A Dirac's Delta Function

Total Page:16

File Type:pdf, Size:1020Kb

A Dirac's Delta Function A Dirac’s delta Function The 1-d delta function, δ(x), is defined through a limiting procedure so that δ(x)=0 for x = 0 (A.1a) and δ(0) = ∞. The meaning of this last relation, taking into account (A.1a), is that b dxf(x)δ(x)=f(0) (A.1b) a for any well-behaved function f(x) and for any pair a, b such that a<0 <b. Some functions producing in the limit the delta function, as defined by (A.1a, A.1b), are given below: ⎧ ⎪ 1 ε ⎪ lim , Lorentzian, (A.2a) ⎪ → + 2 2 ⎪ ε 0 π x + ε ⎪ 2 ⎪ 1 x ⎨⎪ lim √ exp − , Gaussian, (A.2b) σ→0+ σ 2π 2σ2 δ(x)= ⎪ sin(x/ε) ⎪ lim , Dirichlet, (A.2c) ⎪ ε→0+ πx ⎪ ⎪ |L| ⎪ 1 ⎩⎪ lim eikxdk, Fourier. (A.2d) →∞ L 2π −|L| From the definition of the delta function it follows that δ (x − x ) δ(g(x)) = n , (A.3) |g (x )| n n where g(x) is a well-behaved function of x, the sum runs over the real roots, { } ≡ xn ,ofg(x) = 0, and g (xn) (dg/dx)x=xn =0forallxn. The derivative of the delta function is properly defined as δ (x)=0,x= 0 (A.4a) 338 A Dirac’s delta Function and b dxf(x)δ(x)=−f (0) (A.4b) a for any well-behaved function f(x) and for any pair a, b such that a<0 <b. The integral of the delta function is the so-called theta function dθ(x) δ(x)= , (A.5a) dx where 1 , for x>0 , θ(x) ≡ 0 , for x<0 . (A.5b) The definition of the delta function is generalized to more than one inde- pendent cartesian variable: δ (r)=δ (x1) δ (x2) ...δ (xn)(A.6) in an n-dimensional space. More explicitly we have 1 δ (r − r)=δ (x − x) δ (y − y)= δ (r − r) δ (φ − φ) , 2-d , (A.7) r δ (r − r)=δ (x − x) δ (y − y) δ (z − z) 1 = δ (r − r) δ (cos θ − cos θ) δ (φ − φ) r2 1 = δ (r − r) δ (θ − θ) δ (φ − φ) , 3-d . (A.8) r2 sin θ The completeness relation (1.4) of Chap. 1 allows us to obtain many series or integral representations of the delta function. For example, the Fourier representation of the d-dimensional delta function is, according to (1.4) and the footnote on p. 3, ∗ 1 ik · (r−r) δ (r − r )= φk (r) φk (r )= e −→ Ω Ω→∞ k k 1 Ω · r−r 1 k · r−r = dkeik ( ) = dkei ( ) . (A.9) Ω (2π)d (2π)d Equation (A.9) generalizes in d-dimensions (A.2d). Similarly, the delta function over the spherical angles θ and φ can be expressed as a sum over spherical harmonics, Ym(θ, φ), according to (1.4): 1 δ (cos θ − cos θ) δ (φ − φ)= δ (θ − θ) δ (φ − φ) sin θ ∞ ∗ = Ym(θ, φ)Ym (θ ,φ) , (A.10) =0 m=− A Dirac’s delta Function 339 where the spherical harmonics are given in terms of the associated Legendre |m| polynomials, P , by the following expression: 0 (|m|+m) (2 +1)( −|m|)! | | Y (θ, φ)=i(−1) /2 P m (cos θ)eimφ . (A.11) m 4π( + |m|)! Equation (A.10) is valid because the set Ym(θ, φ) is orthonormal π 2π ∗ δ δ sin θdθ dφYm(θ, φ)Y m (θ, φ)= mm . 