Second Quantization: Quantum Fields

Total Page:16

File Type:pdf, Size:1020Kb

Second Quantization: Quantum Fields Second Quantization: Quantum Fields Bosons and Fermions Let Xj stand for the coordinate and spin subscript (if any) of the j-th particle, so that the vector of state jΨi of N particles has the form jΨi ≡ Ψ(X1;X2;:::;XN ): The postulate of indistinguishability states that the Hilbert space of par- ticles of the same kind1 is restricted to the states featuring permutational symmetry: Ψ(:::;Xi;:::Xj;:::) = ±Ψ(:::;Xj;:::Xi;:::); 8 Xi $ Xj: (1) The sign-plus particles are called bosons and the sign-minus particles are called fermions. Single-particle-based basis states Let fφs(X)g be an ONB of vectors of states for one particle. Then, the basis in the N-particle Hilbert space can be chosen as follows (fermions and bosons are labelled with the subscript F and B, respectively): φs (X1) φs (X2) : : : φs (XN ) 1 1 1 1 φ (X ) φ (X ) : : : φ (X ) s2 1 s2 2 s2 N ΨF (X1;X2;:::;XN ) = p ; (2) N! ······························ φsN (X1) φsN (X2) : : : φsN (XN ) X ΨB(X1;X2;:::;XN ) = C φs1 (X1)φs2 (X2) ··· φsN (XN ): (3) nontrv prm In the fermionic case, we have a compact determinant form, the so-called Slater determinant. For bosons, the summation is over all non-trivial2 per- mutations of pairs of coordinates. The bosonic normalization constant C can 1Referred to as identical particles. 2The permutation is trivial if it does not change the form of the term. This happens if and only if the the corresponding functions φ are the same. 1 be conveniently expressed in terms of the occupation numbers of correspond- ing single-particle states: n ! n ! n ! ··· jCj2 = 1 2 3 : (4) N! Problem 1. Derive Eq. (4) [after reading the next section where the notion of occupation numbers is introduced]. Fock nomenclature and Fock space With the representations (2) and (3) for the basis states, it is quite clear that each basis state is exhaustively characterized by the set of N single-particle states, fφs1 ; φs2 ; : : : ; φsN g involved in the N-particle (anty-)symmetric basis state. For fermions, all the N states must be different.3 For bosons, there are no restrictions. Extremely powerful is the Fock nomenclature for the above-mentioned sets of single-particle states. In this approach, one globally enumerates all the single-particle states of a given set fφsg and then uses a string of natural numbers, (n(1); n(2); n(3);:::), such that 1 X n(s) = N; (5) s=1 to label all the N-particle basis states. The convention is that n(s) = 0, if (s) the state φs does not participate in the given basis state, while 1 ≤ n ≤ N (s) stands for the multiplicity of the state φs. Obviously, for fermions, n can be either 0 or 1. The next important step is to abandon the constraint (5) thus allowing the basis states with different numbers of particles.4 This way we arrive at the Fock space. Each vector of the Fock space can be represented as a linear combination of the Fock basis states. Furthermore, each Fock basis vector can be viewed as a direct product of single-mode (Fock) basis vectors: jn(1); n(2); n(3);:::i = jn(1)ijn(2)ijn(3)i · · · : (6) 3Otherwise, the Slater determinant is identically equal to zero. 4By definition, the basis states with different total numbers of particles are orthogonal to each other. 