d 0 * p a 1 1 * r h o ) / r + g a m m a * * 2 * r h o * * 2 - & ( 8 . d 0 * b 4 * * 2 * d * g a m m a * * 4 * r h o * * 2 ) / f 1 * * 2 - & ( 4 . d 0 * b 4 * * 2 * d * * 2 * g a m m a * * 4 * r h o * * 2 ) / f 1 * * 2 - & ( 8 . d 0 * b 4 * d * g a m m a * * 2 * p a 0 1 * r h o * * 2 ) / ( f 1 * r ) - & ( 2 . d 0 * b 2 * c * * 2 * d * g a m m a * * 4 * r h o * * 2 ) / ( f 1 * ( 1 + & c * g a m m a * * 2 * r h o * * 2 ) * * 1 . 5 d 0 ) - & ( 2 . d 0 * b 2 * * 2 * c * * 2 * d * g a m m a * * 4 * r h o * * 2 ) / ( f 1 * * 2 * w * * 2 ) - & ( 1 . d 0 * b 2 * * 2 * c * * 2 * d * * 2 * g a m m a * * 4 * r h o * * 2 ) / ( f 1 * * 2 * w * * 2 ) + & ( 4 . d 0 * b 2 * c * d * g a m m a * * 2 ) / ( f 1 * w ) - & ( 4 . d 0 * b 2 * c * d * g a m m a * * 2 * p a 1 0 * r h o ) / ( f 1 * w ) - & ( 8 . d 0 * b 2 * b 4 * c * d * g a m m a * * 4 * r h o * * 2 ) / ( f 1 * * 2 * w ) - & ( 4 . d 0 * b 2 * b 4 * c * d * * 2 * g a m m a * * 4 * r h o * * 2 ) / ( f 1 * * 2 * w ) - & ( 4 . d 0 * b 2 * c * d * g a m m a * * 2 * p a 0 1 * r h o * * 2 ) / ( f 1 * r * w ) - & ( 4 . d 0 * b 3 * d * g a m m a * * 2 * ( b 2 * c * ( 2 + d ) * g a m m a * * 2 * r h o * * 2 + 2 * b 4 * ( 2 . d 0 + & d ) * g a m m a * * 2 * r h o * * 2 * w + f 1 * ( - 3 + 2 * p a 0 1 * r + & 2 * p a 1 0 * r h o ) * w ) ) / ( f 1 * * 2 * w ) - ( 2 . d 0 * b 1 * d * ( b 2 * c * ( 2 + & d ) * g a m m a * * 2 * r h o * * 2 + 2 . d 0 * ( f 1 * ( - 1 . d 0 + p a 0 1 * r + p a 1 0 * r h o ) + ( 2 + & d ) * g a m m a * * 2 * ( b 3 * r * * 2 + b 4 * r h o * * 2 ) ) * w ) ) / ( f 1 * * 2 * r * w ) ) / 4 . d 0 c E V A R N = V V i * P s i 0 * P s i 0 * r h o R E T U R N E N D c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c D E N O M I N A T O R B E G I N c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c H - a t o m i n a M a g n e t i c F i e l d ( N = 1 c e n t e r s ) c * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * * c c T h i s S U B R O U T I N E c a l c u l a t e s t h e I N T E G R A N D o f t h e E N E R G Y N U M E R A T O R c c N O T E S c c - - U s e s c y l i n d r i c a l c o o r d i n a t e s c - - S u b r o u t i n e t o b e u s e d b y t h e N A G i n t e g r a t i o n r o u t i n e c D 0 1 F C F N A G - l i b c D O U B L E P R E C I S I O N F U N C T I O N E V A R D ( N D I M , X D ) I M P L I C I T D O U B L E P R E C I S I O N ( A - H , O - Z ) D I M E N S I O N X D ( N D I M ) c . . . C O M M O N B L O C K S W A V E F U N C T I O N S P A R A M E T E R S C O M M O N / P A R M / B , a 0 , a 1 , a 2 , a 3 , a 4 , a 5 , b 1 , b 2 , b 3 , b 4 , c , d c . . . C O M M O N B L O C K F O R A B B R E V I A T I O N S C O M M O N / A B B R / B q 2 , S Q R B , F C T R c . . I n t r i n s i c F u n c t i o n s . . I N T R I N S I C D S Q R T , D E X P , D C O S c E V A R D = 0 . 0 D 0 Z E R O = 0 . 