Building Invariants from Spinors

Total Page:16

File Type:pdf, Size:1020Kb

Building Invariants from Spinors Building Invariants from Spinors In the previous section, we found that there are two independent 2 × 2 irreducible rep- resentations of the Lorentz group (more specifically the proper orthochronous Lorentz group that excludes parity and time reversal), corresponding to the generators 1 i J i = σi Ki = − σi 2 2 or 1 i J i = σi Ki = σi 2 2 where σi are the Pauli matrices (with eigenvalues ±1). This means that we can have a fields ηa or χa with two components such that under infinitesimal rotations i i δη = ϵ (σi)η δχ = ϵ (σi)χ 2 2 while under infinitesimal boosts 1 1 δη = ϵ (σi)η δχ = −ϵ (σi)χ : 2 2 We also saw that there is no way to define a parity transformation in a theory with η or χ alone, but if both types of field are included, a consistent way for the fields to transform under parity is η $ χ. In such a theory, we can combine these two fields into a four-component object ! ηa α = ; (1) χa which we call a Dirac spinor. We would now like to understand how to build Lorentz- invariant actions using Dirac spinor fields. Help from quantum mechanics Consider a spin half particle in quantum mechanics. We can write the general state of such a particle (considering only the spin degree of freedom) as j i = 1 j "i + − 1 j #i 2 2 ! 1 2 or, in vector notation for the Jz basis, = . Under an small rotation, the − 1 2 infinitesimal change in the two components of is given by i δ = ϵ (σi) (2) 2 which follows from the general result that the change in any state under an infinitesimal rotation about the i-axis is δj i = ϵiJ ij i : The rule (2) is exactly the same as the rules for how the two-component spinor fields transform under rotations. Thus, we can understand how to make rotationally invariant quantities with spinor fields if we know how to make rotationally invariant quantities in this quantum mechanics system. In quantum mechanics, a basic quantity that is invariant under any symmetry transformation is the inner product h 1j 2i between state vectors. Translating to the spin-z basis, we can say that if 1 and 2 are the explicit two-component vectors representing any two states, that under a rotation, the quantity y 1 2 (3) y is invariant. If we consider more generally the four quantities a b, what we have seen is that one linear combination of these four quantities is invariant upon performing a rotation. It is easy to show that the remaining three linearly independent quantities are mixed together by rotations in the same way as the three components of a vector. Explicitly, if we write y i 1σ 2 ; these three quantities rotate like an ordinary vector when we perform a rotation. This is simplest to see in the case where 2 = 1, in which case the three quantities are simply the expectation values of J i for i = x; y; z. Back to field theory Since the quantity (3) is invariant and since fields η and χ transform in exactly the y same way as 1 and 2 under rotations, it follows immediately that the quantities η η, ηyχ, χyη, χyχ are all invariant under rotations. We can check this explicitly using the transformation rules given above. We can also use these rules to check which of these four quantities is also invariant under a boost. As an example, under a small boost, we have δ(χyχ) = δχyχ + χyδχ 1 1 = −ϵ χy(σi)yχ − ϵ χy(σi)χ 2 2 = −ϵχyσiχ (4) y y where we have used that σi = σi. Thus, χ χ is not invariant under boosts. Similarly, we find that ηyη is not invariant under boosts, but ηyχ and χyη are both invariant. Recalling that a parity transformation acts as η $ χ we find that these two quantities are not independently invariant under parity transformations, but the combination ηyχ + χyη is invariant under rotations, parity transformations, and boosts. This is our first exam- ple of a possible Lorentz-invariant and parity invariant term built from spinor fields.1 1Here, we really mean that this quantity transforms like a scalar field. We can write this directly in terms of the Dirac spinor field (1), by noting that it ∗ is a linear combination of the quantities α β. Specifically, we have y y ∗ y η χ + χ η = Mαβ α β M where M is the 4 × 4 matrix ! 