Second Quantization

Total Page:16

File Type:pdf, Size:1020Kb

Second Quantization Chapter 1 Second Quantization 1.1 Creation and Annihilation Operators in Quan- tum Mechanics We will begin with a quick review of creation and annihilation operators in the non-relativistic linear harmonic oscillator. Let a and a† be two operators acting on an abstract Hilbert space of states, and satisfying the commutation relation a,a† = 1 (1.1) where by “1” we mean the identity operator of this Hilbert space. The operators a and a† are not self-adjoint but are the adjoint of each other. Let α be a state which we will take to be an eigenvector of the Hermitian operators| ia†a with eigenvalue α which is a real number, a†a α = α α (1.2) | i | i Hence, α = α a†a α = a α 2 0 (1.3) h | | i k | ik ≥ where we used the fundamental axiom of Quantum Mechanics that the norm of all states in the physical Hilbert space is positive. As a result, the eigenvalues α of the eigenstates of a†a must be non-negative real numbers. Furthermore, since for all operators A, B and C [AB, C]= A [B, C] + [A, C] B (1.4) we get a†a,a = a (1.5) − † † † a a,a = a (1.6) 1 2 CHAPTER 1. SECOND QUANTIZATION i.e., a and a† are “eigen-operators” of a†a. Hence, a†a a = a a†a 1 (1.7) − † † † † a a a = a a a +1 (1.8) Consequently we find a†a a α = a a†a 1 α = (α 1) a α (1.9) | i − | i − | i Hence the state aα is an eigenstate of a†a with eigenvalue α 1, provided a α = 0. Similarly,| ai† α is an eigenstate of a†a with eigenvalue α−+1, provided a†| αi 6 = 0. This also implies| i that | i 6 1 α 1 = a α (1.10) | − i √α | i 1 α +1 = a† α (1.11) | i √α +1 | i Let us assume that an α =0, n Z+ (1.12) | i 6 ∀ ∈ Hence, an α is an eigenstate of a†a with eigenvalue α n. However, α n< 0 if α<n, which| i contradicts our earlier result that all these− eigenvalues must− be non-negative real numbers. Hence, for a given α there must exist an integer n such that an α = 0 but an+1 α = 0, where n Z+. Let | i 6 | i ∈ 1 α n = an α a†a α n = (α n) α n (1.13) | − i an α | i ⇒ | − i − | − i k | ik where an α n = √α n (1.14) k | − ik − But 1 a α n = an+1 α =0 α = n (1.15) | − i − an α | i ⇒ k | ik In other words the allowed eigenvalues of a†a are the non-negative integers. Let us now define the ground state 0 , as the state annihilated by a, | i a 0 = 0 (1.16) | i Then, an arbitrary state n is | i 1 n = (a†)n 0 (1.17) | i √n! | i which has the inner product n m = δ n! (1.18) h | i n,m 1.1. CREATION AND ANNIHILATION OPERATORS IN QUANTUM MECHANICS3 In summary, we found that creation and annihilation operators obey a† n = √n +1 n +1 (1.19) | i | i a n = √n n 1 (1.20) | i | − i a†a n = n n (1.21) | i | i (1.22) and thus their matrix elements are m a† n = √n +1 δ m a n = √nδ (1.23) h | | i m,n+1 h | | i m,n−1 1.1.1 The Linear harmonic Oscillator The Hamiltonian of the Linear Harmonic Oscillator is P 2 1 H = + mω2X2 (1.24) 2m 2 where X and P , the coordinate and momentum Hermitian operators satisfy canonical commutation relations, [X, P ]= i~ (1.25) We now define the creation and annihilation operators a† and a as 1 mω P a = X + i (1.26) √2 ~ √ ~ r mω 1 mω P a† = X i (1.27) √ ~ − √ ~ 2 r mω which satisfy a,a† = 1 (1.28) Since ~ X = a + a† (1.29) r2mω mω~ a a† P = − (1.30) r 2 i the Hamiltonian takes the simple form 1 H = ~ω a†a + (1.31) 2 The eigenstates of the Hamiltonian are constructed easily using our results since all eigenstates of a†a are eigenstates of H. Thus, the eigenstates of H are the eigenstates of a†a, 1 H n = ~ω n + n (1.32) | i 2 | i 4 CHAPTER 1. SECOND QUANTIZATION with eigenvalues 1 E = ~ω n + (1.33) n 2 where n =0, 1,.... The ground state 0 is the state annihilated by a, | i 1 mω P a 0 X + i 0 = 0 (1.34) | i≡ 2 ~ √ ~ | i r mω Since d x P φ = i~ x φ (1.35) h | | i − dxh | i we find that