31
Part1B Advanced Physics Classical Dynamics Lecture Handout: 1 Lecturer: J. Ellis University of Cambridge Department of Physics Lent Term 2005 Version 1(b) Please send comments or corrections to Dr .J. Ellis, [email protected]

Classical Dynamics -

  • Upload
    others

  • View
    15

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Classical Dynamics -

Part1B Advanced Physics

Classical Dynamics

Lecture Handout: 1

Lecturer: J. Ellis

University of Cambridge Department of Physics

Lent Term 2005

Version 1(b) Please send comments or corrections to Dr .J. Ellis, [email protected]

Page 2: Classical Dynamics -

2

Page 3: Classical Dynamics -

i

Contents

Contents ...................................................................................................................................i Acknowledgments .................................................................................................................... iii 1 Frames of reference, Newtonian and Lagrangian Mechanics ............................................1

1.1 Review of Newtonian Mechanics...............................................................................1 1.1.1 Newtonian Mechanics is:....................................................................................1 1.1.2 The use of vectors...............................................................................................1 1.1.3 Vector basis sets and frames of reference ..........................................................2 1.1.4 Mechanics ...........................................................................................................3 1.1.5 Statics..................................................................................................................3 1.1.6 Kinematics ..........................................................................................................3 1.1.7 Dynamics ............................................................................................................3 1.1.8 Force ...................................................................................................................4 1.1.9 Mass....................................................................................................................4 1.1.10 Momentum..........................................................................................................4 1.1.11 Motion of a collection of particles: the centre of mass.......................................5 1.1.12 Motion of a collection of particles: rotations about the centre of Mass .............5 1.1.13 Torque.................................................................................................................6 1.1.14 Angular momentum............................................................................................6 1.1.15 Relation between G and J ...................................................................................6 1.1.16 Work and Energy................................................................................................7 1.1.17 Potential Energy and Conservative Forces .........................................................7

1.2 The Energy Method ....................................................................................................9 1.3 Lagrangian Mechanics..............................................................................................12

1.3.1 The need for Lagrangian mechanics.................................................................12 1.3.2 Constraints ........................................................................................................13 1.3.3 The Lagrangian mechanics ’recipe’...................................................................13 1.3.4 Examples 1: Time independent Lagrangians....................................................15 1.3.5 Derivation of the Lagrangian mechanics formalism ........................................17 1.3.6 The Hamiltonian ...............................................................................................18 1.3.7 Examples 2: Time dependent Lagrangians.......................................................19 1.3.8 Conservation of energy and the Hamiltonian ...................................................20 1.3.9 Invariance of L and conservation laws .............................................................22

1.4 Non-Inertial Frames..................................................................................................23 1.4.1 Transformations from stationary to rotating frames.........................................23 1.4.2 Derivation of fictitious forces...........................................................................23 1.4.3 Centrifugal force...............................................................................................25 1.4.4 Coriolis force on Earth’s surface ......................................................................26

Page 4: Classical Dynamics -

ii

Page 5: Classical Dynamics -

iii

Classical Dynamics

J. Ellis Frames of reference, Newtonian and Lagrangian mechanics: Revision of Newton’s laws. Frames of reference. Rotating frames, centrifugal and Coriolis forces. Energy method. Use of Lagrangian dynamics to analyse complex dynamical problems, Euler-Lagrange equations. Normal modes: Analysis of motion of many-particle systems in terms of normal modes. Degrees of freedom of a system, matrix notation, orthogonality of the eigenvectors, zero frequency and degenerate modes. Use of symmetries to find normal modes. Modes of molecules and continuous systems (standing waves). Orbits: Kepler’s laws. Effective potentials and the radial equation. Circular and elliptic orbits. Escape velocity, transfer orbits and gravitational slingshot of space probes. Parabolic and hyperbolic orbits. Reduced mass. Rigid body dynamics: Angular velocity and angular momentum as vectors, conservation of angular momentum. Moment of inertia tensor, principal axes. Gyroscopes. Books An excellent book, but very expensive so check your College library: Principles of Dynamics, D.T. Greenwood (2nd edition Prentice & Hall 1988). Also good are: Classical Mechanics, T.W.B. Kibble and F.H. Berkshire (Imperial College Press) Classical Mechanics, V.D. Barger and M.G. Olsson (McGaw-Hill 1995). (limited coverage of normal modes). Comprehensive and recommended for 3rd year courses, but above the level of this course: Classical Mechanics, H. Goldstein, C.P. Poole and J.L.Safko (3rd edition Addison Wesley 2001/3) For the relevant maths: Mathematical Methods for Physics and Engineering, Riley R.F., Hobson M.P. and Bence S.J. (CUP 1997) Acknowledgments The lecture handouts for this course have drawn on those of the previous lecturers, in particular those of Profs Webber and Mackay and Dr. S. Kenderdine.

