Physics 106a, Caltech 16 October, Lecture 5: Hamilton s Principle with Constraints. Examples

Similar documents
Forces of Constraint & Lagrange Multipliers

Constrained motion and generalized coordinates

Physics 106a, Caltech 4 December, Lecture 18: Examples on Rigid Body Dynamics. Rotating rectangle. Heavy symmetric top

Physics 351, Spring 2015, Homework #5. Due at start of class, Friday, February 20, 2015 Course info is at positron.hep.upenn.

Lecture 4. Alexey Boyarsky. October 6, 2015

Lesson 8. Luis Anchordoqui. Physics 168. Thursday, October 11, 18

Chapter 10: Dynamics of Rotational Motion

Solution Set Two. 1 Problem #1: Projectile Motion Cartesian Coordinates Polar Coordinates... 3

Assignments VIII and IX, PHYS 301 (Classical Mechanics) Spring 2014 Due 3/21/14 at start of class

Physics 201 Lecture 16

ESM 3124 Intermediate Dynamics 2012, HW6 Solutions. (1 + f (x) 2 ) We can first write the constraint y = f(x) in the form of a constraint

Lecture 27: Generalized Coordinates and Lagrange s Equations of Motion

Classical Mechanics Review (Louisiana State University Qualifier Exam)

Problem Set III Solutions

Chapter 4. Oscillatory Motion. 4.1 The Important Stuff Simple Harmonic Motion

Projectile Motion and 2-D Dynamics

Physics 351 Wednesday, February 14, 2018

Physics 6010, Fall 2016 Constraints and Lagrange Multipliers. Relevant Sections in Text:

Physics 131: Lecture 21. Today s Agenda

8.012 Physics I: Classical Mechanics Fall 2008

Physics 131: Lecture 21. Today s Agenda

STEP Support Programme. Mechanics STEP Questions

28. Rolling Sphere. Introduction. Holonomic Constraints and non-holonomic Constraints. x origin, and its radius a. We see immediately that its

Nonholonomic Constraints Examples

Lagrangian Dynamics: Derivations of Lagrange s Equations

Concept Question: Normal Force

Euler-Lagrange's equations in several variables

Lecture 10 Mechanical Energy Conservation; Power

Marion and Thornton. Tyler Shendruk October 1, Hamilton s Principle - Lagrangian and Hamiltonian dynamics.

8.012 Physics I: Classical Mechanics Fall 2008

Lagrangian Dynamics: Generalized Coordinates and Forces

Lagrangian and Hamiltonian Mechanics (Symon Chapter Nine)

Problem Goldstein 2-12

Part A Atwood Machines Please try this link:

8.012 Physics I: Classical Mechanics Fall 2008

Lecture 13: Forces in the Lagrangian Approach

CP1 REVISION LECTURE 3 INTRODUCTION TO CLASSICAL MECHANICS. Prof. N. Harnew University of Oxford TT 2017

End-of-Chapter Exercises

Chapter 10 Notes: Lagrangian Mechanics

Advanced Dynamics. - Lecture 4 Lagrange Equations. Paolo Tiso Spring Semester 2017 ETH Zürich

This module requires you to read a textbook such as Fowles and Cassiday on material relevant to the following topics.

8.012 Physics I: Classical Mechanics Fall 2008

MASSACHUSETTS INSTITUTE OF TECHNOLOGY Physics Department 8.01 Physics I Fall Term 2009 Review Module on Solving N equations in N unknowns

Physics 351, Spring 2017, Homework #12. Due at start of class, Friday, April 14, 2017

PHYSICS 221, FALL 2011 EXAM #2 SOLUTIONS WEDNESDAY, NOVEMBER 2, 2011

Classical Mechanics and Electrodynamics

PHYSICS 221, FALL 2009 EXAM #1 SOLUTIONS WEDNESDAY, SEPTEMBER 30, 2009

Rotation review packet. Name:

LAGRANGIAN AND HAMILTONIAN

Classical Mechanics and Electrodynamics

Assignment 9. to roll without slipping, how large must F be? Ans: F = R d mgsinθ.

Lecture 10. Example: Friction and Motion

Phys 1401: General Physics I

PHYSICS 221, FALL 2010 EXAM #1 Solutions WEDNESDAY, SEPTEMBER 29, 2010

PHYS 705: Classical Mechanics. Examples: Lagrange Equations and Constraints

Chapter 10. Rotation

Forces Part 1: Newton s Laws

Description: Using conservation of energy, find the final velocity of a "yo yo" as it unwinds under the influence of gravity.

Analytical Dynamics: Lagrange s Equation and its Application A Brief Introduction

Physics for Scientists and Engineers 4th Edition, 2017

28. Pendulum phase portrait Draw the phase portrait for the pendulum (supported by an inextensible rod)

Applying Newton s Laws

Exam Question 6/8 (HL/OL): Circular and Simple Harmonic Motion. February 1, Applied Mathematics: Lecture 7. Brendan Williamson.

Question 1: A particle starts at rest and moves along a cycloid whose equation is. 2ay y a

31 ROTATIONAL KINEMATICS

Newton s Laws and Free-Body Diagrams General Physics I

Physics 351, Spring 2015, Final Exam.

Rotational motion problems

Normal Force. W = mg cos(θ) Normal force F N = mg cos(θ) F N

Constraints. Noninertial coordinate systems

Rotational Kinematics

Practice. Newton s 3 Laws of Motion. Recall. Forces a push or pull acting on an object; a vector quantity measured in Newtons (kg m/s²)

Mechanics II. Which of the following relations among the forces W, k, N, and F must be true?

4) Vector = and vector = What is vector = +? A) B) C) D) E)

Physics 351, Spring 2017, Homework #2. Due at start of class, Friday, January 27, 2017

1 Motion of a single particle - Linear momentum, work and energy principle

Physics 101: Lecture 08. Common Incorrect Forces (Spooky Rules!) Items below are NOT forces Acceleration: F Net = ma Centripetal Acceleration

Physics 351, Spring 2017, Homework #3. Due at start of class, Friday, February 3, 2017

14.9 Worked Examples. Example 14.2 Escape Velocity of Toro

UNIVERSITY OF SASKATCHEWAN Department of Physics and Engineering Physics

Chapters 5-6. Dynamics: Forces and Newton s Laws of Motion. Applications

Webreview Torque and Rotation Practice Test

Two Hanging Masses. ) by considering just the forces that act on it. Use Newton's 2nd law while

Phys101 Second Major-152 Zero Version Coordinator: Dr. W. Basheer Monday, March 07, 2016 Page: 1

It will be most difficult for the ant to adhere to the wheel as it revolves past which of the four points? A) I B) II C) III D) IV

Lecture 17 - Gyroscopes

Rotation. PHYS 101 Previous Exam Problems CHAPTER

Friction is always opposite to the direction of motion.

Homework 1. Due Thursday, January 21

Topic 1: Newtonian Mechanics Energy & Momentum

Physics Exam 2 October 11, 2007

Mechanics Lecture Notes

Physics 2211 M Quiz #2 Solutions Summer 2017

A. B. C. D. E. v x. ΣF x

Torque and Rotation Lecture 7

Solving high order nonholonomic systems using Gibbs-Appell method

HW 6 Mathematics 503, Mathematical Modeling, CSUF, June 24, 2007

Lecture 11 - Advanced Rotational Dynamics

Columbia University Department of Physics QUALIFYING EXAMINATION

Reading Quiz. Chapter 5. Physics 111, Concordia College

Transcription:

Physics 106a, Caltech 16 October, 2018 Lecture 5: Hamilton s Principle with Constraints We have been avoiding forces of constraint, because in many cases they are uninteresting, and the constraints can be built in to carefully chosen generalized coordinates that can be varied independently. Motion along the constraint directions are zero so we simply leave those coordinates out. If we start with M particles in 3 dimensions, or 3M degrees of freedom, but can mae use of N c constraints, we can reduce the problem to N dof = 3M N c degrees of freedom and thus N dof equations of motion. This is method 1 discussed below. But sometimes, it s not possible (or desirable) to do this: The constraints may be too complicated to allow for carefully chosen generalized coordinates. Sometimes, it s important to now the forces of constraints; for example, we may need to now the tension on a pendulum, to choose a strong enough material. Or, we need to now the forces of constraint that eep a roller coaster car on its trac, to ensure that the trac is strong enough to not brea, and to ensure that the car doesn t fly off the trac. In these cases, choose coordinates that are easy to identify, even if they don t vary independently (because of the constraints); eg, the 3M Cartesian coordinates of M bodies in 3 dimensions. To this one adds N c constraint equations, and, as we ll see, N c Lagrange multipliers. We will end up with N dof = 3M + N c equations of motion. It will thus be more complicated, but we will be able to address the two concerns above. Examples We ll illustrate the following ideas with four examples: The simple pendulum in 2-dimensions. Here, the constraint x 2 + y 2 = l reduces the problem from two cartesian coordinates (x and y) to one generalized coordinate φ satisfying the constraint. Or, if we need to now the tension on the pendulum string, we will use Lagrange multipliers. The Atwood machine (two masses at the two ends of a massless rope slung over a massless, frictionless pulley) is similarly simple to solve by choosing the right generalized coordinate, but again, if we need to now the tension on the pendulum string, we will use Lagrange multipliers. A mass sliding on a circular wire (constraint: r = R), or a mass rolling or sliding on a bowling ball (constraint: r R), or a mass being swung around on a string (constraint: r R). The bicycle wheel on an inclined plane (Hand & Finch Fig 2.6). Here, the constraints are non-holonomic; they are constraints on the velocity, not position of the wheel, so they can t be integrated without first nowing the full solution. Here, we need to use Lagrange multipliers whether we re interested in the constraint forces or not. 1

