Lagrangian Mechanics

Total Page:16

File Type:pdf, Size:1020Kb

Lagrangian Mechanics Chapter 1 Lagrangian Mechanics Our introduction to Quantum Mechanics will be based on its correspondence to Classical Mechanics. For this purpose we will review the relevant concepts of Classical Mechanics. An important concept is that the equations of motion of Classical Mechanics can be based on a variational principle, namely, that along a path describing classical motion the action integral assumes a minimal value (Hamiltonian Principle of Least Action). 1.1 Basics of Variational Calculus The derivation of the Principle of Least Action requires the tools of the calculus of variation which we will provide now. Definition: A functional S[ ] is a map M S[]: R ; = ~q(t); ~q :[t0; t1] R R ; ~q(t) differentiable (1.1) F! F f ⊂ ! g from a space of vector-valued functions ~q(t) onto the real numbers. ~q(t) is called the trajec- tory of a systemF of M degrees of freedom described by the configurational coordinates ~q(t) = (q1(t); q2(t); : : : qM (t)). In case of N classical particles holds M = 3N, i.e., there are 3N configurational coordinates, namely, the position coordinates of the particles in any kind of coordianate system, often in the Cartesian coordinate system. It is important to note at the outset that for the description of a d classical system it will be necessary to provide information ~q(t) as well as dt ~q(t). The latter is the velocity vector of the system. Definition: A functional S[ ] is differentiable, if for any ~q(t) and δ~q(t) where 2 F 2 F d = δ~q(t); δ~q(t) ; δ~q(t) < , δ~q(t) < , t; t [t0; t1] R (1.2) F f 2 F j j jdt j 8 2 ⊂ g a functional δS[ ; ] exists with the properties · · (i) S[~q(t) + δ~q(t)] = S[~q(t)] + δS[~q(t); δ~q(t)] + O(2) (ii) δS[~q(t); δ~q(t)] is linear in δ~q(t). (1.3) δS[ ; ] is called the differential of S[ ]. The linearity property above implies · · δS[~q(t); α1 δ~q1(t) + α2 δ~q2(t)] = α1 δS[~q(t); δ~q1(t)] + α2 δS[~q(t); δ~q2(t)] : (1.4) 1 2 Lagrangian Mechanics Note: δ~q(t) describes small variations around the trajectory ~q(t), i.e. ~q(t) + δ~q(t) is a `slightly' different trajectory than ~q(t). We will later often assume that only variations of a trajectory ~q(t) are permitted for which δ~q(t0) = 0 and δ~q(t1) = 0 holds, i.e., at the ends of the time interval of the trajectories the variations vanish. It is also important to appreciate that δS[ ; ] in conventional differential calculus does not corre- spond to a differentiated function, but rather· · to a differential of the function which is simply the df differentiated function multiplied by the differential increment of the variable, e.g., df = dx dx or, M @f in case of a function of M variables, df = dxj. Pj=1 @xj We will now consider a particular class of functionals S[ ] which are expressed through an integral d over the the interval [t0; t1] where the integrand is a function L(~q(t); dt ~q(t); t) of the configuration d vector ~q(t), the velocity vector dt ~q(t) and time t. We focus on such functionals because they play a central role in the so-called action integrals of Classical Mechanics. d _ In the following we will often use the notation for velocities and other time derivatives dt ~q(t) = ~q(t) dxj and dt =x _ j. Theorem: Let t1 S[~q(t)] = Z dt L(~q(t); ~q_(t); t) (1.5) t0 where L( ; ; ) is a function differentiable in its three arguments. It holds · · · t1 t M M 1 8 @L d @L 9 @L δS[~q(t); δ~q(t)] = Z dt δqj(t) + δqj(t) : (1.6) <X @q − dt @q_ = X @q_ t0 j=1 j j j=1 j t0 : ; For a proof we can use conventional differential calculus since the functional (1.6) is expressed in terms of `normal' functions. We attempt to evaluate t1 S[~q(t) + δ~q(t)] = Z dt L(~q(t) + δ~q(t); ~q_(t) + δ~q_(t); t) (1.7) t0 through Taylor expansion and identification of terms linear in δqj(t), equating these terms with δS[~q(t); δ~q(t)]. For this purpose we consider M _ _ _ @L @L 2 L(~q(t) + δ~q(t); ~q(t) + δ~q(t); t) = L(~q(t); ~q(t); t) + δqj + δq_j + O( ) (1.8) X @q @q_ j=1 j j d _ We note then using dt f(t)g(t) = f(t)g(t) + f(t)_g(t) @L d @L d @L δq_j = δqj δqj : (1.9) @q_j dt @q_j − dt @q_j This yields for S[~q(t) + δ~q(t)] t1 M t1 M @L d @L d @L 2 S[~q(t)] + Z dt δqj + Z dt δqj + O( ) (1.10) X @q − dt @q_ X dt @q_ t0 j=1 j j t0 j=1 j From this follows (1.6) immediately. 