A Vectors, Tensors, and Their Operations

Total Page:16

File Type:pdf, Size:1020Kb

A Vectors, Tensors, and Their Operations A Vectors, Tensors, and Their Operations A.1 Vectors and Tensors A spatial description of the fluid motion is a geometrical description, of which the essence is to ensure that relevant physical quantities are invariant under artificially introduced coordinate systems. This is realized by tensor analysis (cf. Aris 1962). Here we introduce the concept of tensors in an informal way, through some important examples in fluid mechanics. A.1.1 Scalars and Vectors Scalars and vectors are geometric entities independent of the choice of coor- dinate systems. This independence is a necessary condition for an entity to represent some physical quantity. A scalar, say the fluid pressure p or den- sity ρ, obviously has such independence. For a vector, say the fluid velocity u, although its three components (u1,u2,u3) depend on the chosen coordi- nates, say Cartesian coordinates with unit basis vectors (e1, e2, e3), as a single geometric entity the one-form of ei, u = u1e1 + u2e2 + u3e3 = uiei,i=1, 2, 3, (A.1) has to be independent of the basis vectors. Note that Einstein’s convention has been used in the last expression of (A.1): unless stated otherwise, a repeated index always implies summation over the dimension of the space. The inner (scalar) and cross (vector) products of two vectors are familiar. If θ is the angle between the directions of a and b, these operations give a · b = |a||b| cos θ, |a × b| = |a||b| sin θ. While the inner product is a projection operation, the cross-product produces a vector perpendicular to both a and b with magnitude equal to the area of the parallelogram spanned by a and b.Thusa×b determines a vectorial area 694 A Vectors, Tensors, and Their Operations with unit vector n normal to the (a, b) plane, whose direction follows from a to b by the right-hand rule. In particular, the inner and cross-products of Cartesian basis vectors satisfy e · e = δ i j ij , (A.2a,b) ei × ej = ek, i,j,k =1, 2, 3 and cycles where 1if i = j, δ = ij 0if i = j is the Kroneker symbol. The gradient operator ∂ ∇ = ei (A.3) ∂xi is also a vector, which as a single entity is invariant under any coordinate transformations. Thus, the pressure gradient ∇p is a vector; the divergence and curl of velocity, ∇·u = ϑ, ∇×u = ω, are the dilatation scalar and vorticity vector, respectively. A.1.2 Tensors Scalar and vectors can be considered as special tensors of rank (or order) zero and one respectively. In general, as the immediate extension of vectors, in an n-dimensional Euclidean space atensorT of rank m is a geometric entity independent of the choice of coordinate systems, and has nm components with respect to a given coordinate system. When n = 3, like the one-form (A.1) for m a vector, these 3 components Tij...k constitute the coefficients of an m-form of the given base vectors, i.e., T = Tij...keiej ...ek, i,j,...,k =1, 2, 3, and obey certain transformation rule to keep the invariance of T.1 The neces- sity of introducing tensors of ranks higher than one can be illustrated by the following simple example. Assume the fluid velocity at a spatial point x is u, and consider the velocity change at any neighboring point x +dx, see Fig. A.1. To the first-order of dx, there is du =(dx ·∇)u =dxju,j , (A.4a) where dxj (j =1, 2, 3) are Cartesian coefficients of dx. Here and throughout the book we have used a simple notation (·),j to indicate the derivative with 1 In many books a tensor is defined by requiring their components to obey this transformation rule, which is the inverse of that of the basis vector, to ensure the invariance of the tensor. A.1Vectors and Tensors 695 du u(x) x u(xϩdx) dx xϩdx Fig. A.1. A schematic interpretation of (A.4a) respect to xj; sometimes we also denote ∂/∂xi by ∂i.Sou has three directional derivatives ∂u u,j ≡ ,j=1, 2, 3, ∂xj each being a vector. Thus we may further expand them as u,j = ekuk,j,j,k=1, 2, 3. (A.4b) Then, just as p,j is the jth component of pressure gradient vector ∇p,the vector u,j is the jth component of a geometric entity, velocity gradient, defined as ∇u = eju,j = ejekuk,j,j,k=1, 2, 3, (A.5) where summation is to be taken twice, once for j and once for k. This is a two-form of ei (i =1, 2, 3), an example of tensors of rank 2, can also be directly obtained from (A.1) and (A.3). Note that in (A.5) the index j implies the components