0 11 M = ≡ γ0 : 11 0 We have chosen to call this matrix γ0 for reasons that will become apparent. Using this notation, our invariant quantity is yγ0 : Making the further definition that ¯ = yγ0 ; we can write the invariant quantity simply as ¯ Other quantities with two spinors For constructing actions, it will also be useful to understand how other quantities built from two spinor fields transform under Lorentz transformations. The sixteen quan- ∗ 2 tities α β, will mix together when we perform a Lorentz transformation We have already seen that one linear combination, ¯ is invariant under Lorentz transforma- tions. To understand how the remaining quantities transform, we begin by recalling the transformation (4) for the rotationally invariant quantity χyχ under a boost in the i-direction: δ(χyχ) = −ϵχyσiχ : This transformation rule looks just like the transformation rule δt = −ϵxi for the time coordinate under a boost in the i-direction, if we associate t ! χyχ and xi ! χyσiχ. This suggests that the four quantities (χyχ, χyσiχ) transform as a four- vector under under rotations and boosts. We can verify this by checking that the 2To see this, note that under a Lorentz transformation, ∗ ! ∗ ∗ ≡ ∗ α β M(Λ)ασM(Λ)βτ σ τ M(αβ);(στ) σ τ : The matrix M defines some representation of the Lorentz group known as the TENSOR PRODUCT representation of M and M ∗, since M = M y ⊗ M. This representation is reducible, which means that ∗ we can split the sixteen quantities α β into subsets such that Lorentz transformations only mix the quantities in a subset. quantities χyσiχ transform under boosts in the same way as the spatial components of a four-vector. In a similar way, we find that the four quantities (ηyη; −ηyσiη) transform as a four-vector under rotations and boosts. A true four-vector should also transform under parity transformations in the same way as the coordinates, that is (t; xi) ! (t; −xi). From the two potential four-vectors we have just found, one linear combination has this property, namely ! ηyη + χyχ −ηyσiη + χyσiχ y Each of these four quantities is a linear combination of the quantities α β. Using the definition of γ0 above, and defining three new 4 × 4 matrices γi as ! 0 σi γi = −σi 0 it is easy to check that ηyη + χyχ = y = ¯γ0 ; while −ηyσiη + χyσiχ = ¯γi ; In summary, we have found so far that ¯ is a scalar quantity while the four quantities ¯γµ (µ = 0; i) transform as a four-vector under Lorentz transformations. Building tensors with gamma matrices The fact that ¯ is a scalar quantity under Lorentz transformations implies that if transforms like ! MD(Λ) under a general Lorentz transformation (`D' stands for Dirac spinor), then ¯ must transform as ¯ ! ¯ −1 MD (Λ) : With these transformation rules, it follows that under a Lorentz transformation ¯ µ ! ¯ −1 µ γ MD (Λ)γ MD(Λ) : But we found earlier that ¯γµ transforms as a vector quantity, which means that ¯ µ µ ¯ ν γ ! Λ ν γ : For these two transformation rules to agree with each other for any Λ, it must be that −1 µ µ ν MD (Λ)γ MD(Λ) = Λ νγ : Using this, it immediately follows that quantities with additional γ matrices also transform like tensors. We have − ¯ µ1 ··· µn ! ¯ 1 µ1 ··· µn γ γ MD (Λ)γ γ MD(Λ) − − ¯ 1 µ1 ··· 1 µn = MD (Λ)γ MD(Λ) MD (Λ)γ MD(Λ) µ1 ··· µn ¯ ν1 ··· νn = Λ ν1 Λ νn γ γ : These tensors cannot all be independent of one another, since they are all linear com- ∗ binations of only sixteen independent quantities α β. The reason is that products of γ matrices can be simplified using the relations µ ν ν µ µν γ γ + γ γ = 2η 114×4 ; (5) which can be checked from the definition of γµ. So for example, the tensor ¯γµγν ; appears to have 16 independent components, but if we split it into a symmetric part 1 Sµν = ¯ (γµγν + γνγµ) 2 and an antisymmetric part 1 Aµν = ¯ (γµγν − γνγµ) ; 2 then the six independent components of this antisymmetric tensor are independent of the scalar and vector quantities constructed above, but the symmetric part can be rewritten using (5) as Sµν = 2ηµν ¯ : In a similar way, for tensors built using more γ matrices only the completely antisym- metric combinations give new independent terms. In four dimensions, the quantity 1 ¯ (γµγνγλ − γνγµγλ + :::) 6 has four independent components, while the tensor 1 ¯ (γµγνγλγσ − γνγµγλγσ + :::) 24 just has a single independent component (for this to be non-zero, we must have fµ, ν; λ, σg = f0; 1; 2; 3g otherwise everything cancels). Explicitly, we have 1 ¯ (γµγνγλγσ − γνγµγλγσ + :::) = ϵµνλσ ¯γ0γ1γ2γ3 : 24 It is standard to define γ5 = iγ0γ1γ2γ3 so that the single independent quantity that we get by including an antisymmetric product of four gamma matrices between ¯ and may be written ¯γ5 This quantity transforms like a scalar under rotations and boosts, but changes its sign under parity transformations. This type of representation of the Lorentz group (including parity) is known as a PSEUDOSCALAR.
