Matrix Algebra of Spin-L/2 and Spin-L Operators

Total Page:16

File Type:pdf, Size:1020Kb

Matrix Algebra of Spin-L/2 and Spin-L Operators APPENDIX 1 Matrix Algebra of Spin-l/2 and Spin-l Operators It is frequently convenient to work with the matrix representation of spin operators in the eigenbase of the Zeeman Hamiltonian. Some results for spin-1/2 and spin-l systems are given in this Appendix. Eigenvectors Eigenvectors are represented as column matrices (kets) and row matrices (bras), while operators are square matrices. The Ill() and 1m states for spin-t/2 are represented by: Ill() == G) and 1{3) == G) (1) while, for spin 1: (2) 11) ~ m' 10) ~ m' and 1-1) ~ m Operators The matrix representation of spin operators in this eigenbase may be obtained by applying the results of angular momentum theory: IzII, m) = mil, m) I+II,m) = J(I - m)(I + m + 1) II,m + 1), 1+ = Ix + iIy (3) LII,m) = J(I + m)(I - m + 1)II,m -1), L = Ix - iIy J2 II,m) =1(1 + l)II,m). Operators 351 The matrix representation of the operator 12 is thus 1(1 + 1) times the unit matrix of dimension (21 + 1). The other operators are given by: Spin 1/2 Spin 1 (4) L G~) 1 0 0) 12 ~(l 0) 2 ( 0 1 0 4 0 1 001 The element ij of these matrices represents (iIOlj), Ii) being the ket corre­ sponding to m = mi (ml = I, m2 = I - 1, m3 = I - 2, ... ). The matrix representation of eigenvectors and operators in a composite spin system is obtained by working in the appropriate direct product space. Thus, for two spins 1/2, laa) ~ G)®G) ~ m (5) Ipa) ~ (~)®G) ~ m For such an IS spin system, we have in the direct product space: 352 Appendix 1. Matrix Algebra of Spin-lj2 and Spin-l Operators la) IP) (al (0 2Ix = (PI 1 ~ )@ ls= G ~)@(~ ~) laa) lap) IPa) IPP) (aal 0 1 (aPi 0 0 = (pal 0 0 (6) (PPI l! 0 ~J la) IP) 2S - 1 x (al ( 0 1) - C ~ x - I @ (PI 1 0 - 0 ~)@G ~) laa) lap) IPa) IPP} (aal 1 0 (aPI 0 0 (7) = (pal 0 0 (PPI ~~ 0 !J Note that the order of the spins is chosen arbitrarily; the choice made, however, must be maintained right through. The direct product of two ma­ trices A and B is given by: Ai:iB A12B ...... Ai:nBj C=A@B= .: .: (8) l . Ani B An2B ...... Ann B The direct product of an (m x m) matrix with an (n x n) matrix is thus an (mn x mn) matrix. Note that the direct product of matrices is therefore very different from matrix multiplication. Also listed below are the matrix representations of some higher powers of spin operators. These results may be checked by usual matrix multiplication. S~nlp S~nl ~ (H J ~ ( ~ ~ -~) -1 0 1 !(1 0) 4 0 1 (H ~) (9) Operators 353 Spin 1/2 Spin 1 0 [Ix, IyJ+ 0 0 i -i) (~ 0 -1 [Iy, IzJ + 0 0 fi (! -1 !) 1 [Iz,IxJ+ 0 1 (0 0 fl~ -1 -!) Various recursions then follow for i = x, y, z: If = iI7- 2 (I = 1/2, n ~ 2) (10) If = 17- 2 (I = 1, n ~ 3) For spin 1, the following relations also hold, with i = x, y, z: IJ)i =0 (i of. j) (11 ) IJf + If Ii = Ii (i of. j) In general, for a spin I, we have the Cayley identity: [Iz - IJ[Iz - (I - I)J[Iz - (I - 2)J ... [Iz + IJ =0 and I~I+l) = I~I+l) = 0; (12) /-i ) I~ ( . Il . (Iz - j) I~= 0, i = 1, 2, ... , I. J= -(/-,) Finally, we give below the matrix representation of various bilinear operator products for a two-spin-l/2 situation. The results may be verified by matrix multiplication, working with the representation of the individual single-spin operators in the direct product space of the two spins 1/2. 