Page 6: Classical Dynamics -

1

1 Frames of reference, Newtonian and Lagrangian Mechanics

1.1 Review of Newtonian Mechanics

1.1.1 Newtonian Mechanics is: - Non- Relativistic, i.e. 18 ms103 −=<< xcv (speed of light) - Classical, i.e. Js106.6 34−=>> xhEt (Planck’s constant) and so assumes that: - the mass of objects is independent of velocity, time or frame of reference, - measurements of length and time are independent of the frame of reference and, like measurements of mass, are made by comparison with an arbitrary standard, - all parameters can be known precisely.

1.1.2 The use of vectors - Vectors have both magnitude and direction. - If a physical quantity has both magnitude and direction and adds, resolves and behaves under coordinate transformations like a vector, it can be usefully represented by a vector. - Obvious vectors: position, velocity, acceleration, force, momentum - Not such obvious vectors: quantities that ‘happen’ to behave as vectors and so can be represented as vectors such as angular velocity, angular acceleration, torque, and angular momentum.

Page 7: Classical Dynamics -

2

Cylindrical polars x

ρy ρ

φ

φ

1.1.3 Vector basis sets and frames of reference - Vectors ‘exist’ in space and have properties that are independent of any basis set used to represent them. The same is true of functions of vectors, such as dot and cross products, and derivatives of such as div, grad and curl. - Vectors may be represented with various basis sets, e.g:

Cartesian coordinates (basis set of unit vectors independent of position:)

( )zyxr ,,= , where zzyyxxr ˆˆˆ ++= , ( )zyxv ,,=

Cylindrical polar coordinates (basis set varies with position, r ):

( )zr ,ρ= , where zzr ˆˆ+= ρρ

( ) zzzv ˆˆˆ,, ++== φφρρρφρρ

- Coordinate axes (e.g. x , y , z ) and a clock provide a frame of reference, with respect to

which every event can be labelled by its space and time coordinates ( r , t).

- A frame of reference in which Newton’ s laws hold is an inertial frame. - If a frame S (with coordinates ( r , t) ) is an inertial frame and frame S’ ( with coordinates

( ’r , t’ ) ) is in uniform motion relative to S, then S’ is also an inertial frame, because

accelerations are the same in both frames. The equations relating the coordinates of the two frames are know as the Galilean transformation: tt =’ Rrr −=’

tuRR += 0

where R is the position of the origin of the S’ frame in the S frame at time t and u is the

velocity of S’ with respect to S.

x

z y

Cartesian basis set

Page 8: Classical Dynamics -

3

The velocities and accelerations are related by: urRrr −=−= ’

⇒ m

Frr == ’ ⇒ Newton’ s Laws hold in S’ if true in S.

1.1.4 Mechanics = Statics (absence of motion)

+ Kinematics (description of motion, using vectors for position and velocity) + Dynamics (prediction of motion, and involves forces and/or energy)

1.1.5 Statics - Forces and torques balance.

1.1.6 Kinematics - motion can be described with respect to different frames of reference, and ‘kinematics’ includes a treatment of the ‘fictitious’ forces that need to be introduced to describe motion in an accelerating frame. - definition of velocity and acceleration

Velocity: t

rv

t δδ

δ 0lim

→=

Acceleration: t

va

t δδ

δ 0lim

→=

1.1.7 Dynamics - motion can be described with respect to different frames of reference, and ‘kinematics’ includes a treatment of the ‘fictitious’ forces that need to be introduced to describe motion in an accelerating frame.

vv

v

vv δ+

Page 9: Classical Dynamics -

4

- basic principle of dynamics, Newton’ s second law [N2]:

1.1.8 Force - N2 is not a definition of force – force is defined as a push or pull that changes or tends to change a body’ s state of rest or relative motion. Force can be measured by various means, the simplest being to compare it (using a balance of some kind) with the force of gravity on a set of standard weights.

1.1.9 Mass - N2 is, however, a definition of inertial mass, which was shown by experiment (summarised in Newton’ s Law of Gravitation) to be proportional to gravitational mass.

1.1.10 Momentum - Momentum is defined as the product of mass and velocity:

vmp =

Fdt

pdp == (N2 again)

- If a force acts on a particle from time it to ft , then the time integral of the force (the

impulse) is equal to the change in momentum:

∫=−f

i

t

tif

dttFpp )(

applied force mass of body

Fam =

Page 10: Classical Dynamics -

5

- The impulse of a force can be extremely useful in problems such as collisions where the force varies rapidly in an unknown way, but all that is of interest is the ‘net effect’ of the force.

