Quantum Statistics (Quantenstatistik)

Total Page:16

File Type:pdf, Size:1020Kb

Quantum Statistics (Quantenstatistik) Chapter 7 Quantum Statistics (Quantenstatistik) 7.1 Density of states of massive particles Our goal: discover what quantum effects bring into the statistical description of systems. Formulate conditions under which QM-specific effects are not essential and classical descriptions can be used. Consider a single particle in a box with infinitely high walls U ε3 ε2 Figure 7.1: Levels of a single particle in a 3D box. At the ε1 box walls the potential is infinite and the wave-function is L zero. A one-particle Hamiltonian has only the momentum p^2 ~2 H^ = = − r2 (7.1) 2m 2m The infinite potential at the box boundaries imposes the zero-wave-function boundary condition. Hence the eigenfunctions of the Hamiltonian are k = sin(kxx) sin(kyy) sin(kzz) (7.2) ~ 2π + k = ~n; ~n = (n1; n2; n3); ni 2 Z (7.3) L ( ) ~ ~ 2π 2 ~p = ~~k; ϵ = k2 = ~n2;V = L3 (7.4) 2m 2m L 4π2~2 ) ϵ = ~n2 (7.5) 2mV 2=3 Let us introduce spin s (it takes into account possible degeneracy of the energy levels) and change the summation over ~n to an integration over ϵ X X X X X X Z X Z 1 ··· = ··· = · · · ∼ d3~n ··· = dn4πn2 ::: (7.6) 0 ν s ~k s ~n s s 45 46 CHAPTER 7. QUANTUM STATISTICS (QUANTENSTATISTIK) (2mϵ)1=2 V 1=3 1 1 V m3=2 n = V 1=3 ; dn = 2mdϵ, 4πn2dn = p ϵ1=2dϵ (7.7) 2π~ 2π~ 2 (2mϵ)1=2 2π2~3 X = gs = 2s + 1 spin-degeneracy (7.8) s Z X g V m3=2 1 ··· = ps dϵϵ1=2 ::: (7.9) 2~3 ν 2π 0 Number of states in a given energy range is higher at high energies than at low energies. helium atom electron photon −18 −14 −4 E111, eV 1:5 × 10 1:1 × 10 1:1 × 10 −14 −11 Teqiuv,K 1:2 × 10 8:7 × 10 0.83 Table 7.1: The lowest energy as computed for a 1 cm3 cubic box. The lowest energy occurs for the state 111 with nx = ny = nz = 1. While the lowest single particle energy is strictly speaking somewhat greater than zero, it is extremely small and, for all practical purposes, can be taken as zero. However, at very low cryogenic temperatures, photons in the lowest energy state, or \ground state", may have a relatively more significant energy. It can normally be assumed that the lowest energy is zero for noninteracting particles in a box at room temperature. In most cases, the spacing between the lowest and second lowest energy is comparable to the lowest energy, and thus similarly negligible (see, however, Bose-Einstein condensation!) Why do we then need quantum mechanics at all? Classical non-quantum physics too would predict a continuum of energies for the particles. And it too would predict the energy to start from zero. The energy of a noninteracting particle is all kinetic energy; classical physics has that zero if the particle is at rest and positive otherwise. There are several reasons quantum mechanics is essential: (i) It predictsp that there are more single- particle states in a given energy range at high energy than at low energy ( ϵ). (ii) The wave function (anti)symmetrization requirements cannot be accommodated using classical physics (bosons and fermions have the same statistics). 7.2 Quantum Gas, Bose-Einstein and Fermi-Dirac Statistics Let us consider N non-interacting identical particles. XN H^ = H^p (7.10) p=1 All H^p have the same set of states. These are given by a one-particle problem H^ jνi = ϵ jνi (7.11) |{z}sp ν single-particle Hamiltonian Any two configurations that differ only by an interchange of two or more identical particles are regarded as one and the same state. Thus the state of the system is determined by giving number of particles nν with energy ϵν . 7.2. QUANTUM GAS 47 Our goal is to calculate the partition function Z with the condition that X nν = N (7.12) ν The values admitted for every nν may be nν = 0; 1; 2; 3; 4 ::: Bose-Einstein case (7.13) nν = 0; 1 Fermi-Dirac case (7.14) Bose-Einstein statistics must be used for particles with integer spin (bosons, for example He4, photons, phonons) and Fermi-Dirac statistics for particles with half-integer spin (fermions, e.g. electrons, neutrons, protons). For Bose particles, any number of particles may occupy a given state. For Fermi particles, however, there can be only one particle at most in each state (Pauli exclusion principle). Total energy of the system is X Etot = |{z}nν ϵν (7.15) ν Xnumber of particles in a state ν nν = N (7.16) ν Hence the canonical partition function ! X0 X X Y Y X