EOS 352 Continuum Dynamics Conservation of angular momentum

Similar documents
Physics 110. Electricity and Magnetism. Professor Dine. Spring, Handout: Vectors and Tensors: Everything You Need to Know

A Primer on Three Vectors

Computational Fluid Dynamics Prof. Dr. Suman Chakraborty Department of Mechanical Engineering Indian Institute of Technology, Kharagpur

Vectors a vector is a quantity that has both a magnitude (size) and a direction

PLANAR KINETIC EQUATIONS OF MOTION (Section 17.2)

Classical Mechanics in Hamiltonian Form

Chapter 3. Forces, Momentum & Stress. 3.1 Newtonian mechanics: a very brief résumé

16. Rotational Dynamics

d 1 µ 2 Θ = 0. (4.1) consider first the case of m = 0 where there is no azimuthal dependence on the angle φ.

2.20 Fall 2018 Math Review

CIRCULAR MOTION AND ROTATION

Physics 142 Energy in Mechanics Page 1. Energy in Mechanics

Lecture Notes 3

Vector & Fourier Analysis PH1202

Page 52. Lecture 3: Inner Product Spaces Dual Spaces, Dirac Notation, and Adjoints Date Revised: 2008/10/03 Date Given: 2008/10/03

Motion under the Influence of a Central Force

( ) by D n ( y) 1.1 Mathematical Considerations. π e nx Dirac s delta function (distribution) a) Dirac s delta function is defined such that

The Principle of Least Action

Vector, Matrix, and Tensor Derivatives

Chapter 0. Preliminaries. 0.1 Things you should already know

Rotational motion of a rigid body spinning around a rotational axis ˆn;

Differential Equations

Review of Vector Analysis in Cartesian Coordinates

Caltech Ph106 Fall 2001

Second quantization: where quantization and particles come from?

Tensor Analysis in Euclidean Space

A primer on matrices

Lecture 9 - Rotational Dynamics

Math 123, Week 2: Matrix Operations, Inverses

Lecture 3: 1. Lecture 3.

Curves in the configuration space Q or in the velocity phase space Ω satisfying the Euler-Lagrange (EL) equations,

MECH 5312 Solid Mechanics II. Dr. Calvin M. Stewart Department of Mechanical Engineering The University of Texas at El Paso

r CM = ir im i i m i m i v i (2) P = i

Lecture 13: Forces in the Lagrangian Approach

Navier-Stokes Equation: Principle of Conservation of Momentum

Vector calculus. Appendix A. A.1 Definitions. We shall only consider the case of three-dimensional spaces.

Symmetries 2 - Rotations in Space

STEP Support Programme. STEP 2 Matrices Topic Notes

A primer on matrices

Mathematics for Graphics and Vision

Chapter 11. Angular Momentum

Classical Mechanics. Luis Anchordoqui

Elementary Linear Algebra

Physics 141 Dynamics 1 Page 1. Dynamics 1

Chapter 3 Stress, Strain, Virtual Power and Conservation Principles

Introduction to Seismology Spring 2008

Physical Conservation and Balance Laws & Thermodynamics

Physics 111. Tuesday, November 2, Rotational Dynamics Torque Angular Momentum Rotational Kinetic Energy

Tensors, and differential forms - Lecture 2

Numerical Methods for Inverse Kinematics

1 Matrices and matrix algebra

Physics 209 Fall 2002 Notes 5 Thomas Precession

[variable] = units (or dimension) of variable.

- Marine Hydrodynamics. Lecture 14. F, M = [linear function of m ij ] [function of instantaneous U, U, Ω] not of motion history.

Dielectrics. Lecture 20: Electromagnetic Theory. Professor D. K. Ghosh, Physics Department, I.I.T., Bombay

Chapter 7. Kinematics. 7.1 Tensor fields

Introduction to tensors and dyadics

Forces of Constraint & Lagrange Multipliers

Mathematics that Every Physicist should Know: Scalar, Vector, and Tensor Fields in the Space of Real n- Dimensional Independent Variable with Metric

for changing independent variables. Most simply for a function f(x) the Legendre transformation f(x) B(s) takes the form B(s) = xs f(x) with s = df

Thermodynamics (Classical) for Biological Systems Prof. G. K. Suraishkumar Department of Biotechnology Indian Institute of Technology Madras

Rotational Kinetic Energy

CHAPTER 9. Microscopic Approach: from Boltzmann to Navier-Stokes. In the previous chapter we derived the closed Boltzmann equation:

Introduction to Tensor Notation

Introduction to Polar Coordinates in Mechanics (for AQA Mechanics 5)

1/3/2011. This course discusses the physical laws that govern atmosphere/ocean motions.