1 D - 1 0 N C E N = 1 c c E L E C T R O N C O O R D I N A T E S r h o = X D ( 1 ) z = X D ( 2 ) c c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c D i s t a n c e s f o r m t h e e l e c t r o n t o t h e c e n t e r : c - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - - c r = D S Q R T ( r h o * * 2 + z * * 2 ) c c D e f i n i t i o n s : c - - - - - - - - - - - - - - - - - - - - - - - - - - - - g a m m a = B ! m a g n e t i c f i e l d i n a . u . c c D e f i n i t i o n s : w = d s q r t ( 1 . d 0 + c * g a m m a * * 2 * r h o * * 2 ) c p h i = ( a 0 + a 1 * r + a 3 * g a m m a * * 2 * r * * 2 + a 4 * g a m m a * * 2 * r h o * * 2 + & a 5 * g a m m a * * 2 * r * r h o * * 2 + a 2 * w ) / d s q r t ( 1 + b 1 * r + b 3 * g a m m a * * 2 * r * * 2 + & b 4 * g a m m a * * 2 * r h o * * 2 + b 2 * w ) c p f = ( 1 . d 0 + b 1 * r + b 3 * g a m m a * * 2 * r * * 2 + b 4 * g a m m a * * 2 * r h o * * 2 + b 2 * w ) * * ( - d / 2 . d 0 ) P s i 0 = p f * d e x p ( - p h i ) c c E n e r g y D e n o m i n a t o r E V A R D = P s i 0 * P s i 0 * r h o c c R E T U R N E N D c c E N D O F M A I N P R O G R A M c c Appendix H Lagrange-Mesh Method for Hydrogen Atom in Magnetic Field: Code Below it is presented in explicit form the Mathematica Notebook which implements the La- grange Mesh Method for the hydrogen atom in presence of a uniform and constant magnetic field. The program is optimized to run it directly from the terminal/console. (*This program caculates the energy , wave function , \ quadrupo lemoment and cusp parameter of the first \ bound states of the Hydrogen atom in magnetic field *) (* Load module: module load Mathematica /11.2 *) (* Usage: math -noprompt -run < file.txt *) (* Author: Juan Carlos del Valle Nov 19 2019( revisited )*) (* ******************************************** *) (* Initial Data Here*) (* ******************************************** *) (*----------------------------------------------------*) (* Magnetic Quantum Number *) m=0; (*----------------------------------------------------*) (*Basis *) Nr=80; Ntheta =200; (*----------------------------------------------------*) (* Scaling Parameter *) h=0.005; (*----------------------------------------------------*) (* Potential *) V[r_ ,u_]:= - 2/r + 1/4 gamma^2 (1-u^2)r^2 (*----------------------------------------------------*) (* Parity [+1 ,-1]*) p=1; (*----------------------------------------------------*) (* Magnetic Field *) gamma =4000; (*----------------------------------------------------*) (* Tolerance in Float Numbers *) dx =0.000000000000000001; (*----------------------------------------------------*) (* ******************************************** *) (* MAIN PROGRAM *) (* ******************************************** *) (* Matrix Elements *) sols1=NSolve[LaguerreL[Nr,x]==0,x,WorkingPrecision - >50]; k=0; sols1 /.{ r__Rule}:>Set@@@ ({r}/.var:x->Subscript[var ,++k]); Clear[k] (* R a c e s de los polimonios de Legendre *) sols2=NSolve[LegendreP[Ntheta+Abs[m],Abs[m],u]==0&&\ 0<=u<1,u,WorkingPrecision - >50]; k=0; sols2 /.