Second Quantization

Total Page:16

File Type:pdf, Size:1020Kb

Second Quantization Appendix A Second Quantization An exact description of an interacting many-body system requires the solving of the corresponding many-body Schrodinger¨ equations. The formalism of second quantization leads to a substantial simplification in the description of such a many- body system, but it should be noted that it is only a reformulation of the original Schrodinger¨ equation but not yet a solution. The essential step in the second quanti- zation is the introduction of so-called creation and annihilation operators. By doing this, we eliminate the need for the laborious construction, respectively, of the sym- metrized or the anti-symmetrized N-particle wavefunctions from the single-particle wavefunctions. The entire statistics is contained in fundamental commutation rela- tions of these operators. Forces and interactions are expressed in terms of these “creation” and “annihilation” operators. How does one handle an N-particle system? In case the particles are distinguish- able, that is, if they are enumerable, then the method of description follows directly the general postulates of quantum mechanics: H(i) {|φ(i) } 1 : Hilbert space of the ith particle with the orthonormal basis α : (i) (i) φα |φβ =δαβ (A.1) HN : Hilbert space of the N-particle system H = H(1) ⊗ H(2) ⊗···⊗H(N) N 1 1 1 (A.2) with the basis {|φN }: |φ =|φ(1)φ(2) ···φ(N) =|φ(1) |φ(2) ···|φ(N) (A.3) N α1 α2 αN α1 α2 αN An arbitrary N-particle state |ψN , |ψ = c(α ···α )|φ(1)φ(2) ···φ(N) (A.4) N 1 N α1 α2 αN α1···αN underlies the same statistical interpretation as in the case of a 1-particle system. The dynamics of the N-particle system results from the formally unchanged Schrodinger¨ equation: W. Nolting, A. Ramakanth, Quantum Theory of Magnetism, 491 DOI 10.1007/978-3-540-85416-6 BM2, C Springer-Verlag Berlin Heidelberg 2009 492 A Second Quantization ∂|ψ i N = H |ψ (A.5) ∂t N The handling of the many-body problem in quantum mechanics, in the case of distinguishable particles, confronts exactly the same difficulties as in the classical physics, simply because of the large number of degrees of freedom. There are no extra, typically quantum mechanical complications. A.1 Identical Particles What are “identical particles”? To avoid misunderstandings let us strictly separate “particle properties” from “measured quantities of particle observables”. “Parti- cle properties” as e.g. mass, spin, charge, magnetic moment, etc. are in principle unchangeable intrinsic characteristics of the particle. The “measured values of par- ticle observables” as e.g. position, momentum, angular momentum, spin projection, etc., on the other hand, can always change with time. We define “identical particles” in the quantum mechanical sense as particles which agree in all their particle prop- erties. Identical particles are therefore particles, which behave exactly in the same manner under the same physical conditions, i.e. no measurement can differentiate one from the other. Identical particles also exist in classical mechanics. However, if the initial conditions are known, the state of the system for all times is determined by the Hamilton’s equations of motion. Thus the particles are always identifiable! In quantum mechanics, on the contrary, there is the principle of indistinguishability, which says that, identical particles are intrinsically indistinguishable. In quantum mechanics, in some sense, identical particles lose their individuality. This originates from the fact that there do not exist sharp particle trajectories (only “spreading” wave packets!). The regions, where the probability of finding the particle is unequal to zero, overlap for different particles. Any question whose answer requires the observation of one single particle is physically meaningless. We face now the problem, that, essentially for calculational purposes, we can- not avoid, to number the particles. Then, however, the numbering must be so that all physically relevant statements, that are made, are absolutely invariant under a change of this particle labelling. How to manage this is the subject of the following considerations. We introduce the permutation operator P, which, in the N-particle state, inter- changes the particle indices: P|φ(1)φ(2) ···φ(N) =|φ(i1)φ(i2) ···φ(iN ) (A.6) α1 α2 αN α1 α2 αN Every permutation operator can be written as a product of transposition operators Pij. On application of Pij to a state of identical