Recommended publications
  • Quaternions and Cli Ord Geometric Algebras
    Quaternions and Cliord Geometric Algebras Robert Benjamin Easter First Draft Edition (v1) (c) copyright 2015, Robert Benjamin Easter, all rights reserved. Preface As a rst rough draft that has been put together very quickly, this book is likely to contain errata and disorganization. The references list and inline citations are very incompete, so the reader should search around for more references. I do not claim to be the inventor of any of the mathematics found here. However, some parts of this book may be considered new in some sense and were in small parts my own original research. Much of the contents was originally written by me as contributions to a web encyclopedia project just for fun, but for various reasons was inappropriate in an encyclopedic volume. I did not originally intend to write this book. This is not a dissertation, nor did its development receive any funding or proper peer review. I oer this free book to the public, such as it is, in the hope it could be helpful to an interested reader. June 19, 2015 - Robert B. Easter. (v1) [email protected] 3 Table of contents Preface . 3 List of gures . 9 1 Quaternion Algebra . 11 1.1 The Quaternion Formula . 11 1.2 The Scalar and Vector Parts . 15 1.3 The Quaternion Product . 16 1.4 The Dot Product . 16 1.5 The Cross Product . 17 1.6 Conjugates . 18 1.7 Tensor or Magnitude . 20 1.8 Versors . 20 1.9 Biradials . 22 1.10 Quaternion Identities . 23 1.11 The Biradial b/a .
    [Show full text]
  • 1 Parity 2 Time Reversal
    Even Odd Symmetry Lecture 9 1 Parity The normal modes of a string have either even or odd symmetry. This also occurs for stationary states in Quantum Mechanics. The transformation is called partiy. We previously found for the harmonic oscillator that there were 2 distinct types of wave function solutions characterized by the selection of the starting integer in their series representation. This selection produced a series in odd or even powers of the coordiante so that the wave function was either odd or even upon reflections about the origin, x = 0. Since the potential energy function depends on the square of the position, x2, the energy eignevalue was always positive and independent of whether the eigenfunctions were odd or even under reflection. In 1-D parity is the symmetry operation, x → −x. In 3-D the strong interaction is invarient under the symmetry of parity. ~r → −~r Parity is a mirror reflection plus a rotation of 180◦, and transforms a right-handed coordinate system into a left-handed one. Our Macroscopic world is clearly “handed”, but “handedness” in fundamental interactions is more involved. Vectors (tensors of rank 1), as illustrated in the definition above, change sign under Parity. Scalars (tensors of rank 0) do not. One can then construct, using tensor algebra, new tensors which reduce the tensor rank and/or change the symmetry of the tensor. Thus a dual of a symmetric tensor of rank 2 is a pseudovector (cross product of two vectors), and a scalar product of a pseudovector and a vector creates a pseudoscalar. We will construct bilinear forms below which have these rotational and reflection char- acteristics.
    [Show full text]
  • Arxiv:Math/0212058V2 [Math.DG] 18 Nov 2003 Nosbrusof Subgroups Into Rdc C.[Oc 00 Et .].Eape O Uhsae a Spaces Such for Examples 3.2])
    THE SPINOR BUNDLE OF RIEMANNIAN PRODUCTS FRANK KLINKER Abstract. In this note we compare the spinor bundle of a Riemannian mani- fold (M = M1 ×···×MN ,g) with the spinor bundles of the Riemannian factors (Mi,gi). We show that - without any holonomy conditions - the spinor bundle of (M,g) for a special class of metrics is isomorphic to a bundle obtained by tensoring the spinor bundles of (Mi,gi) in an appropriate way. For N = 2 and an one dimensional factor this construction was developed in [Baum 1989a]. Although the fact for general factors is frequently used in (at least physics) literature, a proof was missing. I would like to thank Shahram Biglari, Mario Listing, Marc Nardmann and Hans-Bert Rademacher for helpful comments. Special thanks go to Helga Baum, who pointed out some difficulties arising in the pseudo-Riemannian case. We consider a Riemannian manifold (M = M MN ,g), which is a product 1 ×···× of Riemannian spin manifolds (Mi,gi) and denote the projections on the respective factors by pi. Furthermore the dimension of Mi is Di such that the dimension of N M is given by D = i=1 Di. The tangent bundle of M is decomposed as P ∗ ∗ (1) T M = p Tx1 M p Tx MN . (x0,...,xN ) 1 1 ⊕···⊕ N N N We omit the projections and write TM = i=1 TMi. The metric g of M need not be the product metric of the metrics g on M , but L i i is assumed to be of the form c d (2) gab(x) = Ai (x)gi (xi)Ai (x), T Mi a cd b for D1 + + Di−1 +1 a,b D1 + + Di, 1 i N ··· ≤ ≤ ··· ≤ ≤ In particular, for those metrics the splitting (1) is orthogonal, i.e.