0 0 B Dirac’s bra and ket Notation In quantum mechanics, the states of a particle are described by eigenfunctions, φn (r), of any hermitian operator, L, corresponding to a physical quantity such as its momentum, its Hamiltonian, etc. The same physical states can equally well be described by the Fourier transforms, φn (k), of φn (r)orbyanyother representation of φn (r) such as the set {an,} of the expansion φn (r)= an,ψ (r)(B.1) in terms of a complete orthonormal set of functions {ψ}. It is very convenient to introduce for a physical state a single symbol that will be independent of any particular representation and such that the various representations, φn (r), φn (k), and {an,}, can be reproduced easily whenever they are needed. This symbol is denoted by |n or by |φn and is called a ket. Thus State of a particle characterized by quantum number(s) n ⇐⇒ | φn or |n To each ket |φn corresponds one and only one conjugate entity denoted by φn| and called a bra. We define also the inner product of two kets, |φ and |ψ, (in that order) as a complex quantity denoted by φ | ψ and such that ∗ φ | ψ = ψ | φ . (B.2) In particular, for an orthonormal set we have φn | φm = δnm . (B.3) Physically, the inner product φ | ψ gives the probability amplitude for a par- ticle being in state |ψ to be observed in state |φ. A complete set of kets, {|φn}, defines an abstract Hilbert space. Some usual complete sets of kets for a spinless particle are the following: (a) The position kets {|r}, where r runs over all positions in the domain Ω; |r corresponds to the state where the particle is completely localized at the point r. 342 B Dirac’s bra and ket Notation (b) The momentum kets {|k}; |k corresponds to the state where the particle has momentum k. (c) The Bloch kets {|k,n}; |k,n corresponds to the state where the particle has crystal momentum k and belongs to the nth band. (d) The kets {|φ}; |φ is an eigenstate of the Hamiltonian. The indexes fully characterize this eigenstate. According to the above, the inner product r | ψ gives the probability ampli- tude for a particle being in the state |ψ to be observed at the point r; but this probability amplitude is nothing other than ψ (r). Hence ψ (r)=r | ψ . (B.4) Similarly, ψ(k)=k | ψ . (B.5) If in (B.4) we choose |ψ = |k, then we obtain 1 · r r | k = √ eik . (B.6) Ω The rhs is the r-representation of a normalized plane wave; if in (B.4) we choose |ψ = |r,thenwehave r | r = δ (r − r) . (B.7) Equation (B.7) follows from the orthonormality of the set {|r}. From (B.4) and (B.7) we obtain ψ (r)=r | ψ = drδ (r − r) r | ψ = dr r | rr | ψ = r| dr |rr | ψ . (B.8) Since (B.8) is valid for every r| and every |ψ, we end up with the complete- ness relation for the set {|r}: dr |rr| =1. (B.9) Similarly, we obtain the relation Ω 1= |kk|−→ dk |kk| . (B.10) Ω→∞ d k (2π) Using (B.9) and (B.10) we can easily transform from ψ(r)toψ(k)andvice versa 1 · r ψ(k)=k | ψ = dr k | rr | ψ = dr √ e−ik ψ (r) Ω 1 · r = √ dre−ik ψ (r) , (B.11) Ω B Dirac’s bra and ket Notation 343 1 · r ψ (r)=r | ψ = r | kk | ψ = √ eik ψ(k) k Ω k √ Ω · r = dkeik ψ(k) . (B.12) (2π)d Notice that the definition of the inner product coincides with the usual defi- nition in the r-representation: φ | ψ = dr φ | rr | ψ = drφ∗ (r) ψ (r) . (B.13) The form of any operator in the Dirac notation follows by calculating its matrix elements. For example, taking into account that the position opera- tor, rˆ, satisfies by definition the relation r| rˆ = r| r and that the momentum operator, pˆ, satisfies the relation k| pˆ = k| k,wehave r | rˆ | ψ = r r | ψ = rψ (r) , (B.14) k | pˆ | ψ = k k | ψ = kψ (k) . (B.15) Using (B.10) and (B.15) we obtain the matrix element r | pˆ | ψ: 1 · r r | pˆ | ψ = r | kk | pˆ | ψ = √ eik kψ(k) k k Ω ∂ · r 1 ∂ 1 k · r = −i eik √ ψ (k)=−i √ ei ψ(k) k ∂r Ω ∂r k Ω ∂ = −i ψ (r) . (B.16) ∂r The last relation follows from (B.12). In a similar way one can show that ∂ k | rˆ | ψ =i ψ(k) . (B.17) ∂k The relations (B.14)–(B.17) can be generalized for operators that are func- tions of rˆ or pˆ. Since by definition the function f Aˆ of any operator Aˆ is an operator with the same eigenstates as Aˆ and eigenvalues f (Ai), where Ai are the eigenvalues of Aˆ,wehave r | f (rˆ) | ψ = f ∗ (r) ψ (r) , (B.18) k | g (pˆ) | ψ = g∗ (k) ψ(k) , (B.19) ∂ r | g (pˆ) | ψ = g∗ −i ψ (r) , (B.20) ∂r ∂ k | f (rˆ) | ψ = f ∗ i ψ (k) . (B.21) ∂k 344 B Dirac’s bra and ket Notation For hermitian operators f ∗ (r)=f (r)andg∗ (k)=g (k). As an application of (B.18) and (B.20) we have 8 9 pˆ2 2 r + V (rˆ) ψ = − ∇2 + V (r) ψ (r) . (B.22) 2m 2m r For particles with spin 1/2 we have to specify whether the spin is up (↑)or down (↓). Thus a complete set is {|r, ↑ , |r, ↓} or {|k, ↑ , |k, ↓}. Assuming that the spin quantization axis is the z-axis, we have for the spin operator s = x0sx + y0sy + z0sz: 1 σ = +1 for spin up , s |r,σ = σ |r,σ , (B.23) z 2 σ = −1forspindown, r, ↑|s | ψ 1 r, ↑|ψ = σ , (B.24) r, ↓|s | ψ 2 r, ↓|ψ where σ is the vector defined by the three Pauli matrices 01 0 −i 10 σ = x + y + z .
Recommended publications
  • Spheres in Infinite-Dimensional Normed Spaces Are Lipschitz Contractible
    PROCEEDINGS OF THE AMERICAN MATHEMATICAL SOCIETY Volume 88. Number 3, July 1983 SPHERES IN INFINITE-DIMENSIONAL NORMED SPACES ARE LIPSCHITZ CONTRACTIBLE Y. BENYAMINI1 AND Y. STERNFELD Abstract. Let X be an infinite-dimensional normed space. We prove the following: (i) The unit sphere {x G X: || x II = 1} is Lipschitz contractible. (ii) There is a Lipschitz retraction from the unit ball of JConto the unit sphere. (iii) There is a Lipschitz map T of the unit ball into itself without an approximate fixed point, i.e. inffjjc - Tx\\: \\x\\ « 1} > 0. Introduction. Let A be a normed space, and let Bx — {jc G X: \\x\\ < 1} and Sx = {jc G X: || jc || = 1} be its unit ball and unit sphere, respectively. Brouwer's fixed point theorem states that when X is finite dimensional, every continuous self-map of Bx admits a fixed point. Two equivalent formulations of this theorem are the following. 1. There is no continuous retraction from Bx onto Sx. 2. Sx is not contractible, i.e., the identity map on Sx is not homotopic to a constant map. It is well known that none of these three theorems hold in infinite-dimensional spaces (see e.g. [1]). The natural generalization to infinite-dimensional spaces, however, would seem to require the maps to be uniformly-continuous and not merely continuous. Indeed in the finite-dimensional case this condition is automatically satisfied. In this article we show that the above three theorems fail, in the infinite-dimen- sional case, even under the strongest uniform-continuity condition, namely, for maps satisfying a Lipschitz condition.