1.1: Variational Calculus 3 We now consider the question for which functions the functionals of the type (1.5) assume extreme values. For this purpose we define Definition: An extremal of a differentiable functional S[ ] is a function qe(t) with the property δS[~qe(t); δ~q(t)] = 0 for all δ~q(t) . (1.11) 2 F The extremals ~qe(t) can be identified through a condition which provides a suitable differential equation for this purpose. This condition is stated in the following theorem. Theorem: Euler{Lagrange Condition For the functional defined through (1.5), it holds in case δ~q(t0) = δ~q(t1) = 0 that ~qe(t) is an extremal, if and only if it satisfies the conditions (j = 1; 2;:::;M) d @L @L = 0 (1.12) dt @q_j − @qj The proof of this theorem is based on the property Lemma: If for a continuous function f(t) f :[t0; t1] R R (1.13) ⊂ ! holds t1 Z dt f(t)h(t) = 0 (1.14) t0 for any continuous function h(t) with h(t0) = h(t1) = 0, then 2 F f(t) 0 on [t0; t1]. (1.15) ≡ We will not provide a proof for this Lemma. The proof of the above theorem starts from (1.6) which reads in the present case t M 1 8 @L d @L 9 δS[~q(t); δ~q(t)] = Z dt δqj(t) : (1.16) <X @q − dt @q_ = t0 j=1 j j : ; This property holds for any δqj with δ~q(t) . According to the Lemma above follows then (1.12) 2 F for j = 1; 2;:::M. On the other side, from (1.12) for j = 1; 2;:::M and δqj(t0) = δqj(t1) = 0 follows according to (1.16) the property δS[~qe(t); ] 0 and, hence, the above theorem. · ≡ An Example As an application of the above rules of the variational calculus we like to prove the well-known result 2 that a straight line in R is the shortest connection (geodesics) between two points (x1; y1) and (x2; y2). Let us assume that the two points are connected by the path y(x); y(x1) = y1; y(x2) = y2. The length of such path can be determined starting from the fact that the incremental length ds in going from point (x; y(x)) to (x + dx; y(x + dx)) is dy dy ds = r(dx)2 + ( dx)2 = dxr1 + ( )2 : (1.17) dx dx 4 Lagrangian Mechanics The total path length is then given by the integral x1 dy s = Z dx r1 + ( )2 : (1.18) x0 dx s is a functional of y(x) of the type (1.5) with L(y(x); dy ) = 1 + (dy=dx)2. The shortest path dx p is an extremal of s[y(x)] which must, according to the theorems above, obey the Euler{Lagrange dy condition. Using y0 = dx the condition reads d @L d y = 0 ! = 0 : (1.19) 2 dx @y0 dx 1 + (y ) p 0 From this follows y = 1 + (y )2 = const and, hence, y = const. This in turn yields y(x) = 0 p 0 0 ax + b. The constants a and b are readily identified through the conditons y(x1) = y1 and y(x2) = y2. One obtains y1 y2 y(x) = − (x x2) + y2 : (1.20) x1 x2 − − Exercise 1.1.1: Show that the shortest path between two points on a sphere are great circles, i.e., circles whose centers lie at the center of the sphere. 1.2 Lagrangian Mechanics The results of variational calculus derived above allow us now to formulate the Hamiltonian Prin- ciple of Least Action of Classical Mechanics and study its equivalence to the Newtonian equations of motion. Threorem: Hamiltonian Principle of Least Action The trajectories ~q(t) of systems of particles described through the Newtonian equations of motion d @U (mjq_j) + = 0 ; j = 1; 2;:::M (1.21) dt @qj are extremals of the functional, the so-called action integral, t1 S[~q(t)] = Z dt L(~q(t); ~q_(t); t) (1.22) t0 where L(~q(t); ~q_(t); t) is the so-called Lagrangian M 1 _ 2 L(~q(t); ~q(t); t) = mjq_ U(q1; q2; : : : ; qM ) : (1.23) X 2 j − j=1 Presently we consider only velocity{independent potentials.