{ r__Rule}:>Set@@@ ({r}/.var:u->Subscript[var ,++k]); Do[Subscript[u, -i]=- Subscript[u, i],{i,1,Ntheta /2}] (* Radial Kinetic Elements *) Do[For[i=1,i{"Arnoldi","Criteria"\ ->"RealPart",MaxIterations - >10000000}][[1]]; (* Coefficients for EigenFunctions *) CF=Eigenvectors[Ham+IdentityMatrix [( Ntheta Nr)/2],-1,\ Method ->{"Arnoldi","Criteria"->"RealPart",MaxIterations\ - >10000000}]; Do[Subscript[c, (1-r)*Nr+s,r]= Rationalize[CF [[1]][[s]],dx]\ , {r,1,Ntheta /2}, {s,Nr*(r-1)+1,r*Nr}]; (* Quadrupole Moment *) z2=h*h*2* Sum[Subscript[c, i, j]^2* Subscript[x, i]^2*\ Subscript[u, j]^2,{i,1,Nr},{j, 1,Ntheta /2}]; r2=h*h*Sum[Subscript[c, i, j]^2* Subscript[x, i]^2*\ (1- Subscript[u, j]^2), {i, 1,Nr},{j, 1,Ntheta /2}]; Q=z2-r2; Export["B" <> ToString[N[gamma ]] <> "_cf.dat", CF [[1]]]; (*Cusp Parameter *) f[r_, i_] := ((-1)^i*( LaguerreL[Nr, r]/(r - Subscript\ [x, i]))* Exp[-(r/2)])/ Subscript[x, i]^2^( -1) df[y_, i_] := D[(( -1)^i*( LaguerreL[Nr, r]/(r - \ Subscript[x, i]))* Exp[-(r/2)])/ Subscript[x, i]^2^(-1) , r]\ /. r -> y CuspN=Sum[f[0, i]*df[0, n]* Subscript[c, i, j]*\ Subscript[c, n, j],{i,1,Nr}, {j,1,Ntheta /2}, {n,1,Nr}]; CuspD=Sum[f[0, i]*f[0, n]* Subscript[c, i, j]* Subscript\ [c, n, j], {i,1,Nr}, {j,1,Ntheta /2}, {n,1,Nr}]; Cusp= CuspN/(h*CuspD); (* OUTPUT *) fileName = "B"<>ToString[N[gamma]]<>".dat"; file = OpenWrite[fileName , PageWidth -> Infinity ]; WriteString[file ," *******************************\n "]; WriteString[file ,"Hydrogen Atom in Magnetic Field \n "]; WriteString[file ," in Lagrange Mesh Method \n "]; WriteString[file ,"********************************\n "]; WriteString[file ," \n "]; WriteString[file ," Q. number m = "<>ToString[m]"\n" ]; WriteString[file ," \n "]; WriteString[file ,"................................\n "]; WriteString[file ,<>ToString[Nr]<>"x"<>ToString[Ntheta]"\n"]; WriteString[file ," Scaling " <>ToString[N[h]]<>"\n"]; WriteString[file ,"................................\n "]; WriteString[file ," \n "]; WriteString[file ,"Energy = "<>ToString[epsilon]<>"\n\"]; WriteString[file ," \n "]; WriteString[file ,"QM="<>ToString[NumberForm[Q,16]]<>"\\ n"]; WriteString[file ," \n "]; WriteString[file ," Cusp Par. = "<>ToString[Cusp]"\n "]; WriteString[file ,"................................\n "]; Close[file] Appendix I Plots of Variational Parameters: Hydrogen Atom in Magnetic Field Plots of the optimal variational parameters {α0,1,2,3,4, β1,2,3} for Approximant (12.0.3) as func- tions of log(1 + γ2) are shown below. 0 3 6 9 12 15 18 2 3 4 5 6 0 3 6 9 12 15 18 1 2 3 4 0 3 6 9 12 15 18 2 4 6 8 10 12 0 3 6 9 12 15 18 0.05 0.10 0.15 0.20 0.25 Figure I.1: Optimal variational parameters {α0, α1, α2, α3} of the Approximant (12.0.3) as functions of log(1 + γ2). 145 146 Appendix I. Plots of Variational Parameters: Hydrogen Atom in Magnetic Field 0 3 6 9 12 15 18 0.2 0.4 0.6 0.8 1.0 1.2 0 3 6 9 12 15 18 2 4 6 8 0 3 6 9 12 15 18 4 8 12 16 20 0 3 6 9 12 15 18 0.5 1.0 1.5 2.0 Figure I.2: Optimal variational parameters {α4, β1, β2, β3} of the Approximant (12.0.3) as func- tions of log(1 + γ2). Bibliography [1] L. D. Landau and M. Lifshitz, Quantum Mechanics, Non-Relativistic Theory. Pergamon, 1977. [2] J. J. Sakurai and J. Napolitano, Modern quantum mechanics; 2nd ed. San Francisco, CA: Addison-Wesley, 2011. [3] D. Griffiths, Introduction of Quantum Mechanics. Prentice Hall, Inc., 1995. [4] P. M. Morse and H. Feshbach, “Methods of theoretical physics,” American