of ∇ and goes first, while the index k implies the components of u and goes after j. The order inverse in uk,j is only apparent, because this component of ∇u is merely an abbreviation of ∂uk/∂xj. Since ∇u is independent of the magnitude and direction of dx, in comparison with (A.4a) it is a more general description of the velocity variation at the neighborhood of a point. In physical and numerical experiments one always uses certain coordinate systems, and the direct output is always components of tensors. Using com- ponent operations is also convenient in deriving formulas, and sometimes only one typical component is sufficient. For example, for the velocity gradient we may just write down a representative component uk,j. But writing the final equations in coordinate-independent tensorial form enables seeing the physical objectivity of these equations clearly. Any tensor of rank 2 can be expressed by a square matrix, but not vise versa. Like any square matrix, the deformation tensor, or generally any tensor of rank 2, can always be decomposed into a symmetric part and an anti symmetric part, each being a tensor of rank 2 ∇u = D + Ω, (A.6a) 696 A Vectors, Tensors, and Their Operations where 1 D ≡ (u + u )=D (A.6b) ij 2 j,i i,j ji is called the strain-rate tensor, having six independent components, and 1 Ω ≡ (u − u )=−Ω (A.6c) ij 2 j,i i,j ij is the spin tensor or vorticity tensor, having three independent components. As illustrated by (A.6a), the addition (as well as differentiation and inte- gration) of tensors is the same as that of vectors. The multiplication of tensors, in general, results in a tensor of higher rank. A new operation of tensors is contraction by summing a pair of components with the same index (“dummy index”). This reduces the rank by two. The simplest example is the inner product of vectors: a · b = aibi or ∇·u = ϑ. Similarly, for two tensors of rank 2 there is Aij Bjk = Cik. A.1.3 Unit Tensor and Permutation Tensor The most fundamental symmetric tensor of rank 2 is the unit tensor I ≡ eiejδij (A.7) of which the components in any Cartesian coordinates are invariably the Kro- necker delta. The contraction of two unit tensors gives δij δjk = δik or I · I = I. Making contraction again (denoted by double dots) yields I : I = δii = n (dimension of the space). Clearly, the contraction of two tensors of rank 2 is the same as the multipli- cation of two matrices. With a few exceptions in Sect. 3.2, all tensors we need in this book but one will be of ranks no more than 2. Like (A.2b), the cross-product of vectors in three-dimensional space can be expressed by (a × b)i = ei · (ejaj × ekbk)=ajbkei · (ej × ek)=ijkajbk, where ijk are the Cartesian components of a tensor of rank 3 E ≡ eiejekijk. (A.8a) A.1Vectors and Tensors 697 To find these components, we use the familiar rule of vector product e · e e · e e · e δ δ δ i 1 i 2 i 3 i1 i2 i3 ≡ · × · · · ijk ei (ej ek)= ej e1 ej e2 ej e3 = δj1 δj2 δj3 ek · e1 ek · e2 ek · e3 δk1 δk2 δk3 (A.8b) 1if(ijk) = (123), (231), (312), = −1if(ijk) = (132), (213), (321), 0. otherwise. Therefore, these components in any Cartesian coordinates are invariably the permutation symbol ijk, completely anti symmetric under exchange of any pair of its three indices. For this reason, tensor E is called permutation tensor. It is used frequently whenever there is a cross-product. If more than one cross- products appear, one need the multiplication of two permutation tensors. In a three-dimensional space, from (A.8b) there is δil δim δin ijklmn = δjl δjm δjn . (A.9a) δkl δkm δkn Contraction with respect to k,nyields ijklmk = δilδjm − δimδjl. (A.9b) Continuing the contraction with respect to j, m then gives ijkljk =(3− 1)δil =2δil, (A.9c) and continuing again ijkijk = 3(3 − 1) = 6. (A.9d) One often needs to handle two-dimensional flow. If a and b are vectors in theflowplanesuchthata × b is along the third direction e3 normal to the plane, then in ijkajbk one of the i, j, k must be 3 due to (A.8b), with the other two varying between 1 and 2. In this case (A.9b) yields δjm δjn 3jk3lm = , j,k,m,n=1, 2. (A.10a) δkm δkn Contracting with respect to k,nyields 3jk3mk =(2− 1)δjm = δjm, (A.10b) and continuing again, 3jk3jk = 2(2 − 1) = 2. (A.10c) 698 A Vectors, Tensors, and Their Operations The particular importance of the permutation tensor in vorticity and vortex dynamics lies in the fact that ωi = ijkuk,j = ijkΩjk, (A.11a) due to (A.3) and (A.6).