Recommended publications
  • Branched Hamiltonians and Supersymmetry
    Branched Hamiltonians and Supersymmetry Thomas Curtright, University of Miami Wigner 111 seminar, 12 November 2013 Some examples of branched Hamiltonians are explored, as recently advo- cated by Shapere and Wilczek. These are actually cases of switchback poten- tials, albeit in momentum space, as previously analyzed for quasi-Hamiltonian dynamical systems in a classical context. A basic model, with a pair of Hamiltonian branches related by supersymmetry, is considered as an inter- esting illustration, and as stimulation. “It is quite possible ... we may discover that in nature the relation of past and future is so intimate ... that no simple representation of a present may exist.” – R P Feynman Based on work with Cosmas Zachos, Argonne National Laboratory Introduction to the problem In quantum mechanics H = p2 + V (x) (1) is neither more nor less difficult than H = x2 + V (p) (2) by reason of x, p duality, i.e. the Fourier transform: ψ (x) φ (p) ⎫ ⎧ x ⎪ ⎪ +i∂/∂p ⎪ ⇐⇒ ⎪ ⎬⎪ ⎨⎪ i∂/∂x p − ⎪ ⎪ ⎪ ⎪ ⎭⎪ ⎩⎪ This equivalence of (1) and (2) is manifest in the QMPS formalism, as initiated by Wigner (1932), 1 2ipy/ f (x, p)= dy x + y ρ x y e− π | | − 1 = dk p + k ρ p k e2ixk/ π | | − where x and p are on an equal footing, and where even more general H (x, p) can be considered. See CZ to follow, and other talks at this conference. Or even better, in addition to the excellent books cited at the conclusion of Professor Schleich’s talk yesterday morning, please see our new book on the subject ... Even in classical Hamiltonian mechanics, (1) and (2) are equivalent under a classical canonical transformation on phase space: (x, p) (p, x) ⇐⇒ − But upon transitioning to Lagrangian mechanics, the equivalence between the two theories becomes obscure.
    [Show full text]
  • Chapter 5 ANGULAR MOMENTUM and ROTATIONS
    Chapter 5 ANGULAR MOMENTUM AND ROTATIONS In classical mechanics the total angular momentum L~ of an isolated system about any …xed point is conserved. The existence of a conserved vector L~ associated with such a system is itself a consequence of the fact that the associated Hamiltonian (or Lagrangian) is invariant under rotations, i.e., if the coordinates and momenta of the entire system are rotated “rigidly” about some point, the energy of the system is unchanged and, more importantly, is the same function of the dynamical variables as it was before the rotation. Such a circumstance would not apply, e.g., to a system lying in an externally imposed gravitational …eld pointing in some speci…c direction. Thus, the invariance of an isolated system under rotations ultimately arises from the fact that, in the absence of external …elds of this sort, space is isotropic; it behaves the same way in all directions. Not surprisingly, therefore, in quantum mechanics the individual Cartesian com- ponents Li of the total angular momentum operator L~ of an isolated system are also constants of the motion. The di¤erent components of L~ are not, however, compatible quantum observables. Indeed, as we will see the operators representing the components of angular momentum along di¤erent directions do not generally commute with one an- other. Thus, the vector operator L~ is not, strictly speaking, an observable, since it does not have a complete basis of eigenstates (which would have to be simultaneous eigenstates of all of its non-commuting components). This lack of commutivity often seems, at …rst encounter, as somewhat of a nuisance but, in fact, it intimately re‡ects the underlying structure of the three dimensional space in which we are immersed, and has its source in the fact that rotations in three dimensions about di¤erent axes do not commute with one another.
