Chapter 2 Principles of Quantum Statistics

Total Page:16

File Type:pdf, Size:1020Kb

Chapter 2 Principles of Quantum Statistics Chapter 2 Principles of Quantum Statistics In this chapter we will study one of the most important assumptions of the quantum the- ory, the symmetrization postulate, and its consequence for the statistical properties of an ensemble of particles at thermal equilibrium. The three representative distribution func- tions, Maxwell-Boltzmann, Bose-Einstein and Fermi-Dirac distributions, are introduced here. Then, we will describe the other important assumption of the quantum theory, the non-commutability of conjugate observables, and its consequence for the uncertainty product in determining a certain pair of observables. The discussions in this chapter pro- vide the foundation of thermal noise and quantum noise. Most of the discussions in this chapter follow the excellent texts on stastical mechanics by F. Rief [1] and on quantum mechanics by C. Cohen-Tannoudji et al. [2]. 2.1 Symmetrization Postulate of Quantum Mechanics 2.1.1 Quantum indistinguishability Let us consider a following thought experiment. Two identical quantum particles are incident upon a 50%{50% particle beam splitter and we measure the output by two particle counters. 1 or 2 ? 50-50% beam splitter Figure 2.1: Collision of two identical particles. To analyze this problem, we label the two particles as particle 1 and particle 2 based on 1 the click event in the detectors. One possible input state is j1;R; 2;Li ¢ ¢ ¢ particle 1 is in the right input port and particle 2 is in the left input port. Since identical quantum particles are \indistinguishable", the following input state is equally possible : j1;L; 2;Ri ¢ ¢ ¢ particle 1 is in the left input port and particle 2 is in the right input port. Most general input state is thus constructed as the linear superposition of the two : c0j1;R; 2;Li + c1j1;L; 2;Ri 2 2 (2.1) (jc0j + jc1j = 1 if the two states are orthogonal.) As shown below, an experimental result is dependent on the choice of the two c-numbers c0 and c1. Therefore, the above mathematical state has an inherent ambiguity in predicting an experimental result and we need a fundamental assumption of the theory, \symmetriza- tion postulate"[2]. 2.1.2 Statement of the postulate A physical state which represents a real system consisting of identical quantum particles is either symmetric or anti-symmetric with respect to the permutation of any two particles. A particle which obeys the former is called a boson and a particle which obeys the latter is called a fermion. For instance, the quantum states of two identical bosons and fermions are 1 Boson : p [j1;R; 2;Li + j1;L; 2;Ri]; 2 (2.2) 1 Fermion : p [j1;R; 2;Li ¡ j1;L; 2;Ri]; 2 if the two states are orthogonal. To demonstrate the important conseque of this new postulate, next let us calculate the probability of ¯nding the two particles in the same output port (L). This probability consists of two possibilities. 2 2 1 1 2 + direct term exchange term Figure 2.2: Two possibilities for an output state j1;L; 2;Li. 3 à ! 1 1 The beam splitter scattering matrix is U^ = p1 in the basis of the two trans- 2 ¡1 1 " # jk i mission modes R!L . Therefore, the probability is jkL!Ri The probability for obtaining j1;R; 2;Ri state is identical to (2.3). We can calculate the probability of ¯nding one particle in each output port. The probability of obtaining j1;L; 2;Ri state consists of two possibilities: 2 1 1 2 direct term exchange term Figure 2.3: Two possibilities for an output state j1;L; 2;Ri. The probability for obtaining j1;R; 2;Li state is identical to (2.4). The scattering patterns of identical bosons and fermions are shown in Fig. 2.4. 