2 Factorization (6) is central for casting the theory of many-body identical- particle systems in the form of quantum field theory (QFT). Elementary subsystems of which the quantum field consists are not the individual parti- cles. Rather, these are the single-particle modes. For each mode s, there is a complete set of orthonormal states fjn(s)ig, (s) (s) hn1 jn2 i = δ (s) (s) : (7) n1 n2 Recalling that n(s) = 0; 1 for fermions and n(s) = 0; 1; 2;::: for bosons, we see that, nomenclature-wise, the Fock space of fermions can be viewed as a set of two-level systems, while the Fock space of bosons corresponds to that of a set of quantum harmonic oscillators. The analogy between the quantum field of bosons and a set of independent quantum harmonic oscillators becomes essentially direct for a gas of noninteracting particles.5 Within the QFT picture, the number of particles is an observable rather than a fixed external parameter. The operator of the total number of par- ticles, N^, is constructed as a sum of operators of occupation of numbers for each mode: ^ X N = n^s; (8) s (s) (s) (s) n^sjn i = n jn i: (9) In this formalism, the occupation numbers of the mode s are the eigenvalues of the operatorn ^s. Ideal gas Consider now a noninteracting system. Choose fφs(X)g to be an ONB of single-particle energy eigenstates, with s corresponding single-particle en- ergy. Then, in the Fock representation, the Hamiltonian reads: ^ X H = sn^s: (10) s Note a qualitative difference between (10) and (8). While (10) requires that (i) the system be ideal and (ii) the set fφs(X)g be the set of single-particle energy eigenstates, none of the two requirements is relevant to (8). 5No matter whether the system of bosons is uniform or placed in an external potential. 3 Given that, by construction, different operatorsn ^s commute with each other, we conclude that the gas of ideal fermions is equivalent to a set of non-interacting two-level systems, while the ideal bosons are equivalent to a set of non-interacting harmonic oscillators. In both cases, the single-particle energies s play the role of inter-level separations. Gibbs distribution. In particular, the above observation means that the (grand canonical) Gibbs statistical operator6 factorizes into a product of (grand canonical) Gibbs statistical operators for each single-particle mode: ^ 0 Y ^ 0 e−βH = e−βHs : (11) s Indeed, ^ 0 ^ ^ X ^ 0 ^ 0 H = H − µN = Hs; Hs = ~sn^s; ~s = s − µ. (12) s This way we reduce the problem of equilibrium statistics of ideal fermions or bosons to the equilibrium statistics of (respectively) a two-level system or a quantum harmonic oscillator. In the Fock representation, an explicit expression for the statistical oper- ator for the mode s is ^ 0 X (s) e−βHs = jn(s)i e−β~sn hn(s)j: (13) n(s) Up to a difference in the value of ~s, we have the same expression for each mode, so that the subscript s becomes redundant and we can write (^ρ ≡ ^ 0 e−βHs ) ρ^ = X jni e−βn~ hnj: (14) n The normalization constant7 z = Trρ ^ = X e−βn~ (15) n has a very deep statistical-mechanical meaning. This is the so-called partition function of the single mode. The partition function, Z, for the whole (ideal gas) system is then nothing but a product (here we restore the mode label) Y Z = zs: (16) s 6For clarity, we omit the normalization factor. 7The normalization rule isρ ^ ! ρ/z^ . 4 Formula (16) is central for the statistical mechanics of ideal Fermi and Bose gases.8 Problem 2. Calculate single-mode z as a function of (, µ, T ). The average occupation number of the mode, 1 1 n¯ = Trρ ^n^ = X n e−βn~ z z n is given by 1 n¯ = (Fermions); (17) eβ~ + 1 1 n¯ = (Bose): (18) eβ~ − 1 Problem 3. Derive (17) and (18). Hint. The following mathematical trick might be useful. @ X n eλn = X eλn: n @λ n In the limit β~ 1 (that is ~ T ), the difference between bosons and fermions disappears, since the contribution of n > 1 is exponentially suppressed: ρ^ ≈ j0ih0j + j1i e−/T~ h1j; n¯ ≈ e−/T~ (~ T ): (19) This \ultra quantum" (in terms of the statistics of occupation numbers) regime actually corresponds to the regime of Boltzmann gas of classical par- ticles! −~s=T Problem 4. Show thatn ¯s = e corresponds to the Maxwell distribu- tion of classical particles, once s is understood as the momentum eigenvalue, 8In statistical mechanics, one extracts all the thermodynamic quantities from Z, by simply calculating partial derivatives of the function T ln Z with respect to temperature, or/and chemical potential, or/and volume. 