    [Show full text]
  • Rotational Invariance in Critical Planar Lattice Models
    Rotational invariance in critical planar lattice models Hugo Duminil-Copin∗y, Karol Kajetan Kozlowski,§ Dmitry Krachun,y Ioan Manolescu,z Mendes Oulamara∗ December 25, 2020 Abstract In this paper, we prove that the large scale properties of a number of two- dimensional lattice models are rotationally invariant. More precisely, we prove that the random-cluster model on the square lattice with cluster-weight 1 ≤ q ≤ 4 ex- hibits rotational invariance at large scales. This covers the case of Bernoulli percola- tion on the square lattice as an important example. We deduce from this result that the correlations of the Potts models with q 2 f2; 3; 4g colors and of the six-vertex height function with ∆ 2 [−1; −1=2] are rotationally invariant at large scales. Contents 1 Introduction2 1.1 Motivation....................................2 1.2 Definition of the random-cluster model and distance between percolation configurations..................................3 1.3 Main results for the random-cluster model..................5 1.4 Applications to other models.........................7 2 Proof Roadmap9 2.1 Random-cluster model on isoradial rectangular graphs............9 2.2 Universality among isoradial rectangular graphs and a first version of the coupling...................................... 11 2.3 The homotopy topology and the second and third versions of the coupling. 13 2.4 Harvesting integrability on the torus..................... 16 ∗Institut des Hautes Études Scientifiques and Université Paris-Saclay yUniversité de Genève §ENS Lyon zUniversity of Fribourg 1 3 Preliminaries 19 3.1 Definition of the random-cluster model.................... 19 3.2 Elementary properties of the random-cluster model............. 20 3.3 Uniform bounds on crossing probabilities..................
    [Show full text]
  • Rotational Invariance of Two-Level Group Codes Over Dihedral and Dicyclic Groups
    Sddhangl, Vol. 23, Part 1, February 1998, pp. 45-56. © Printed in India. Rotational invariance of two-level group codes over dihedral and dicyclic groups JYOTI BALI ~ and B SUNDAR RAJAN 2 i Department of Electrical Engineering, Indian Institute of Technology, Hauz Khas, New Delhi 110016, India 2Department of Electrical Communication Engineering, Indian Institute of Science, Bangalore 560 012, India e-mail: [email protected]; [email protected] Abstract. Phase-rotational invariance properties for two-level constructed, (using a binary code and a code over a residue class integer ring as component codes) Euclidean space codes (signal sets) in two and four dimensions are discussed. The label codes are group codes over dihedral and dicyclic groups respectively. A set of necessary and sufficient conditions on the component codes is obtained for the resulting signal sets to be rotationally invariant to several phase angles. Keywords. Multilevel codes; group codes; dihedral groups; coded modulation. 1. Introduction It is well known (Viterbi & Omura 1979; Bendetto et al 1987) that digital communication over Additive White Gaussian Noise (AWGN) channel can be modelled as transmission of a point from a finite set of points, called signal set, of a finite dimensional vector space and the Maximum Likelihood soft decoding then becomes choosing the closest point in the signal set, in the sense of Euclidean distance, from the received point in the space. The probability of error performance to a large extent is dominated by the minimum of the pairwise distances of the signal points. The problem of signal set design for AWGN channel then is choosing a specified number of points in a space of specified dimensions in such a way that the minimum distance is the maximum possible.
    [Show full text]
  • The Lagrangian and Hamiltonian Mechanical Systems
    THE LAGRANGIAN AND HAMILTONIAN MECHANICAL SYSTEMS ALEXANDER TOLISH Abstract. Newton's Laws of Motion, which equate forces with the time- rates of change of momenta, are a convenient way to describe mechanical systems in Euclidean spaces with cartesian coordinates. Unfortunately, the physical world is rarely so cooperative|physicists often explore systems that are neither Euclidean nor cartesian. Different mechanical formalisms, like the Lagrangian and Hamiltonian systems, may be more effective at describing such phenomena, as they are geometric rather than analytic processes. In this paper, I shall construct Lagrangian and Hamiltonian mechanics, prove their equivalence to Newtonian mechanics, and provide examples of both non- Newtonian systems in action. Contents 1. The Calculus of Variations 1 2. Manifold Geometry 3 3. Lagrangian Mechanics 4 4. Two Electric Pendula 4 5. Differential Forms and Symplectic Geometry 6 6. Hamiltonian Mechanics 7 7. The Double Planar Pendulum 9 Acknowledgments 11 References 11 1. The Calculus of Variations Lagrangian mechanics applies physics not only to particles, but to the trajectories of particles. We must therefore study how curves behave under small disturbances or variations. Definition 1.1. Let V be a Banach space. A curve is a continuous map : [t0; t1] ! V: A variation on the curve is some function h of t that creates a new curve +h.A functional is a function from the space of curves to the real numbers. p Example 1.2. Φ( ) = R t1 1 +x _ 2dt, wherex _ = d , is a functional. It expresses t0 dt the length of curve between t0 and t1. Definition 1.3.