0 0 0 1 0 1 i [ 0 0 1 IxSx = 4 Y 1 0 IxS =4 0 -1 0 -~J ~~ 0 0 ~J 1 0 0 0 0 0 0 i 0 -1 1 0 1 IySx = 4 IySy = 4 0 -~1 0 -~J I~ 0 0 U0 0 354 Appendix 1. Matrix Algebra of Spin-1/2 and Spin-I Operators 0 0 1 0 1 0 -1 0 1 0 0 IzSx = (13) IzSz = 4 ~ 0 -1 4 0 0 ~ 0 0 ~J ~ 0 -1 -!J -1 0 0 1 i 1 0 0 1 0 0 IzSy = 4 ~o~ IxSz = 4 0 0 0 0 -] 0 -1 ~J l~ -1 0 0 -1 i 0 0 0 IySz = 4 ~ 0 0 [ -1 0 ~J From the above, it is clear that IxSx, IySy, IxSy, and IySx represent zero,- and double-quantum coherences in the spin-I/2 IS system. Operator representations for an IS system with I = 1, S = 1/2 involve (6 x 6) matrices. For example, 0 1 0) (0 1 j2Ix = (1 0 I ® Is = 1 0 o 1 0 0 I 11a) lIP) lOa) lOP) 1-1a) 1-1P) (lal 0 0 0 0 o (lPI 0 0 0 1 0 o (Oal 1 0 0 0 o (14) =(OPI 0 1 0 0 0 1 ( -lal 0 0 0 0 o ( -lPI 0 0 0 0 o (1 0 0) (15) 2Sx = II ® G~) ~ ~ ~ ~ 09 (~ ~) 0 0 0 0 1 0 0 0 0 0 0 0 0 (16) 0 0 1 0 0 0 0 ~1 0 0 0 0 0 0 0 1 ~j Operators 355 0 0 -1 0 0 0 0 0 0 -1 0 0 1 0 0 0 -1 0 j2Iy = i (17) 0 1 0 0 0 -1 0 0 1 0 0 0 0 0 0 0 0 0 -1 0 0 0 0 1 0 0 0 0 0 0 0 0 -1 0 0 2Sy= i (18) 0 0 1 0 0 0 0 0 0 0 0 -1 0 0 0 0 1 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 0 1= (19) z 0 0 0 0 0 0 0 0 0 0 -1 0 0 0 0 0 0 -1 1 0 0 0 0 0 0 -1 0 0 0 0 0 0 1 0 0 0 2Sz= (20) 0 0 0 -1 0 0 0 0 0 0 0 0 0 0 0 0 -1 APPENDIX 2 The Hausdorff Formula It is frequently necessary to evaluate expressions of the form: RA = exp( -iXt) A exp(iXt) which describe the evolution of an operator A for t seconds, under the action of a time-independent Hamiltonian x. The Baker-Campbell-Hausdorff for­ mula consolidates the series expansion of exponential operators to evaluate such operator evolutions: RA == exp (-iXt) A exp (iXt) (it)2 = A - (it) [X, A] + T![X, [X,A]] (1) (it)3 - 3! [X, [X, [X, A]]] + ... Note that this formula involves successive commutators with the Hamiltonian. An NMR problem has a well-defined, finite dimensional operator space associated with it and so the number of independent commutators is limited and the series expansion quickly leads to an operator recursion. It may be noted especially that it is crucial to work with the appropriate basis set of linearly independent operators, and to insure that successive commutators are expressed in this basis set, so that the operator recursions are not lost sight of. Suitable basis set operators for problems involving spin-l/2 and spin-l systems have been discussed in Chapter 1. We discuss below briefly some cases of interest. Appendix 2. The Hausdorff Formula 357 CASE 1. [£', A] = O. It follows that RA = A since all the commutators vanish. In other words, A is invariant to evolution under a Hamiltonian with which it commutes. For example consider a spin-1 system, with £' = tv; - 1(1 + 1)) and A = [/x,!y] + . [£', A] ~ [1;, Ixly + Iylx] = [1;, Ixly] + [1;, Iylx] = ilz(l; - In + i(l; - I;)Iz - iIAI; - I;) - iU; - I;)Iz (2) = 2i(lAI; - I;) + (I; - I;)Iz) = 2i(lz - Iz) = 0 where we have employed the properties of spin-1 operators discussed in Appendix 1, and some properties of commutators to be discussed later in this Appendix. Thus, RA = [1x'!y]+. CASE 2. [£', A] = aA. In this case, RA = A exp ( - iat). As an example, consider £' = M z and A = I + = Ix + ily • Then, [£', A] = iMy + Mx = L\A. Thus, RA = 1+ exp (-iL\t). CASE 3. [£', A] = B, [£', B] = kA, A =1= aB. This has been termed a problem of order 2. We find, RA = ( A + (i~{2 kA + (i~t k2A + ... ) _ (itB + (i~;3 kB + ... ) 2 4 2 3 _ A t k + t k _ ) _ iB (t _ t k + ) - (1 _ 2! 4! ... 3! ... (3) = A cos}kt - ~iB sin}kt Problems of order 2 have been encountered frequently in Chapters 1 and 3 and the corresponding equations of motion, R A , have been given there. We consider here a few representative examples. (1) A = Ix, £' = 2nJIzSz. [£', A] = i2nJIySz = B; (4) Thus, RA = Ix cos nJt + 2IySz sin nJt. (5) (2) A = Ix, £' = M z· [£',A] = iMy = B; (6) Thus RA = Ix cos M + Iy sin M. (7) 358 Appendix 2. The Hausdorff Formula Problems of order higher than two are encountered, for example, in AMXN systems evolving under a strong coupling Hamiltonian as discussed in Chapter 1 and lead to multiple frequencies of evolution. Commutator Algebra We give here some of the standard results of commutator algebra that are useful in evaluating the Hausdorff formula for specific Yf and A.