Z = exp −β nν ϵν = exp (−βnν ϵν ) = exp (−βnν ϵν ) = (7.17) n1;n2;n38 ν nν ν ν nν < 1 Y 1 + exp(−βϵ ) + exp(−2βϵ ) + ··· = - bosons ν ν 1 − exp(−βϵ ) : ν (7.18) ν 1 + exp(−βϵν ) - fermions P Unfortunately this is NOT correct, since we have an auxiliary condition ν nν = N = const. It is in fact difficult to perform this summation in a canonical ensemble. The trick is to switch to the grandcanonical description: let us connect our system to a \particle-bath" which has chemical potential ! µ such that hNi = N. The grandcanonical partition function removes the restriction on the total number of particles, i.e. all nν are allowed now. Hence ! ! X1 X X X ZG = exp(βµN)Z(N) = exp βµ nν exp −β nν (ϵν − µ) = (7.19) N=0 nν ν ν 8 < 1 Y X Y - Bose-Einstein statistics − − 1 − exp(−β(ϵ − µ)) exp ( β(ϵν µ)nν ) = : ν (7.20) ν nν ν 1 + exp(−β(ϵν − µ)) - Fermi-Dirac statistics 7.2.1 Average occupation numbers The grand potential 8X <> ln [1 − exp(−β(ϵν − µ))] - bosons 1 1 − νX Φ = ln ZG = > (7.21) β β :− ln [1 + exp(−β(ϵν − µ))] - fermions ν 48 CHAPTER 7. QUANTUM STATISTICS (QUANTENSTATISTIK) Recalling thermodynamics Φ = E − TS − µN (7.22) dΦ = T dS − P dV + µdN − (T dS + SdT ) − (µdN + Ndµ) = −P dV − SdT − Ndµ (7.23) 8X > 1 ( ) <> - bosons @Φ exp(β(ϵν − µ)) − 1 h i − νX N = = > 1 (7.24) @µ V;T :>− - fermions exp(β(ϵ − µ)) + 1 ν ν Recalling that * + X hNi = nν (7.25) ν we obtain the average occupation numbers for the energy levels 8 1 <> − − - Bose-Einstein statistics h i exp(β(ϵν µ)) 1 nν = > 1 (7.26) : - Fermi-Dirac statistics exp(β(ϵν − µ)) + 1 7.2.2 The role of the constraint on the number of particles We have now calculated N(µ). For practical reasons we however need µ(N). We will address two points here: (i) when this inversion possible? (ii) if the number of particles is not constrained, can we still use our results? Case 1: The number of particles is known and kept fixed to N. In this case 8X > 1 BE <> = N (µ) - bosons exp(β(ϵ − µ)) − 1 Xν ν N = > 1 (7.27) :> = N FD(µ) - fermions exp(β(ϵ − µ)) + 1 ν ν By inverting these relations we can obtain µ. X 1 Question: For bosons, has many plus-minus infinity points, and therefore the exp(β(ϵ − µ)) − 1 ν ν ! equation N BE(µ) = N has many solutions. Why do we still talk about a unique chemical potential for bosons? To answer this, let us have a look at the number of particles in the ground state, n0. Assuming ϵ0 = 0 1 n = ≥ 0 ) µ ≤ 0 (7.28) 0 exp(−βµ) − 1 BE Hence, for bosons, there is a restriction on the sign of the chemical potential. For fermions, there are no restrictions. Case 2: For some particular types of bosons (e.g. photons, phonons), it is possible to have generation and annihilation of particles. Hence, N is not fixed and, strictly speaking, the grandcanonical description cannot be used. However, we know that if some internal variable is unconstrained, the thermodynamic potential of interest will adjust that variable in such a way that it attains its minimum, e.g. for free energy F = F (T; V; N) - for a system with fixed N (7.29) dF = −SdT − P dV + µdN (7.30) 7.2. QUANTUM GAS 49 if N is not known, it will adjust itself in such a away that F reaches its minimum, i.e. ( ) @F 0 = = µ - for systems with a number of particles which is not fixed (7.31) @N T;V We can still use the grandcanonical result, but with µ = 0. You can also compare this to our \incorrect" treatment at the beginning of the lecture. It was exactly the case with µ = 0. 7.2.3 Equation of state of a 3D quantum gas The grand potential X Φ = −T ln ZG = −T σ ln [1 + σ exp(−β(ϵν − µ))] (7.32) ν 1 nν = − − (7.33) (exp( β(ϵν µ)) + σ +1 fermions σ = (7.34) −1 bosons Changing the sum to an integral Z V m3=2 1 Φ = −T σ p ϵ1=2dϵ ln [1 + σ exp(−β(ϵ − µ))] (7.35) 2 3 2π ~ 0 Integration by part removes the logarithm Z 1 Z 1 3=2 1 3=2 1=2 − − ϵ − − ϵ σβ ϵ dϵ ln [1 + σ exp( β(ϵ µ))] = ln [1 + σ exp( β(ϵ µ))] + dϵ − − 0 3=2 0 0 3=2 σ + exp( β(ϵ µ)) (7.36) Hence Z 3=2 1 X 2 V m 1=2 1 2 2 Φ = − p dϵϵ ϵ = − ϵν nν = − E = −PV (7.37) 2 3 − − |{z} 3 2π ~ 0 σ + exp( β(ϵ µ)) 3 3 ν thermodynamics 2 PV = E (7.38) 3 Let us compare this to the classical ideal gas 2 2 3 2 PV = NT; E = × NT = NT ) PV = E (7.39) 3 3 2 3 The classical expression (in terms of P , V , E) is valid for weakly interacting bosons and fermions, even if the (anti)symmetrization is taken into account! 7.2.4 Classical limit In the classical limit, all states are only \weakly" occupied.