    [Show full text]
  • Examples of Manifolds
    Examples of Manifolds Example 1 (Open Subset of IRn) Any open subset, O, of IRn is a manifold of dimension n. One possible atlas is A = (O, ϕid) , where ϕid is the identity map. That is, ϕid(x) = x. n Of course one possible choice of O is IR itself. Example 2 (The Circle) The circle S1 = (x,y) ∈ IR2 x2 + y2 = 1 is a manifold of dimension one. One possible atlas is A = {(U , ϕ ), (U , ϕ )} where 1 1 1 2 1 y U1 = S \{(−1, 0)} ϕ1(x,y) = arctan x with − π < ϕ1(x,y) <π ϕ1 1 y U2 = S \{(1, 0)} ϕ2(x,y) = arctan x with 0 < ϕ2(x,y) < 2π U1 n n n+1 2 2 Example 3 (S ) The n–sphere S = x =(x1, ··· ,xn+1) ∈ IR x1 +···+xn+1 =1 n A U , ϕ , V ,ψ i n is a manifold of dimension . One possible atlas is 1 = ( i i) ( i i) 1 ≤ ≤ +1 where, for each 1 ≤ i ≤ n + 1, n Ui = (x1, ··· ,xn+1) ∈ S xi > 0 ϕi(x1, ··· ,xn+1)=(x1, ··· ,xi−1,xi+1, ··· ,xn+1) n Vi = (x1, ··· ,xn+1) ∈ S xi < 0 ψi(x1, ··· ,xn+1)=(x1, ··· ,xi−1,xi+1, ··· ,xn+1) n So both ϕi and ψi project onto IR , viewed as the hyperplane xi = 0. Another possible atlas is n n A2 = S \{(0, ··· , 0, 1)}, ϕ , S \{(0, ··· , 0, −1)},ψ where 2x1 2xn ϕ(x , ··· ,xn ) = , ··· , 1 +1 1−xn+1 1−xn+1 2x1 2xn ψ(x , ··· ,xn ) = , ··· , 1 +1 1+xn+1 1+xn+1 are the stereographic projections from the north and south poles, respectively.
    [Show full text]
  • The Fundamental Theorem of Calculus for Lebesgue Integral
    Divulgaciones Matem´aticasVol. 8 No. 1 (2000), pp. 75{85 The Fundamental Theorem of Calculus for Lebesgue Integral El Teorema Fundamental del C´alculo para la Integral de Lebesgue Di´omedesB´arcenas([email protected]) Departamento de Matem´aticas.Facultad de Ciencias. Universidad de los Andes. M´erida.Venezuela. Abstract In this paper we prove the Theorem announced in the title with- out using Vitali's Covering Lemma and have as a consequence of this approach the equivalence of this theorem with that which states that absolutely continuous functions with zero derivative almost everywhere are constant. We also prove that the decomposition of a bounded vari- ation function is unique up to a constant. Key words and phrases: Radon-Nikodym Theorem, Fundamental Theorem of Calculus, Vitali's covering Lemma. Resumen En este art´ıculose demuestra el Teorema Fundamental del C´alculo para la integral de Lebesgue sin usar el Lema del cubrimiento de Vi- tali, obteni´endosecomo consecuencia que dicho teorema es equivalente al que afirma que toda funci´onabsolutamente continua con derivada igual a cero en casi todo punto es constante. Tambi´ense prueba que la descomposici´onde una funci´onde variaci´onacotada es ´unicaa menos de una constante. Palabras y frases clave: Teorema de Radon-Nikodym, Teorema Fun- damental del C´alculo,Lema del cubrimiento de Vitali. Received: 1999/08/18. Revised: 2000/02/24. Accepted: 2000/03/01. MSC (1991): 26A24, 28A15. Supported by C.D.C.H.T-U.L.A under project C-840-97. 76 Di´omedesB´arcenas 1 Introduction The Fundamental Theorem of Calculus for Lebesgue Integral states that: A function f :[a; b] R is absolutely continuous if and only if it is ! 1 differentiable almost everywhere, its derivative f 0 L [a; b] and, for each t [a; b], 2 2 t f(t) = f(a) + f 0(s)ds: Za This theorem is extremely important in Lebesgue integration Theory and several ways of proving it are found in classical Real Analysis.