Foundations of Solid Mechanics. James R. Rice. Harvard University. January, 1998

4.1 Important Notes on Notation

particle p = m v F ext = d P = M d v cm dt

Basics of Special Relativity

Magnetostatics and the vector potential

Basic Quantum Mechanics Prof. Ajoy Ghatak Department of Physics Indian Institute of Technology, Delhi

Sometimes the domains X and Z will be the same, so this might be written:

HAMILTON S PRINCIPLE

The Matrix Representation of a Three-Dimensional Rotation Revisited

Mathematical Preliminaries

Physics 6303 Lecture 2 August 22, 2018

Contravariant and Covariant as Transforms

Rotational Kinematics

Finishing Chapter 26 on dipoles.. Electric Potential Energy of: Point Charges Dipoles Electric Potential: V Voltage: ΔV

Classical Mechanics Solutions 1.

Physics Fall Mechanics, Thermodynamics, Waves, Fluids. Lecture 20: Rotational Motion. Slide 20-1

Electromagnetic energy and momentum

Draft:S Ghorai. 1 Body and surface forces. Stress Principle

Surface force on a volume element.

Math review. Math review

= 1 and 2 1. T =, and so det A b d

Vectors, metric and the connection

A short review of continuum mechanics

,, rectilinear,, spherical,, cylindrical. (6.1)

Chapter 2 The Continuum Equations

Torque and Rotation Lecture 7

2. FLUID-FLOW EQUATIONS SPRING 2019

1 Dirac Notation for Vector Spaces

Kinematics (special case) Dynamics gravity, tension, elastic, normal, friction. Energy: kinetic, potential gravity, spring + work (friction)

ES.182A Topic 44 Notes Jeremy Orloff

Lecture 35: The Inertia Tensor

Control Volume. Dynamics and Kinematics. Basic Conservation Laws. Lecture 1: Introduction and Review 1/24/2017

Lecture 1: Introduction and Review

Physics 141 Energy 1 Page 1. Energy 1

Transcription:

EOS 352 Continuum Dynamics Conservation of angular momentum c Christian Schoof. Not to be copied, used, or revised without explicit written permission from the copyright owner The copyright owner explicitly opts out of UBC policy # 81. ermission to use this document is only granted on a case-by case basis. The document is never shared under the terms of UBC policy # 81. March 15, 2017 Overview These notes cover the following Angular momentum for point particles Cross products in subscript notation Conservation of angular momentum for continua Torques due to surface and body forces Reduction to a symmetric stress tensor Angular momentum for point particles Angular momentum for a collection of point particles is a distinct quantity from linear momentum. It is easily possible to have zero total linear momentum but a finite angular momentum. The definition of angular momentum L for a single particle of mass m, velocity u and position vector r = (x, y, z) = (x 1, x 2, x 3 ) relative to some fixed coordinate system is L = r mu = r p. where p is the ordinary linear momentum of the particle. For a collection of particles A, B, C etc, with positions r A, r B,... and linear momenta p A, p B,... total angular momentum is L tot = r A p A + r B p B +... = r p 1

x 2 u R m m R x 1 u Figure 1: Rotation of two equal masses about a their common of centre of mass. The system has zero total linear momentum but a finite angular momentum. Often, angular momentum is associated with rotations. Take two particles A and B of equal mass m connected by a string of length 2R, rotating in the x 1 x 2 -plane about their common centre of mass at (0, 0, 0). When the string is at an angle θ to the x 1 -axis, the particles will have positions and linear velocities and r A = R(cos(θ), sin(θ), 0), u A = Rω( sin(θ), cos(θ), 0), r B = R( cos(θ), sin(θ), 0), u B = Rω(sin(θ), cos(θ), 0). where ω = dθ/ is rotation rate or angular velocity. Using the definition above, their angular momenta are and r A mu A = mr 2 ω(0, 0, cos(θ) 2 + sin(θ) 2 ) = mr 2 ω(0, 0, 1), r B mu B = mr 2 ω(0, 0, cos(θ) 2 + sin(θ) 2 ) = mr 2 ω(0, 0, 1). so total angular momentum is L tot = 2mR 2 ω(0, 0, 1). 2