Journal of Physics, vol. 22, no. 6, pp. 410–413, 1954. [5] A. V. Turbiner and J. C. López-Vieyra, Notas del Curso: La Ecuación de Schrödinger: Teoŕıa y práctica . Unpublished, 2014. [6] T. Kato, “On the eigenfunctions of many-particle systems in quantum mechanics,” Com- mun. Pure Appl. Math., vol. 10, 1957. [7] A. V. Turbiner, “A new approach to the eigenvalue problem in quantum mechanics: Convergent perturbation theory for rising potentials,” J. Phys. A, vol. 14, no. 7, p. 1641, 1981. [8] A. V. Turbiner, “The eigenvalue spectrum in quantum mechanics and the nonlinearization procedure,” Phys. Usp., vol. 27, no. 9, p. 668, 1984. [9] J. C. del Valle and A. V. Turbiner, “Radial anharmonic oscillator: Perturbation theory, new semiclassical expansion, approximating eigenfunctions. I. Generalities, cubic anhar- monicity case,” Int. J. Mod. Phys. A, vol. 34, no. 26, p. 1950143, 2019. [10] J. C. del Valle and A. V. Turbiner, “Radial anharmonic oscillator: Perturbation theory, new semiclassical expansion, approximating eigenfunctions. II. Quartic and sextic anhar- monicity cases,” International Journal of Modern Physics A, vol. 35, no. 01, p. 2050005, 2020. [11] A. V. Turbiner, M. A. Escobar-Ruiz, and J. C. del Valle, Work in progress. [12] W. Heisenberg, “Über quantentheoretische Umdeutung kinematischer und mechanischer Beziehungen,” Z. Physik, vol. 33, no. 1, pp. 879–893, 1925. 147 148 Bibliography [13] E. Schrödinger, “An Undulatory Theory of the Mechanics of Atoms and Molecules,” Phys. Rev., vol. 28, pp. 1049–1070, 1926. [14] E. C. Brinsmade, J. B.; Kemble, “The occurrence of harmonics in the infra-red absorption spectra of diatomic gases,” Proc. Natl. Acad. Sci. U.S.A., vol. 3, 1917. [15] J. L. Dunham, “Intensities in the harmonic band of hydrogen chloride,” Phys. Rev. (Series I), vol. 34, 1929. [16] P. M. Morse and E. C. G. Stueckelberg, “Diatomic molecules according to the wave mechanics I: Electronic levels of the hydrogen molecular ion,” Phys. Rev. (Series I), vol. 33, 1929. [17] P. M. Morse, “Diatomic molecules according to the wave mechanics. II. Vibrational lev- els,” Phys. Rev., vol. 34, pp. 57–64, 1929. [18] E. C. Kemble, “On the energy required to split HCI into atomic ions,” J. Opt. Soc. Am. (1917-1983), vol. 12, no. 1, p. 1, 1926. [19] E. Fues, “Das Eigenschwingungsspektrum zweiatomiger Moleküle in der Undulations- mechanik,” Ann. Phys., vol. 385, no. 12, pp. 367–396, 1926. [20] C. M. Bender and T. T. Wu, “Anharmonic Oscillator,” Phys. Rev., vol. 184, pp. 1231– 1260, 1969. [21] C. M. Bender, “Generalized anharmonic oscillator,” J. Math. Phys., vol. 11, 1970. [22] C. M. Bender and T. T. Wu, “Anharmonic Oscillator. II. A Study of Perturbation Theory in Large Order,” Phys. Rev. D, vol. 7, pp. 1620–1636, 1973. [23] A. V. Turbiner, “Anharmonic oscillator and double-well potential: approximating eigen- functions,” Lett. Math. Phys., vol. 74, no. 2, pp. 169–180, 2005. [24] A. V. Turbiner, “Double well potential: perturbation theory, tunneling, WKB, (beyond instantons),” Int. J. Mod. Phys. A, vol. 25, pp. 647–658, 2010. [25] H. Taşeli and