m 2 m 1 y Holonomic constraints First set the problem up in terms of 3M coordinates q that completely specify the system before constraints are introduced. (These might be the Cartesian components of the M position vectors r i or might be some other choice of coordinates.) Now introduce N c holonomic constraints G j (q 1, q 2... q 3M, t) = 0 j = 1... N c. (1) In Hamilton s principle we choose to loo at path variations that are consistent with the constraints, so that S = 3M Ldt δs = δq dt = 0 for {δq } consistent with constraints (2) δq with /δq the variational derivative δq L q d dt We cannot put /δq to zero, since the δq are not independent. There are two methods to do the constrained optimization. Method 1: Reduced number of coordinates ( ) L. (3) q To do the constrained optimization we can find some reduced number N = 3M N c of generalized coordinates { q, = 1... N} such that we can vary them independently and each variation is consistent with the constraints. These are the generalized coordinates consistent with the constraints we have used before. One choice might be a reduced set N of the original q with the other N c of the {q } varying to maintain the constraints. Varying with respect to these coordinates gives δs = N δ q δ q dt = 0 (4) and since the δ q are independent and may be chosen arbitrarily each /δ q = 0, giving the N equations of motion for the constrained problem as before L d ( ) L = 0. (5) q dt q Also, since variations of the q are consistent with the constraints, the Lagrangian can be evaluated without including the constraint forces. 2

Method 2: Lagrange multipliers Alternatively we can do the constrained optimization using the method of Lagrange multipliers. In this approach we find the stationary value of a modified action S with the constraints added using Lagrange multipliers λ j (t), which may depend on t, but not on any of the coordinates {q }. If the λ j s are time-independent, this is called static equilibrium. N c S = L + λ j (t)g j dt, (6) t i j=1 now treating all q as effectively independent. Requiring S to be stationary, δ S tf 3M = N c G j + λ j δq dt = 0; (7) δq q then gives the 3M + N c equations δq + N c j=1 j=1 λ j G j q = 0 = 1... 3M (8) S λ j = 0 G j = 0, j = 1... N c (9) for the same number 3M + N c of unnowns at each time: q (t), λ j (t). Why does this wor? Note that S = S for paths satisfying the constraints (G = 0): maing S stationary for path variations satisfying the constraints is certainly the same as maing S stationary for such variations. How about the Lagrange multiplier terms? Here s an outline of the argument, using the example of a pendulum in 2-dimensions. In Cartesian coordinates, the 2D pendulum has Lagrangian, and constraint: L = 1 2 m ( ẋ 2 + ẏ 2) mgy, and G(x, y) = x 2 + y 2 l = 0. (10) One constraint means we have one Lagrange multiplier, λ. Requiring S to be stationary, f [( ) ( ) ] δ S = δx + λ G δx + x δy + λ G δy dt = 0. (11) y i Because of the constraint, δx and δy cannot be varied independently, and thus the two terms in parentheses can t be set to zero separately. Or can they? The tric: we haven t yet specified λ. Let s choose it to be such that the first term in parentheses is equal to zero. We are left only with the term containing δy. Now we can assert that the path y can be varied independently (so that δx is constrained, but it no longer appears in the integral!). Thus the term multiplying δy (which also contains the λ that is now specified from the first term) must be equal to zero. Got it? Continuing with this example, we have: δx : δy : δx = mẍ; δy = mÿ mg; G G x = x x 2 + y 2 = x l y = y x 2 + y 2 = y l (12) (13) 3