We see that angular momentum increases with angular velocity and the length of the string. Through its definition as a cross product, it is also perpendicular to the plane of rotation: the vector L is aligned with the axis of rotation (in this case, the z- or x 3 -axis). Lastly, we can compute the total linear momentum of the system as mu A + mu B = 0, so the system has zero total linear momentum but non-zero angular momentum. This is because the system overall has no linear motion (i.e., its centre of mass is not moving), but it does have a non-zero rotation, which is what angular momentum describes. Conservation of angular momentum for point particles Newton s second law states that, for a particle dp = F (1) where F is the total force on particle. We can think of this as the sum of forces F exerted on by other particles, dp = F. (2) Newton s third law, which really ensures conservation of momentum, requires that the forces between a particle and a particle are equal and opposite, or F = F, (3) which incidentally (with = ) ensures that F = 0, so a particle cannot exert a net force on itself. The rate of change of total momentum of the system is then d p tot = d = = p d p F = (F AB + F AC +...) + (F BA + F BC +...) + (F CA + F CB +...) +... = 0 3

as all the terms in the sum can be paired up with terms that are equal and opposite, for instance F AC + F CA = 0 etc. A more mathematically elegant way of doing this would be to realize that and are dummy indices in the sum, so we must have F = F QQ Q Q swapping the index for Q and for Q. But equally, we should have F QQ = F Q Q swapping Q for and Q for. Hence, noting also that the order of summation does not matter, F = F = F, that is, we have just swapped the dummy indices and. But also F = F as F = F. But the last two equations taken together imply that F = F, which is possible only if F = 0. It follows from the above that the total linear momentum p tot cannot change over time and is therefore conserved. Total angular momentum is also a conserved quantity. The reason for this is that forces between these particles are generally parallel to the line joining them. To understand this, consider the equivalent of Newton s second law for angular momentum. For a single particle, have dl = d(r mu) = dr d(mu) mu + r But r is the position vector of the particle, so the rate of change of r is the particle velocity u, u = dr. Hence dl d(mu) = u mu + r. But the cross product of u with itself vanishes while d(mu)/ = F is the force on the particle, so dl = r F. (4) 4

F AC r -r A B r A r -r A C r B F CA r -r B C r C Figure 2: Multiple particles A, B, C, their position vectors and the vectors joining them. If the force F AC exerted by C on A is not aligned with the vector r A r C joining them (as suggested by the image), then neither is the equal and opposite force F CA exerted by C on A, the two forces generate a net torque. As shown in this figure, they would cause an anticlockwise rotation and thereby change the angular momentum of the system. However, angular momentum is conserved if the forces one particle exerts on another are aligned with the vector joining them. 5

The quantity on the right-hand side of this equation is usually known as the torque acting on the particle, denoted by T = r F. Now, for a particle in a collection of particles, we can write dl = r F = r F = r F The rate of change of total angular momentum is therefore dl tot = dl = r F r F where the right-hand side is the sum of torques on the particles. But, by the same process of swapping the dummy indices and as above, we can show that r F = r F = r F because the forces F and F are equal and opposite. Now, using this, we can write ( ) r F = 1 r F + r F 2 ( ) = 1 r F r F 2 = 1 (r r ) F. 2 But r r is the vector that joins particle to particle. If the force F is parallel to the direction from to, it follows that r r and F are parallel and hence their cross product is zero. it follows that 1 2 (r r ) F = 0 6

and therefore that dl tot = 0 so that angular momentum is conserved. All the basic forces in nature (gravitational, electrostatic,... ) are such that forces act along the line connecting two particles, and consequently angular momentum is conserved. Cross products and angular momentum in subscript notation Angular momentum is defined as a cross product of two vectors. Taking a cross product is not a straightforward to write in subscript notation as most other vector operations. Specifically, if r = (x 1, x 2, x 3 ) and u = (u 1, u 2, u 3 ), then L = (L 1, L 2, L 3 ) = r mu is of the form L 1 = m(x 2 u 3 x 3 u 2 ) L 2 = m(x 3, u 1 x 1 u 3 ) L 3 = m(x 1 u 2 x 2 u 1 ), (5) and there does not appear to be a simple formula for the vector component L i. The reason why cross products are difficult to write in succinct subscript notation is that they actually are slightly unusual constructs in vector algebra: they require a right-hand rule to define their direction: in a left-handed world, the would point in the opposite direction. Nature of course does not really discriminate between leftand right-handed rotations in this way, and in particular, does not care that the x 1 -, x 2 - and x 3 -axes by convention form a right-handed coordinate system. Crossproducts are sometimes called pseudovectors because they change direction in going from left-handed to right-handed coordinate systems. In subscript notation, it is actually much easier to write them as so-called antisymmetric tensors with two indices. Suppose that we were to define a tensor L ij through L ij = m(x i u j x j u i ). (6) Then the components of L ij could be written out explicitly in the form L 11 = L 22 = L 33 = 0, L 12 = m(x 1 u 2 x 2 u 1 ) = L 21, L 13 = m(x 1 u 3 x 3 u 1 ) = L 31, or in matrix form L = L 23 = m(x 2 u 3 x 3 u 2 ) = L 32, (7) 0 m(x 1 u 2 x 2 u 1 ) m(x 1 u 3 x 3 u 1 ) m(x 1 u 2 x 2 u 1 ) 0 m(x 2 u 3 x 3 u 2 ) m(x 1 u 3 x 3 u 1 ) m(x 2 u 3 x 3 u 2 0 7