ψ (x)= x 0 satisfies 0 h | i ~ d x + ψ (x) = 0 (1.36) mω dx 0 whose (normalized) solution is mω mω 1/4 x2 ψ (x)= e− 2~ (1.37) 0 π~ The wave functions ψ (x) of the excited states n are n | i 1 † n ψn(x) = x n = x (a ) 0 h | i √n! h | | i 1 mω n/2 ~ d n = x ψ0(x) (1.38) √n! 2~ − mω dx Creation and annihilation operators are very useful. Let us consider for instance the anharmonic oscillator whose Hamiltonian is P 2 1 H = + mω2X2 + λX4 (1.39) 2m 2 Let us compute the eigenvalues En to lowest order in perturbation theory in powers of λ. The first order shift ∆En is ∆E = λ n X4 n + O(λ2) n h | | i ~ 2 = λ n (a + a†)4 n + . 2mω h | | i ~ 2 = λ n a†a†aa n + other terms with two a′s and two a†′s + . 2mω h | | i ~ 2 = λ 6n2 +6n +3 + . (1.40) 2mω 1.1. CREATION AND ANNIHILATION OPERATORS IN QUANTUM MECHANICS5 1.1.2 Many Harmonic Oscillators It is trivial to extend these ideas to the case of many harmonic oscillators, which is a crude model of an elastic solid. Consider a system of N identical linear harmonic oscillators of mass M and frequency ω, with coordinates Q and { i} momenta Pi , where i = 1,...,N. These operators satisfy the commutation relations { } [Qj,Qk] = [Pj , Pk]=0, [Qj , Pk]= i~δjk (1.41) where j, k =1,...,N. The Hamiltonian is N N P 2 1 H = i + V Q Q (1.42) 2M 2 ij i j i=1 i i,j=1 X X where Vij is a symmetric positive definite matrix, Vij = Vji. We will find the spectrum (and eigenstates) of this system by changing vari- ables to normal modes and using creation and annihilation operators for the normal modes. To this end we will first rescale coordinates and momenta so as to absorb the particle mass M: xi = MiQi (1.43) Pi pi = p (1.44) √Mi 1 Uij = Vij (1.45) MiMj which also satisfy p [xj ,pj]= i~δjk (1.46) and N N p2 1 H = i + U x x (1.47) 2 2 ij i j i=1 i,j=1 X X We now got to normal mode variables by means of an orthogonal transformation −1 Cjk, i.e. [C ]jk = Ckj , N N xk = Ckj xj , pk = Ckj pj j=1 j=1 X X Ne eN CkiCji = δkj , CikCij = δkj (1.48) i=1 i=1 X X Sine the matrix Uij is real symmetric and positive definite, its eigenvalues, which 2 2 we will denote by ωk (with k = 1,...,N), are all non-negative, ωk 0. The eigenvalue equation is ≥ N 2 CkiCℓj Uij = ωkδkℓ, (no sum over k) (1.49) i,j=1 X 6 CHAPTER 1. SECOND QUANTIZATION Since the transformation is orthogonal, it preserves the commutation relations [xj , xk] = [pj , pk]=0, [xj , pk]= i~δjk (1.50) and the Hamiltoniane ise now diagonale e e e 1 N H = p2 + ω2 x2 (1.51) 2 j j j i=1 X We now define creation and annihilatione operatorse for the normal modes 1 i aj = √ωj xj + pj √ ~ ω 2 √ j † 1 i a = √ωj xej pej j √ ~ − ω 2 √ j ~ e † e xj = aj + aj s2ωj e ~ωj p = i a a† (1.52) j − 2 j − j r where, once again, e † † † [aj,ak] aj ,ak =0, aj ,ak = δjk (1.53) h i h i and the normal mode Hamiltonian takes the standard form N 1 H = ~ω a†a + (1.54) j j j 2 j=1 X † The eigenstates of the Hamiltonian are labelled by the eigenvalues of aj aj for each normal mode j, n ,...,n n . Hence, | 1 N i≡|{ j }i N 1 H n ,...,n = ~ω n + n ,...,n (1.55) | 1 N i j j 2 | 1 N i j=1 X where † N (a )nj n ,...,n = j 0,..., 0 (1.56) | 1 N i | i j=1 nj ! Y The ground state of the system, whichp we will denote by 0 , is the state in which all normal modes are in their ground state, | i 0 0,..., 0 (1.57) | i ≡ | i 1.1. CREATION AND ANNIHILATION OPERATORS IN QUANTUM MECHANICS7 Thus, the ground state 0 is annihilated by the annihilation operators of all normal modes, | i a 0 =0, j (1.58) j | i ∀ and the ground state energy of the system Egnd is N 1 E = ~ω (1.59) gnd 2 j j=1 X The energy of the excited states is N E(n1,...,nN )= ~ωjnj + Egnd (1.60) j=1 X We can now regard the state 0 as the vacuum state and the excited states | i n1,...,nN as a state with nj excitations (or particles) of type j.