5 s ≡ p. Note that it is not sufficient to simply identify the exponential with Maxwell's exponential. It is equally important to demonstrate that the sum- mation over s can be replaced with the integration dp; that is, it is crucial to make sure that there is no nontrivial Jacobian associated with going from P R s(:::) to (:::)dp. Hint. Since, in a macroscopic system, the answer is in- dependent of the system's shape and the boundary conditions, it makes sense to choose the most convenient ones. Namely, a finite-size box of rectangular shape with periodic boundary conditions. For bosons, there is yet another classical limit. It takes place at ~ T and corresponds to the classical limit of corresponding harmonic oscillator. In terms of the QFT perspective, this limit corresponds to the classical-field behavior of corresponding mode. Second quantized form of operators I: Statement of the problem Consistently with the distinguishability postulate, all the relevant opera- tors/observables should be permutational symmetric. The class of permu- tational symmetric operators/observables further splits into single-particle, F^(1), two-particle, F^(2), etc., subclasses: N N N ^(1) X ^(1) ^(2) 1 X X ^(2) F = f (Xj); F = f (Xj;Xk);::: j=1 2 j=1 k=1(6=j) (20) Second quantization is the formalism allowing one to represent the operators F^(1); F^(2);:::; etc.
Recommended publications
  • 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]
  • 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.
    [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]
  • 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]
  • Chapter 7 Second Quantization 7.1 Creation and Annihilation Operators
    Chapter 7 Second Quantization Creation and annihilation operators. Occupation number. Anticommutation relations. Normal product. Wick's theorem. One-body operator in second quantization. Hartree- Fock potential. Two-particle Random Phase Approximation (RPA). Two-particle Tamm- Dankoff Approximation (TDA). 7.1 Creation and annihilation operators In Fig. 3 of the previous Chapter it is shown the single-particle levels generated by a double-magic core containing A nucleons. Below the Fermi level (FL) all states hi are occupied and one can not place a particle there. In other words, the A-particle state j0i, with all levels hi occupied, is the ground state of the inert (frozen) double magic core. Above the FL all states are empty and, therefore, a particle can be created in one of the levels denoted by pi in the Figure. In Second Quantization one introduces the y j i creation operator cpi such that the state pi in the nucleus containing A+1 nucleons can be written as, j i y j i pi = cpi 0 (7.1) this state has to be normalized, that is, h j i h j y j i h j j i pi pi = 0 cpi cpi 0 = 0 cpi pi = 1 (7.2) from where it follows that j0i = cpi jpii (7.3) and the operator cpi can be interpreted as the annihilation operator of a particle in the state jpii. This implies that the hole state hi in the (A-1)-nucleon system is jhii = chi j0i (7.4) Occupation number and anticommutation relations One notices that the number y nj = h0jcjcjj0i (7.5) is nj = 1 is j is a hole state and nj = 0 if j is a particle state.