particles, results, according to the principle of indistinguishability, in a state, which, at the most, is different from the initial state by an unimportant phase factor λ = exp(iη): A.1 Identical Particles 493 P |···φ(i) ···φ( j) ··· = |···φ( j) ···φ(i) ··· ij αi α j αi α j = λ |···φ(i) ···φ( j) ··· (A.7) αi α j 2 = λ =± Since Pij 1, it is necessary that 1. Therefore, a system of identical particles must be either symmetric or antisymmetric against interchange of particles! This defines two different Hilbert spaces: H(+) |ψ (+) N : the space of symmetric states N (+) (+) Pij |ψN =|ψN (A.8) H(−) |ψ (−) N : the space of antisymmetric states N (−) (−) Pij |ψN =−|ψN (A.9) In these spaces, Pij are Hermitian and unitary! What are the properties the observables must have for a system of identical par- ticles? They must necessarily depend on the coordinates of all the particles A = A(1, 2, ··· , N) (A.10) and must commute with all the transpositions (permutations) , = Pij A − 0 (A.11) This is valid specially for the Hamiltonian H and therefore also for the time evolu- tion operator i U t, t = exp − H t − t H = H t ( 0) ( 0) ;( ( )) (A.12) , = Pij U − 0 (A.13) That means the symmetry character of an N-particle state remains unchanged for all times! Which Hilbert space out of H(+) and H(−) is applicable for which type of particles is established in relativistic quantum field theory. We will take over the spin-statistics theorem from there, without any proof. H(+) N : Space of symmetric states of N identical particles with integral spin. These particles are called Bosons. H(−) N : Space of antisymmetric states of N identical particles with half-integral spin. These particles are called Fermions. 494 A Second Quantization A.2 Continuous Fock Representation A.2.1 Symmetrized Many-Particle States H(ε) Let N be the Hilbert space of a system of N identical particles. Here + : Bosons ε = (A.14) − : Fermions Let φ be a 1-particle observable (or a set of 1-particle observables) with a continuous spectrum, φα being a particular eigenvalue corresponding to the 1-particle eigenstate |φα : φ |φα =φα |φα (A.15) φα|φβ =δ(φα − φβ )(≡ δ(α − β)) (A.16) dφα|φα φα|=1l i n H1 (A.17) H(ε) A basis of N is constructed from the following (anti-) symmetrized N-particle states: ) * (ε) 1 p (1) (2) (N) |φα ···φα = ε P |φ |φ ···|φ (A.18) 1 N α1 α2 αN N! P p is the number of transpositions in P. The summation runs over all the possible permutations P. The sequence of the 1-particle states in the ket on the left-hand side of (A.18) is called the standard ordering. It is arbitrary, but has to be fixed right at the beginning. Interchange of two 1-particle symbols on the left means only a constant factor ε on the right. One can easily prove the following relations for the above introduced N-particle states: Scalar product: (ε) φ ···|φ ··· (ε) β1 α1 ) * 1 pα = ε Pα δ(φβ − φα ) ···δ(φβ − φα ) (A.19) N! 1 1 N N Pα The index α shall indicate that Pα permutes only the φα’s. Completeness relation: ··· φ ··· φ |φ ··· (ε)(ε) φ ···|= d β1 d βN β1 β1 1l (A.20) A.2 Continuous Fock Representation 495 This leads to a formal representation of an observable of the N-particle system, which will be used later a few times: = ··· φ ··· φ φ ··· φ ∗ A d α1 d αN d β1 d βN ∗|φ ··· (ε)(ε) φ ···| |φ ··· (ε)(ε) φ ···| α1 α1 A β1 β1 (A.21) A.2.2 Construction Operators H(ε) | We want to build up the basis states of N , step by step, from the vacuum state 0 ( 0|0 =1) with the help of the operator † ≡ † cφα cα These operators are defined by their action on the states: √ † (ε) (ε) c |0 = 1|φα ∈ H (A.22) α1 1 1 √ † (ε) (ε) (ε) c |φα = 2|φα φα ∈ H (A.23) α2 1 2 1 2 Or, in general √ † (ε) (ε) c |φα ···φα = N + 1 |φβ φα ···φα (A.24) β 1 N 1 N ∈H(ε) ∈H(ε) N N+1 † cβ is called the creation operator.Itcreates an extra particle in the N-particle state. The relation (A.24) is obviously reversible: (ε) 1 † † † |φα φα φα ···φα = √ cα cα ···cα |0 (A.25) 1 2 3 N N! 1 2 N |φ ··· (ε) | The N-particle state αi can be built up from the vacuum state 0 by apply- ing a sequence of N creation operators, where the order of the operators is to be strictly obeyed. For a product of two creation operators, it follows from (A.24) † † (ε) (ε) c c |φα ···φα = N(N − 1)|φα φα φα ···φα (A.26) α1 α2 3 N 1 2 3 N If the sequence of the operators is reversed, then we have † † (ε) c c |φα ···φα α2 α1 3 N = − |φ φ φ ···φ (ε) N(N 1) α2 α1 α3 αN = ε − |φ φ φ ···φ (ε) N(N 1) α1 α2 α3 αN (A.27) 496 A Second Quantization The last step follows because of (A.18).