    [Show full text]
  • Introductory Lectures on Quantum Field Theory
    Introductory Lectures on Quantum Field Theory a b L. Álvarez-Gaumé ∗ and M.A. Vázquez-Mozo † a CERN, Geneva, Switzerland b Universidad de Salamanca, Salamanca, Spain Abstract In these lectures we present a few topics in quantum field theory in detail. Some of them are conceptual and some more practical. They have been se- lected because they appear frequently in current applications to particle physics and string theory. 1 Introduction These notes are based on lectures delivered by L.A.-G. at the 3rd CERN–Latin-American School of High- Energy Physics, Malargüe, Argentina, 27 February–12 March 2005, at the 5th CERN–Latin-American School of High-Energy Physics, Medellín, Colombia, 15–28 March 2009, and at the 6th CERN–Latin- American School of High-Energy Physics, Natal, Brazil, 23 March–5 April 2011. The audience on all three occasions was composed to a large extent of students in experimental high-energy physics with an important minority of theorists. In nearly ten hours it is quite difficult to give a reasonable introduction to a subject as vast as quantum field theory. For this reason the lectures were intended to provide a review of those parts of the subject to be used later by other lecturers. Although a cursory acquaintance with the subject of quantum field theory is helpful, the only requirement to follow the lectures is a working knowledge of quantum mechanics and special relativity. The guiding principle in choosing the topics presented (apart from serving as introductions to later courses) was to present some basic aspects of the theory that present conceptual subtleties.
    [Show full text]
  • On the Spinor Representation
    Eur. Phys. J. C (2017) 77:487 DOI 10.1140/epjc/s10052-017-5035-y Regular Article - Theoretical Physics On the spinor representation J. M. Hoff da Silva1,a, C. H. Coronado Villalobos1,2,b, Roldão da Rocha3,c, R. J. Bueno Rogerio1,d 1 Departamento de Física e Química, Universidade Estadual Paulista, Guaratinguetá, SP, Brazil 2 Instituto de Física, Universidade Federal Fluminense, Av. Gal. Milton Tavares de Souza s/n, Gragoatá, Niterói, RJ 24210-346, Brazil 3 Centro de Matemática, Computação e Cognição, Universidade Federal do ABC-UFABC, Santo André 09210-580, Brazil Received: 16 February 2017 / Accepted: 30 June 2017 / Published online: 21 July 2017 © The Author(s) 2017. This article is an open access publication Abstract A systematic study of the spinor representation at least in principle, related to a set of fermionic observ- by means of the fermionic physical space is accomplished ables. Our aim in this paper is to delineate the importance and implemented. The spinor representation space is shown of the representation space, by studying its properties, and to be constrained by the Fierz–Pauli–Kofink identities among then relating them to their physical consequences. As one the spinor bilinear covariants. A robust geometric and topo- will realize, the representation space is quite complicated due logical structure can be manifested from the spinor space, to the constraints coming out the Fierz–Pauli–Kofink iden- wherein the first and second homotopy groups play promi- tities. However, a systematic study of the space properties nent roles on the underlying physical properties, associated ends up being useful to relate different domains (subspaces) to fermionic fields.