    [Show full text]
  • Triangular Numbers /, 3,6, 10, 15, ", Tn,'" »*"
    TRIANGULAR NUMBERS V.E. HOGGATT, JR., and IVIARJORIE BICKWELL San Jose State University, San Jose, California 9111112 1. INTRODUCTION To Fibonacci is attributed the arithmetic triangle of odd numbers, in which the nth row has n entries, the cen- ter element is n* for even /?, and the row sum is n3. (See Stanley Bezuszka [11].) FIBONACCI'S TRIANGLE SUMS / 1 =:1 3 3 5 8 = 2s 7 9 11 27 = 33 13 15 17 19 64 = 4$ 21 23 25 27 29 125 = 5s We wish to derive some results here concerning the triangular numbers /, 3,6, 10, 15, ", Tn,'" »*". If one o b - serves how they are defined geometrically, 1 3 6 10 • - one easily sees that (1.1) Tn - 1+2+3 + .- +n = n(n±M and (1.2) • Tn+1 = Tn+(n+1) . By noticing that two adjacent arrays form a square, such as 3 + 6 = 9 '.'.?. we are led to 2 (1.3) n = Tn + Tn„7 , which can be verified using (1.1). This also provides an identity for triangular numbers in terms of subscripts which are also triangular numbers, T =T + T (1-4) n Tn Tn-1 • Since every odd number is the difference of two consecutive squares, it is informative to rewrite Fibonacci's tri- angle of odd numbers: 221 222 TRIANGULAR NUMBERS [OCT. FIBONACCI'S TRIANGLE SUMS f^-O2) Tf-T* (2* -I2) (32-22) Ti-Tf (42-32) (52-42) (62-52) Ti-Tl•2 (72-62) (82-72) (9*-82) (Kp-92) Tl-Tl Upon comparing with the first array, it would appear that the difference of the squares of two consecutive tri- angular numbers is a perfect cube.
    [Show full text]
  • Degenerate Eigenvalue Problem 32.1 Degenerate Perturbation
    Physics 342 Lecture 32 Degenerate Eigenvalue Problem Lecture 32 Physics 342 Quantum Mechanics I Wednesday, April 23rd, 2008 We have the matrix form of the first order perturbative result from last time. This carries over pretty directly to the Schr¨odingerequation, with only minimal replacement (the inner product and finite vector space change, but notationally, the results are identical). Because there are a variety of quantum mechanical systems with degenerate spectra (like the Hydrogen 2 eigenstates, each En has n associated eigenstates) and we want to be able to predict the energy shift associated with perturbations in these systems, we can copy our arguments for matrices to cover matrices with more than one eigenvector per eigenvalue. The punch line of that program is that we can use the non-degenerate perturbed energies, provided we start with the \correct" degenerate linear combinations. 32.1 Degenerate Perturbation N×N Going back to our symmetric matrix example, we have A IR , and 2 again, a set of eigenvectors and eigenvalues: A xi = λi xi. This time, suppose that the eigenvalue λi has a set of M associated eigenvectors { that is, suppose a set of eigenvectors yj satisfy: A yj = λi yj j = 1 M (32.1) −! 1 of 9 32.1. DEGENERATE PERTURBATION Lecture 32 (so this represents M separate equations) that are themselves orthonormal1. Clearly, any linear combination of these vectors is also an eigenvector: M M X X A βk yk = λi βk yk: (32.2) k=1 k=1 M PM Define the general combination of yi to be z βk yk, also an f gi=1 ≡ k=1 eigenvector of A with eigenvalue λi.