Recommended publications
  • Quantum Field Theory with No Zero-Point Energy
    Published for SISSA by Springer Received: March 16, 2018 Revised: May 20, 2018 Accepted: May 27, 2018 Published: May 30, 2018 JHEP05(2018)191 Quantum field theory with no zero-point energy John R. Klauder Department of Physics, University of Florida, Gainesville, FL 32611-8440, U.S.A. Department of Mathematics, University of Florida, Gainesville, FL 32611-8440, U.S.A. E-mail: [email protected] Abstract: Traditional quantum field theory can lead to enormous zero-point energy, which markedly disagrees with experiment. Unfortunately, this situation is built into con- ventional canonical quantization procedures. For identical classical theories, an alternative quantization procedure, called affine field quantization, leads to the desirable feature of having a vanishing zero-point energy. This procedure has been applied to renormalizable and nonrenormalizable covariant scalar fields, fermion fields, as well as general relativity. Simpler models are offered as an introduction to affine field quantization. Keywords: Nonperturbative Effects, Renormalization Regularization and Renormalons, Models of Quantum Gravity ArXiv ePrint: 1803.05823 Open Access, c The Authors. 3 https://doi.org/10.1007/JHEP05(2018)191 Article funded by SCOAP . Contents 1 Introduction 1 2 Nonrenormalizable models exposed 2 3 Ultralocal scalar model: a nonrenormalizable example 3 4 Additional studies using affine variables 5 JHEP05(2018)191 1 Introduction The source of zero-point energy for a free scalar field is readily canceled by normal order- ing of the Hamiltonian operator, which leaves every term with an annihilation operator. Unfortunately, for a non-free scalar field, say with a quartic potential term, normal order- ing cannot eliminate the zero-point energy since there is always a term composed only of creation operators.
    [Show full text]
  • Stochastic Quantization of Fermionic Theories: Renormalization of the Massive Thirring Model
    Instituto de Física Teórica IFT Universidade Estadual Paulista October/92 IFT-R043/92 Stochastic Quantization of Fermionic Theories: Renormalization of the Massive Thirring Model J.C.Brunelli Instituto de Física Teórica Universidade Estadual Paulista Rua Pamplona, 145 01405-900 - São Paulo, S.P. Brazil 'This work was supported by CNPq. Instituto de Física Teórica Universidade Estadual Paulista Rua Pamplona, 145 01405 - Sao Paulo, S.P. Brazil Telephone: 55 (11) 288-5643 Telefax: 55(11)36-3449 Telex: 55 (11) 31870 UJMFBR Electronic Address: [email protected] 47553::LIBRARY Stochastic Quantization of Fermionic Theories: 1. Introduction Renormalization of the Massive Thimng Model' The stochastic quantization method of Parisi-Wu1 (for a review see Ref. 2) when applied to fermionic theories usually requires the use of a Langevin system modified by the introduction of a kernel3 J. C. Brunelli (1.1a) InBtituto de Física Teórica (1.16) Universidade Estadual Paulista Rua Pamplona, 145 01405 - São Paulo - SP where BRAZIL l = 2Kah(x,x )8(t - ?). (1.2) Here tj)1 tp and the Gaussian noises rj, rj are independent Grassmann variables. K(xty) is the aforementioned kernel which ensures the proper equilibrium limit configuration for Accepted for publication in the International Journal of Modern Physics A. mas si ess theories. The specific form of the kernel is quite arbitrary but in what follows, we use K(x,y) = Sn(x-y)(-iX + ™)- Abstract In a number of cases, it has been verified that the stochastic quantization procedure does not bring new anomalies and that the equilibrium limit correctly reproduces the basic (jJsfsini g the Langevin approach for stochastic processes we study the renormalizability properties of the models considered4.