1.1.11 Motion of a collection of particles: the centre of mass

- The centre of mass R of a collection of particles (positions ri ) is defined as:

∑=i

ii rmRM

where the total mass M is given by:

∑=i

imM

Using N2, the acceleration of the centre of mass is given by:

∑∑∑∑∑ +===i j

iji

ioi

ii

ii FFFrmRM

where Fi is the total force on the ith particle, Fio is the external force on the ith particle and Fij is the force on the ith particle from the jth . Since jiij FF −= by Newton’ s 3rd law, the term

∑∑i j

ijF sums to zero, and the acceleration of the centre of mass is given by the sum of the

external forces Fo:

oi

ioi

ii FFrmRM === ∑∑

Since oPRM = , the total momentum, in the absence of an external force momentum is

conserved ( 0== oo PF ).

1.1.12 Motion of a collection of particles: rotations about the centre of Mass

- The motion of a collection of particles can be considered in two parts – the motion of the centre of mass, and the movement (or rotation if the particles make a rigid body) of the collection of particles about the centre of mass. To help handle motion about a point, it is useful to define the quantities torque and angular momentum.

Page 11: Classical Dynamics -

6

1.1.13 Torque - The ‘effectiveness’ of a force at producing rotations about a particular point, A, is given by the moment (or torque) G of the force about that point:

FrG ×=

where r is the position of the point of action of the force relative to the reference point, A. Since torque is defined as the cross product of two vectors, it itself is automatically a vector.

1.1.14 Angular momentum - The angular momentum of a particle about a particular point, A, is given by the moment of the linear momentum about that point: prJ ×=

1.1.15 Relation between G and J - For a single particle, (index i)direct application of N2 to the definitions of G and J gives: iiiiiiiiiii GGFrvmvprprJ =+=×+×=×+×= 0

i.e. torque is the rate of change of angular momentum - For a collection of particles with no forces between them, the rate of change of angular momentum of each particle can be summed and equated to the sum of the external torques acting on each particle. o

ii

ii GGJJ === ∑∑

where J is the total angular velocity of the system and Go is the total external couple on the system. - If now two of the particles have are interacting, then the contribution to the total couple will be:

( ) 1221212121 FrrFrFrG ×−=×+×=∆

Page 12: Classical Dynamics -

7

( 2112 FF −= by N3). If we assume that the force between the two particles is directed along

the line between them then 12F is parallel to ( )21 rr − and 0=∆G . It follows that the

internal forces do not affect the result that the rate of change of total angular momentum of a system is given by the sum of the externally applied torques. - Angular momentum was ‘invented’ because it is a useful tool in rotational mechanics. The relation oGJ = enables forces applied to a body to be related to the its rotational

acceleration, and hence solve the dynamics.

1.1.16 Work and Energy - Work done by a force = force x (distance moved parallel to force)

rFrFW δθδδ .cos ==

- Work done by a force moving along path P

∫∫ ==2

1

2

1

..t

t

r

r

vFrdFW dt

(using N2) ∫=2

1

.t

t

vdt

vdm dt ( )∫=

2

1

2

2

1t

t

vdt

dm dt

21

22 2

1

2

1vmvm −=

= increase in KINETIC ENERGY (T)

1.1.17 Potential Energy and Conservative Forces - If the work done, W, by a force in moving along a path from a point 1r to a point 2r is

independent of the path taken, then the force is know as a conservative force and it is possible to relate the work done to a function ( )rU which takes the form of a potential energy:

( ) ( )21 rUrUW −=

(The equation defines U so that if work is done by the force, U decreases)

θ Frr δ+

r

1r

2r P

Page 13: Classical Dynamics -

8

- equating expressions for the work done in terms of kinetic and potential energies gives:

( ) ( )2121

22 2

1

2

1rUrUmvmvW −=−=

⇒ ( ) ( ) 222

211 2

1

2

1mvrUmvrU +=+ Conservation of Energy ( VTE += )

- ’Conservation of energy’ is simply the integral form of N2. - Other forms of ’energy’ (heat, electostatic etc) are ’fixed’ so as to fit in with this simple idea. - The force may be recovered from the potential energy by differentiating:

1D dx

xdUxF

)()( −=

3D ( )rUrF ∇−=)(

(∇ is the gradient vector operator which is given in Cartesian coordinates by

∂∂

∂∂

∂∂=∇

z

U

y

U

x

UU ,, )

- Central forces are conservative. Suppose: rrfrF ˆ)()( =

and since drrdr =.

we have ∫∫ =2

1

2

1

)(.r

r

r

r

drrfrdF

since the work done by the force in going from 1r to 2r is a scalar integral of )(rf , the work

done is independent of path and the force is conservative - How can we know if a force is conservative, without performing integrals over all paths? A force is conservative if:

0. =∫ rdF

for any closed path. Stokes theorem states that integral of a vector round a closed loop is equal to the integral of the curl of that vector across any surface bounded by that loop. Since

for a conservative force 0. =∫ rdF for all loops then 0. =×∇∫ SdF for any integration over

any area and so for a conservative field:

Page 14: Classical Dynamics -

9

0=×∇ F

So, if the curl of a force is everywhere zero, the force is conservative and there exists a scalar function )(rφ such that:

φ∇−=F

1.2 The Energy Method - It is often much easier to write an expression for the energy of a mechanical system the to work out the forces acting on the various bodies, and the ‘energy method’ enables single variable problems to be solved, provided energy is conserved. Example 1.2.1 Simple pendulum (one variable) Total energy of system: VTE +=

Kinetic energy: ( ) 222

2

1

2

1 θθ mllmT ==

Potential energy: ( ) )cos1( θθ −== mglmgyV

−+= )cos1(

2

1 22 θθ mglmldt

d

dt

dE

0sin2 =+= θθθθ mglml

so, either the system is stationary ( 0=θ ) or: θθ sin2 mglml −=

θθ sinl

g−=

- Note: (1) the equation of motion is derived from the energy (2) no forces were considered

(3) one constraint (conservation of energy) was used – so we could solve for only one degree of freedom

m

l

θ

Page 15: Classical Dynamics -

10

Example1.2.2 1D diatomic molecule

( )212

222

211 2

1

2

1

2

1xxkxmxmE −++=

Try differentiating w.r.t. time:

( )( )1212222111 xxxxkxxmxxmdt

dE −−++=

- because there are now two variables – cannot simply divide by the derivative of one of them as in Example 1.2.1 - need an extra constraint to allow one of the variables to be eliminated. - since no external force – centre of mass moves at constant velocity, so use coordinates X0 , the position of the centre of mass and X, the separation between the masses:

21

22110 mm

xmxmX

++= 12 xxX −= 21 mmM +=

since there is no external force:

00 =X

The substitution into the expression for the energy gives:

2220 2

1

2

1

2

1kXXXME ++= µ (µ is the reduced mass,

+

=22

21

mm

mmµ )

and taking the derivative with respect to time gives:

000 =++= XkXXXXXMdt

dE µ

m1 m2

x1 x2

k

Page 16: Classical Dynamics -

11

since 00 =X and if we take the case where 0≠X we can divide by X to give the equation of

motion:

kXX += µ0

and we have the solution:

VtAX +=0 , ( )εω += taX cos

from which ( )tx1 and ( )tx2 can be deduced. (A, V, a and ε are constants determined by the

state of the molecule at 0=t , and ω is the angular frequency of the simple harmonic motion

of X, µω k= .)

Page 17: Classical Dynamics -

12

1.3 Lagrangian Mechanics

1.3.1 The need for Lagrangian mechanics - A straight forward application of N2 to a dynamics problem results in very complex algebra, except in the simplest of cases. Consider the following example: Example 1.3.1.1 A double pendulum Two point masses hung from a single point by two threads of each of length l and tensions T1, and T2 as indicated. (The x and y coordinates are measured w.r.t. the equilibrium position -

012 == θθ .)

Equations of motion:

112211 sinsin θθ TTxm −=

gmTTym 1221111 coscos −−= θθ

2222 sinθTxm −=

gmTym 22222 cos −= θ

Constraints needed to eliminate T1, T2, θ1, θ2,

( ) 21

21

2 xyll +−=

( ) ( )212

212

2 xxyyll −++−=

lx1

1sin =θ

( )l

xx 122sin −=θ

If we can we find a way of deriving the dynamics from the energy, the problem would be much simpler since one can choose ‘easy’ variables, and far fewer equations are needed:

( )( )122122

21

22

21

21 cos2

2

1

2

1.K.E θθθθθθθ −+++= lmlm

( ) ( )21211 coscos2cos1.P.E θθθ −−+−= glmglm

(Note – to find the the velocity of m2 the 'θ1', and 'θ2' contributions were added using the 'cos' rule.)

θ1

l, T1

m2 x2

y2

x1 m1 θ2

y1

l, T2

Page 18: Classical Dynamics -

13

1.3.2 Constraints A constraint is something that links otherwise independent coordinates. As can be seen from Example 1.3.1.1, constraints greatly increase the complexity of dynamics problems, but Lagrangian mechanics can handle them very effectively. There are many types of constraints on a mechanical system, but, for the purposes of Lagrangian mechanics, they can be divided into two types: Holonomic constraints in which the constraint can be expressed as equations connecting the coordinates of the particles (and possibly time). Examples are the two strings in the double pendulum (Example 1.3.1.1), a bead moving along a wire, or particle constrained to stay on a surface. Nonholonomic constraints are those that cannot be expressed by such equations – examples include the walls that constrain particles to lie inside a box, or a particle sliding over a horizontal, solid cylinder, since there comes a point when the particle falls off the cylinder that can only be found once the solution to the motion is known.