    [Show full text]
  • [Math.FA] 3 Dec 1999 Rnfrneter for Theory Transference Introduction 1 Sas Ihnrah U Twl Etetdi Eaaepaper
    Transference in Spaces of Measures Nakhl´eH. Asmar,∗ Stephen J. Montgomery–Smith,† and Sadahiro Saeki‡ 1 Introduction Transference theory for Lp spaces is a powerful tool with many fruitful applications to sin- gular integrals, ergodic theory, and spectral theory of operators [4, 5]. These methods afford a unified approach to many problems in diverse areas, which before were proved by a variety of methods. The purpose of this paper is to bring about a similar approach to spaces of measures. Our main transference result is motivated by the extensions of the classical F.&M. Riesz Theorem due to Bochner [3], Helson-Lowdenslager [10, 11], de Leeuw-Glicksberg [6], Forelli [9], and others. It might seem that these extensions should all be obtainable via transference methods, and indeed, as we will show, these are exemplary illustrations of the scope of our main result. It is not straightforward to extend the classical transference methods of Calder´on, Coif- man and Weiss to spaces of measures. First, their methods make use of averaging techniques and the amenability of the group of representations. The averaging techniques simply do not work with measures, and do not preserve analyticity. Secondly, and most importantly, their techniques require that the representation is strongly continuous. For spaces of mea- sures, this last requirement would be prohibitive, even for the simplest representations such as translations. Instead, we will introduce a much weaker requirement, which we will call ‘sup path attaining’. By working with sup path attaining representations, we are able to prove a new transference principle with interesting applications. For example, we will show how to derive with ease generalizations of Bochner’s theorem and Forelli’s main result.
    [Show full text]
  • 2 Hilbert Spaces You Should Have Seen Some Examples Last Semester
    2 Hilbert spaces You should have seen some examples last semester. The simplest (finite-dimensional) ex- C • A complex Hilbert space H is a complete normed space over whose norm is derived from an ample is Cn with its standard inner product. It’s worth recalling from linear algebra that if V is inner product. That is, we assume that there is a sesquilinear form ( , ): H H C, linear in · · × → an n-dimensional (complex) vector space, then from any set of n linearly independent vectors we the first variable and conjugate linear in the second, such that can manufacture an orthonormal basis e1, e2,..., en using the Gram-Schmidt process. In terms of this basis we can write any v V in the form (f ,д) = (д, f ), ∈ v = a e , a = (v, e ) (f , f ) 0 f H, and (f , f ) = 0 = f = 0. i i i i ≥ ∀ ∈ ⇒ ∑ The norm and inner product are related by which can be derived by taking the inner product of the equation v = aiei with ei. We also have n ∑ (f , f ) = f 2. ∥ ∥ v 2 = a 2. ∥ ∥ | i | i=1 We will always assume that H is separable (has a countable dense subset). ∑ Standard infinite-dimensional examples are l2(N) or l2(Z), the space of square-summable As usual for a normed space, the distance on H is given byd(f ,д) = f д = (f д, f д). • • ∥ − ∥ − − sequences, and L2(Ω) where Ω is a measurable subset of Rn. The Cauchy-Schwarz and triangle inequalities, • √ (f ,д) f д , f + д f + д , | | ≤ ∥ ∥∥ ∥ ∥ ∥ ≤ ∥ ∥ ∥ ∥ 2.1 Orthogonality can be derived fairly easily from the inner product.
    [Show full text]
  • Introduction Into Quaternions for Spacecraft Attitude Representation
    Introduction into quaternions for spacecraft attitude representation Dipl. -Ing. Karsten Groÿekatthöfer, Dr. -Ing. Zizung Yoon Technical University of Berlin Department of Astronautics and Aeronautics Berlin, Germany May 31, 2012 Abstract The purpose of this paper is to provide a straight-forward and practical introduction to quaternion operation and calculation for rigid-body attitude representation. Therefore the basic quaternion denition as well as transformation rules and conversion rules to or from other attitude representation parameters are summarized. The quaternion computation rules are supported by practical examples to make each step comprehensible. 1 Introduction Quaternions are widely used as attitude represenation parameter of rigid bodies such as space- crafts. This is due to the fact that quaternion inherently come along with some advantages such as no singularity and computationally less intense compared to other attitude parameters such as Euler angles or a direction cosine matrix. Mainly, quaternions are used to • Parameterize a spacecraft's attitude with respect to reference coordinate system, • Propagate the attitude from one moment to the next by integrating the spacecraft equa- tions of motion, • Perform a coordinate transformation: e.g. calculate a vector in body xed frame from a (by measurement) known vector in inertial frame. However, dierent references use several notations and rules to represent and handle attitude in terms of quaternions, which might be confusing for newcomers [5], [4]. Therefore this article gives a straight-forward and clearly notated introduction into the subject of quaternions for attitude representation. The attitude of a spacecraft is its rotational orientation in space relative to a dened reference coordinate system.