Recommended publications
  • Vectors and Beyond: Geometric Algebra and Its Philosophical
    dialectica Vol. 63, N° 4 (2010), pp. 381–395 DOI: 10.1111/j.1746-8361.2009.01214.x Vectors and Beyond: Geometric Algebra and its Philosophical Significancedltc_1214 381..396 Peter Simons† In mathematics, like in everything else, it is the Darwinian struggle for life of ideas that leads to the survival of the concepts which we actually use, often believed to have come to us fully armed with goodness from some mysterious Platonic repository of truths. Simon Altmann 1. Introduction The purpose of this paper is to draw the attention of philosophers and others interested in the applicability of mathematics to a quiet revolution that is taking place around the theory of vectors and their application. It is not that new math- ematics is being invented – on the contrary, the basic concepts are over a century old – but rather that this old theory, having languished for many decades as a quaint backwater, is being rediscovered and properly applied for the first time. The philosophical importance of this quiet revolution is not that new applications for old mathematics are being found. That presumably happens much of the time. Rather it is that this new range of applications affords us a novel insight into the reasons why vectors and their mathematical kin find application at all in the real world. Indirectly, it tells us something general but interesting about the nature of the spatiotemporal world we all inhabit, and that is of philosophical significance. Quite what this significance amounts to is not yet clear. I should stress that nothing here is original: the history is quite accessible from several sources, and the mathematics is commonplace to those who know it.
    [Show full text]
  • Recent Books on Vector Analysis. 463
    1921.] RECENT BOOKS ON VECTOR ANALYSIS. 463 RECENT BOOKS ON VECTOR ANALYSIS. Einführung in die Vehtoranalysis, mit Anwendungen auf die mathematische Physik. By R. Gans. 4th edition. Leip­ zig and Berlin, Teubner, 1921. 118 pp. Elements of Vector Algebra. By L. Silberstein. New York, Longmans, Green, and Company, 1919. 42 pp. Vehtor analysis. By C. Runge. Vol. I. Die Vektoranalysis des dreidimensionalen Raumes. Leipzig, Hirzel, 1919. viii + 195 pp. Précis de Calcul géométrique. By R. Leveugle. Preface by H. Fehr. Paris, Gauthier-Villars, 1920. lvi + 400 pp. The first of these books is for the most part a reprint of the third edition, which has been reviewed in this BULLETIN (vol. 21 (1915), p. 360). Some small condensation of results has been accomplished by making deductions by purely vector methods. The author still clings to the rather common idea that a vector is a certain triple on a certain set of coordinate axes, so that his development is rather that of a shorthand than of a study of expressions which are not dependent upon axes. The second book is a quite brief exposition of the bare fundamentals of vector algebra, the notation being that of Heaviside, which Silberstein has used before. It is designed primarily to satisfy the needs of those studying geometric optics. However the author goes so far as to mention a linear vector operator and the notion of dyad. The exposition is very simple and could be followed by high school students who had had trigonometry. The third of the books mentioned is an elementary exposition of vectors from the standpoint of Grassmann.