1.3.3 The Lagrangian mechanics ’recipe’ For systems that have: (1) holonomic constraints (and certain extensions to some special nonholonomic cases) (2) forces can be derived from a scalar potential1 (which is normally a function of the position of the particles, but in special cases, such as magnetic forces on charged particles, may be a function of velocity as well) (3) workless constraints (constraint forces perpendicular to allowed motions – all the examples of constraints in section 1.3.2 are workless) the following 'recipe' can be used to derive the equations that govern the dynamics of the system: Step 1: Decide on the number of degrees of freedom that the system has (Nf) Step 2: Choose Nf coordinates ( iq , where the index i runs from 1 to Nf ) that can be

independently varied 2

Step 3: Write down the kinetic energy (T) and potential energy (V) in terms of the coordinates

iq and their time derivatives iq and define a quantity, the Lagrangian (L) as 3:

VTL −=

Page 19: Classical Dynamics -

14

Step 4: For each coordinate iq , an equation of motion is given by:

0=∂∂−

∂∂

ii q

L

q

L

dt

d

(the Euler-Lagrange equations 4)

Step 5: Solve the Nf equations obtained in (4) to give the solution to the problem Comments: (1) The recipe cannot be used if particles move against frictional forces, since these cannot be described by a scalar potential. (2) If the problem is in D dimensions, with N particles, and k constraints, there are ND possible independent coordinates, of which k can be eliminated by applying the constraint conditions, giving kNDN f −= . If it makes the problem easier to set up or solve, more

coordinates than Nf can be used, but the constraints that would otherwise have been used to eliminate the extra ones must then be included in the Euler-Lagrange equations by the method of Lagrange multipliers (outside the scope of this course). (3) In defining the state of a system, the iq and iq coordinates can be specified

independently- i.e. you can choose freely where a particle is and how fast it is moving. (4) A conjugate momentum can be defined that ’belongs’ with each coordinate iq :

ii q

Lp

∂∂=

and the Euler-Lagrange equations become:

(The Euler-Lagrange equations are also known simply as the Lagrange equations)

Rate of change of conjugate momentum

Generalised force

ii q

Lp

dt

d

∂∂=

Page 20: Classical Dynamics -

15

1.3.4 Examples 1: Time independent Lagrangians Example 1.3.4.1 Simple Harmonic Motion Consider a mass m attached by a spring of constant k to a fixed support.

22

2

1

2

1kxxmVTL −=−=

The Euler-Lagrange Equation gives:

0=∂∂−

∂∂

x

L

x

L

dt

d

02

1

2

1 22 =

∂∂−

∂∂

kxx

xmxdt

d

( ) 0=+ kxxmdt

d

(note how x has been treated as an independent variable, so, e.g. 0=∂∂x

x

)

kxxm −= Example 1.3.4.2 Double pendulum (see figure of Example 0.1)

V-TL =

( )( ) ( ) ( )21211122122

21

22

21

21 coscos2cos1cos2

2

1

2

1 θθθθθθθθθθ −−−−−−+++= glmglmlmlm

Making the small angle approximation:

( ) ( )22

212

21121

22

21

22

21

21 2

1

2

12

2

1

2

1 θθθθθθθθ +−−+++= glmglmlmlmL

The Euler-Lagrange Equations give:

θ1: ( ) 0121122

212

212

1 =++++ θθθθθ glmglmlmlmlmdt

d

( ) ( ) 021122121 =++++ mmglmlmm θθθ 1.3.4.1

Page 21: Classical Dynamics -

16

θ2: ( ) 02212

222

2 =++ θθθ glmlmlmdt

d

( ) 022122 =++ θθθ gmlm 1.3.4.2

Equations 1.3.4.1 and 1.3.4.2 are the equations of motion of the system, derived with much less effort than a direct application of N2 to the problem. To solve: for simplicity let 21 mm = and guess a solution in which ( ) ( )tt 12 αθθ = . Substituting

into Equations 1.3.4.1 and 1.3.4.2 gives: ( ) 022 11 =++ θθα gl 1.3.4.3

( ) 01 11 =++ θαθα gl 1.3.4.4

Since equations 1.3.4.3 and 1.3.4.4 are now equivalent, the ratio of the coefficients of 1θ and

1θ must be the same in both equations, giving:

( )

ααα 2

1

2 =++

⇒ 2±=α

If we now look for oscillating solutions of the form ( )εωθ += tAcos1 we have:

( ) ( ) 0cos]22[ 2 =+++− εωαω tAgl

and ( )αω

+=

2

22

l

g

so there are two modes of oscillation,

1st mode: 1θ and 2θ oscillate in phase, 12 2θθ = and ( )22

2

+=

l

2nd mode: 1θ and 2θ oscillate in antiphase, 12 2θθ −= and ( )22

2

−=

l

note: - The Lagrangian method yields the equations of motion, which still need to be solved - This is a good example of a system showing normal modes –one for each coordinate