    [Show full text]
  • SECOND QUANTIZATION Lecture Notes with Course Quantum Theory
    SECOND QUANTIZATION Lecture notes with course Quantum Theory Dr. P.J.H. Denteneer Fall 2008 2 SECOND QUANTIZATION x1. Introduction and history 3 x2. The N-boson system 4 x3. The many-boson system 5 x4. Identical spin-0 particles 8 x5. The N-fermion system 13 x6. The many-fermion system 14 1 x7. Identical spin- 2 particles 17 x8. Bose-Einstein and Fermi-Dirac distributions 19 Second Quantization 1. Introduction and history Second quantization is the standard formulation of quantum many-particle theory. It is important for use both in Quantum Field Theory (because a quantized field is a qm op- erator with many degrees of freedom) and in (Quantum) Condensed Matter Theory (since matter involves many particles). Identical (= indistinguishable) particles −! state of two particles must either be symmetric or anti-symmetric under exchange of the particles. 1 ja ⊗ biB = p (ja1 ⊗ b2i + ja2 ⊗ b1i) bosons; symmetric (1a) 2 1 ja ⊗ biF = p (ja1 ⊗ b2i − ja2 ⊗ b1i) fermions; anti − symmetric (1b) 2 Motivation: why do we need the \second quantization formalism"? (a) for practical reasons: computing matrix elements between N-particle symmetrized wave functions involves (N!)2 terms (integrals); see the symmetrized states below. (b) it will be extremely useful to have a formalism that can handle a non-fixed particle number N, as in the grand-canonical ensemble in Statistical Physics; especially if you want to describe processes in which particles are created and annihilated (as in typical high-energy physics accelerator experiments). So: both for Condensed Matter and High-Energy Physics this formalism is crucial! (c) To describe interactions the formalism to be introduced will be vastly superior to the wave-function- and Hilbert-space-descriptions.
    [Show full text]
  • Second Quantization∗
    Second Quantization∗ Jörg Schmalian May 19, 2016 1 The harmonic oscillator: raising and lowering operators Lets first reanalyze the harmonic oscillator with potential m!2 V (x) = x2 (1) 2 where ! is the frequency of the oscillator. One of the numerous approaches we use to solve this problem is based on the following representation of the momentum and position operators: r x = ~ ay + a b 2m! b b r m ! p = i ~ ay − a : (2) b 2 b b From the canonical commutation relation [x;b pb] = i~ (3) follows y ba; ba = 1 y y [ba; ba] = ba ; ba = 0: (4) Inverting the above expression yields rm! i ba = xb + pb 2~ m! r y m! i ba = xb − pb (5) 2~ m! ∗Copyright Jörg Schmalian, 2016 1 y demonstrating that ba is indeed the operator adjoined to ba. We also defined the operator y Nb = ba ba (6) which is Hermitian and thus represents a physical observable. It holds m! i i Nb = xb − pb xb + pb 2~ m! m! m! 2 1 2 i = xb + pb − [p;b xb] 2~ 2m~! 2~ 2 2 1 pb m! 2 1 = + xb − : (7) ~! 2m 2 2 We therefore obtain 1 Hb = ! Nb + : (8) ~ 2 1 Since the eigenvalues of Hb are given as En = ~! n + 2 we conclude that the eigenvalues of the operator Nb are the integers n that determine the eigenstates of the harmonic oscillator. Nb jni = n jni : (9) y Using the above commutation relation ba; ba = 1 we were able to show that p a jni = n jn − 1i b p y ba jni = n + 1 jn + 1i (10) y The operator ba and ba raise and lower the quantum number (i.e.
    [Show full text]
  • Casimir-Polder Interaction in Second Quantization
    Universitat¨ Potsdam Institut fur¨ Physik und Astronomie Casimir-Polder interaction in second quantization Dissertation zur Erlangung des akademischen Grades doctor rerum naturalium (Dr. rer. nat.) in der Wissen- schaftsdisziplin theoretische Physik. Eingereicht am 21. Marz¨ 2011 an der Mathematisch – Naturwissen- schaftlichen Fakultat¨ der Universitat¨ Potsdam von Jurgen¨ Schiefele Betreuer: PD Dr. Carsten Henkel This work is licensed under a Creative Commons License: Attribution - Noncommercial - Share Alike 3.0 Germany To view a copy of this license visit http://creativecommons.org/licenses/by-nc-sa/3.0/de/ Published online at the Institutional Repository of the University of Potsdam: URL http://opus.kobv.de/ubp/volltexte/2011/5417/ URN urn:nbn:de:kobv:517-opus-54171 http://nbn-resolving.de/urn:nbn:de:kobv:517-opus-54171 Abstract The Casimir-Polder interaction between a single neutral atom and a nearby surface, arising from the (quantum and thermal) fluctuations of the electromagnetic field, is a cornerstone of cavity quantum electrodynamics (cQED), and theoretically well established. Recently, Bose-Einstein condensates (BECs) of ultracold atoms have been used to test the predic- tions of cQED. The purpose of the present thesis is to upgrade single-atom cQED with the many-body theory needed to describe trapped atomic BECs. Tools and methods are developed in a second-quantized picture that treats atom and photon fields on the same footing. We formulate a diagrammatic expansion using correlation functions for both the electromagnetic field and the atomic system. The formalism is applied to investigate, for BECs trapped near surfaces, dispersion interactions of the van der Waals-Casimir-Polder type, and the Bosonic stimulation in spontaneous decay of excited atomic states.