It should be clear that the matrix L has precisely three independent entries. Further, comparing (7) with (5), it should be clear that, if we know the components (L 1, L 2, L 3 ) of the vector L, then we immediately know all the components L ij of the matrix L, as we have L 1 = L 23 = L 32, L 2 = L 13 = L 31, L 3 = L 12 = L 21, (8) and L 11 = L 22 = L 33 = 0. In other words, there is no information contained in L that is not also contained in L or vice versa: if we know one of these objects, we can reconstruct the other. So instead of using the cross product L = r mu as the definition of angular momentum, we can equally use the tensor L ij. Similarly, we can define torque as an antisymmetric tensor T ij = x i F j x j F i, (9) and just as we did above in (8), we can show that the cross product vector T = (T 1, T 2, T 3 ) relates to the matrix T through T 1 = T 23 = T 32, T 2 = T 13 = T 31, T 3 = T 12 = T 21, (10) But (4) in component form reads dl i / = T i. With (8) and (10), this becomes dl ij = T ij. (11) or d(m(x i u j x j u i )) = x i F j x j F i. Our next task is to re-write this for a continuum. Conservation of angular momentum for a continuum In the tensor notation above, the angular momentum contained in a small volume δv around position (x 1, x 2, x 3 ) and moving with velocity (u 1, u 2, u 3 ) is δl ij = δm(x i u j x j u i ) = ρ(x i u j x j u i )δv and the total angular momentum in a Lagrangian volume V (t) can therefore be found by summing L ij = ρ(x i u j x j u i ) dv V (t) The fundamental assumption made in continuum mechanics in general is that torques on the volume arise entirely because of surface and body forces. Recall that the surface force on a surface element δs can be written as δf i = σ ik n k δs, 8

where we have chosen k as the dummy index to avoid confusion with the fixed indices i and j that appear above. The associated torque is therefore δt ij = x i δf j x j δf i = (x i σ jk n k x j σ ik n k )δs. Similarly, the net force on a volume element δv due to a body force f i is with associated torque δf i = f i δv δt ij = x i δf j x j δf i = (x i f j x j f i )δv. The total torque can therefore be found by summing T ij = (x i σ jk x j σ ik )n k ds + (x i f j x j f i ) dv S(t) V (t) Substituting these into (11) gives d ρ(x i u j x j u i ) dv = (x i σ jk x j σ ik )n k ds + V (t) S(t) V (t) (x i f j x j f i ) dv We can now follow the same procedure as before to turn this into a differential equation, using Reynolds transport theorem and the divergence theorem. The only difference is that the conserved quantity L ij and hence the associated density ρ(x i u j x j u i ) now have two indices i and j, rather than one index i for momentum p i and momentum density ρu i. All this requires is somewhat more careful handling of the repeated indices that appear in the dot products in Reynolds transport theorem and the divergence theorem these must now not be i or j, so we will use k below. Applying Reynolds transport theorem to the time derivative, V (t) t [ρ(x iu j x j u i )] dv + [ρ(x i u j x j u i )] u k n k ds = S(t) (x i σ jk x j σ ik )n k ds + (x i f j x j f i ) dv S(t) Next, using the divergence theorem, [ ] t [ρ(x iu j x j u i )]+ [ρ(x i u j x j u i )u k ] [x i σ jk x j σ ik ] (x i f j x j f i ) dv = 0 V (t) And the usual argument about V (t) being arbitrary allows us to conclude that the integrand must be zero, t [ρ(x iu j x j u i )] + [ρ(x i u j x j u i )u k ] 9 V (t) [x i σ jk x j σ ik ] (x i f j x j f i ) = 0. (12)