4 boson fermion 1 1 ψ =[]1,;,,;,RRLL 2 − 1 2 ψ =[]1,;,,;,RLLR 2 − 1 2 out 2 out 2 Figure 2.4: The scattering behaviours of two identical bosons and fermions at 50-50% beam splitter. Comments : 1. The (2, 0) and (0, 2) output characteristics of bosons are called a \bunching" e®ect. The constructive interference enhances the probability of ¯nding two particles in the same state. This is the ultimate origin of ¯nal state stimulation in lasers, Bose Einstein condensation and superconductivity. 2. The deterministic (1, 1) output characteristics of fermions are called \anti-bunching" e®ect. The destructive interference between the direct and exchange terms sup- presses the probability of ¯nding more than two particles in the same state. This is the manifestation of Pauli's exclusion principle. 1 2.1.3 Symmetrization postulate for spin ¡ 2 particles 1 Next let us consider a following thought experiment. Two spin ¡ 2 particles are incident upon a 50%{50% beam splitter. The scattering matrix of a beam splitter is assumed to be spin-independent. If the two particles are in spin triplet states (symmetric spin states), the orbital states are symmetric for bosons and anti-symmetric for fermions to satisfy the symmetrization postulate for overall states : 8 > j "i j "i > 1 2 <> 1 j #i1j #i2 Boson : p [jRi1jLi2 + jLi1jRi2] ­ 2 > > 1 : p [j "i1j #i2 + j #i1j "i2] 2 8 > j "i j "i > 1 2 <> 1 j #i1j #i2 Fermion : p [jRi1jLi2 ¡ jLi1jRi2] ­ (2.5) 2 > > 1 : p [j "i1j #i2 + j #i1j "i2] 2 5 Their collision characteristics are the same as those for spin-less particles mentioned above. However, if the two particles are in a spin singlet state (anti-symmetric spin state), the orbital states are anti-symmetric for bosons and symmetric for fermions to satisfy the symmetrization postulate for overall states : 1 1 Boson : p [jRi1jLi2 ¡ jLi1jRi2] ­ p [j "i1j #i2 ¡ j #i1j "i2] 2 2 (2.6) 1 1 Fermion : p [jRi1jLi2 + jLi1jRi2] ­ p [j "i1j #i2 ¡ j #i1j "i2] 2 2 Now, the bosons feature a fermionic collision, i.e. (1, 1) output, and the fermions feature a bosonic collision, i.e. (2, 0) or (0, 2) output. This example illustrates a very important 1 nature of two spin ¡ 2 fermions which can occupy the same state. Conversely if two identical fermions occupy the same orbital state, their spin state is always a spin singlet state. 1 ¡jxi1jxi2 ­ p [j "i1j #i2 ¡ j #i1j "i2] (2.7) 2 AKA ¢¢¸ same orbital state 2.2 Thermodynamic Partition Functions 2.2.1 A small system in contact with a large heat reservoir Let us consider an ideal gas of non-interacting identical particles in a volume V and at a P temperature T . We assume the total number of particles, r nr = N, is constant, where r designates a microscopic state and nr is the number of particles in that state. The total energy of such an ideal gas system is X ER = "rnr ; (2.8) r where "r is the kinetic energy of a microscopic state r and R designates a macroscopic state represented by the occupation number of each microscopic state, fn1; n2; ¢ ¢ ¢ ; nr; ¢ ¢ ¢g. The thermodynamic partition function Z is de¯ned by X Z = e¡¯ER ; (2.9) R P where R stands for summation over all possible macroscopic states, ¯ = 1=kBT is a temperature parameter and kB is a Boltzmann constant. Next let us obtain the probability of ¯nding a particular macroscopic state R. For this purpose we assume that our system is a small system A in thermal contact to a large heat reservoir