    [Show full text]
  • The Vorticity Equation in a Rotating Stratified Fluid
    Chapter 7 The Vorticity Equation in a Rotating Stratified Fluid The vorticity equation for a rotating, stratified, viscous fluid » The vorticity equation in one form or another and its interpretation provide a key to understanding a wide range of atmospheric and oceanic flows. » The full Navier-Stokes' equation in a rotating frame is Du 1 +∧fu =−∇pg − k +ν∇2 u Dt ρ T where p is the total pressure and f = fk. » We allow for a spatial variation of f for applications to flow on a beta plane. Du 1 +∧fu =−∇pg − k +ν∇2 u Dt ρ T 1 2 Now uu=(u⋅∇ ∇2 ) +ω ∧ u ∂u 1 2 1 2 +∇2 ufuku +bω +g ∧=- ∇pgT − +ν∇ ∂ρt di take the curl D 1 2 afafafωω+=ffufu +⋅∇−+∇⋅+∇ρ∧∇+ ωpT ν∇ ω Dt ρ2 Dω or =−uf ⋅∇ +.... Dt Note that ∧ [ ω + f] ∧ u] = u ⋅ (ω + f) + (ω + f) ⋅ u - (ω + f) ⋅ u, and ⋅ [ω + f] ≡ 0. Terminology ωa = ω + f is called the absolute vorticity - it is the vorticity derived in an a inertial frame ω is called the relative vorticity, and f is called the planetary-, or background vorticity Recall that solid body rotation corresponds with a vorticity 2Ω. Interpretation D 1 2 afafafωω+=ffufu +⋅∇−+∇⋅+∇ρ∧∇+ ωpT ν∇ ω Dt ρ2 Dω is the rate-of-change of the relative vorticity Dt −⋅∇uf: If f varies spatially (i.e., with latitude) there will be a change in ω as fluid parcels are advected to regions of different f. Note that it is really ω + f whose total rate-of-change is determined.
    [Show full text]
  • Geometric-Algebra Adaptive Filters Wilder B
    1 Geometric-Algebra Adaptive Filters Wilder B. Lopes∗, Member, IEEE, Cassio G. Lopesy, Senior Member, IEEE Abstract—This paper presents a new class of adaptive filters, namely Geometric-Algebra Adaptive Filters (GAAFs). They are Faces generated by formulating the underlying minimization problem (a deterministic cost function) from the perspective of Geometric Algebra (GA), a comprehensive mathematical language well- Edges suited for the description of geometric transformations. Also, (directed lines) differently from standard adaptive-filtering theory, Geometric Calculus (the extension of GA to differential calculus) allows Fig. 1. A polyhedron (3-dimensional polytope) can be completely described for applying the same derivation techniques regardless of the by the geometric multiplication of its edges (oriented lines, vectors), which type (subalgebra) of the data, i.e., real, complex numbers, generate the faces and hypersurfaces (in the case of a general n-dimensional quaternions, etc. Relying on those characteristics (among others), polytope). a deterministic quadratic cost function is posed, from which the GAAFs are devised, providing a generalization of regular adaptive filters to subalgebras of GA. From the obtained update rule, it is shown how to recover the following least-mean squares perform calculus with hypercomplex quantities, i.e., elements (LMS) adaptive filter variants: real-entries LMS, complex LMS, that generalize complex numbers for higher dimensions [2]– and quaternions LMS. Mean-square analysis and simulations in [10]. a system identification scenario are provided, showing very good agreement for different levels of measurement noise. GA-based AFs were first introduced in [11], [12], where they were successfully employed to estimate the geometric Index Terms—Adaptive filtering, geometric algebra, quater- transformation (rotation and translation) that aligns a pair of nions.