Recommended publications
  • 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]
  • 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]
  • Grand-Canonical Ensemble
    PHYS4006: Thermal and Statistical Physics Lecture Notes (Unit - IV) Open System: Grand-canonical Ensemble Dr. Neelabh Srivastava (Assistant Professor) Department of Physics Programme: M.Sc. Physics Mahatma Gandhi Central University Semester: 2nd Motihari-845401, Bihar E-mail: [email protected] • In microcanonical ensemble, each system contains same fixed energy as well as same number of particles. Hence, the system dealt within this ensemble is a closed isolated system. • With microcanonical ensemble, we can not deal with the systems that are kept in contact with a heat reservoir at a given temperature. 2 • In canonical ensemble, the condition of constant energy is relaxed and the system is allowed to exchange energy but not the particles with the system, i.e. those systems which are not isolated but are in contact with a heat reservoir. • This model could not be applied to those processes in which number of particle varies, i.e. chemical process, nuclear reactions (where particles are created and destroyed) and quantum process. 3 • So, for the method of ensemble to be applicable to such processes where number of particles as well as energy of the system changes, it is necessary to relax the condition of fixed number of particles. 4 • Such an ensemble where both the energy as well as number of particles can be exchanged with the heat reservoir is called Grand Canonical Ensemble. • In canonical ensemble, T, V and N are independent variables. Whereas, in grand canonical ensemble, the system is described by its temperature (T),volume (V) and chemical potential (μ). 5 • Since, the system is not isolated, its microstates are not equally probable.
    [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]
  • The Grand Canonical Ensemble
    University of Central Arkansas The Grand Canonical Ensemble Stephen R. Addison Directory ² Table of Contents ² Begin Article Copyright °c 2001 [email protected] Last Revision Date: April 10, 2001 Version 0.1 Table of Contents 1. Systems with Variable Particle Numbers 2. Review of the Ensembles 2.1. Microcanonical Ensemble 2.2. Canonical Ensemble 2.3. Grand Canonical Ensemble 3. Average Values on the Grand Canonical Ensemble 3.1. Average Number of Particles in a System 4. The Grand Canonical Ensemble and Thermodynamics 5. Legendre Transforms 5.1. Legendre Transforms for two variables 5.2. Helmholtz Free Energy as a Legendre Transform 6. Legendre Transforms and the Grand Canonical Ensem- ble 7. Solving Problems on the Grand Canonical Ensemble Section 1: Systems with Variable Particle Numbers 3 1. Systems with Variable Particle Numbers We have developed an expression for the partition function of an ideal gas. Toc JJ II J I Back J Doc Doc I Section 2: Review of the Ensembles 4 2. Review of the Ensembles 2.1. Microcanonical Ensemble The system is isolated. This is the ¯rst bridge or route between mechanics and thermodynamics, it is called the adiabatic bridge. E; V; N are ¯xed S = k ln ­(E; V; N) Toc JJ II J I Back J Doc Doc I Section 2: Review of the Ensembles 5 2.2. Canonical Ensemble System in contact with a heat bath. This is the second bridge between mechanics and thermodynamics, it is called the isothermal bridge. This bridge is more elegant and more easily crossed. T; V; N ¯xed, E fluctuates.