Page 22: Classical Dynamics -

17

Example 1.3.4.3 Double Pulley

( ) ( )23

22

21 2

1

2

1

2

1xymxymxmT ++−+=

( ) )(321 yxgmxygmgxmV ++−−−=

( ) ( )( ) )(

2

1

2

1

2

1

321

23

22

21

yxgmxygmgxm

xymxymxmVTL

+−−++

++−+=−=

using Euler-Lagrange equations:

for x: ( ) ( )( ) gmgmgmxymxymxmdt

d321321 −−=++−−

( ) ( ) ( )gmmmymmxmmm 32123321 −−=−+++

for y: ( ) ( )( ) gmgmxymxymdt

d3232 −=++−

( ) ( ) ( )gmmymmxmm 322323 −=++−

so for m1=4kg, m2=1kg, m3=3kg, and g=10ms-2 we have:

028 =+ yx 22042 −−=+ msyx

and 2

710 −−= msx 2

740 −= msy

Note the tensions of the strings do not appear in the Lagrangian which simplifies the calculation, but by the same token they cannot be derived from the Lagrangian.

1.3.5 Derivation of the Lagrangian mechanics formalism

- The Euler–Lagrange equations are essentially Newton's second law written in generalised coordinates. - The details of the derivation are rather involved, but effectively one starts with N2 expressed for each Cartesian coordinate of the system, and transforms coordinates form Cartesian to generalised ones, using the fact that the forces can be written as derivatives of a scalar potential, i.e. it is a coordinate transformation.

m1

m2

m3

x

y

Page 23: Classical Dynamics -

18

1.3.6 The Hamiltonian - Lagrange’s equations are based on iq and iq as independent coordinates - remember you

can choose freely where a particle is and how fast it is moving. - You may wish to use the conjugate momentum ip as variables instead of the velocities iq .

If you do, then the same physics is contained within, and expressed by a new function, H, the Hamiltonian, given by:

∑=

−∂∂=

fN

i ii L

q

LqH

1

1.3.6.1

- A similar variable change – a so called 'canonical transformation' - is found in classical thermodynamics. All the thermodynamical information about a system is contained, for example, in the Gibbs free energy, provided you use P and T as base variables, i.e. within the equation:

SdTVdPdG −= . 1.3.6.2 If you wish to use V and T as base variables instead, then in order to have a function that completely describes the system in these new variables, you need to subtract PV from G to obtain the Helmholz free energy, PVGF −= , which expresses dF as an exact derivative with respect to V and T:

SdTPdVVdPPdVdGdF −−=−−=

If one notes from equation 1.3.6.2 that P

GV

∂∂= then we have:

P

GPGF

∂∂−= 1.3.6.3

which illustrates why H is derived from L in the way shown in equation 1.3.6.1. - If the potential V is not a function of velocity and if the equations that convert generalised coordinates to Cartesian ones do not contain time explicitly, then it can be shown that the Hamiltonian is equal to the total energy of the system. (For the purists this is a sufficient but not a necessary condition – there are some velocity dependent potentials, such as the one used to describe magnetic forces on charged particles for which the Hamiltonian is still equal to the energy!) c.f. - Example 1.3.4.1

Page 24: Classical Dynamics -

19

EVTkxxmkxxmxmxH =+=+=+−= 2222

2

1

2

1

2

1

2

1

1.3.7 Examples 2: Time dependent Lagrangians Example 1.3.7.1 System with SHM driver – driven by a time dependent potential. Mass on spring driven by moving support. For this system:

2

2

1xmT = ( )2

0cos2

1taxkV ω−=

and ( )20

2 cos2

1

2

1taxkxmL ω−−=

The Euler-Lagrange equation gives:

( ) ( ) 0cos 0 =−+ taxkxmdt

d ω

( ) 0cos 0 =−+ taxkxm ω

The Hamilton for the system is time dependent, and in this case equal to the total energy:

( )20

22 cos2

1

2

1taxkxmLxmL

x

LxH ω−+=−=−

∂∂=

Example 1.3.7.2 wire on bead – a system driven by making the actual Cartesian coordinates of the system depend on the generalised one (q the distance of the bead along the wire) in a way that depends on time:

θcosqz = , ( )φθ tqx cossin= , ( )φθ tqy sinsin=

( ) 222 sin2

1

2

1 φθ qmqmT +=

θcosmgqV =

x

k

m

ta 0cosω

θ

m

q

fixedφ

Page 25: Classical Dynamics -

20

( ) θφθ cossin2

1

2

1 222 mgqqmqmVTL −+=−=

The Euler-Lagrange equation gives: 0cossin 22 =+− θφθ mgmqqm

θφθ cossin 22 gqq −=

To understand the behaviour of the system, consider how the acceleration along the wire depends on q:

So for 0qq > mass starts at rest, and then

is flung out with an ever increasing acceleration.