Recommended publications
  • 1 Notation for States
    The S-Matrix1 D. E. Soper2 University of Oregon Physics 634, Advanced Quantum Mechanics November 2000 1 Notation for states In these notes we discuss scattering nonrelativistic quantum mechanics. We will use states with the nonrelativistic normalization h~p |~ki = (2π)3δ(~p − ~k). (1) Recall that in a relativistic theory there is an extra factor of 2E on the right hand side of this relation, where E = [~k2 + m2]1/2. We will use states in the “Heisenberg picture,” in which states |ψ(t)i do not depend on time. Often in quantum mechanics one uses the Schr¨odinger picture, with time dependent states |ψ(t)iS. The relation between these is −iHt |ψ(t)iS = e |ψi. (2) Thus these are the same at time zero, and the Schrdinger states obey d i |ψ(t)i = H |ψ(t)i (3) dt S S In the Heisenberg picture, the states do not depend on time but the op- erators do depend on time. A Heisenberg operator O(t) is related to the corresponding Schr¨odinger operator OS by iHt −iHt O(t) = e OS e (4) Thus hψ|O(t)|ψi = Shψ(t)|OS|ψ(t)iS. (5) The Heisenberg picture is favored over the Schr¨odinger picture in the case of relativistic quantum mechanics: we don’t have to say which reference 1Copyright, 2000, D. E. Soper [email protected] 1 frame we use to define t in |ψ(t)iS. For operators, we can deal with local operators like, for instance, the electric field F µν(~x, t).
    [Show full text]
  • Path Integrals in Quantum Mechanics
    Path Integrals in Quantum Mechanics Dennis V. Perepelitsa MIT Department of Physics 70 Amherst Ave. Cambridge, MA 02142 Abstract We present the path integral formulation of quantum mechanics and demon- strate its equivalence to the Schr¨odinger picture. We apply the method to the free particle and quantum harmonic oscillator, investigate the Euclidean path integral, and discuss other applications. 1 Introduction A fundamental question in quantum mechanics is how does the state of a particle evolve with time? That is, the determination the time-evolution ψ(t) of some initial | i state ψ(t ) . Quantum mechanics is fully predictive [3] in the sense that initial | 0 i conditions and knowledge of the potential occupied by the particle is enough to fully specify the state of the particle for all future times.1 In the early twentieth century, Erwin Schr¨odinger derived an equation specifies how the instantaneous change in the wavefunction d ψ(t) depends on the system dt | i inhabited by the state in the form of the Hamiltonian. In this formulation, the eigenstates of the Hamiltonian play an important role, since their time-evolution is easy to calculate (i.e. they are stationary). A well-established method of solution, after the entire eigenspectrum of Hˆ is known, is to decompose the initial state into this eigenbasis, apply time evolution to each and then reassemble the eigenstates. That is, 1In the analysis below, we consider only the position of a particle, and not any other quantum property such as spin. 2 D.V. Perepelitsa n=∞ ψ(t) = exp [ iE t/~] n ψ(t ) n (1) | i − n h | 0 i| i n=0 X This (Hamiltonian) formulation works in many cases.