A0. We also assume the interaction between A and A0 is extremely small, so that their energies are additive. The energy of A is not ¯xed and depends on the speci¯c 6 macroscopic state R. It is the total energy of the combined system A and A0 which has a 0 constant value ET . This constant total energy is split into those of A and A : 0 ET = ER + E ; (2.10) where E0 is the energy of A0. According to the fundamental postulate of statistical mechanics, the probability of ¯nding A in a speci¯c macroscopic state R is proportioned to the number of states 0 0 0 ­ (ET ¡ ER) accessible to A when its energy lies in a range ±E near E = ET ¡ ER. Hence 0 0 PR = C ­ (ET ¡ ER) ; (2.11) where C0 is a constant independent of R and determined by the normalization condition P R pR = 1. 0 Since A is a much smaller system than A ;ER ¿ ET and (2.11) can be approximated 0 0 by expanding the logarithm of ­ (ET ¡ ER) about E = ET : · ¸ 0 0 @ 0 0 ln­ (ET ¡ ER) ' ln­ (ET ) ¡ 0 ln­ (E ) ¢ ER : (2.12) @E ET The derivative in (2.12) is the de¯nition of the temperature parameter of the heat reservoir according to the statistical mechanics [1], · ¸ @ 0 0 0 ln­ (E ) ´ ¯ = 1=kBT: (2.13) @E ET The above result means that the heat reservoir A0 is much larger than A so that its temperature T remains unchanged by such a small amount of energy it gives to A. Thus, we obtain 0 0 ¡¯ER ­ (ET ¡ ER) = ­ (ET ) e : (2.14) 0 since ­ (ET ) is a constant independent of R, (2.11) becomes ¡¯ER PR = Ce ; (2.15) where C = C0­0 (E ) can be determined by the above mentioned normalization condition P T R pR = 1. The probability of ¯nding the gas in a speci¯c macroscopic state R = fn1; n2; ¢ ¢ ¢ ; nr; ¢ ¢ ¢g is ¯nally given by e¡¯ER e¡¯ER PR = X = : (2.16) e¡¯ER0 Z R0 If the above is true, such a system is called a canonical ensemble[1; 3].
Recommended publications
  • Canonical Ensemble
    ME346A Introduction to Statistical Mechanics { Wei Cai { Stanford University { Win 2011 Handout 8. Canonical Ensemble January 26, 2011 Contents Outline • In this chapter, we will establish the equilibrium statistical distribution for systems maintained at a constant temperature T , through thermal contact with a heat bath. • The resulting distribution is also called Boltzmann's distribution. • The canonical distribution also leads to definition of the partition function and an expression for Helmholtz free energy, analogous to Boltzmann's Entropy formula. • We will study energy fluctuation at constant temperature, and witness another fluctuation- dissipation theorem (FDT) and finally establish the equivalence of micro canonical ensemble and canonical ensemble in the thermodynamic limit. (We first met a mani- festation of FDT in diffusion as Einstein's relation.) Reading Assignment: Reif x6.1-6.7, x6.10 1 1 Temperature For an isolated system, with fixed N { number of particles, V { volume, E { total energy, it is most conveniently described by the microcanonical (NVE) ensemble, which is a uniform distribution between two constant energy surfaces. const E ≤ H(fq g; fp g) ≤ E + ∆E ρ (fq g; fp g) = i i (1) mc i i 0 otherwise Statistical mechanics also provides the expression for entropy S(N; V; E) = kB ln Ω. In thermodynamics, S(N; V; E) can be transformed to a more convenient form (by Legendre transform) of Helmholtz free energy A(N; V; T ), which correspond to a system with constant N; V and temperature T . Q: Does the transformation from N; V; E to N; V; T have a meaning in statistical mechanics? A: The ensemble of systems all at constant N; V; T is called the canonical NVT ensemble.