    [Show full text]
  • Josiah Willard Gibbs
    GENERAL ARTICLE Josiah Willard Gibbs V Kumaran The foundations of classical thermodynamics, as taught in V Kumaran is a professor textbooks today, were laid down in nearly complete form by of chemical engineering at the Indian Institute of Josiah Willard Gibbs more than a century ago. This article Science, Bangalore. His presentsaportraitofGibbs,aquietandmodestmanwhowas research interests include responsible for some of the most important advances in the statistical mechanics and history of science. fluid mechanics. Thermodynamics, the science of the interconversion of heat and work, originated from the necessity of designing efficient engines in the late 18th and early 19th centuries. Engines are machines that convert heat energy obtained by combustion of coal, wood or other types of fuel into useful work for running trains, ships, etc. The efficiency of an engine is determined by the amount of useful work obtained for a given amount of heat input. There are two laws related to the efficiency of an engine. The first law of thermodynamics states that heat and work are inter-convertible, and it is not possible to obtain more work than the amount of heat input into the machine. The formulation of this law can be traced back to the work of Leibniz, Dalton, Joule, Clausius, and a host of other scientists in the late 17th and early 18th century. The more subtle second law of thermodynamics states that it is not possible to convert all heat into work; all engines have to ‘waste’ some of the heat input by transferring it to a heat sink. The second law also established the minimum amount of heat that has to be wasted based on the absolute temperatures of the heat source and the heat sink.
    [Show full text]
  • Scaling the Vorticity Equation
    Vorticity Equation in Isobaric Coordinates To obtain a version of the vorticity equation in pressure coordinates, we follow the same procedure as we used to obtain the z-coordinate version: ∂ ∂ [y-component momentum equation] − [x-component momentum equation] ∂x ∂y Using the p-coordinate form of the momentum equations, this is: ∂ ⎡∂v ∂v ∂v ∂v ∂Φ ⎤ ∂ ⎡∂u ∂u ∂u ∂u ∂Φ ⎤ ⎢ + u + v +ω + fu = − ⎥ − ⎢ + u + v +ω − fv = − ⎥ ∂x ⎣ ∂t ∂x ∂y ∂p ∂y ⎦ ∂y ⎣ ∂t ∂x ∂y ∂p ∂x ⎦ which yields: d ⎛ ∂u ∂v ⎞ ⎛ ∂ω ∂u ∂ω ∂v ⎞ vorticity equation in ()()ζ p + f = − ζ p + f ⎜ + ⎟ − ⎜ − ⎟ dt ⎝ ∂x ∂y ⎠ p ⎝ ∂y ∂p ∂x ∂p ⎠ isobaric coordinates Scaling The Vorticity Equation Starting with the z-coordinate form of the vorticity equation, we begin by expanding the total derivative and retaining only nonzero terms: ∂ζ ∂ζ ∂ζ ∂ζ ∂f ⎛ ∂u ∂v ⎞ ⎛ ∂w ∂v ∂w ∂u ⎞ 1 ⎛ ∂p ∂ρ ∂p ∂ρ ⎞ + u + v + w + v = − ζ + f ⎜ + ⎟ − ⎜ − ⎟ + ⎜ − ⎟ ()⎜ ⎟ ⎜ ⎟ 2 ⎜ ⎟ ∂t ∂x ∂y ∂z ∂y ⎝ ∂x ∂y ⎠ ⎝ ∂x ∂z ∂y ∂z ⎠ ρ ⎝ ∂y ∂x ∂x ∂y ⎠ Next, we will evaluate the order of magnitude of each of these terms. Our goal is to simplify the equation by retaining only those terms that are important for large-scale midlatitude weather systems. 1 Scaling Quantities U = horizontal velocity scale W = vertical velocity scale L = length scale H = depth scale δP = horizontal pressure fluctuation ρ = mean density δρ/ρ = fractional density fluctuation T = time scale (advective) = L/U f0 = Coriolis parameter β = “beta” parameter Values of Scaling Quantities (midlatitude large-scale motions) U 10 m s-1 W 10-2 m s-1 L 106 m H 104 m δP (horizontal) 103 Pa ρ 1 kg m-3 δρ/ρ 10-2 T 105 s -4 -1 f0 10 s β 10-11 m-1 s-1 2 ∂ζ ∂ζ ∂ζ ∂ζ ∂f ⎛ ∂u ∂v ⎞ ⎛ ∂w ∂v ∂w ∂u ⎞ 1 ⎛ ∂p ∂ρ ∂p ∂ρ ⎞ + u + v + w + v = − ζ + f ⎜ + ⎟ − ⎜ − ⎟ + ⎜ − ⎟ ()⎜ ⎟ ⎜ ⎟ 2 ⎜ ⎟ ∂t ∂x ∂y ∂z ∂y ⎝ ∂x ∂y ⎠ ⎝ ∂x ∂z ∂y ∂z ⎠ ρ ⎝ ∂y ∂x ∂x ∂y ⎠ ∂ζ ∂ζ ∂ζ U 2 , u , v ~ ~ 10−10 s−2 ∂t ∂x ∂y L2 ∂ζ WU These inequalities appear because w ~ ~ 10−11 s−2 the two terms may partially offset ∂z HL one another.