R. Eid, “Eigenvalues of the two-dimensional Schrödinger equation with nonseparable potentials,” Int. J. Quantum Chem., vol. 59, no. 3, pp. 183–201, 1996. [26] E. J. Weniger, “A convergent renormalized strong coupling perturbation expansion for the ground state energy of the quartic, sextic, and octic anharmonic oscillator,” Ann. Phys., vol. 246, no. 1, pp. 133–165, 1996. [27] S. Graffi, V. Grecchi, and B. Simon, “Borel Summability: Application to the anharmonic oscillator,” Phys. Lett. B, vol. B 32, pp. 631–634, 1970. [28] H. J. Silverstone, J. G. Harris, J. Č́ıžek, and J. Paldus, “Asymptotics of high-order perturbation theory for the one-dimensional anharmonic oscillator by quasisemiclassical methods,” Phys. Rev. A, vol. 32, pp. 1965–1980, 1985. Bibliography 149 [29] V. Eletsky and V. Popov, “On perturbation theory for the gr2Nanharmonic oscillator,” Phys. Lett. B, vol. 94, no. 1, pp. 65 – 68, 1980. [30] F. M. Fernández and R. Guardiola, “The strong coupling expansion for anharmonic os- cillators,” J. Phys. A, vol. 30, no. 20, p. 7187, 1997. [31] J. Loeffel, A. Martin, B. Simon, and A. Wightman, “Pade approximants and the anhar- monic oscillator,” Phys. Lett. B, vol. 30, no. 9, pp. 656 – 658, 1969. [32] M. R. M. Witwit, “Energy levels of a two-dimensional anharmonic oscillator: Hill deter- minant approach,” Pramana, vol. 41, no. 6, pp. 493–502, 1993. [33] S. N. Biswas et al., “The Hill determinant: an application to the anharmonic oscillator,” Phys. Rev. D, vol. 4, pp. 3617–3620, 1971. [34] M. Tater and A. V. Turbiner, “Failure of the Hill determinant method for a sextic anhar- monic oscillator,” J. Phys. A, vol. 26, pp. 697–710, 1993. [35] P. Lu, “WKB Calculation of the Energy Levels of Anharmonic Oscillators,” J. Chem. Phys., vol. 47, no. 2, pp. 815–817, 1967. [36] F. Hioe, D. Macmillen, and E. Montroll, “Quantum theory of anharmonic oscillators: Energy levels of a single and a pair of coupled oscillators with quartic coupling,” Phys. Rep., vol. 43, no. 7, pp. 305 – 335, 1978. [37] E. Yukalova and V. Yukalov, “Self-similar approximation for an anharmonic oscillator of arbitrary dimensionality,” Phys. Lett. A, vol. 175, no. 1, pp. 27 – 35, 1993. [38] N. Ari and M. Demiralp, “Energy levels of a two-dimensional anharmonic oscillator: Characteristic function approach,” J. Math. Phys., vol. 26, no. 6, pp. 1179–1185, 1985. [39] A. D. Dolgov and V. S. Popov, “The anharmonic oscillator and its dependence on space dimensions,” Phys. Lett., vol. 86B, pp. 185–188, 1979. [40] D. J. Doren and D. R. Herschbach, “Spatial dimension as an expansion parameter in quantum mechanics,” Phys. Rev. A, vol. 34, pp. 2654–2664, 1986. [41] C. Müller, Spherical Harmonics. Springer-Verlag, 1966. [42] C. Efthimiou and C. Frye, Spherical Harmonics in p Dimensions. World Scientific, 2014. [43] K. D. Granzow, “N-Dimensional total orbital angular-momentum operator,” J. Math. Phys., vol. 4, no. 7, pp. 897–900, 1963. [44] M. A. Escobar-Ruiz, E. Shuryak, and A. V. Turbiner, “Fluctuations in quantum mechan- ics and field theories from a new version of semiclassical theory.,” Phys. Rev. D, vol. 93, p. 105039, 2016. [45] M. A. Escobar-Ruiz, E. Shuryak, and A. V. Turbiner, “Fluctuations in quantum mechan- ics and field theories from a new version of semiclassical theory. II,” Phys. Rev. D, vol. 96, p. 045005, 2017. 