Reduction to a symmetric stress tensor Equation (12) looks exceedingly complicated, but we can in fact reduce it to something very simply by using the momentum conservation equation (ρu i ) t + (ρu iu k ) = σ ik + f i (13) All we need to do is apply the product rule to the derivatives in (12) and recognize a few facts about differentiating the terms x i and x j that appear in some of the derivatives. We will apply the product rule to each term in (12) in turn. Take the first term, t [ρ(x iu j x j u i )] = (ρu jx i ) (ρu ix j ) t t = (ρu j) x i x i + ρu j t t (ρu i) t x j ρu i x j t But now you have to remember what the partial derivative /t stands for: differentiating with respect to t while holding x 1, x 2 and x 3 constant. It follows that x i t = x j t = 0 as this amounts to differentiating constants. Hence t [ρ(x iu j x j u i )] = (ρu j) x i (ρu i) x j (14) t t Next, take the second term in (12), [ρ(x i u j u k x j u i u k )] = (ρu ju k x i ) (ρu iu k x j ) = (ρu ju k ) x i x i + ρu j u k (ρu iu k ) x j ρu i u k x j (15) Again we have to take a closer look at the partial derivatives x i / and x j /. The partial derivative with respect to x k is a derivative taken with all the other x i s (i.e., with i k) held constant, as well as with t held constant. The derivative of x i with respect to x k is therefore zero if i k, and one if i = k. To see this, we can write the derivatives out explicitly for all the possible combinations of i and k, = 1, = 0, = 0, = 1, 10 x 3 = 0, x 3 = 0,

x 3 = 0, x 3 = 0, In matrix form, this becomes the identity matrix x 3 x 3 x 3 x 3 x 3 = x 3 x 3 = 1. 1 0 0 0 1 0 0 0 1 To facilitate writing this in subscript notation, we define a special tensor known as the Kronecker delta by { 1 if i = j, δ ij = (16) 0 otherwise, so that x i = δ ik, x j = δ jk (17) Now consider a sum over a repeated index involving the Kronecker delta, say δ ij a j = 3 δ ij a j = δ i1 a 1 + δ i2 a 2 + δ i3 a 3. j=1 Only one of the three terms on the right can be non-zero, because the Kronecker delta is zero except if both its indices are the same. Take i = 1. Then δ i1 a 1 = a 1 but δ i2 a 2 = δ i3 a 3 = 0, and δ ij a j = a 1. Similarly, if i = 2, we get δ ij a j = a 2, and if i = 3, we get δ ij a j = a 3, This can be summarized by saying δ ij a j = a i. We can do the same thing for tensors with more indices. For instance, δ ij σ jk = σ ik by the same argument as above: δ ij σ jk = 3 j=1 δ ijσ jk, and the only term in the sum that is not zero is that for which j = i. Exercise 1 Recall that a repeated index indicates summation. In standard matrix notation, how would you write a i = δ ij a j? How would you write σ ik = δ ij σ ij? With these properties of the Kronecker delta in hand, (15) becomes [ρ(x i u j u k x j u i u k )] = (ρu ju k ) x i + ρu j u k δ ik (ρu iu k ) x j ρu i u k δ jk = (ρu ju k ) x i + ρu j u i (ρu iu k ) x j ρu i u j = (ρu ju k ) x i (ρu iu k ) x j (18) as ρu j u i ρu i u j = 0. 11

Similarly, we can deal with the third term in (12): [x i σ jk x j σ ik ] = σ jk x i + σ jk x i σ ik x j σ ik x j Substituting (14), (18) and (19) in (12), we get = σ jk x i + σ jk δ ik σ ik x j = σ ik δ jk = σ jk x i + σ ji σ ik x j σ ij (19) (ρu j ) x i (ρu i) x j + (ρu ju k ) x i (ρu iu k ) x j σ jk x i σ ji + σ ik x j +σ ij (x i f j x j f i ) = 0 t t (20) Now this does not look any simpler than the original, but recall that we wanted to use the momentum conservation equation (13). Hence we group terms we can recognize from (13) together. This allows us to rewrite (20) as [ (ρuj ) + (ρu ju k ) σ jk f j ] [ (ρui ) x i + (ρu iu k ) σ ik ] f i x j σ ji +σ ij = 0 t t (21) But if we replace all the i s by j s in (13), we can recognize that the first expression in square brackets is zero. Similarly, the second expression in square brackets is zero, and we are left rather prosaically with σ ji + σ ij = 0 as an alternative form of (12). In other words, conservation of angular momentum boils down to the stress tensor σ ij being symmetric, σ ij = σ ji. 12