    [Show full text]
  • Magnetism, Magnetic Properties, Magnetochemistry
    Magnetism, Magnetic Properties, Magnetochemistry 1 Magnetism All matter is electronic Positive/negative charges - bound by Coulombic forces Result of electric field E between charges, electric dipole Electric and magnetic fields = the electromagnetic interaction (Oersted, Maxwell) Electric field = electric +/ charges, electric dipole Magnetic field ??No source?? No magnetic charges, N-S No magnetic monopole Magnetic field = motion of electric charges (electric current, atomic motions) Magnetic dipole – magnetic moment = i A [A m2] 2 Electromagnetic Fields 3 Magnetism Magnetic field = motion of electric charges • Macro - electric current • Micro - spin + orbital momentum Ampère 1822 Poisson model Magnetic dipole – magnetic (dipole) moment [A m2] i A 4 Ampere model Magnetism Microscopic explanation of source of magnetism = Fundamental quantum magnets Unpaired electrons = spins (Bohr 1913) Atomic building blocks (protons, neutrons and electrons = fermions) possess an intrinsic magnetic moment Relativistic quantum theory (P. Dirac 1928) SPIN (quantum property ~ rotation of charged particles) Spin (½ for all fermions) gives rise to a magnetic moment 5 Atomic Motions of Electric Charges The origins for the magnetic moment of a free atom Motions of Electric Charges: 1) The spins of the electrons S. Unpaired spins give a paramagnetic contribution. Paired spins give a diamagnetic contribution. 2) The orbital angular momentum L of the electrons about the nucleus, degenerate orbitals, paramagnetic contribution. The change in the orbital moment
    [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]
  • Spin Statistics, Partition Functions and Network Entropy
    This is a repository copy of Spin Statistics, Partition Functions and Network Entropy. White Rose Research Online URL for this paper: https://eprints.whiterose.ac.uk/118694/ Version: Accepted Version Article: Wang, Jianjia, Wilson, Richard Charles orcid.org/0000-0001-7265-3033 and Hancock, Edwin R orcid.org/0000-0003-4496-2028 (2017) Spin Statistics, Partition Functions and Network Entropy. Journal of Complex Networks. pp. 1-25. ISSN 2051-1329 https://doi.org/10.1093/comnet/cnx017 Reuse Items deposited in White Rose Research Online are protected by copyright, with all rights reserved unless indicated otherwise. They may be downloaded and/or printed for private study, or other acts as permitted by national copyright laws. The publisher or other rights holders may allow further reproduction and re-use of the full text version. This is indicated by the licence information on the White Rose Research Online record for the item. Takedown If you consider content in White Rose Research Online to be in breach of UK law, please notify us by emailing [email protected] including the URL of the record and the reason for the withdrawal request. [email protected] https://eprints.whiterose.ac.uk/ IMA Journal of Complex Networks (2017)Page 1 of 25 doi:10.1093/comnet/xxx000 Spin Statistics, Partition Functions and Network Entropy JIANJIA WANG∗ Department of Computer Science,University of York, York, YO10 5DD, UK ∗[email protected] RICHARD C. WILSON Department of Computer Science,University of York, York, YO10 5DD, UK AND EDWIN R. HANCOCK Department of Computer Science,University of York, York, YO10 5DD, UK [Received on 2 May 2017] This paper explores the thermodynamic characterization of networks using the heat bath analogy when the energy states are occupied under different spin statistics, specified by a partition function.
    [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]
  • Entropy: from the Boltzmann Equation to the Maxwell Boltzmann Distribution
    Entropy: From the Boltzmann equation to the Maxwell Boltzmann distribution A formula to relate entropy to probability Often it is a lot more useful to think about entropy in terms of the probability with which different states are occupied. Lets see if we can describe entropy as a function of the probability distribution between different states. N! WN particles = n1!n2!....nt! stirling (N e)N N N W = = N particles n1 n2 nt n1 n2 nt (n1 e) (n2 e) ....(nt e) n1 n2 ...nt with N pi = ni 1 W = N particles n1 n2 nt p1 p2 ...pt takeln t lnWN particles = "#ni ln pi i=1 divide N particles t lnW1particle = "# pi ln pi i=1 times k t k lnW1particle = "k# pi ln pi = S1particle i=1 and t t S = "N k p ln p = "R p ln p NA A # i # i i i=1 i=1 ! Think about how this equation behaves for a moment. If any one of the states has a probability of occurring equal to 1, then the ln pi of that state is 0 and the probability of all the other states has to be 0 (the sum over all probabilities has to be 1). So the entropy of such a system is 0. This is exactly what we would have expected. In our coin flipping case there was only one way to be all heads and the W of that configuration was 1. Also, and this is not so obvious, the maximal entropy will result if all states are equally populated (if you want to see a mathematical proof of this check out Ken Dill’s book on page 85).