    [Show full text]
  • Lagrangian Mechanics
    Alain J. Brizard Saint Michael's College Lagrangian Mechanics 1Principle of Least Action The con¯guration of a mechanical system evolving in an n-dimensional space, with co- ordinates x =(x1;x2; :::; xn), may be described in terms of generalized coordinates q = (q1;q2;:::; qk)inak-dimensional con¯guration space, with k<n. The Principle of Least Action (also known as Hamilton's principle) is expressed in terms of a function L(q; q_ ; t)knownastheLagrangian,whichappears in the action integral Z tf A[q]= L(q; q_ ; t) dt; (1) ti where the action integral is a functional of the vector function q(t), which provides a path from the initial point qi = q(ti)tothe¯nal point qf = q(tf). The variational principle ¯ " à !# ¯ Z d ¯ tf @L d @L ¯ j 0=±A[q]= A[q + ²±q]¯ = ±q j ¡ j dt; d² ²=0 ti @q dt @q_ where the variation ±q is assumed to vanish at the integration boundaries (±qi =0=±qf), yields the Euler-Lagrange equation for the generalized coordinate qj (j =1; :::; k) à ! d @L @L = ; (2) dt @q_j @qj The Lagrangian also satis¯esthesecond Euler equation à ! d @L @L L ¡ q_j = ; (3) dt @q_j @t and thus for time-independent Lagrangian systems (@L=@t =0)we¯nd that L¡q_j @L=@q_j is a conserved quantity whose interpretationwill be discussed shortly. The form of the Lagrangian function L(r; r_; t)isdictated by our requirement that Newton's Second Law m Är = ¡rU(r;t)describing the motion of a particle of mass m in a nonuniform (possibly time-dependent) potential U(r;t)bewritten in the Euler-Lagrange form (2).
    [Show full text]
  • Lagrangian Mechanics - Wikipedia, the Free Encyclopedia Page 1 of 11
    Lagrangian mechanics - Wikipedia, the free encyclopedia Page 1 of 11 Lagrangian mechanics From Wikipedia, the free encyclopedia Lagrangian mechanics is a re-formulation of classical mechanics that combines Classical mechanics conservation of momentum with conservation of energy. It was introduced by the French mathematician Joseph-Louis Lagrange in 1788. Newton's Second Law In Lagrangian mechanics, the trajectory of a system of particles is derived by solving History of classical mechanics · the Lagrange equations in one of two forms, either the Lagrange equations of the Timeline of classical mechanics [1] first kind , which treat constraints explicitly as extra equations, often using Branches [2][3] Lagrange multipliers; or the Lagrange equations of the second kind , which Statics · Dynamics / Kinetics · Kinematics · [1] incorporate the constraints directly by judicious choice of generalized coordinates. Applied mechanics · Celestial mechanics · [4] The fundamental lemma of the calculus of variations shows that solving the Continuum mechanics · Lagrange equations is equivalent to finding the path for which the action functional is Statistical mechanics stationary, a quantity that is the integral of the Lagrangian over time. Formulations The use of generalized coordinates may considerably simplify a system's analysis. Newtonian mechanics (Vectorial For example, consider a small frictionless bead traveling in a groove. If one is tracking the bead as a particle, calculation of the motion of the bead using Newtonian mechanics) mechanics would require solving for the time-varying constraint force required to Analytical mechanics: keep the bead in the groove. For the same problem using Lagrangian mechanics, one Lagrangian mechanics looks at the path of the groove and chooses a set of independent generalized Hamiltonian mechanics coordinates that completely characterize the possible motion of the bead.