    [Show full text]
  • Tensor-Tensor Products with Invertible Linear Transforms
    Tensor-Tensor Products with Invertible Linear Transforms Eric Kernfelda, Misha Kilmera, Shuchin Aeronb aDepartment of Mathematics, Tufts University, Medford, MA bDepartment of Electrical and Computer Engineering, Tufts University, Medford, MA Abstract Research in tensor representation and analysis has been rising in popularity in direct response to a) the increased ability of data collection systems to store huge volumes of multidimensional data and b) the recognition of potential modeling accuracy that can be provided by leaving the data and/or the operator in its natural, multidi- mensional form. In comparison to matrix theory, the study of tensors is still in its early stages. In recent work [1], the authors introduced the notion of the t-product, a generalization of matrix multiplication for tensors of order three, which can be extended to multiply tensors of arbitrary order [2]. The multiplication is based on a convolution-like operation, which can be implemented efficiently using the Fast Fourier Transform (FFT). The corresponding linear algebraic framework from the original work was further developed in [3], and it allows one to elegantly generalize all classical algorithms from linear algebra. In the first half of this paper, we develop a new tensor-tensor product for third-order tensors that can be implemented by us- ing a discrete cosine transform, showing that a similar linear algebraic framework applies to this new tensor operation as well. The second half of the paper is devoted to extending this development in such a way that tensor-tensor products can be de- fined in a so-called transform domain for any invertible linear transform.
    [Show full text]
  • The Pattern Calculus for Tensor Operators in Quantum Groups Mark Gould and L
    The pattern calculus for tensor operators in quantum groups Mark Gould and L. C. Biedenharn Citation: Journal of Mathematical Physics 33, 3613 (1992); doi: 10.1063/1.529909 View online: http://dx.doi.org/10.1063/1.529909 View Table of Contents: http://scitation.aip.org/content/aip/journal/jmp/33/11?ver=pdfcov Published by the AIP Publishing Articles you may be interested in Irreducible tensor operators in the regular coaction formalisms of compact quantum group algebras J. Math. Phys. 37, 4590 (1996); 10.1063/1.531643 Tensor operators for quantum groups and applications J. Math. Phys. 33, 436 (1992); 10.1063/1.529833 Tensor operators of finite groups J. Math. Phys. 14, 1423 (1973); 10.1063/1.1666196 On the structure of the canonical tensor operators in the unitary groups. I. An extension of the pattern calculus rules and the canonical splitting in U(3) J. Math. Phys. 13, 1957 (1972); 10.1063/1.1665940 Irreducible tensor operators for finite groups J. Math. Phys. 13, 1892 (1972); 10.1063/1.1665928 Reuse of AIP Publishing content is subject to the terms: https://publishing.aip.org/authors/rights-and-permissions. Downloaded to IP: 130.102.42.98 On: Thu, 29 Sep 2016 02:42:08 The pattern calculus for tensor operators in quantum groups Mark Gould Department of Applied Mathematics, University of Queensland,St. Lucia, Queensland,Australia 4072 L. C. Biedenharn Department of Physics, Duke University, Durham, North Carolina 27706 (Received 28 February 1992; accepted for publication 1 June 1992) An explicit algebraic evaluation is given for all q-tensor operators (with unit norm) belonging to the quantum group U&(n)) and having extremal operator shift patterns acting on arbitrary U&u(n)) irreps.
    [Show full text]
  • Motion of the Reduced Density Operator
    Motion of the Reduced Density Operator Nicholas Wheeler, Reed College Physics Department Spring 2009 Introduction. Quantum mechanical decoherence, dissipation and measurements all involve the interaction of the system of interest with an environmental system (reservoir, measurement device) that is typically assumed to possess a great many degrees of freedom (while the system of interest is typically assumed to possess relatively few degrees of freedom). The state of the composite system is described by a density operator ρ which in the absence of system-bath interaction we would denote ρs ρe, though in the cases of primary interest that notation becomes unavailable,⊗ since in those cases the states of the system and its environment are entangled. The observable properties of the system are latent then in the reduced density operator ρs = tre ρ (1) which is produced by “tracing out” the environmental component of ρ. Concerning the specific meaning of (1). Let n) be an orthonormal basis | in the state space H of the (open) system, and N) be an orthonormal basis s !| " in the state space H of the (also open) environment. Then n) N) comprise e ! " an orthonormal basis in the state space H = H H of the |(closed)⊗| composite s e ! " system. We are in position now to write ⊗ tr ρ I (N ρ I N) e ≡ s ⊗ | s ⊗ | # ! " ! " ↓ = ρ tr ρ in separable cases s · e The dynamics of the composite system is generated by Hamiltonian of the form H = H s + H e + H i 2 Motion of the reduced density operator where H = h I s s ⊗ e = m h n m N n N $ | s| % | % ⊗ | % · $ | ⊗ $ | m,n N # # $% & % &' H = I h e s ⊗ e = n M n N M h N | % ⊗ | % · $ | ⊗ $ | $ | e| % n M,N # # $% & % &' H = m M m M H n N n N i | % ⊗ | % $ | ⊗ $ | i | % ⊗ | % $ | ⊗ $ | m,n M,N # # % &$% & % &'% & —all components of which we will assume to be time-independent.