    [Show full text]
  • Perturbation Theory and Exact Solutions
    PERTURBATION THEORY AND EXACT SOLUTIONS by J J. LODDER R|nhtdnn Report 76~96 DISSIPATIVE MOTION PERTURBATION THEORY AND EXACT SOLUTIONS J J. LODOER ASSOCIATIE EURATOM-FOM Jun»»76 FOM-INST1TUUT VOOR PLASMAFYSICA RUNHUIZEN - JUTPHAAS - NEDERLAND DISSIPATIVE MOTION PERTURBATION THEORY AND EXACT SOLUTIONS by JJ LODDER R^nhuizen Report 76-95 Thisworkwat performed at part of th«r«Mvchprogmmncof thcHMCiattofiafrccmentof EnratoniOTd th« Stichting voor FundtmenteelOiutereoek der Matctk" (FOM) wtihnnmcWMppoft from the Nederhmdie Organiutic voor Zuiver Wetemchap- pcigk Onderzoek (ZWO) and Evntom It it abo pabHtfMd w a the* of Ac Univenrty of Utrecht CONTENTS page SUMMARY iii I. INTRODUCTION 1 II. GENERALIZED FUNCTIONS DEFINED ON DISCONTINUOUS TEST FUNC­ TIONS AND THEIR FOURIER, LAPLACE, AND HILBERT TRANSFORMS 1. Introduction 4 2. Discontinuous test functions 5 3. Differentiation 7 4. Powers of x. The partie finie 10 5. Fourier transforms 16 6. Laplace transforms 20 7. Hubert transforms 20 8. Dispersion relations 21 III. PERTURBATION THEORY 1. Introduction 24 2. Arbitrary potential, momentum coupling 24 3. Dissipative equation of motion 31 4. Expectation values 32 5. Matrix elements, transition probabilities 33 6. Harmonic oscillator 36 7. Classical mechanics and quantum corrections 36 8. Discussion of the Pu strength function 38 IV. EXACTLY SOLVABLE MODELS FOR DISSIPATIVE MOTION 1. Introduction 40 2. General quadratic Kami1tonians 41 3. Differential equations 46 4. Classical mechanics and quantum corrections 49 5. Equation of motion for observables 51 V. SPECIAL QUADRATIC HAMILTONIANS 1. Introduction 53 2. Hamiltcnians with coordinate coupling 53 3. Double coordinate coupled Hamiltonians 62 4. Symmetric Hamiltonians 63 i page VI. DISCUSSION 1. Introduction 66 ?.
    [Show full text]
  • Chapter 8 Stability Theory
    Chapter 8 Stability theory We discuss properties of solutions of a first order two dimensional system, and stability theory for a special class of linear systems. We denote the independent variable by ‘t’ in place of ‘x’, and x,y denote dependent variables. Let I ⊆ R be an interval, and Ω ⊆ R2 be a domain. Let us consider the system dx = F (t, x, y), dt (8.1) dy = G(t, x, y), dt where the functions are defined on I × Ω, and are locally Lipschitz w.r.t. variable (x, y) ∈ Ω. Definition 8.1 (Autonomous system) A system of ODE having the form (8.1) is called an autonomous system if the functions F (t, x, y) and G(t, x, y) are constant w.r.t. variable t. That is, dx = F (x, y), dt (8.2) dy = G(x, y), dt Definition 8.2 A point (x0, y0) ∈ Ω is said to be a critical point of the autonomous system (8.2) if F (x0, y0) = G(x0, y0) = 0. (8.3) A critical point is also called an equilibrium point, a rest point. Definition 8.3 Let (x(t), y(t)) be a solution of a two-dimensional (planar) autonomous system (8.2). The trace of (x(t), y(t)) as t varies is a curve in the plane. This curve is called trajectory. Remark 8.4 (On solutions of autonomous systems) (i) Two different solutions may represent the same trajectory. For, (1) If (x1(t), y1(t)) defined on an interval J is a solution of the autonomous system (8.2), then the pair of functions (x2(t), y2(t)) defined by (x2(t), y2(t)) := (x1(t − s), y1(t − s)), for t ∈ s + J (8.4) is a solution on interval s + J, for every arbitrary but fixed s ∈ R.