    [Show full text]
  • Arxiv:1512.08882V1 [Cond-Mat.Mes-Hall] 30 Dec 2015
    Topological Phases: Classification of Topological Insulators and Superconductors of Non-Interacting Fermions, and Beyond Andreas W. W. Ludwig Department of Physics, University of California, Santa Barbara, CA 93106, USA After briefly recalling the quantum entanglement-based view of Topological Phases of Matter in order to outline the general context, we give an overview of different approaches to the classification problem of Topological Insulators and Superconductors of non-interacting Fermions. In particular, we review in some detail general symmetry aspects of the "Ten-Fold Way" which forms the foun- dation of the classification, and put different approaches to the classification in relationship with each other. We end by briefly mentioning some of the results obtained on the effect of interactions, mainly in three spatial dimensions. I. INTRODUCTION Based on the theoretical and, shortly thereafter, experimental discovery of the Z2 Topological Insulators in d = 2 and d = 3 spatial dimensions dominated by spin-orbit interactions (see [1,2] for a review), the field of Topological Insulators and Superconductors has grown in the last decade into what is arguably one of the most interesting and stimulating developments in Condensed Matter Physics. The field is now developing at an ever increasing pace in a number of directions. While the understanding of fully interacting phases is still evolving, our understanding of Topological Insulators and Superconductors of non-interacting Fermions is now very well established and complete, and it serves as a stepping stone for further developments. Here we will provide an overview of different approaches to the classification problem of non-interacting Fermionic Topological Insulators and Superconductors, and exhibit connections between these approaches which address the problem from very different angles.
    [Show full text]
  • Chapter 4 Introduction to Many-Body Quantum Mechanics
    Chapter 4 Introduction to many-body quantum mechanics 4.1 The complexity of the quantum many-body prob- lem After learning how to solve the 1-body Schr¨odinger equation, let us next generalize to more particles. If a single body quantum problem is described by a Hilbert space of dimension dim = d then N distinguishable quantum particles are described by theH H tensor product of N Hilbert spaces N (N) N ⊗ (4.1) H ≡ H ≡ H i=1 O with dimension dN . As a first example, a single spin-1/2 has a Hilbert space = C2 of dimension 2, but N spin-1/2 have a Hilbert space (N) = C2N of dimensionH 2N . Similarly, a single H particle in three dimensional space is described by a complex-valued wave function ψ(~x) of the position ~x of the particle, while N distinguishable particles are described by a complex-valued wave function ψ(~x1,...,~xN ) of the positions ~x1,...,~xN of the particles. Approximating the Hilbert space of the single particle by a finite basis set with d basis functions, the N-particle basisH approximated by the same finite basis set for single particles needs dN basis functions. This exponential scaling of the Hilbert space dimension with the number of particles is a big challenge. Even in the simplest case – a spin-1/2 with d = 2, the basis for N = 30 spins is already of of size 230 109. A single complex vector needs 16 GByte of memory and will not fit into the memory≈ of your personal computer anymore.