    [Show full text]
  • Quantum Field Theory*
    Quantum Field Theory y Frank Wilczek Institute for Advanced Study, School of Natural Science, Olden Lane, Princeton, NJ 08540 I discuss the general principles underlying quantum eld theory, and attempt to identify its most profound consequences. The deep est of these consequences result from the in nite number of degrees of freedom invoked to implement lo cality.Imention a few of its most striking successes, b oth achieved and prosp ective. Possible limitation s of quantum eld theory are viewed in the light of its history. I. SURVEY Quantum eld theory is the framework in which the regnant theories of the electroweak and strong interactions, which together form the Standard Mo del, are formulated. Quantum electro dynamics (QED), b esides providing a com- plete foundation for atomic physics and chemistry, has supp orted calculations of physical quantities with unparalleled precision. The exp erimentally measured value of the magnetic dip ole moment of the muon, 11 (g 2) = 233 184 600 (1680) 10 ; (1) exp: for example, should b e compared with the theoretical prediction 11 (g 2) = 233 183 478 (308) 10 : (2) theor: In quantum chromo dynamics (QCD) we cannot, for the forseeable future, aspire to to comparable accuracy.Yet QCD provides di erent, and at least equally impressive, evidence for the validity of the basic principles of quantum eld theory. Indeed, b ecause in QCD the interactions are stronger, QCD manifests a wider variety of phenomena characteristic of quantum eld theory. These include esp ecially running of the e ective coupling with distance or energy scale and the phenomenon of con nement.
    [Show full text]
  • Arxiv:2012.13347V1 [Physics.Class-Ph] 15 Dec 2020
    KPOP E-2020-04 Bra-Ket Representation of the Inertia Tensor U-Rae Kim, Dohyun Kim, and Jungil Lee∗ KPOPE Collaboration, Department of Physics, Korea University, Seoul 02841, Korea (Dated: June 18, 2021) Abstract We employ Dirac's bra-ket notation to define the inertia tensor operator that is independent of the choice of bases or coordinate system. The principal axes and the corresponding principal values for the elliptic plate are determined only based on the geometry. By making use of a general symmetric tensor operator, we develop a method of diagonalization that is convenient and intuitive in determining the eigenvector. We demonstrate that the bra-ket approach greatly simplifies the computation of the inertia tensor with an example of an N-dimensional ellipsoid. The exploitation of the bra-ket notation to compute the inertia tensor in classical mechanics should provide undergraduate students with a strong background necessary to deal with abstract quantum mechanical problems. PACS numbers: 01.40.Fk, 01.55.+b, 45.20.dc, 45.40.Bb Keywords: Classical mechanics, Inertia tensor, Bra-ket notation, Diagonalization, Hyperellipsoid arXiv:2012.13347v1 [physics.class-ph] 15 Dec 2020 ∗Electronic address: [email protected]; Director of the Korea Pragmatist Organization for Physics Educa- tion (KPOP E) 1 I. INTRODUCTION The inertia tensor is one of the essential ingredients in classical mechanics with which one can investigate the rotational properties of rigid-body motion [1]. The symmetric nature of the rank-2 Cartesian tensor guarantees that it is described by three fundamental parameters called the principal moments of inertia Ii, each of which is the moment of inertia along a principal axis.