    [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]
  • International Centre for Theoretical Physics
    INTERNATIONAL CENTRE FOR THEORETICAL PHYSICS OUTLINE OF A NONLINEAR, RELATIVISTIC QUANTUM MECHANICS OF EXTENDED PARTICLES Eckehard W. Mielke INTERNATIONAL ATOMIC ENERGY AGENCY UNITED NATIONS EDUCATIONAL, SCIENTIFIC AND CULTURAL ORGANIZATION 1981 MIRAMARE-TRIESTE IC/81/9 International Atomic Energy Agency and United Nations Educational Scientific and Cultural Organization 3TERKATIOHAL CENTRE FOR THEORETICAL PHTSICS OUTLINE OF A NONLINEAR, KELATIVXSTTC QUANTUM HECHAHICS OF EXTENDED PARTICLES* Eckehard W. Mielke International Centre for Theoretical Physics, Trieste, Italy. MIRAHABE - TRIESTE January 198l • Submitted for publication. ABSTRACT I. INTRODUCTION AMD MOTIVATION A quantum theory of intrinsically extended particles similar to de Broglie's In recent years the conviction has apparently increased among most theory of the Double Solution is proposed, A rational notion of the particle's physicists that all fundamental interactions between particles - including extension is enthroned by realizing its internal structure via soliton-type gravity - can be comprehended by appropriate extensions of local gauge theories. solutions of nonlinear, relativistic wave equations. These droplet-type waves The principle of gauge invarisince was first formulated in have a quasi-objective character except for certain boundary conditions vhich 1923 by Hermann Weyl[l57] with the aim to give a proper connection between may be subject to stochastic fluctuations. More precisely, this assumption Dirac's relativistic quantum theory [37] of the electron and the Maxwell- amounts to a probabilistic description of the center of a soliton such that it Lorentz's theory of electromagnetism. Later Yang and Mills [167] as well as would follow the conventional quantum-mechanical formalism in the limit of zero Sharp were able to extend the latter theory to a nonabelian gauge particle radius.
    [Show full text]
  • Lecture 7: Ensembles
    Matthew Schwartz Statistical Mechanics, Spring 2019 Lecture 7: Ensembles 1 Introduction In statistical mechanics, we study the possible microstates of a system. We never know exactly which microstate the system is in. Nor do we care. We are interested only in the behavior of a system based on the possible microstates it could be, that share some macroscopic proporty (like volume V ; energy E, or number of particles N). The possible microstates a system could be in are known as the ensemble of states for a system. There are dierent kinds of ensembles. So far, we have been counting microstates with a xed number of particles N and a xed total energy E. We dened as the total number microstates for a system. That is (E; V ; N) = 1 (1) microstatesk withsaXmeN ;V ;E Then S = kBln is the entropy, and all other thermodynamic quantities follow from S. For an isolated system with N xed and E xed the ensemble is known as the microcanonical 1 @S ensemble. In the microcanonical ensemble, the temperature is a derived quantity, with T = @E . So far, we have only been using the microcanonical ensemble. 1 3 N For example, a gas of identical monatomic particles has (E; V ; N) V NE 2 . From this N! we computed the entropy S = kBln which at large N reduces to the Sackur-Tetrode formula. 1 @S 3 NkB 3 The temperature is T = @E = 2 E so that E = 2 NkBT . Also in the microcanonical ensemble we observed that the number of states for which the energy of one degree of freedom is xed to "i is (E "i) "i/k T (E " ).
    [Show full text]
  • Singularity Development and Supersymmetry in Holography
    Singularity development and supersymmetry in holography Alex Buchel Department of Applied Mathematics Department of Physics and Astronomy University of Western Ontario London, Ontario N6A 5B7, Canada Perimeter Institute for Theoretical Physics Waterloo, Ontario N2J 2W9, Canada Abstract We study the effects of supersymmetry on singularity development scenario in hologra- phy presented in [1] (BBL). We argue that the singularity persists in a supersymmetric extension of the BBL model. The challenge remains to find a string theory embedding of the singularity mechanism. May 23, 2017 arXiv:1705.08560v1 [hep-th] 23 May 2017 Contents 1 Introduction 2 2 DG model in microcanonical ensemble 4 2.1 Grandcanonicalensemble(areview) . 6 2.2 Microcanonicalensemble .......................... 8 2.3 DynamicsofDG-Bmodel ......................... 10 2.3.1 QNMs and linearized dynamics of DG-B model . 13 2.3.2 Fully nonlinear evolution of DG-B model . 15 3 Supersymmetric extension of BBL model 16 3.1 PhasediagramandQNMs ......................... 18 3.2 DynamicsofsBBLhorizons ........................ 21 3.2.1 Dynamics of symmetric sBBL sector and its linearized symmetry breakingfluctuations . .. .. 23 3.2.2 UnstablesBBLdynamics. 25 4 Conclusions 27 A Numerical setup 29 A.1 DG-Bmodel................................. 29 A.2 sBBLmodel................................. 30 1 Introduction Horizons are ubiquitous in holographic gauge theory/string theory correspondence [2,3]. Static horizons are dual to thermal states of the boundary gauge theory [4], while their long-wavelength near-equilibrium dynamics encode the effective boundary hydrodynamics of the theory [5]. Typically, dissipative effects in the hydrodynamics (due to shear and bulk viscosities) lead to an equilibration of a gauge theory state — a slightly perturbed horizon in a dual gravitational description settles to an equi- librium configuration.
    [Show full text]