    [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 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [Show full text]
  • Representations of U(2O) and the Value of the Fine Structure Constant
    Symmetry, Integrability and Geometry: Methods and Applications Vol. 1 (2005), Paper 028, 8 pages Representations of U(2∞) and the Value of the Fine Structure Constant William H. KLINK Department of Physics and Astronomy, University of Iowa, Iowa City, Iowa, USA E-mail: [email protected] Received September 28, 2005, in final form December 17, 2005; Published online December 25, 2005 Original article is available at http://www.emis.de/journals/SIGMA/2005/Paper028/ Abstract. A relativistic quantum mechanics is formulated in which all of the interactions are in the four-momentum operator and Lorentz transformations are kinematic. Interactions are introduced through vertices, which are bilinear in fermion and antifermion creation and annihilation operators, and linear in boson creation and annihilation operators. The fermion- antifermion operators generate a unitary Lie algebra, whose representations are fixed by a first order Casimir operator (corresponding to baryon number or charge). Eigenvectors and eigenvalues of the four-momentum operator are analyzed and exact solutions in the strong coupling limit are sketched. A simple model shows how the fine structure constant might be determined for the QED vertex. Key words: point form relativistic quantum mechanics; antisymmetric representations of infinite unitary groups; semidirect sum of unitary with Heisenberg algebra 2000 Mathematics Subject Classification: 22D10; 81R10; 81T27 1 Introduction With the exception of QCD and gravitational self couplings, all of the fundamental particle interaction Hamiltonians have the form of bilinears in fermion and antifermion creation and annihilation operators times terms linear in boson creation and annihilation operators. For example QED is a theory bilinear in electron and positron creation and annihilation operators and linear in photon creation and annihilation operators.
    [Show full text]
  • Angular Momentum 23.1 Classical Description
    Physics 342 Lecture 23 Angular Momentum Lecture 23 Physics 342 Quantum Mechanics I Monday, March 31st, 2008 We know how to obtain the energy of Hydrogen using the Hamiltonian op- erator { but given a particular En, there is degeneracy { many n`m(r; θ; φ) have the same energy. What we would like is a set of operators that allow us to determine ` and m. The angular momentum operator L^, and in partic- 2 ular the combination L and Lz provide precisely the additional Hermitian observables we need. 23.1 Classical Description Going back to our Hamiltonian for a central potential, we have p · p H = + U(r): (23.1) 2 m It is clear from the dependence of U on the radial distance only, that angular momentum should be conserved in this setting. Remember the definition: L = r × p; (23.2) and we can write this in indexed notation which is slightly easier to work with using the Levi-Civita symbol, defined as (in three dimensions) 8 < 1 for (ijk) an even permutation of (123). ijk = −1 for an odd permutation (23.3) : 0 otherwise Then we can write 3 3 X X Li = ijk rj pk = ijk rj pk (23.4) j=1 k=1 1 of 9 23.2. QUANTUM COMMUTATORS Lecture 23 where the sum over j and k is implied in the second equality (this is Einstein summation notation). There are three numbers here, Li for i = 1; 2; 3, we associate with Lx, Ly, and Lz, the individual components of angular momentum about the three spatial axes.
    [Show full text]
  • Orthogonalization of Fermion K-Body Operators and Representability
    Orthogonalization of fermion k-Body operators and representability Bach, Volker Rauch, Robert April 10, 2019 Abstract The reduced k-particle density matrix of a density matrix on finite- dimensional, fermion Fock space can be defined as the image under the orthogonal projection in the Hilbert-Schmidt geometry onto the space of k- body observables. A proper understanding of this projection is therefore intimately related to the representability problem, a long-standing open problem in computational quantum chemistry. Given an orthonormal basis in the finite-dimensional one-particle Hilbert space, we explicitly construct an orthonormal basis of the space of Fock space operators which restricts to an orthonormal basis of the space of k-body operators for all k. 1 Introduction 1.1 Motivation: Representability problems In quantum chemistry, molecules are usually modeled as non-relativistic many- fermion systems (Born-Oppenheimer approximation). More specifically, the Hilbert space of these systems is given by the fermion Fock space = f (h), where h is the (complex) Hilbert space of a single electron (e.g. h = LF2(R3)F C2), and the Hamiltonian H is usually a two-body operator or, more generally,⊗ a k-body operator on . A key physical quantity whose computation is an im- F arXiv:1807.05299v2 [math-ph] 9 Apr 2019 portant task is the ground state energy . E0(H) = inf ϕ(H) (1) ϕ∈S of the system, where ( )′ is a suitable set of states on ( ), where ( ) is the Banach space ofS⊆B boundedF operators on and ( )′ BitsF dual. A directB F evaluation of (1) is, however, practically impossibleF dueB F to the vast size of the state space .