    [Show full text]
  • Symmetry and Tensors Rotations and Tensors
    Symmetry and tensors Rotations and tensors A rotation of a 3-vector is accomplished by an orthogonal transformation. Represented as a matrix, A, we replace each vector, v, by a rotated vector, v0, given by multiplying by A, 0 v = Av In index notation, 0 X vm = Amnvn n Since a rotation must preserve lengths of vectors, we require 02 X 0 0 X 2 v = vmvm = vmvm = v m m Therefore, X X 0 0 vmvm = vmvm m m ! ! X X X = Amnvn Amkvk m n k ! X X = AmnAmk vnvk k;n m Since xn is arbitrary, this is true if and only if X AmnAmk = δnk m t which we can rewrite using the transpose, Amn = Anm, as X t AnmAmk = δnk m In matrix notation, this is t A A = I where I is the identity matrix. This is equivalent to At = A−1. Multi-index objects such as matrices, Mmn, or the Levi-Civita tensor, "ijk, have definite transformation properties under rotations. We call an object a (rotational) tensor if each index transforms in the same way as a vector. An object with no indices, that is, a function, does not transform at all and is called a scalar. 0 A matrix Mmn is a (second rank) tensor if and only if, when we rotate vectors v to v , its new components are given by 0 X Mmn = AmjAnkMjk jk This is what we expect if we imagine Mmn to be built out of vectors as Mmn = umvn, for example. In the same way, we see that the Levi-Civita tensor transforms as 0 X "ijk = AilAjmAkn"lmn lmn 1 Recall that "ijk, because it is totally antisymmetric, is completely determined by only one of its components, say, "123.
    [Show full text]
  • On Manifolds of Tensors of Fixed Tt-Rank
    ON MANIFOLDS OF TENSORS OF FIXED TT-RANK SEBASTIAN HOLTZ, THORSTEN ROHWEDDER, AND REINHOLD SCHNEIDER Abstract. Recently, the format of TT tensors [19, 38, 34, 39] has turned out to be a promising new format for the approximation of solutions of high dimensional problems. In this paper, we prove some new results for the TT representation of a tensor U ∈ Rn1×...×nd and for the manifold of tensors of TT-rank r. As a first result, we prove that the TT (or compression) ranks ri of a tensor U are unique and equal to the respective seperation ranks of U if the components of the TT decomposition are required to fulfil a certain maximal rank condition. We then show that d the set T of TT tensors of fixed rank r forms an embedded manifold in Rn , therefore preserving the essential theoretical properties of the Tucker format, but often showing an improved scaling behaviour. Extending a similar approach for matrices [7], we introduce certain gauge conditions to obtain a unique representation of the tangent space TU T of T and deduce a local parametrization of the TT manifold. The parametrisation of TU T is often crucial for an algorithmic treatment of high-dimensional time-dependent PDEs and minimisation problems [33]. We conclude with remarks on those applications and present some numerical examples. 1. Introduction The treatment of high-dimensional problems, typically of problems involving quantities from Rd for larger dimensions d, is still a challenging task for numerical approxima- tion. This is owed to the principal problem that classical approaches for their treatment normally scale exponentially in the dimension d in both needed storage and computa- tional time and thus quickly become computationally infeasable for sensible discretiza- tions of problems of interest.
    [Show full text]
  • Majorana Spinors
    MAJORANA SPINORS JOSE´ FIGUEROA-O'FARRILL Contents 1. Complex, real and quaternionic representations 2 2. Some basis-dependent formulae 5 3. Clifford algebras and their spinors 6 4. Complex Clifford algebras and the Majorana condition 10 5. Examples 13 One dimension 13 Two dimensions 13 Three dimensions 14 Four dimensions 14 Six dimensions 15 Ten dimensions 16 Eleven dimensions 16 Twelve dimensions 16 ...and back! 16 Summary 19 References 19 These notes arose as an attempt to conceptualise the `symplectic Majorana{Weyl condition' in 5+1 dimensions; but have turned into a general discussion of spinors. Spinors play a crucial role in supersymmetry. Part of their versatility is that they come in many guises: `Dirac', `Majorana', `Weyl', `Majorana{Weyl', `symplectic Majorana', `symplectic Majorana{Weyl', and their `pseudo' counterparts. The tra- ditional physics approach to this topic is a mixed bag of tricks using disparate aspects of representation theory of finite groups. In these notes we will attempt to provide a uniform treatment based on the classification of Clifford algebras, a work dating back to the early 60s and all but ignored by the theoretical physics com- munity. Recent developments in superstring theory have made us re-examine the conditions for the existence of different kinds of spinors in spacetimes of arbitrary signature, and we believe that a discussion of this more uniform approach is timely and could be useful to the student meeting this topic for the first time or to the practitioner who has difficulty remembering the answer to questions like \when do symplectic Majorana{Weyl spinors exist?" The notes are organised as follows.