Recommended publications
  • Introduction to the Many-Body Problem
    INTRODUCTION TO THE MANY-BODY PROBLEM UNIVERSITY OF FRIBOURG SPRING TERM 2010 Dionys Baeriswyl General literature A. A. Abrikosov, L. P. Gorkov and I. E. Dzyaloshinski, Methods of Quantum Field Theory in Statistical Physics, Prentice-Hall 1963. A. L. Fetter and J. D. Walecka, Quantum Theory of Many-Particle Systems, McGraw-Hill 1971. G. D. Mahan, Many-Particle Physics, Plenum Press 1981. J. W. Negele and H. Orland, Quantum Many Particle Systems, Perseus Books 1998. Ph. A. Martin and F. Rothen, Many-Body Problems and Quantum Field Theory, Springer-Verlag 2002. H. Bruus and K. Flensberg, Many-Body Quantum Theory in Condensed Matter Physics, Oxford University Press 2004. Contents 1 Second quantization 2 1.1 Many-particlestates ......................... 2 1.2 Fockspace............................... 3 1.3 Creationandannihilationoperators . 5 1.4 Quantumfields ............................ 7 1.5 Representationofobservables . 10 1.6 Wick’stheorem ............................ 16 2 Many-boson systems 19 2.1 Bose-Einstein condensation in a trap . 19 2.2 TheweaklyinteractingBosegas. 20 2.3 TheGross-Pitaevskiiequation . 23 3 Many-electron systems 26 3.1 Thejelliummodel........................... 26 3.2 Hartree-Fockapproximation . 27 3.3 TheWignercrystal .......................... 28 3.4 TheHubbardmodel ......................... 29 4 Magnetism 34 4.1 Exchange ............................... 34 4.2 Magnetic order in the Heisenberg model . 36 5 Electrons and phonons 39 5.1 Theharmoniccrystal.. .. .. 39 5.2 Electron-phononinteraction . 41 5.3 Phonon-inducedattraction
    [Show full text]
  • Identical Particles
    8.06 Spring 2016 Lecture Notes 4. Identical particles Aram Harrow Last updated: May 19, 2016 Contents 1 Fermions and Bosons 1 1.1 Introduction and two-particle systems .......................... 1 1.2 N particles ......................................... 3 1.3 Non-interacting particles .................................. 5 1.4 Non-zero temperature ................................... 7 1.5 Composite particles .................................... 7 1.6 Emergence of distinguishability .............................. 9 2 Degenerate Fermi gas 10 2.1 Electrons in a box ..................................... 10 2.2 White dwarves ....................................... 12 2.3 Electrons in a periodic potential ............................. 16 3 Charged particles in a magnetic field 21 3.1 The Pauli Hamiltonian ................................... 21 3.2 Landau levels ........................................ 23 3.3 The de Haas-van Alphen effect .............................. 24 3.4 Integer Quantum Hall Effect ............................... 27 3.5 Aharonov-Bohm Effect ................................... 33 1 Fermions and Bosons 1.1 Introduction and two-particle systems Previously we have discussed multiple-particle systems using the tensor-product formalism (cf. Section 1.2 of Chapter 3 of these notes). But this applies only to distinguishable particles. In reality, all known particles are indistinguishable. In the coming lectures, we will explore the mathematical and physical consequences of this. First, consider classical many-particle systems. If a single particle has state described by position and momentum (~r; p~), then the state of N distinguishable particles can be written as (~r1; p~1; ~r2; p~2;:::; ~rN ; p~N ). The notation (·; ·;:::; ·) denotes an ordered list, in which different posi­ tions have different meanings; e.g. in general (~r1; p~1; ~r2; p~2)6 = (~r2; p~2; ~r1; p~1). 1 To describe indistinguishable particles, we can use set notation.