    [Show full text]
  • Ten1perature, Least Action, and Lagrangian Mechanics L
    14 Auguq Il)<)5 2) 291 PHYSICS LETTERS A 's. Rev. ELSEVIER Physics Letters A 204 (1995) 133-135 l) 759. 7. to be Ten1perature, least action, and Lagrangian mechanics l. 843. William G. Hoover Department ofApplied Science, Ulliuersity ofCalifornia at Davis / Livermore alld Lawrence Licermore Nalional Laboratory, LiueTlrJOre, CA 94551-7808, USA Received 28 March 1995; revised manuscript received 2 June 1995; accepted for publication 2 June J995 Communicated by A.R. Bishop Abstract Temperature is considered from two different mechanical standpoints. The more approach uses Hamilton's least action principle. From it we derive thermostatting forces identical to those found using Gauss' Principle. This approach applies to equilibrium and nonequilibrium systems. The Lagrangian approach is different, and less useful, applying only at equilibrium. Keywords: Least action principle; Nonequilibrium; Lagrangian mechanics 1. Introduction When the context is clear we use just a single q or q to represent whole sets. Feynman repeatedly emphasized the generality of Lagrangian mechanics is particularly well-suited Hamilton's principle of least action [1], the physical to the treatment of constrained systems. Constraints, law from which both kinds of mechanics, classical used first to represent mechanical linkages and cou­ and quantum, follow. The principle states that the plings, have more recently been used in molecular observed trajectory {qed} is that which minimizes simulations, where they maintain the shape of a the action integral, polyatomic molecule and reduce the number of dy­ namical equations to be integrated. The usual ap­ ofL(q,q,t)dt of(K <P)dt=O, proach is to add a "Lagrange multiplier" to the Lagrangian in such a way as to impose the constraint where K is the kinetic energy, <P is the potential without affecting energy conservation [2,3].
    [Show full text]
  • Analytical Mechanics: Lagrangian and Hamiltonian Formalism
    MATH-F-204 Analytical mechanics: Lagrangian and Hamiltonian formalism Glenn Barnich Physique Théorique et Mathématique Université libre de Bruxelles and International Solvay Institutes Campus Plaine C.P. 231, B-1050 Bruxelles, Belgique Bureau 2.O6. 217, E-mail: [email protected], Tel: 02 650 58 01. The aim of the course is an introduction to Lagrangian and Hamiltonian mechanics. The fundamental equations as well as standard applications of Newtonian mechanics are derived from variational principles. From a mathematical perspective, the course and exercises (i) constitute an application of differential and integral calculus (primitives, inverse and implicit function theorem, graphs and functions); (ii) provide examples and explicit solutions of first and second order systems of differential equations; (iii) give an implicit introduction to differential manifolds and to realiza- tions of Lie groups and algebras (change of coordinates, parametrization of a rotation, Galilean group, canonical transformations, one-parameter subgroups, infinitesimal transformations). From a physical viewpoint, a good grasp of the material of the course is indispensable for a thorough understanding of the structure of quantum mechanics, both in its operatorial and path integral formulations. Furthermore, the discussion of Noether’s theorem and Galilean invariance is done in a way that allows for a direct generalization to field theories with Poincaré or conformal invariance. 2 Contents 1 Extremisation, constraints and Lagrange multipliers 5 1.1 Unconstrained extremisation . .5 1.2 Constraints and regularity conditions . .5 1.3 Constrained extremisation . .7 2 Constrained systems and d’Alembert principle 9 2.1 Holonomic constraints . .9 2.2 d’Alembert’s theorem . 10 2.3 Non-holonomic constraints .