    [Show full text]
  • ROTATION: a Review of Useful Theorems Involving Proper Orthogonal Matrices Referenced to Three- Dimensional Physical Space
    Unlimited Release Printed May 9, 2002 ROTATION: A review of useful theorems involving proper orthogonal matrices referenced to three- dimensional physical space. Rebecca M. Brannon† and coauthors to be determined †Computational Physics and Mechanics T Sandia National Laboratories Albuquerque, NM 87185-0820 Abstract Useful and/or little-known theorems involving33× proper orthogonal matrices are reviewed. Orthogonal matrices appear in the transformation of tensor compo- nents from one orthogonal basis to another. The distinction between an orthogonal direction cosine matrix and a rotation operation is discussed. Among the theorems and techniques presented are (1) various ways to characterize a rotation including proper orthogonal tensors, dyadics, Euler angles, axis/angle representations, series expansions, and quaternions; (2) the Euler-Rodrigues formula for converting axis and angle to a rotation tensor; (3) the distinction between rotations and reflections, along with implications for “handedness” of coordinate systems; (4) non-commu- tivity of sequential rotations, (5) eigenvalues and eigenvectors of a rotation; (6) the polar decomposition theorem for expressing a general deformation as a se- quence of shape and volume changes in combination with pure rotations; (7) mix- ing rotations in Eulerian hydrocodes or interpolating rotations in discrete field approximations; (8) Rates of rotation and the difference between spin and vortici- ty, (9) Random rotations for simulating crystal distributions; (10) The principle of material frame indifference (PMFI); and (11) a tensor-analysis presentation of classical rigid body mechanics, including direct notation expressions for momen- tum and energy and the extremely compact direct notation formulation of Euler’s equations (i.e., Newton’s law for rigid bodies).
    [Show full text]
  • Spin in Relativistic Quantum Theory∗
    Spin in relativistic quantum theory∗ W. N. Polyzou, Department of Physics and Astronomy, The University of Iowa, Iowa City, IA 52242 W. Gl¨ockle, Institut f¨urtheoretische Physik II, Ruhr-Universit¨at Bochum, D-44780 Bochum, Germany H. Wita la M. Smoluchowski Institute of Physics, Jagiellonian University, PL-30059 Krak´ow,Poland today Abstract We discuss the role of spin in Poincar´einvariant formulations of quan- tum mechanics. 1 Introduction In this paper we discuss the role of spin in relativistic few-body models. The new feature with spin in relativistic quantum mechanics is that sequences of rotationless Lorentz transformations that map rest frames to rest frames can generate rotations. To get a well-defined spin observable one needs to define it in one preferred frame and then use specific Lorentz transformations to relate the spin in the preferred frame to any other frame. Different choices of the preferred frame and the specific Lorentz transformations lead to an infinite number of possible observables that have all of the properties of a spin. This paper provides a general discussion of the relation between the spin of a system and the spin of its elementary constituents in relativistic few-body systems. In section 2 we discuss the Poincar´egroup, which is the group relating inertial frames in special relativity. Unitary representations of the Poincar´e group preserve quantum probabilities in all inertial frames, and define the rel- ativistic dynamics. In section 3 we construct a large set of abstract operators out of the Poincar´egenerators, and determine their commutation relations and ∗During the preparation of this paper Walter Gl¨ockle passed away.