    [Show full text]
  • Physical Biology
    Physical Biology Kevin Cahill Department of Physics and Astronomy University of New Mexico, Albuquerque, NM 87131 i For Ginette, Mike, Sean, Peter, Mia, James, Dylan, and Christopher Contents Preface page v 1 Atoms and Molecules 1 1.1 Atoms 1 1.2 Bonds between atoms 1 1.3 Oxidation and reduction 5 1.4 Molecules in cells 6 1.5 Sugars 6 1.6 Hydrocarbons, fatty acids, and fats 8 1.7 Nucleotides 10 1.8 Amino acids 15 1.9 Proteins 19 Exercises 20 2 Metabolism 23 2.1 Glycolysis 23 2.2 The citric-acid cycle 23 3 Molecular devices 27 3.1 Transport 27 4 Probability and statistics 28 4.1 Mean and variance 28 4.2 The binomial distribution 29 4.3 Random walks 32 4.4 Diffusion as a random walk 33 4.5 Continuous diffusion 34 4.6 The Poisson distribution 36 4.7 The gaussian distribution 37 4.8 The error function erf 39 4.9 The Maxwell-Boltzmann distribution 40 Contents iii 4.10 Characteristic functions 41 4.11 Central limit theorem 42 Exercises 45 5 Statistical Mechanics 47 5.1 Probability and the Boltzmann distribution 47 5.2 Equilibrium and the Boltzmann distribution 49 5.3 Entropy 53 5.4 Temperature and entropy 54 5.5 The partition function 55 5.6 The Helmholtz free energy 57 5.7 The Gibbs free energy 58 Exercises 59 6 Chemical Reactions 61 7 Liquids 65 7.1 Diffusion 65 7.2 Langevin's Theory of Brownian Motion 67 7.3 The Einstein-Nernst relation 71 7.4 The Poisson-Boltzmann equation 72 7.5 Reynolds number 73 7.6 Fluid mechanics 76 8 Cell structure 77 8.1 Organelles 77 8.2 Membranes 78 8.3 Membrane potentials 80 8.4 Potassium leak channels 82 8.5 The Debye-H¨uckel equation 83 8.6 The Poisson-Boltzmann equation in one dimension 87 8.7 Neurons 87 Exercises 89 9 Polymers 90 9.1 Freely jointed polymers 90 9.2 Entropic force 91 10 Mechanics 93 11 Some Experimental Methods 94 11.1 Time-dependent perturbation theory 94 11.2 FRET 96 11.3 Fluorescence 103 11.4 Some mathematics 107 iv Contents References 109 Index 111 Preface Most books on biophysics hide the physics so as to appeal to a wide audi- ence of students.
    [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]
  • 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]
  • Spontaneous Symmetry Breaking and Mass Generation As Built-In Phenomena in Logarithmic Nonlinear Quantum Theory
    Vol. 42 (2011) ACTA PHYSICA POLONICA B No 2 SPONTANEOUS SYMMETRY BREAKING AND MASS GENERATION AS BUILT-IN PHENOMENA IN LOGARITHMIC NONLINEAR QUANTUM THEORY Konstantin G. Zloshchastiev Department of Physics and Center for Theoretical Physics University of the Witwatersrand Johannesburg, 2050, South Africa (Received September 29, 2010; revised version received November 3, 2010; final version received December 7, 2010) Our primary task is to demonstrate that the logarithmic nonlinearity in the quantum wave equation can cause the spontaneous symmetry break- ing and mass generation phenomena on its own, at least in principle. To achieve this goal, we view the physical vacuum as a kind of the funda- mental Bose–Einstein condensate embedded into the fictitious Euclidean space. The relation of such description to that of the physical (relativis- tic) observer is established via the fluid/gravity correspondence map, the related issues, such as the induced gravity and scalar field, relativistic pos- tulates, Mach’s principle and cosmology, are discussed. For estimate the values of the generated masses of the otherwise massless particles such as the photon, we propose few simple models which take into account small vacuum fluctuations. It turns out that the photon’s mass can be naturally expressed in terms of the elementary electrical charge and the extensive length parameter of the nonlinearity. Finally, we outline the topological properties of the logarithmic theory and corresponding solitonic solutions. DOI:10.5506/APhysPolB.42.261 PACS numbers: 11.15.Ex, 11.30.Qc, 04.60.Bc, 03.65.Pm 1. Introduction Current observational data in astrophysics are probing a regime of de- partures from classical relativity with sensitivities that are relevant for the study of the quantum-gravity problem [1,2].
    [Show full text]