    [Show full text]
  • An S-Matrix for Massless Particles Arxiv:1911.06821V2 [Hep-Th]
    An S-Matrix for Massless Particles Holmfridur Hannesdottir and Matthew D. Schwartz Department of Physics, Harvard University, Cambridge, MA 02138, USA Abstract The traditional S-matrix does not exist for theories with massless particles, such as quantum electrodynamics. The difficulty in isolating asymptotic states manifests itself as infrared divergences at each order in perturbation theory. Building on insights from the literature on coherent states and factorization, we construct an S-matrix that is free of singularities order-by-order in perturbation theory. Factorization guarantees that the asymptotic evolution in gauge theories is universal, i.e. independent of the hard process. Although the hard S-matrix element is computed between well-defined few particle Fock states, dressed/coherent states can be seen to form as intermediate states in the calculation of hard S-matrix elements. We present a framework for the perturbative calculation of hard S-matrix elements combining Lorentz-covariant Feyn- man rules for the dressed-state scattering with time-ordered perturbation theory for the asymptotic evolution. With hard cutoffs on the asymptotic Hamiltonian, the cancella- tion of divergences can be seen explicitly. In dimensional regularization, where the hard cutoffs are replaced by a renormalization scale, the contribution from the asymptotic evolution produces scaleless integrals that vanish. A number of illustrative examples are given in QED, QCD, and N = 4 super Yang-Mills theory. arXiv:1911.06821v2 [hep-th] 24 Aug 2020 Contents 1 Introduction1 2 The hard S-matrix6 2.1 SH and dressed states . .9 2.2 Computing observables using SH ........................... 11 2.3 Soft Wilson lines . 14 3 Computing the hard S-matrix 17 3.1 Asymptotic region Feynman rules .
    [Show full text]
  • 5 the Dirac Equation and Spinors
    5 The Dirac Equation and Spinors In this section we develop the appropriate wavefunctions for fundamental fermions and bosons. 5.1 Notation Review The three dimension differential operator is : ∂ ∂ ∂ = , , (5.1) ∂x ∂y ∂z We can generalise this to four dimensions ∂µ: 1 ∂ ∂ ∂ ∂ ∂ = , , , (5.2) µ c ∂t ∂x ∂y ∂z 5.2 The Schr¨odinger Equation First consider a classical non-relativistic particle of mass m in a potential U. The energy-momentum relationship is: p2 E = + U (5.3) 2m we can substitute the differential operators: ∂ Eˆ i pˆ i (5.4) → ∂t →− to obtain the non-relativistic Schr¨odinger Equation (with = 1): ∂ψ 1 i = 2 + U ψ (5.5) ∂t −2m For U = 0, the free particle solutions are: iEt ψ(x, t) e− ψ(x) (5.6) ∝ and the probability density ρ and current j are given by: 2 i ρ = ψ(x) j = ψ∗ ψ ψ ψ∗ (5.7) | | −2m − with conservation of probability giving the continuity equation: ∂ρ + j =0, (5.8) ∂t · Or in Covariant notation: µ µ ∂µj = 0 with j =(ρ,j) (5.9) The Schr¨odinger equation is 1st order in ∂/∂t but second order in ∂/∂x. However, as we are going to be dealing with relativistic particles, space and time should be treated equally. 25 5.3 The Klein-Gordon Equation For a relativistic particle the energy-momentum relationship is: p p = p pµ = E2 p 2 = m2 (5.10) · µ − | | Substituting the equation (5.4), leads to the relativistic Klein-Gordon equation: ∂2 + 2 ψ = m2ψ (5.11) −∂t2 The free particle solutions are plane waves: ip x i(Et p x) ψ e− · = e− − · (5.12) ∝ The Klein-Gordon equation successfully describes spin 0 particles in relativistic quan- tum field theory.
    [Show full text]
  • The Critical Casimir Effect in Model Physical Systems
    University of California Los Angeles The Critical Casimir Effect in Model Physical Systems A dissertation submitted in partial satisfaction of the requirements for the degree Doctor of Philosophy in Physics by Jonathan Ariel Bergknoff 2012 ⃝c Copyright by Jonathan Ariel Bergknoff 2012 Abstract of the Dissertation The Critical Casimir Effect in Model Physical Systems by Jonathan Ariel Bergknoff Doctor of Philosophy in Physics University of California, Los Angeles, 2012 Professor Joseph Rudnick, Chair The Casimir effect is an interaction between the boundaries of a finite system when fluctua- tions in that system correlate on length scales comparable to the system size. In particular, the critical Casimir effect is that which arises from the long-ranged thermal fluctuation of the order parameter in a system near criticality. Recent experiments on the Casimir force in binary liquids near critical points and 4He near the superfluid transition have redoubled theoretical interest in the topic. It is an unfortunate fact that exact models of the experi- mental systems are mathematically intractable in general. However, there is often insight to be gained by studying approximations and toy models, or doing numerical computations. In this work, we present a brief motivation and overview of the field, followed by explications of the O(2) model with twisted boundary conditions and the O(n ! 1) model with free boundary conditions. New results, both analytical and numerical, are presented. ii The dissertation of Jonathan Ariel Bergknoff is approved. Giovanni Zocchi Alex Levine Lincoln Chayes Joseph Rudnick, Committee Chair University of California, Los Angeles 2012 iii To my parents, Hugh and Esther Bergknoff iv Table of Contents 1 Introduction :::::::::::::::::::::::::::::::::::::: 1 1.1 The Casimir Effect .