    [Show full text]
  • Notes on Green's Functions Theory for Quantum Many-Body Systems
    Notes on Green's Functions Theory for Quantum Many-Body Systems Carlo Barbieri Department of Physics, University of Surrey, Guildford GU2 7XH, UK June 2015 Literature A modern and compresensive monography is • W. H. Dickhoff and D. Van Neck, Many-Body Theory Exposed!, 2nd ed. (World Scientific, Singapore, 2007). This book covers several applications of Green's functions done in recent years, including nuclear and electroncic systems. Most the material of these lectures can be found here. Two popular textbooks, that cover the almost complete formalism, are • A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Physics (McGraw-Hill, New York, 1971), • A. A. Abrikosov, L. P. Gorkov and I. E. Dzyaloshinski, Methods of Quantum Field Theory in Statistical Physics (Dover, New York, 1975). Other useful books on many-body Green's functions theory, include • R. D. Mattuck, A Guide to Feynmnan Diagrams in the Many-Body Problem, (McGraw-Hill, 1976) [reprinted by Dover, 1992], • J. P. Blaizot and G. Ripka, Quantum Theory of Finite Systems (MIT Press, Cambridge MA, 1986), • J. W. Negele and H. Orland, Quantum Many-Particle Systems (Ben- jamin, Redwood City CA, 1988). 2 Chapter 1 Preliminaries: Basic Formulae of Second Quantization and Pictures Many-body Green's functions (MBGF) are a set of techniques that originated in quantum field theory but have also found wide applications to the many- body problem. In this case, the focus are complex systems such as crystals, molecules, or atomic nuclei. However, many-body Green's functions still share the same language with elementary particles theory, and have several concepts in common.
    [Show full text]
  • Second Quantization I Fermion
    Second quantization for fermions Masatsugu Sei Suzuki Department of Physics SUNY at Binghamton (Date: April 20, 2017) Vladimir Aleksandrovich Fock (December 22, 1898 – December 27, 1974) was a Soviet physicist, who did foundational work on quantum mechanics and quantum electrodynamics. His primary scientific contribution lies in the development of quantum physics, although he also contributed significantly to the fields of mechanics, theoretical optics, theory of gravitation, physics of continuous media. In 1926 he derived the Klein– Gordon equation. He gave his name to Fock space, the Fock representation and Fock state, and developed the Hartree–Fock method in 1930. He made many subsequent scientific contributions, during the rest of his life. Fock developed the electromagnetic methods for geophysical exploration in a book The theory of the study of the rocks resistance by the carottage method (1933); the methods are called the well logging in modern literature. Fock made significant contributions to general relativity theory, specifically for the many body problems. https://en.wikipedia.org/wiki/Vladimir_Fock 1. Introduction What is the definition of the second quantization? In quantum mechanics, the system behaves like wave and like particle (duality). The first quantization is that particles behave like waves. For example, the wave function of electrons is a solution of the Schrödinger equation. Although electromagnetic waves behave like wave, they also behave like particle, as known as photon. This idea is known as second quantization; waves behave like particles. Instead of traditional treatment of a wave function as a solution of the Schrödinger equation in the position representation or representation of some dynamical observable, we will now find a totally different basis formed by number states (or Fock states).
    [Show full text]