150 Bibliography [46] E. Shuryak and A. V. Turbiner, “Transseries for the ground state density and generalized Bloch equation: Double-well potential case,” Phys. Rev. D, vol. 98, p. 105007, 2018. [47] J. C. del Valle and D. J. Nader, “Toward the theory of the Yukawa potential,” J. Math. Phys., vol. 59, no. 10, p. 102103, 2018. [48] R. Lynch and H. Mavromatis, “N-dimensional harmonic oscillator yields monotonic series for the mathematical constant π,” J. Comput. Appl. Math., vol. 30, no. 2, pp. 127–137, 1990. [49] T. I. Banks and C. M. Bender, “Anharmonic oscillator with polynomial self-Interaction,” J. Math. Phys., vol. 13, no. 9, pp. 1320–1324, 1972. [50] F. Dyson, “Divergence of Perturbation Theory in Quantum Electrodynamics,” Phys. Rev., vol. 85, pp. 631–632, 1952. [51] J. Killingbeck, “Renormalised perturbation series,” J. Phys. A, vol. 14, no. 5, p. 1005, 1981. [52] A. V. Turbiner and A. G. Ushveridze, “Anharmonic oscillator: Constructing the strong coupling expansions,” J. Math. Phys., vol. 29, no. 9, pp. 2053–2063, 1988. [53] A. V. Turbiner, “A New Approach to finding levels of energy of bound states in quan- tum mechanics: Convergent perturbation theory,” Soviet Phys. – Pisma ZhETF, vol. 30, pp. 379–383, 1979. [54] F. Alarcón, El estado excitado 2σg del ión molecular H+ 2 en un campo magnético. Bsc thesis, UNAM, 2004. [55] H. Medel, Puntos cŕıticos de sistemas moleculares coulombianos. PhD thesis, UNAM, 2011. [56] D. J. Nader, Existencia de la cadena molecular de hidrógeno H3 y del ión H− 2 en campos magnéticos intensos. PhD thesis, UNAM, 2019. [57] A. Genz and A. Malik, “Remarks on algorithm 006: An adaptive algorithm for numerical integration over an N -dimensional rectangular region,” J. Comput. Appl. Math., vol. 6, no. 4, pp. 295 – 302, 1980. [58] D. Baye, “The Lagrange-mesh method,” Phys. Rep., vol. 565, no. Supplement C, pp. 1– 107, 2015. [59] H. Bateman, Higher Transcendental Functions, vol. 1. McGraw-Hill Book Company, 1953. [60] A. D. Dolgov and V. S. Popov, “Higher orders an structure of the perturbation series for the anharmonic oscillator,” Zh. Eksp. Teor. Fiz., vol. 75, pp. 668–694, 1978. [61] A. V. Turbiner, “On perturbation theory and variational methods in quantummechanics,” Zh. Eksp. Teor. Fiz., vol. 79, pp. 1719 – 1745, 1980. Bibliography 151 [62] A. V. Turbiner and A. G. Ushveridze, “Anharmonic oscillator: Constructing the strong coupling expansions,” J. Math. Phys., vol. 29, no. 9, pp. 2053–2063, 1988. [63] F. M. Fernández and R. Guardiola, “The strong coupling expansion for anharmonic os- cillators,” J. Phys. A, vol. 30, no. 20, p. 7187, 1997. [64] E. J. Weniger, “Construction of the strong coupling expansion for the ground state energy of the quartic, sextic, and octic anharmonic oscillator via a renormalized strong coupling expansion,” Phys. Rev. Lett., vol. 77, pp. 2859–2862, 1996. [65] M. Abramowitz