    [Show full text]
  • Determinants in Geometric Algebra
    Determinants in Geometric Algebra Eckhard Hitzer 16 June 2003, recovered+expanded May 2020 1 Definition Let f be a linear map1, of a real linear vector space Rn into itself, an endomor- phism n 0 n f : a 2 R ! a 2 R : (1) This map is extended by outermorphism (symbol f) to act linearly on multi- vectors f(a1 ^ a2 ::: ^ ak) = f(a1) ^ f(a2) ::: ^ f(ak); k ≤ n: (2) By definition f is grade-preserving and linear, mapping multivectors to mul- tivectors. Examples are the reflections, rotations and translations described earlier. The outermorphism of a product of two linear maps fg is the product of the outermorphisms f g f[g(a1)] ^ f[g(a2)] ::: ^ f[g(ak)] = f[g(a1) ^ g(a2) ::: ^ g(ak)] = f[g(a1 ^ a2 ::: ^ ak)]; (3) with k ≤ n. The square brackets can safely be omitted. The n{grade pseudoscalars of a geometric algebra are unique up to a scalar factor. This can be used to define the determinant2 of a linear map as det(f) = f(I)I−1 = f(I) ∗ I−1; and therefore f(I) = det(f)I: (4) For an orthonormal basis fe1; e2;:::; eng the unit pseudoscalar is I = e1e2 ::: en −1 q q n(n−1)=2 with inverse I = (−1) enen−1 ::: e1 = (−1) (−1) I, where q gives the number of basis vectors, that square to −1 (the linear space is then Rp;q). According to Grassmann n-grade vectors represent oriented volume elements of dimension n. The determinant therefore shows how these volumes change under linear maps.
    [Show full text]
  • Matrices and Tensors
    APPENDIX MATRICES AND TENSORS A.1. INTRODUCTION AND RATIONALE The purpose of this appendix is to present the notation and most of the mathematical tech- niques that are used in the body of the text. The audience is assumed to have been through sev- eral years of college-level mathematics, which included the differential and integral calculus, differential equations, functions of several variables, partial derivatives, and an introduction to linear algebra. Matrices are reviewed briefly, and determinants, vectors, and tensors of order two are described. The application of this linear algebra to material that appears in under- graduate engineering courses on mechanics is illustrated by discussions of concepts like the area and mass moments of inertia, Mohr’s circles, and the vector cross and triple scalar prod- ucts. The notation, as far as possible, will be a matrix notation that is easily entered into exist- ing symbolic computational programs like Maple, Mathematica, Matlab, and Mathcad. The desire to represent the components of three-dimensional fourth-order tensors that appear in anisotropic elasticity as the components of six-dimensional second-order tensors and thus rep- resent these components in matrices of tensor components in six dimensions leads to the non- traditional part of this appendix. This is also one of the nontraditional aspects in the text of the book, but a minor one. This is described in §A.11, along with the rationale for this approach. A.2. DEFINITION OF SQUARE, COLUMN, AND ROW MATRICES An r-by-c matrix, M, is a rectangular array of numbers consisting of r rows and c columns: ¯MM..
    [Show full text]
  • Arxiv:2004.02090V2 [Math.NT]
    ON INDEFINITE AND POTENTIALLY UNIVERSAL QUADRATIC FORMS OVER NUMBER FIELDS FEI XU AND YANG ZHANG Abstract. A number field k admits a binary integral quadratic form which represents all integers locally but not globally if and only if the class number of k is bigger than one. In this case, there are only finitely many classes of such binary integral quadratic forms over k. A number field k admits a ternary integral quadratic form which represents all integers locally but not globally if and only if the class number of k is even. In this case, there are infinitely many classes of such ternary integral quadratic forms over k. An integral quadratic form over a number field k with more than one variables represents all integers of k over the ring of integers of a finite extension of k if and only if this quadratic form represents 1 over the ring of integers of a finite extension of k. 1. Introduction Universal integral quadratic forms over Z have been studied extensively by Dickson [7] and Ross [24] for both positive definite and indefinite cases in the 1930s. A positive definite integral quadratic form is called universal if it represents all positive integers. Ross pointed out that there is no ternary positive definite universal form over Z in [24] (see a more conceptual proof in [9, Lemma 3]). However, this result is not true over totally real number fields. In [20], Maass proved that x2 + y2 + z2 is a positive definite universal quadratic form over the ring of integers of Q(√5).
    [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]