where we have used the constraint for the last steps in the above two lines. We thus have two equations of motion: mẍ = λ x l mÿ = mg + λ y l (14) (15) Now thin about how Newton would describe this. The (massless) pendulum string is under tension T (the force along the string, pulling up) to hold up the mass against gravity. (If the string is massive, the tension can vary along the string). That constraining force has components F c x = T sin θ = T (x/l) and F c y = T cos θ = T (y/l). We thus have exactly the same equations as above, with the identification λ T. Thus, we interpret λ to be (proportional to) the constraining force. This (a) justifies the Lagrange multiplier method, and (b) gives us a method for identifying and quantifying the required constraining force. Exercise: go through this argument with the Atwood machine, derive a formula for the tension on the massless string holding up the two masses. You can refer to textboos (eg. Hand and Finch 2.7) for more discussion of the method. Hand and Finch 2.6 also discusses the application of Lagrange multipliers to the simpler problem of the constrained min/maximization of functions of variables ( static equilibrium ; λ j (t) λ j ), rather than functionals of functions read these discussions if you find the use of Lagrange multipliers unclear. Equation (8) has a geometrical interpretation that the gradient of L is in the plane formed by the normal derivatives to the constraint surfaces, and so has zero component in directions of variation consistent with the constraints. The Lagrange multipliers are related to the (generalized) components of the total constraint force F (c) = j λ j G j q (16) i.e. the second term on the left hand side of Eq. (8) is F (c). I find Hand and Finch a little confusing in the discussion of this point, so here is my version. Derive the generalized equations of motion (Golden rule #1) see Lecture 4 for the 3M coordinates q ( ) d T T = δw = F (nc) dt q q δq + F (c) (17) where we include all the forces in the virtual wor, including the constraint forces, since variations of the 3M coordinates do not necessarily give virtual displacements carefully arranged to be perpendicular to the constraint forces. The non-constraint forces derive from the external potential V that we now F (nc) = V/ q : these terms are transferred to the left hand side and form part of the conventional Lagrangian to give ( ) d L L = F (c) dt q q. (18) Now compare with Eq. (8). Nonholonomic constraints The inclusion of constraints in the Lagrange multiplier approach Eq. (8) involves only the differential form of the constraints, and so nonintegrable differential nonholonomic constraints can also be 4

implemented. For N c constraints of the form 3M g j δq = 0 equivalent to 3M g j q = 0, j = 1... N c, (19) where the coefficients g j may depend on {q l } and t, the constrained optimization is given by setting 3M N c + λ j (t)g j δq dt = 0 (20) δq j=1 where again the second term in the bracets give zero total contribution to the expression for variations consistent with the constraints. As before, with the extra freedom introduced by the λ j, the 3M values of δq can be treated as effectively independent. Examples of the use of Lagrange multipliers 1. Mass m confined to vertical circle of radius R: This might be a particle sliding on a ball in a vertical plane through the center, a mass twirling in a vertical plane on the end of a string, or a bead on a hoop, etc. Use polar coordinates (r, θ) with θ measured from the vertical. Then T = 1 2 m(r2 θ2 + ṙ 2 ) inetic energy (21) V = mg(r r cos θ) potential energy ( V = 0 at top) (22) G(r) = r R = 0 constraint (23) Find the stationary value of the effective action S including the constraint with Lagrange multiplier λ [ ] 1 S = 2 m(r2 θ2 + ṙ 2 ) + mg(r r cos θ) + λ(t)(r R) (24) taing δθ, δr as independent δ S δθ : mgr sin θ d dt (mr2 θ) = 0 (25) δ S δr : mr θ 2 mg cos θ + λ d (mṙ) = 0 dt (26) These are to be solved together with the constraint r R = 0. (27) Equation (25) is the equation of motion for θ, and Eq. (26) with the constraint ()r is constant, ṙ = 0, and r = 0) gives λ: λ = mg cos θ mr θ 2 (28) which is indeed the radial constraint force F r (c), since r θ 2 is the radial acceleration for the circular motion. (In this case G/ r = 1, and so λ is equal to the component of the constraint force. More generally the constraint force is λ multiplied by the derivative of the constraint function.) 5