    [Show full text]
  • Two-State Systems
    1 TWO-STATE SYSTEMS Introduction. Relative to some/any discretely indexed orthonormal basis |n) | ∂ | the abstract Schr¨odinger equation H ψ)=i ∂t ψ) can be represented | | | ∂ | (m H n)(n ψ)=i ∂t(m ψ) n ∂ which can be notated Hmnψn = i ∂tψm n H | ∂ | or again ψ = i ∂t ψ We found it to be the fundamental commutation relation [x, p]=i I which forced the matrices/vectors thus encountered to be ∞-dimensional. If we are willing • to live without continuous spectra (therefore without x) • to live without analogs/implications of the fundamental commutator then it becomes possible to contemplate “toy quantum theories” in which all matrices/vectors are finite-dimensional. One loses some physics, it need hardly be said, but surprisingly much of genuine physical interest does survive. And one gains the advantage of sharpened analytical power: “finite-dimensional quantum mechanics” provides a methodological laboratory in which, not infrequently, the essentials of complicated computational procedures can be exposed with closed-form transparency. Finally, the toy theory serves to identify some unanticipated formal links—permitting ideas to flow back and forth— between quantum mechanics and other branches of physics. Here we will carry the technique to the limit: we will look to “2-dimensional quantum mechanics.” The theory preserves the linearity that dominates the full-blown theory, and is of the least-possible size in which it is possible for the effects of non-commutivity to become manifest. 2 Quantum theory of 2-state systems We have seen that quantum mechanics can be portrayed as a theory in which • states are represented by self-adjoint linear operators ρ ; • motion is generated by self-adjoint linear operators H; • measurement devices are represented by self-adjoint linear operators A.
    [Show full text]
  • Introduction to Vector and Tensor Analysis
    Introduction to vector and tensor analysis Jesper Ferkinghoff-Borg September 6, 2007 Contents 1 Physical space 3 1.1 Coordinate systems . 3 1.2 Distances . 3 1.3 Symmetries . 4 2 Scalars and vectors 5 2.1 Definitions . 5 2.2 Basic vector algebra . 5 2.2.1 Scalar product . 6 2.2.2 Cross product . 7 2.3 Coordinate systems and bases . 7 2.3.1 Components and notation . 9 2.3.2 Triplet algebra . 10 2.4 Orthonormal bases . 11 2.4.1 Scalar product . 11 2.4.2 Cross product . 12 2.5 Ordinary derivatives and integrals of vectors . 13 2.5.1 Derivatives . 14 2.5.2 Integrals . 14 2.6 Fields . 15 2.6.1 Definition . 15 2.6.2 Partial derivatives . 15 2.6.3 Differentials . 16 2.6.4 Gradient, divergence and curl . 17 2.6.5 Line, surface and volume integrals . 20 2.6.6 Integral theorems . 24 2.7 Curvilinear coordinates . 26 2.7.1 Cylindrical coordinates . 27 2.7.2 Spherical coordinates . 28 3 Tensors 30 3.1 Definition . 30 3.2 Outer product . 31 3.3 Basic tensor algebra . 31 1 3.3.1 Transposed tensors . 32 3.3.2 Contraction . 33 3.3.3 Special tensors . 33 3.4 Tensor components in orthonormal bases . 34 3.4.1 Matrix algebra . 35 3.4.2 Two-point components . 38 3.5 Tensor fields . 38 3.5.1 Gradient, divergence and curl . 39 3.5.2 Integral theorems . 40 4 Tensor calculus 42 4.1 Tensor notation and Einsteins summation rule .
    [Show full text]