    [Show full text]
  • Infeld–Van Der Waerden Wave Functions for Gravitons and Photons
    Vol. 38 (2007) ACTA PHYSICA POLONICA B No 8 INFELD–VAN DER WAERDEN WAVE FUNCTIONS FOR GRAVITONS AND PHOTONS J.G. Cardoso Department of Mathematics, Centre for Technological Sciences-UDESC Joinville 89223-100, Santa Catarina, Brazil [email protected] (Received March 20, 2007) A concise description of the curvature structures borne by the Infeld- van der Waerden γε-formalisms is provided. The derivation of the wave equations that control the propagation of gravitons and geometric photons in generally relativistic space-times is then carried out explicitly. PACS numbers: 04.20.Gr 1. Introduction The most striking physical feature of the Infeld-van der Waerden γε-formalisms [1] is related to the possible occurrence of geometric wave functions for photons in the curvature structures of generally relativistic space-times [2]. It appears that the presence of intrinsically geometric elec- tromagnetic fields is ultimately bound up with the imposition of a single condition upon the metric spinors for the γ-formalism, which bears invari- ance under the action of the generalized Weyl gauge group [1–5]. In the case of a spacetime which admits nowhere-vanishing Weyl spinor fields, back- ground photons eventually interact with underlying gravitons. Nevertheless, the corresponding coupling configurations turn out to be in both formalisms exclusively borne by the wave equations that control the electromagnetic propagation. Indeed, the same graviton-photon interaction prescriptions had been obtained before in connection with a presentation of the wave equations for the ε-formalism [6, 7]. As regards such earlier works, the main procedure seems to have been based upon the implementation of a set of differential operators whose definitions take up suitably contracted two- spinor covariant commutators.
    [Show full text]
  • 1 RELATIVE VORTICITY EQUATION Newton's Law in a Rotating Frame in Z-Coordinate
    RELATIVE VORTICITY EQUATION Newton’s law in a rotating frame in z-coordinate (frictionless): ∂U + U ⋅∇U = −2Ω × U − ∇Φ − α∇p ∂t ∂U ⎛ U ⋅ U⎞ + ∇⎜ ⎟ + (∇ × U) × U = −2Ω × U − ∇Φ − α∇p ∂t ⎝ 2 ⎠ Applying ∇ × to both sides, and noting ω ≡ ∇ × U and using identities (the underlying tilde indicates vector): 1 A ⋅∇A = ∇(A ⋅ A) + (∇ × A) × A 2 ∇ ⋅(∇ × A) = 0 ∇ × ∇γ = 0 ∇ × (γ A) = ∇γ × A + γ ∇ × A ∇ × (F × G) = F(∇ ⋅G) − G(∇ ⋅ F) + (G ⋅∇)F − (F ⋅∇)G So, ∂ω ⎡ ⎛ U ⋅ U⎞ ⎤ + ∇ × ⎢∇⎜ ⎟ ⎥ + ∇ × (ω × U) = −∇ × (2Ω × U) − ∇ × ∇Φ − ∇ × (α∇p) ∂t ⎣ ⎝ 2 ⎠ ⎦ ⇓ ∂ω + ∇ × (ω × U) = −∇ × (2Ω × U) − ∇α × ∇p − α∇ × ∇p ∂t Using S to denote baroclinicity vector, S = −∇α × ∇p , then, ∂ω + ω(∇ ⋅ U) − U(∇ ⋅ω) + (U ⋅∇)ω − (ω ⋅∇)U = −2Ω(∇ ⋅ U) + U∇ ⋅(2Ω) − (U ⋅∇)(2Ω) + (2Ω ⋅∇)U + S ∂t ∂ω + ω(∇ ⋅ U) + (U ⋅∇)ω − (ω ⋅∇)U = −2Ω(∇ ⋅ U) − (U ⋅∇)(2Ω) + (2Ω ⋅∇)U + S ∂t A little rearrangement: ∂ω = − (U ⋅∇)(ω + 2Ω) − (ω + 2Ω)(∇ ⋅ U) + [(ω + 2Ω)⋅∇]U + S ∂t tendency advection convergence twisting baroclinicity RELATIVE VORTICITY EQUATION (Vertical Component) Take k ⋅ on the 3-D vorticity equation: ∂ζ = −k ⋅(U ⋅∇)(ω + 2Ω) − k ⋅(ω + 2Ω)(∇ ⋅ U) + k ⋅[(ω + 2Ω)⋅∇]U + k ⋅S ∂t 1 ∇p In isobaric coordinates, k = , so k ⋅S = 0 . Working through all the dot product, you ∇p should get vorticity equation for the vertical component as shown in (A7.1). The same vorticity equations can be easily derived by applying the curl to the horizontal momentum equations (in p-coordinates, for example). ∂ ⎡∂u ∂u ∂u ∂u uv tanφ uω f 'ω ∂Φ ⎤ − ⎢ + u + v + ω = + + + fv − ⎥ ∂y
    [Show full text]
  • Spacetime Algebra As a Powerful Tool for Electromagnetism