    [Show full text]
  • DEGENERACY CURVES, GAPS, and DIABOLICAL POINTS in the SPECTRA of NEUMANN PARALLELOGRAMS P Overfelt
    DEGENERACY CURVES, GAPS, AND DIABOLICAL POINTS IN THE SPECTRA OF NEUMANN PARALLELOGRAMS P Overfelt To cite this version: P Overfelt. DEGENERACY CURVES, GAPS, AND DIABOLICAL POINTS IN THE SPECTRA OF NEUMANN PARALLELOGRAMS. 2020. hal-03017250 HAL Id: hal-03017250 https://hal.archives-ouvertes.fr/hal-03017250 Preprint submitted on 20 Nov 2020 HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. DEGENERACY CURVES, GAPS, AND DIABOLICAL POINTS IN THE SPECTRA OF NEUMANN PARALLELOGRAMS P. L. OVERFELT Abstract. In this paper we consider the problem of solving the Helmholtz equation over the space of all parallelograms subject to Neumann boundary conditions and determining the degeneracies occurring in their spectra upon changing the two parameters, angle and side ratio. This problem is solved numerically using the finite element method (FEM). Specifically for the lowest eleven normalized eigenvalue levels of the family of Neumann parallelograms, the intersection of two (or more) adjacent eigen- value level surfaces occurs in one of three ways: either as an isolated point associated with the special geometries, i.e., the rectangle, the square, or the rhombus, as part of a degeneracy curve which appears to contain an infinite number of points, or as a diabolical point in the Neumann parallelogram spec- trum.
    [Show full text]
  • Physics 403, Spring 2011 Problem Set 2 Due Thursday, February 17
    Physics 403, Spring 2011 Problem Set 2 due Thursday, February 17 1. A Differential Equation for the Legendre Polynomials (15 pts): In this prob- lem, we will derive a second order linear differential equation satisfied by the Legendre polynomials (−1)n dn P (x) = [(1 − x2)n] : n 2nn! dxn (a) Show that Z 1 2 0 0 Pm(x)[(1 − x )Pn(x)] dx = 0 for m 6= n : (1) −1 Conclude that 2 0 0 [(1 − x )Pn(x)] = λnPn(x) 0 for λn a constant. [ Prime here and in the following questions denotes d=dx.] (b) Perform the integral (1) for m = n to find a relation for λn in terms of the leading n coefficient of the Legendre polynomial Pn(x) = knx + :::. What is kn? 2. A Generating Function for the Legendre Polynomials (20 pts): In this problem, we will derive a generating function for the Legendre polynomials. (a) Verify that the Legendre polynomials satisfy the recurrence relation 0 0 0 Pn − 2xPn−1 + Pn−2 − Pn−1 = 0 : (b) Consider the generating function 1 X n g(x; t) = t Pn(x) : n=0 Use the recurrence relation above to show that the generating function satisfies the first order differential equation d (1 − 2xt + t2) g(x; t) = tg(x; t) : dx Determine the constant of integration by using the fact that Pn(1) = 1. (c) Use the generating function to express the potential for a point particle q jr − r0j in terms of Legendre polynomials. 1 3. The Dirac delta function (15 pts): Dirac delta functions are useful mathematical tools in quantum mechanics.
    [Show full text]