and I. A. Stegun, Handbook of Mathematical Functions with Formulas, Graphs, and Mathematical Tables. New York: Dover, 1964. [66] H. Meißner and E. O. Steinborn, “Quartic, sextic, and octic anharmonic oscillators: Pre- cise energies of ground state and excited states by an iterative method based on the generalized Bloch equation,” Phys. Rev. A, vol. 56, pp. 1189–1200, 1997. [67] M. R. M. Witwit, “Energy levels of two-dimensional anharmonic oscillators with sextic and octic perturbations,” J. Math. Phys., vol. 33, no. 12, pp. 4196–4205, 1992. [68] W. H. P. Jordan, “Anwendung der Quantenmechanik auf das Problem der anomalen Zeemaneffekte,” Z. Phys., vol. 37, 1926. [69] L. I. Schiff and H. Snyder, “Theory of the quadratic Zeeman effect,” Phys. Rev. (Series I), vol. 55, 1939. [70] R. H. Garstang, “Atoms in high magnetic fields (white dwarfs),” Rep. Prog. Phys., vol. 40, 1977. [71] J. E. Avron, B. G. Adams, J. Č́ıžek, M. Clay, M. L. Glasser, P. Otto, J. Paldus, and E. Vrscay, “Bender-Wu formula, the SO(4,2) dynamical group, and the Zeeman effect in hydrogen,” Phys. Rev. Lett., vol. 43, pp. 691–693, 1979. [72] C. M. Bender and S. A. Orszag, Advanced Mathematical Methods for Scientists and Engi- neers I: Asymptotic Methods and Perturbation Theory. Springer-Verlag New York, 1 ed., 1999. [73] J. L. Guillou and J. Zinn-Justin, “The hydrogen atom in strong magnetic fields: Sum- mation of the weak field series expansion,” Ann. Phys., vol. 147, no. 1, pp. 57 – 84, 1983. [74] Y. P. Kravchenko, M. A. Liberman, and B. Johansson, “Exact solution for a hydrogen atom in a magnetic field of arbitrary strength,” Phys. Rev. A, vol. 54, pp. 287–305, 1996. [75] D. Baye, M. Vincke, and M. Hesse, “Simple and accurate calculations on a Lagrange mesh of the hydrogen atom in a magnetic field,” J. Phys. B, vol. 41, no. 5, p. 055005, 2008. [76] J. H. Wang and C. S. Hsue, “Calculation of the energy levels of a hydrogen atom in a magnetic field of arbitrary strength by using B splines,” Phys. Rev. A, vol. 52, pp. 4508– 4514, 1995. 152 Bibliography [77] R. Wallis and H. Bowlden, “Theory of impurity photo-ionization spectrum of semicon- ductors in magnetic fields,” J. Phys. Chem. Solids, vol. 7, no. 1, pp. 78 – 89, 1958. [78] Y. Yafet, R. Keyes, and E. Adams, “Hydrogen atom in a strong magnetic field,” J. Phys. Chem. Solids, vol. 1, no. 3, pp. 137 – 142, 1956. [79] A. Galindo and P. Pascual, “Hydrogen atom in a strong magnetic field,” Il Nuovo Cimento B (1971-1996), vol. 34, no. 1, pp. 155–168, 1976. [80] A. V. Turbiner, “The hydrogen atom in an external magnetic field,” J. Phys. A, vol. 17, no. 4, pp. 859–875, 1984. [81] A. Y. Potekhin and A. V. Turbiner, “Hydrogen atom in a magnetic field: The quadrupole moment,” Phys. Rev. A, vol. 63, p. 065402, 2001. [82] H. Kobori and T. Ohyama, “Variational calculations of energy levels associated with 1s, 2p and 3d states of hydrogenic donors for semiconductors in arbitrary magnetic field,” J. Phys. Chem. Solids, vol. 58, no. 12, pp. 2057 – 2064, 1997. [83] E. O. Pokatilov and M. M. Rusanov, “Variational calculation of energy levels of a