    [Show full text]
  • Examples in Lagrangian Mechanics C Alex R
    Lecture 18 Friday - October 14, 2005 Written or last updated: October 14, 2005 P441 – Analytical Mechanics - I Examples in Lagrangian Mechanics c Alex R. Dzierba Sample problems using Lagrangian mechanics Here are some sample problems. I will assign similar problems for the next problem set. Example 1 In Figure 1 we show a box of mass m sliding down a ramp of mass M. The ramp moves without friction on the horizontal plane and is located by coordinate x1. The box also slides without friction on the ramp and is located by coordinate x2 with respect to the ramp. x 2 m x1 M θ Figure 1: A box slides down a ramp without friction and the ramp slides along a horizontal surface without friction. The kinetic energy of the ramp TM is given by: 1 2 TM = Mx˙ 1 (1) 2 and the kinetic energy of the box Tm is given by: 1 2 2 Tm = m x˙ 1 +x ˙ 2 + 2x ˙ 1x˙ 1 cos θ (2) 2 The velocity of the box is obviously derived from the vector sum of the velocity relative to the ramp and the velocity of the ramp. The potential energy of the system is just the potential energy of the box and that leads to: U = −mgx2 sin θ (3) 1 and finally the Lagrangian is given by: 1 2 1 2 2 L(x1, x˙ 1, x2, x˙ 2)= TM + Tm − U = Mx˙ 1 + m x˙ 1 +x ˙ 2 + 2x ˙ 1x˙ 1 cos θ + mgx2 sin θ (4) 2 2 The equations of motion are: d ∂L ∂L = (5) dt ∂x˙ 1 ∂x1 and d ∂L ∂L = (6) dt ∂x˙ 2 ∂x2 Equation 5 leads to: d [m(x ˙ 1 +x ˙ 2 cos θ)+ Mx˙ 1] = 0 (7) dt The RHS of equation 7 is zero because the Lagrangian does not explicitly depend on x1.
    [Show full text]
  • Simple Pendulum Via Lagrangian Mechanics by Frank Owen, 22 May 2014 © All Rights Reserved
    Alpha Omega Engineering, Inc. 872 Toro Street, San Luis Obispo, CA 93401 – (805) 541 8608 (home office), (805) 441 3995 (cell) Simple pendulum via Lagrangian mechanics by Frank Owen, 22 May 2014 © All rights reserved The equation of motion for a simple pendulum of length l, operating in a gravitational field is ̈ This equation can be obtained by applying Newton’s Second Law (N2L) to the pendulum and then writing the equilibrium equation. It is instructive to work out this equation of motion also using Lagrangian mechanics to see how the procedure is applied and that the result obtained is the same. For this example we are using the simplest of pendula, i.e. one with a massless, inertialess link and an inertialess pendulum bob at its end, as shown in Figure 1. Figure 1 – Simple pendulum Lagrangian formulation The Lagrangian function is defined as where T is the total kinetic energy and U is the total potential energy of a mechanical system. 1 To get the equations of motion, we use the Lagrangian formulation ( ) ̇ where q signifies generalized coordinates and F signifies non-conservative forces acting on the mechanical system. For the simplify pendulum, we assume no friction, so no non-conservative forces, so all Fi are 0. The aforementioned equation of motion is in terms of as a coordinate, not in terms of x and y. So we need to use kinematics to get our energy terms in terms of . For T, we need the velocity of the mass. ̇ So ( ̇) ̇ Figure 2 – Height for potential energy The potential energy, U, depends only on the y-coordinate.
    [Show full text]
  • Rotational Invariant Operators Based on Steerable Filter Banks
    Rotational Invariant Operators based on Steerable Filter Banks X. Shi A.L. Ribeiro Castro R. Manduchi R. Montgomery Apple Computer, Inc. Dept. of Mathematics Dept. of Comp. Eng. Dept. of Mathematics [email protected] UC Santa Cruz UC Santa Cruz UC Santa Cruz [email protected] [email protected] [email protected] ABSTRACT rotational invariant descriptors starting from the original N We introduce a technique for designing rotation invariant steerable filters. This requires solving a first order linear PDE. operators based on steerable filter banks. Steerable filters are A closed form solution for this case is provided in this paper. widely used in Computer Vision as local descriptors for texture II. PROBLEM STATEMENT AND SOLUTION analysis. Rotation invariance has been shown to improve A. Steerable Filters as Local Descriptors texture-based classification in certain contexts. Our approach to invariance is based on solving the PDE associated with the A mainstream approach to texture analysis is based on the formulation of invariance in a Lie group framework. representation of texture patches by low–dimensional local descriptors. A linear filter bank is normally used for this I. INTRODUCTION purpose. The vector (descriptor) formed by the outputs of Image texture analysis is normally performed by first ex- N filters at a certain pixel is a rank–N linear mapping of tracting feature vectors (descriptors), designed so as to capture the graylevel profile within a neighborhood of that pixel. The the most relevant visual information. The descriptors, and marginal or joint statistics of the descriptor are then used to the subsequent analysis algorithm, should be invariant under characterize the variability of texture appearance.
    [Show full text]