    Spacetime algebra as a powerful tool for electromagnetism Justin Dressela,b, Konstantin Y. Bliokhb,c, Franco Norib,d aDepartment of Electrical and Computer Engineering, University of California, Riverside, CA 92521, USA bCenter for Emergent Matter Science (CEMS), RIKEN, Wako-shi, Saitama, 351-0198, Japan cInterdisciplinary Theoretical Science Research Group (iTHES), RIKEN, Wako-shi, Saitama, 351-0198, Japan dPhysics Department, University of Michigan, Ann Arbor, MI 48109-1040, USA Abstract We present a comprehensive introduction to spacetime algebra that emphasizes its prac- ticality and power as a tool for the study of electromagnetism. We carefully develop this natural (Clifford) algebra of the Minkowski spacetime geometry, with a particular focus on its intrinsic (and often overlooked) complex structure. Notably, the scalar imaginary that appears throughout the electromagnetic theory properly corresponds to the unit 4-volume of spacetime itself, and thus has physical meaning. The electric and magnetic fields are combined into a single complex and frame-independent bivector field, which generalizes the Riemann-Silberstein complex vector that has recently resurfaced in stud- ies of the single photon wavefunction. The complex structure of spacetime also underpins the emergence of electromagnetic waves, circular polarizations, the normal variables for canonical quantization, the distinction between electric and magnetic charge, complex spinor representations of Lorentz transformations, and the dual (electric-magnetic field exchange) symmetry that produces helicity conservation in vacuum fields. This latter symmetry manifests as an arbitrary global phase of the complex field, motivating the use of a complex vector potential, along with an associated transverse and gauge-invariant bivector potential, as well as complex (bivector and scalar) Hertz potentials.
    [Show full text]
  • Geometric Intuition
    thesis Geometric intuition Richard Feynman once made a statement represent the consequences of rotations in to the effect that the history of mathematics One day, perhaps, 3-space faithfully. is largely the history of improvements in Clifford’s algebra will It is also completely natural not only to add notation — the progressive invention of or subtract multi-vectors, but to multiply or ever more efficient means for describing be taught routinely to divide them — something not possible with logical relationships and making them students in place of ordinary vectors. The result is always another easier to grasp and manipulate. The Romans multi-vector. In the particular case of a were stymied in their efforts to advance vector analysis. multi-vector that is an ordinary vector V, the mathematics by the clumsiness of Roman inverse turns out to be V/v2, where v2 is the numerals for arithmetic calculations. After to j, that is, eiej = −ejei. Another way to put squared magnitude of V. It’s a vector in the Euclid, geometry stagnated for nearly it is that multiplication between parallel same direction but of reciprocal magnitude. 2,000 years until Descartes invented a new vectors is commutative, whereas it is anti- Write out the components for the notation with his coordinates, which made commutative for orthogonal vectors. product of two vectors U and V, and you it easy to represent points and lines in These rules are enough to define find the resultUV = U•V + ĬU × V, with • space algebraically. the algebra, and it’s then easy to work and × being the usual dot and cross product Feynman himself, of course, introduced out various implications.
    [Show full text]