    [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]
  • 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]
  • Magnetic Materials I
    5 Magnetic materials I Magnetic materials I ● Diamagnetism ● Paramagnetism Diamagnetism - susceptibility All materials can be classified in terms of their magnetic behavior falling into one of several categories depending on their bulk magnetic susceptibility χ. without spin M⃗ 1 χ= in general the susceptibility is a position dependent tensor M⃗ (⃗r )= ⃗r ×J⃗ (⃗r ) H⃗ 2 ] In some materials the magnetization is m / not a linear function of field strength. In A [ such cases the differential susceptibility M is introduced: d M⃗ χ = d d H⃗ We usually talk about isothermal χ susceptibility: ∂ M⃗ χ = T ( ∂ ⃗ ) H T Theoreticians define magnetization as: ∂ F⃗ H[A/m] M=− F=U−TS - Helmholtz free energy [4] ( ∂ H⃗ ) T N dU =T dS− p dV +∑ μi dni i=1 N N dF =T dS − p dV +∑ μi dni−T dS−S dT =−S dT − p dV +∑ μi dni i=1 i=1 Diamagnetism - susceptibility It is customary to define susceptibility in relation to volume, mass or mole: M⃗ (M⃗ /ρ) m3 (M⃗ /mol) m3 χ= [dimensionless] , χρ= , χ = H⃗ H⃗ [ kg ] mol H⃗ [ mol ] 1emu=1×10−3 A⋅m2 The general classification of materials according to their magnetic properties μ<1 <0 diamagnetic* χ ⃗ ⃗ ⃗ ⃗ ⃗ ⃗ B=μ H =μr μ0 H B=μo (H +M ) μ>1 >0 paramagnetic** ⃗ ⃗ ⃗ χ μr μ0 H =μo( H + M ) → μ r=1+χ μ≫1 χ≫1 ferromagnetic*** *dia /daɪə mæ ɡˈ n ɛ t ɪ k/ -Greek: “from, through, across” - repelled by magnets. We have from L2: 1 2 the force is directed antiparallel to the gradient of B2 F = V ∇ B i.e.
    [Show full text]
  • 5.1 Two-Particle Systems
    5.1 Two-Particle Systems We encountered a two-particle system in dealing with the addition of angular momentum. Let's treat such systems in a more formal way. The w.f. for a two-particle system must depend on the spatial coordinates of both particles as @Ψ well as t: Ψ(r1; r2; t), satisfying i~ @t = HΨ, ~2 2 ~2 2 where H = + V (r1; r2; t), −2m1r1 − 2m2r2 and d3r d3r Ψ(r ; r ; t) 2 = 1. 1 2 j 1 2 j R Iff V is independent of time, then we can separate the time and spatial variables, obtaining Ψ(r1; r2; t) = (r1; r2) exp( iEt=~), − where E is the total energy of the system. Let us now make a very fundamental assumption: that each particle occupies a one-particle e.s. [Note that this is often a poor approximation for the true many-body w.f.] The joint e.f. can then be written as the product of two one-particle e.f.'s: (r1; r2) = a(r1) b(r2). Suppose furthermore that the two particles are indistinguishable. Then, the above w.f. is not really adequate since you can't actually tell whether it's particle 1 in state a or particle 2. This indeterminacy is correctly reflected if we replace the above w.f. by (r ; r ) = a(r ) (r ) (r ) a(r ). 1 2 1 b 2 b 1 2 The `plus-or-minus' sign reflects that there are two distinct ways to accomplish this. Thus we are naturally led to consider two kinds of identical particles, which we have come to call `bosons' (+) and `fermions' ( ).
    [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]
  • Identical Particles in Classical Physics One Can Distinguish Between
    Identical particles In classical physics one can distinguish between identical particles in such a way as to leave the dynamics unaltered. Therefore, the exchange of particles of identical particles leads to di®erent con¯gurations. In quantum mechanics, i.e., in nature at the microscopic levels identical particles are indistinguishable. A proton that arrives from a supernova explosion is the same as the proton in your glass of water.1 Messiah and Greenberg2 state the principle of indistinguishability as \states that di®er only by a permutation of identical particles cannot be distinguished by any obser- vation whatsoever." If we denote the operator which interchanges particles i and j by the permutation operator Pij we have ^ y ^ hªjOjªi = hªjPij OPijjÃi ^ which implies that Pij commutes with an arbitrary observable O. In particular, it commutes with the Hamiltonian. All operators are permutation invariant. Law of nature: The wave function of a collection of identical half-odd-integer spin particles is completely antisymmetric while that of integer spin particles or bosons is completely symmetric3. De¯ning the permutation operator4 by Pij ª(1; 2; ¢ ¢ ¢ ; i; ¢ ¢ ¢ ; j; ¢ ¢ ¢ N) = ª(1; 2; ¢ ¢ ¢ ; j; ¢ ¢ ¢ ; i; ¢ ¢ ¢ N) we have Pij ª(1; 2; ¢ ¢ ¢ ; i; ¢ ¢ ¢ ; j; ¢ ¢ ¢ N) = § ª(1; 2; ¢ ¢ ¢ ; i; ¢ ¢ ¢ ; j; ¢ ¢ ¢ N) where the upper sign refers to bosons and the lower to fermions. Emphasize that the symmetry requirements only apply to identical particles. For the hydrogen molecule if I and II represent the coordinates and spins of the two protons and 1 and 2 of the two electrons we have ª(I;II; 1; 2) = ¡ ª(II;I; 1; 2) = ¡ ª(I;II; 2; 1) = ª(II;I; 2; 1) 1On the other hand, the elements on earth came from some star/supernova in the distant past.