hydrogen-like system in a magnetic field,” Sov. Phys. Solid State, vol. 10, p. 2458, 1969. [84] D. Larsen, “Shallow donor levels of insb in a magnetic field,” J. Phys. Chem. Solids, vol. 29, no. 2, pp. 271 – 280, 1968. [85] N. Lee, D. Larsen, and B. Lax, “Exciton levels in a magnetic field,” J. Phys. Chem. Solids, vol. 34, no. 6, pp. 1059 – 1067, 1973. [86] P. C. Makado and N. C. McGill, “Energy levels of a neutral hydrogen-like system in a constant magnetic field of arbitrary strength,” J. Phys. C, vol. 19, no. 6, 1986. [87] D. Cabib, E. Fabri, and G. Fiorio, “The ground state of the exciton in a magnetic field,” Solid State Commun., vol. 9, no. 17, pp. 1517 – 1520, 1971. [88] G. A. Arteca, F. M. Fernandez, and E. A. Castro, “Summation of the perturbation series for the energy levels of a hydrogen atom in a magnetic field,” Chem. Phys. Lett., vol. 102, no. 4, pp. 344 – 353, 1983. [89] Z. Chen and S. P. Goldman, “Relativistic and nonrelativistic finite-basis-set calculations of low-lying levels of hydrogenic atoms in intense magnetic fields,” Phys. Rev. A, vol. 45, pp. 1722–1731, 1992. [90] H. C. Praddaude, “Energy levels of hydrogenlike atoms in a magnetic field,” Phys. Rev. A, vol. 6, pp. 1321–1324, 1972. [91] J. C. López-Vieyra and H. O. Pilón, “Hydrogen atom in a magnetic field: electromagnetic transitions of the lowest states,” Rev. Mex. Fis., vol. 54, no. 1, pp. 49 – 57, 2008. [92] W. Rösner, G. Wunner, H. Herold, and H. Ruder, “Hydrogen atoms in arbitrary magnetic fields. I. Energy levels and wavefunctions,” J. Phys. B, vol. 17, no. 1, pp. 29–52, 1984. Bibliography 153 [93] M. V. Ivanov, “The hydrogen atom in a magnetic field of intermediate strength,” J. Phys. B, vol. 21, no. 3, pp. 447–462, 1988. [94] J. Xi, X. He, and B. Li, “Energy levels of the hydrogen atom in arbitrary magnetic fields obtained by using B-spline basis sets,” Phys. Rev. A, vol. 46, pp. 5806–5811, 1992. [95] H. R. G. Wunner, “Energy levels and electromagnetic transitions of atoms in supertrong magnetic fields,” J. Phys. Colloq., vol. 43, 1982. [96] G. Fonte, P. Falsaperla, G. Schiffrer, and D. Stanzial, “Quadratic Zeeman effect for hydrogen: A method for rigorous bound-state error estimates,” Phys. Rev. A, vol. 41, pp. 5807–5813, 1990. [97] C. Stubbins, K. Das, and Y. Shiferaw, “Low-lying energy levels of the hydrogen atom in a strong magnetic field,” J. Phys. B, vol. 37, no. 10, pp. 2201–2209, 2004. [98] H. Nakashima and H. Nakatsuji, “Solving the Schrödinger and Dirac equations for a hydrogen atom in the universe’s strongest magnetic fields with the free complement method,” Astrophys. J., vol. 725, no. 1, pp. 528–533, 2010. [99] H. Hasegawa and R. Howard, “Optical absorption spectrum of hydrogenic atoms in a strong magnetic field,” J. Phys. Chem. Solids, vol. 21, 1961. [100] L. Gorkov and I. Dzyaloshinskii, “Contribution to the theory of the Mott exciton in a strong magnetic field,” Sov. Phys. JETP, vol. 26, pp. 449–451, 1968. [101] V. Pavlov-Verevkin and B. Zhilinskii, “Neutral hydrogen-like system in a magnetic field,” Phys. Lett. A, vol. 78, no. 3, pp. 244–245, 1980.