The Lagrangian for this system does not contain time explicitly, and so, as outlined in section 1.3.8 below, the Hamilton for the system is a constant of the motion, given by :

( ) θφθ cossin2

1

2

1 222 mgqqmqmLqmqLq

LqH +−=−=−

∂∂=

In this case, because the actual Cartesian coordinates of the system depend on the generalised ones in a way that depends on time (a ’rheonomic’ system), H is not equal to the total energy which would be given by:

( ) θφθ cossin2

1

2

1 222 mgqqmqmVTE ++=+=

and given that H is conserved, E, with its different functional form, is not conserved.

1.3.8 Conservation of energy and the Hamiltonian

- The way H changes with time can be deduced by calculating dt

dH:

q

q

θφθ220 sin

cos

gq =unstable

equilibrium

Page 26: Classical Dynamics -

21

∑=

−∂∂=

fN

i ii L

q

LqH

1

∑=

∂∂+

∂∂=

fN

i ii

ii

dt

dL

q

L

dt

dq

q

Lq

dt

dH

1

1.3.6.1

- To work out dt

dL we can start with the expression for dL in terms of changes in the

coordinates of L, which in general are the iq , the iq , and t,:

dtt

Lqd

q

Ldq

q

LdL

ff N

ii

i

N

ii

i ∂∂+

∂∂+

∂∂= ∑∑

== 11

’Dividing’ by dt gives:

t

Lq

q

Lq

q

L

dt

dL ff N

ii

i

N

ii

i ∂∂+

∂∂+

∂∂= ∑∑

== 11

and substituting into equation 1.3.6.1 gives:

∑= ∂

∂−

∂∂−

∂∂=

fN

i iii t

L

q

L

q

L

dt

dq

dt

dH

1

1.3.6.2

The Euler-Lagrange equations mean that the square bracket in equation 1.3.6.2 is always zero, and so:

t

L

dt

dH

∂∂−=

So, if L does not have an explicit time dependence, i.e. if t does not appear explicitly in L

then 0=∂∂

t

L and H is a constant of the motion. If in addition, the potential V is not a function

of velocity and if the equations that convert generalised coordinates to Cartesian ones do not contain time explicitly then the Hamiltonian is equal to the energy of the system, and energy is thus conserved. This astonishing result may be expressed thus: Energy is conserved because of the time invariance of space (you get the same result if you do the experiment now, or a time t later, provided you start with an identical initial conditions).

Page 27: Classical Dynamics -

22

1.3.9 Invariance of L and conservation laws - The conservation of energy is a specific example of a general principle known as Noether’s Theorem: ’If a symmetry exists in the Lagrangian, there is a corresponding constant of the motion’ - What about translational invariance? i.e. an isolated system where the Lagrangian (L’) of the system after moving the whole system by an amount (δx, δy, δz) with respect to a Cartesian frame of reference is equal to the Lagrangian of the system before the move (L). We have:

( ) ( )2,,,,,’ xOz

Lz

y

Ly

x

LxLtqQzzyyxxLL ii δδδδδδδ +

∂∂+

∂∂+

∂∂+=+++=

Since LL =’ for any value of the small displacement the derivativesx

L

∂∂

etc must be zero.

The Euler-Lagrange equations then give:

0=∂∂=

x

Lp

dt

dx (etc)

and xp , yp , zp are constants of the motion – i.e. linear momentum is conserved. i.e.:

Linear momentum is conserved because of the translational invariance of space

- In a similar way the rotational invariance of space leads to the conservation of angular momentum.

Page 28: Classical Dynamics -

23

1.4 Non-Inertial Frames

1.4.1 Transformations from stationary to rotating frames

- If a point on a rigid body has a position vector r measured with respect to the axis of rotation, and if the body is rotating with and angular velocity ω , then the velocity of the point

with respect to stationary coordinates is given by: rv ×= ω0

(vectors with the suffix o are with respect to the stationary frame, S0, those without a suffix are with respect to the rotating frame, S, and for convenience the equations are written for the instant in time when the coordinate axes of S and S0 coincide, so rr =0 and ωω =0 )

- If the point is also moving on the rigid body with a velocity v with respect to coordinates that rotate with the body, then this extra velocity must be added into the velocity 0v :

rvv ×+= ω0

- similarly the time derivative of any vector in S0 is related to that of the same vector quantity, but measured with respect to S by:

Adt

Ad

dt

Ad

SS

×+

=

ω

0

1.4.2 Derivation of fictitious forces - Newton’ s Laws only apply in inertial frames. It is, however, often convenient to solve dynamical problems from the point of view of a rotating (non-inertial) frame such as one fixed to the Earth. If you are making observations in such a frame, and if you insist that you want to use Newton’ s laws to describe what you see, then you have to allow for ‘fictitious’ forces that seem to push objects around in mysterious ways. - In dynamics these ‘forces’ are introduced as a ‘fix’ to make Newton’ s laws work in a frame where the laws are actually invalid, because solving the problem in an ‘proper’ inertial frame is harder than worrying about the fictitious (‘fix’ ) forces in the non-inertial one. They are also introduced because to an observer forces are more apparent than accelerations, and if

Page 29: Classical Dynamics -

24

something appears to accelerate, we have enough of an intuitive feel for Newton’ s laws to ask what force is accelerating them. If you want to hold an object still in a rotating frame, technically what you are doing is accelerating the object towards the axis of rotation. However, what you see is a stationary object and what you feel is the very real centrifugal force (actually a N3 reaction to you pushing and accelerating the object inwards) that tends to fling the object outwards. Furthermore, if you let go of that object, it is a fair question to ask what force produces the apparent acceleration outwards that ensues. - In order to derive an expression for the acceleration of an object as seen in a rotating frame, we can use the formula given above that rates the time derivative of an vector in stationary and rotating frames. Firstly for velocity:

rdt

rd

dt

rd

SS

×+

=

ω

0

0 , i.e. rvv ×+= ω0

The subtle bit of this derivation now follows. The true acceleration of an object (i.e. measured in an inertial frame such as S0 ) is the time derivative of the true velocity, i.e. 0v if you

measure it in S0 and rv ×+ ω if you measure it in S. (note the observer in S thinks the

velocity is just v , but once he is told that actually it is rv ×+ ω he can then use this to work

out the true acceleration in S coordinates.) So, the expression for the true acceleration, a0 , in terms of vectors measured in S is:

( ) ( )rv

dt

rvdv

dt

vd

dt

vda

SSS

×+×+

×+=×+

=

= ωωωω 0

000

0

( )rvva ××+×+×+= ωωωω

( )rva ××+×+= ωωω2

Now, an observer in the rotating frame measures an acceleration a, and as far as he is concerned, since he wishes to believe that Newton’ s laws apply in his non-inertial frame, the total force acting on the object must be given by am , which will have to be composed as

follows:

Total Apparent force Coriolis force True force 0am= centrifugal force

( )rmvmFam ××−×−= ωωω2

Page 30: Classical Dynamics -

25

- The relation between a and a0 can also be derived by expressing the position vector of a point, r0 explicitly in terms of the coordinates (x, y, z) and rotating basis set, ( x , y , z ) in S0

and taking derivatives w.r.t. time, noting that: xx ˆˆ ×= ω , ( )xx ˆˆ ××= ωω etc.

zzyyxxr ˆˆˆ0 ++=

........ˆˆ0 +++= xxxxr (y and z terms similar)

( ) ........ˆˆ ++×+= xxxx ω (y and z terms similar)

( ) ( )[ ] ........ˆˆ2ˆ0 +××+×+= xxxxxxr ωωω (y and z terms similar)

( )rvaa ××+×+= ωωω20 (as previously)

1.4.3 Centrifugal force - Using cylindrical polar coordinates (ρ, φ, z) (see section 1.1.3) with the axis of rotation along the z axis:

zzr ˆˆ+= ρρ , zωω = , ρ is the distance from the axis or rotation

⇒ ( ) φωρρωρω ˆˆˆ =×=× zr

⇒ ( ) ( ) ρρωφρωωω ˆˆˆ 22 =×−=××− zr

i.e. Centrifugal force = ×2ωm (distance from axis of rotation)

and is directed away from the axis of rotation Example: rotation of Earth at equator

64001

22

2

dayRe

πkm ≈ 0.003g

Page 31: Classical Dynamics -

26

1.4.4 Coriolis force on Earth’s surface vmF Cor ×Ω−= 2

=Ω Earth’ s angular velocity

=v velocity in frame S′, fixed to Earth

To derive the Coriolis force at a latitude λ,

set up a Coordinate axes in S′: ’x pointing East

’y pointing North

’z pointing radially outwards at latitude λ.

Expressing the Earths angular velocity in this coordinate system gives:

’sin'cos zy λλ Ω+Ω=Ω

for v in the plane of the Earth’ s surface at this latitude:

’' yvxvv yx +=

⇒ ( )’'sin2 yvxvmF xyCor −Ω= λ sideways force

⇒ ’cos2 zvm xλΩ+ vertical force

- The sideways force is always perpendicular to v in surface plane with a magnitude of

λsin2 vmΩ , independent of the direction of v in the surface plane, and acts to the right in the northern hemisphere, and to the left in the southern hemisphere. - The vertical force is usually small compared to the gravitational attraction of the Earth, and is often neglected.

’z’y

’x

Ω

Equator

N pole

λ