This constraining force is supplied by the material properties of the hoop, bowling ball, rope, or roller coaster. At least, we hope. For a mass sliding on a ball we require F r (c) > 0 (the ball can only push) and for a mass on a string we require F r (c) < 0 (the string can only pull), so this sets limits on when the constrained solution actually matches the physical solution (ie, tells us how strong our constraining material needs to be before it fails). For a bead on a wire F r c) can have either sign, so the constrained solution is valid for all times. For a roller coaster, the engineer must ensure that the required constraining force, supplied by the trac, is sufficient to eep the cart on the trac and eep the trac from collapsing. 2. Nonholonomic constraints: vertical rotating wheel on a sloping plane: See Hand and Finch 2.8 and Fig. 2.6 for the setup, although the y coordinate should point up hill, not down. Note that the xy plane is tilted, not horizontal. I ll do the case of a wheel with all the mass at the rim; Hand and Finch loo at a dis, which changes the moment of inertia calculation. The wheel is assumed not to tilt in the motion. See the figure a few pages bac. The rolling-without-slipping constraints are δx R sin θδφ = 0 or ẋ = R sin θ φ (29) δy R cos θδφ = 0 or ẏ = R cos θ φ (30) where the second pair are called velocity constraints. Then the inetic energy is T = 1 2 m(ẋ2 + ẏ 2 ) + 1 2 mr2 φ2 + 1 4 mr2 θ2 (31) If you substitute ẋ and ẏ using the velocity constraints, as H&F does, you get: T = mr 2 φ2 + 1 4 mr2 θ2 (32) This actually wors; but in general it will not wor (see below); you should not mae such substitutions in the Lagrangian, but only in the equations of motion obtained from the Lagrangian. The potential energy is The Lagrangian is V = mgf(y) with f(y) = y sin α. (33) L = 1 2 m(ẋ2 + ẏ 2 ) + 1 2 mr2 φ2 + 1 4 mr2 θ2 mgf(y). (34) Finding the stationary point of the action for variations subject to the constraints introducing the Lagrange multipliers λ x, λ y gives δx δx + δy Collecting terms ( ) δx + λ x δx + δy + δφ + δφ δθ δθ + λ x (δx R sin θδφ) + λ y (δy R cos θδφ) = 0. (35) ( ) ( ) δy + λ y δy + δφ λ xr sin θ λ y R cos θ δφ + ( ) δθ = 0 (36) δθ 6

Now δx, δy, δφ, δθ can be treated as effectively independent so each () must be zero. This gives d dt (mẋ) + λ x = 0 (37) d dt (mẏ) mgf (y) + λ y = 0 (38) d dt (mr2 φ) λx R sin θ λ y R cos θ = 0 (39) d ( ) 1 dt 2 mr2 θ = 0 (40) These are to be solved together with the constraint equations ẋ = R sin θ φ (41) ẏ = R cos θ φ (42) Using Eqs. (37,38) to evaluate the Lagrange multipliers eliminating ẋ, ẏ with Eqs. (41,42) gives λ x = d dt (mr sin θ φ) (43) λ y = d dt (mr cos θ φ) + mgf (y) (44) When these are substituted into Eq. (39) some miraculous cancellation gives 2mR 2 φ = mgf (y)r cos θ (45) where the factor of 2 on the left hand side comes from the x, y inetic energy adding to the φ inetic energy. This is to be solved with Eq. (40) or mr 2 θ = constant (46) which is the conservation of angular momentum about the axis normal to the plane (θ is ignorable). For a uniform slope f(y) = y sin α, f (y) = sin α, so that y drops out of these two equations, and the constraint equation is not needed. For a ramp of varying slope where y is needed to evaluate the right hand side of Eq. (45), we would also have include the constraint equation ẏ = R cos θ φ (47) in the list of equations to be solved together, i.e. Eqs. (45-47). There is a subtlety in such calculations including nonintegrable differential constraints. You might be tempted, as were Hand and Finch, to use the constraint equations Eqs. (41,42) to simplify the inetic energy to T = mr 2 φ2 + 1 4 mr2 θ2 Using this, the Lagrangian would depend just on φ, θ, y and their time derivatives, and we might plan to find the stationary value of the action varying these variables subject to the single constraint Eq. (42). This procedure using the velocity constraints to express the Lagrangian in terms of fewer variables and using this to form the action is in general (48) 7

incorrect for nonholonomic constraints (it is fine for the coordinate constraints of a holonomic system). You can find a discussion of this rather subtle point in the boo A Mathematical Introduction to Robot Manipulation by Murray, Li, and Shastry, pp 274-6. This is available on Richard Murray s Caltech website and the direct lin to a pdf file of the boo is here. The development is quite mathematical, so you may find it an effort to follow. The boo is also a good indication of the importance of the Lagrangian formulation of mechanics, and of nonholonomic constraints in robotics. For reasons I am not clear about, the reduction procedure (as used by Hand and Finch) gives the correct answer for the rolling wheel problem just considered. 8