    [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]
  • The Conventionality of Parastatistics
    The Conventionality of Parastatistics David John Baker Hans Halvorson Noel Swanson∗ March 6, 2014 Abstract Nature seems to be such that we can describe it accurately with quantum theories of bosons and fermions alone, without resort to parastatistics. This has been seen as a deep mystery: paraparticles make perfect physical sense, so why don't we see them in nature? We consider one potential answer: every paraparticle theory is physically equivalent to some theory of bosons or fermions, making the absence of paraparticles in our theories a matter of convention rather than a mysterious empirical discovery. We argue that this equivalence thesis holds in all physically admissible quantum field theories falling under the domain of the rigorous Doplicher-Haag-Roberts approach to superselection rules. Inadmissible parastatistical theories are ruled out by a locality- inspired principle we call Charge Recombination. Contents 1 Introduction 2 2 Paraparticles in Quantum Theory 6 ∗This work is fully collaborative. Authors are listed in alphabetical order. 1 3 Theoretical Equivalence 11 3.1 Field systems in AQFT . 13 3.2 Equivalence of field systems . 17 4 A Brief History of the Equivalence Thesis 20 4.1 The Green Decomposition . 20 4.2 Klein Transformations . 21 4.3 The Argument of Dr¨uhl,Haag, and Roberts . 24 4.4 The Doplicher-Roberts Reconstruction Theorem . 26 5 Sharpening the Thesis 29 6 Discussion 36 6.1 Interpretations of QM . 44 6.2 Structuralism and Haecceities . 46 6.3 Paraquark Theories . 48 1 Introduction Our most fundamental theories of matter provide a highly accurate description of subatomic particles and their behavior.
    [Show full text]
  • Talk M.Alouani
    From atoms to solids to nanostructures Mebarek Alouani IPCMS, 23 rue du Loess, UMR 7504, Strasbourg, France European Summer Campus 2012: Physics at the nanoscale, Strasbourg, France, July 01–07, 2012 1/107 Course content I. Magnetic moment and magnetic field II. No magnetism in classical mechanics III. Where does magnetism come from? IV. Crystal field, superexchange, double exchange V. Free electron model: Spontaneous magnetization VI. The local spin density approximation of the DFT VI. Beyond the DFT: LDA+U VII. Spin-orbit effects: Magnetic anisotropy, XMCD VIII. Bibliography European Summer Campus 2012: Physics at the nanoscale, Strasbourg, France, July 01–07, 2012 2/107 Course content I. Magnetic moment and magnetic field II. No magnetism in classical mechanics III. Where does magnetism come from? IV. Crystal field, superexchange, double exchange V. Free electron model: Spontaneous magnetization VI. The local spin density approximation of the DFT VI. Beyond the DFT: LDA+U VII. Spin-orbit effects: Magnetic anisotropy, XMCD VIII. Bibliography European Summer Campus 2012: Physics at the nanoscale, Strasbourg, France, July 01–07, 2012 3/107 Magnetic moments In classical electrodynamics, the vector potential, in the magnetic dipole approximation, is given by: μ μ × rˆ Ar()= 0 4π r2 where the magnetic moment μ is defined as: 1 μ =×Id('rl ) 2 ∫ It is shown that the magnetic moment is proportional to the angular momentum μ = γ L (γ the gyromagnetic factor) The projection along the quantification axis z gives (/2Bohrmaμ = eh m gneton) μzB= −gmμ Be European Summer Campus 2012: Physics at the nanoscale, Strasbourg, France, July 01–07, 2012 4/107 Magnetization and Field The magnetization M is the magnetic moment per unit volume In free space B = μ0 H because there is no net magnetization.
    [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]