    [Show full text]
  • Chapter 4 MANY PARTICLE SYSTEMS
    Chapter 4 MANY PARTICLE SYSTEMS The postulates of quantum mechanics outlined in previous chapters include no restrictions as to the kind of systems to which they are intended to apply. Thus, although we have considered numerous examples drawn from the quantum mechanics of a single particle, the postulates themselves are intended to apply to all quantum systems, including those containing more than one and possibly very many particles. Thus, the only real obstacle to our immediate application of the postulates to a system of many (possibly interacting) particles is that we have till now avoided the question of what the linear vector space, the state vector, and the operators of a many-particle quantum mechanical system look like. The construction of such a space turns out to be fairly straightforward, but it involves the forming a certain kind of methematical product of di¤erent linear vector spaces, referred to as a direct or tensor product. Indeed, the basic principle underlying the construction of the state spaces of many-particle quantum mechanical systems can be succinctly stated as follows: The state vector à of a system of N particles is an element of the direct product space j i S(N) = S(1) S(2) S(N) ­ ­ ¢¢¢­ formed from the N single-particle spaces associated with each particle. To understand this principle we need to explore the structure of such direct prod- uct spaces. This exploration forms the focus of the next section, after which we will return to the subject of many particle quantum mechanical systems. 4.1 The Direct Product of Linear Vector Spaces Let S1 and S2 be two independent quantum mechanical state spaces, of dimension N1 and N2, respectively (either or both of which may be in…nite).
    [Show full text]
  • General Quantization
    General quantization David Ritz Finkelstein∗ October 29, 2018 Abstract Segal’s hypothesis that physical theories drift toward simple groups follows from a general quantum principle and suggests a general quantization process. I general- quantize the scalar meson field in Minkowski space-time to illustrate the process. The result is a finite quantum field theory over a quantum space-time with higher symmetry than the singular theory. Multiple quantification connects the levels of the theory. 1 Quantization as regularization Quantum theory began with ad hoc regularization prescriptions of Planck and Bohr to fit the weird behavior of the electromagnetic field and the nuclear atom and to handle infinities that blocked earlier theories. In 1924 Heisenberg discovered that one small change in algebra did both naturally. In the early 1930’s he suggested extending his algebraic method to space-time, to regularize field theory, inspiring the pioneering quantum space-time of Snyder [43]. Dirac’s historic quantization program for gravity also eliminated absolute space-time points from the quantum theory of gravity, leading Bergmann too to say that the world point itself possesses no physical reality [5, 6]. arXiv:quant-ph/0601002v2 22 Jan 2006 For many the infinities that still haunt physics cry for further and deeper quanti- zation, but there has been little agreement on exactly what and how far to quantize. According to Segal canonical quantization continued a drift of physical theory toward simple groups that special relativization began. He proposed on Darwinian grounds that further quantization should lead to simple groups [32]. Vilela Mendes initiated the work in that direction [37].
    [Show full text]
  • Fock Space Dynamics
    TCM315 Fall 2020: Introduction to Open Quantum Systems Lecture 10: Fock space dynamics Course handouts are designed as a study aid and are not meant to replace the recommended textbooks. Handouts may contain typos and/or errors. The students are encouraged to verify the information contained within and to report any issue to the lecturer. CONTENTS Introduction 1 Composite systems of indistinguishable particles1 Permutation group S2 of 2-objects 2 Admissible states of three indistiguishable systems2 Parastatistics 4 The spin-statistics theorem 5 Fock space 5 Creation and annihilation of particles 6 Canonical commutation relations 6 Canonical anti-commutation relations 7 Explicit example for a Fock space with 2 distinguishable Fermion states7 Central oscillator of a linear boson system8 Decoupling of the one particle sector 9 References 9 INTRODUCTION The chapter 5 of [6] oers a conceptually very transparent introduction to indistinguishable particle kinematics in quantum mechanics. Particularly valuable is the brief but clear discussion of para-statistics. The last sections of the chapter are devoted to expounding, physics style, the second quantization formalism. Chapter I of [2] is a very clear and detailed presentation of second quantization formalism in the form that is needed in mathematical physics applications. Finally, the dynamics of a central oscillator in an external eld draws from chapter 11 of [4]. COMPOSITE SYSTEMS OF INDISTINGUISHABLE PARTICLES The composite system postulate tells us that the Hilbert space of a multi-partite system is the tensor product of the Hilbert spaces of the constituents. In other words, the composite system Hilbert space, is spanned by an orthonormal basis constructed by taking the tensor product of the elements of the orthonormal bases of the Hilbert spaces of the constituent systems.
    [Show full text]