classical mechanics fall 2011 chapter 11: coupled...

14
1 Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal Modes 1. Two Masses Coupled By Three Springs There are many interesting systems in which individual oscillators are coupled by some kind of interaction. Because of the coupling, the motion of each oscillator is influenced by others with which it interacts. An example of such a coupled system is a crystalline solid in which each of the many atoms that constitute the crystal interacts with other atoms in the crystal via an interatomic force. We begin our study of coupled oscillators with a much simpler system than a crystal but which nevertheless provides valuable information on the behavior of more complex systems: two objects coupled by three springs. The springs are ideal springs, which obey Hooke’s law for all extensions (and compressions). The leftmost and rightmost springs are attached to rigid supports at one end. The objects, which we take to be carts, move on a frictionless horizontal surface. Since this is a relatively simple system, we will use Newton’s second law to write down the equations of motion. Note that we can find the force exerted on a cart by a spring by considering the displacement of the end of the spring attached to the cart relative to the other end of that spring. This is important in cases in which both ends of a spring are attached to moving carts so that the displacement of each end of the spring changes with time. The equations of motion are given below. Cart 1: m 1 x 1 = k 1 x 1 k 2 ( x 1 x 2 ). Cart 2: m 2 x 2 = k 3 x 2 k 2 ( x 2 x 1 ). We can rewrite the equations of motion as follows: m 1 x 1 = ( k 1 + k 2 ) x 1 + k 2 x 2 (11.1) and m 2 x 2 = k 2 x 1 ( k 2 + k 3 ) x 2 . (11.2) Eqs. (11.1) and (11.2) can be collectively written using a matrix equation: M x = Kx, (11.3) where x is the column matrix

Upload: others

Post on 23-Jan-2020

14 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  1  

Classical Mechanics Fall 2011

Chapter 11: Coupled Oscillators and Normal Modes

1.  Two  Masses  Coupled  By  Three  Springs  

There  are  many  interesting  systems  in  which  individual  oscillators  are  coupled  by  some  kind  of  interaction.  Because  of  the  coupling,  the  motion  of  each  oscillator  is  influenced  by  others  with  which  it  interacts.  An  example  of  such  a  coupled  system  is  a  crystalline  solid  in  which  each  of  the  many  atoms  that  constitute  the  crystal  interacts  with  other  atoms  in  the  crystal  via  an  interatomic  force.  We  begin  our  study  of  coupled  oscillators  with  a  much  simpler  system  than  a  crystal  but  which  nevertheless  provides  valuable  information  on  the  behavior  of  more  complex  systems:  two  objects  coupled  by  three  springs.  

     

The  springs  are  ideal  springs,  which  obey  Hooke’s  law  for  all  extensions  (and  compressions).  The  leftmost  and  rightmost  springs  are  attached  to  rigid  supports  at  one  end.  The  objects,  which  we  take  to  be  carts,  move  on  a  frictionless  horizontal  surface.  Since  this  is  a  relatively  simple  system,  we  will  use  Newton’s  second  law  to  write  down  the  equations  of  motion.  Note  that  we  can  find  the  force  exerted  on  a  cart  by  a  spring  by  considering  the  displacement  of  the  end  of  the  spring  attached  to  the  cart  relative  to  the  other  end  of  that  spring.  This  is  important  in  cases  in  which  both  ends  of  a  spring  are  attached  to  moving  carts  so  that  the  displacement  of  each  end  of  the  spring  changes  with  time.  The  equations  of  motion  are  given  below.  

Cart  1:   m1x1 = −k1x1 − k2 (x1 − x2 ).  

Cart  2:   m2x2 = −k3x2 − k2 (x2 − x1).  

We  can  rewrite  the  equations  of  motion  as  follows:  

  m1x1 = −(k1 + k2 )x1 + k2x2   (11.1)  

and  

  m2x2 = k2x1 − (k2 + k3)x2.   (11.2)  

Eqs.  (11.1)  and  (11.2)  can  be  collectively  written  using  a  matrix  equation:  

  Mx = −Kx,   (11.3)  

where   x  is  the  column  matrix  

Page 2: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  2  

  x =x1x2

⎝⎜⎜

⎠⎟⎟,   (11.4)  

M  is  the  mass  matrix  

 m1 00 m2

⎝⎜⎜

⎠⎟⎟,   (11.5)  

and  K  is  the  spring-­‐constant  matrix  

  K =k1 + k2 −k2−k2 k2 + k3

⎝⎜⎜

⎠⎟⎟.   (11.6)  

Using  matrices  is  a  very  powerful  method  of  doing  mechanics,  as  we  shall  presently  see.    

In-­class  Problem:  Taylor,  Problem  11.2  

Since  the  system  involves  springs  and  masses  and  no  friction,  we  look  for  oscillatory  solutions  to  the  equations  of  motion.  Let  us  try  

 x1 = F1 cosωt +G1 sinωtx2 = F2 cosωt +G2 sinωt.

  (11.7)  

To  facilitate  finding  the  solutions,  let  us  use  complex  numbers  to  represent  the  trial  solutions.  We  have    

 z1 = C1e

i(ωt−δ1 ) = a1eiωt

z2 = C2ei(ωt−δ2 ) = a2e

iωt .   (11.8)  

In  Eq.  (11.8),  a1 = C1e− iδ1  and  a2 = C2e

− iδ2 ,  where  a1, a2,  C1,  and  C2  are  constants.  (a1 and a2  are  complex,  C1  and  C2    are  real.)  The  actual  solutions  (the  displacements   x1  and   x2 )  are  the  real  parts  of   z1  and   z2 .    In  matrix  form,  we  have    

  z =a1a2

⎝⎜⎜

⎠⎟⎟eiωt = aeiωt ,   (11.9)  

with  x = Rez.  Substituting  the  trial  solution  Eq.  (11.9)  in  the  matrix  equation  of  motion  (11.3)  gives    

  −ω 2Maeiωt = −Kaeiωt ,  

which  can  be  rewritten  as    

Page 3: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  3  

  (K −ω 2M)a = 0.   (11.10)  

This  is  an  eigenvalue-­‐type  equation  for  which  solutions  exist  if    

  det(K −ω 2M) = 0.   (11.11)  

The  eigenvalues  are  the  values  of  ω 2 obtained  from  solving  Eq.  (11.11).  After  obtaining  the  eigenvalues,  one  can  then  find  the  column  matrices   a ,  i.e.,  the  eigenvectors,  by  substituting  each  eigenvalue  into  Eq.  (11.10).  One  then  has  the  complete  solution  after  taking  the  real  part  of  z.  Let  us  illustrate  the  technique  by  solving  the  problem  of  two  identical  masses  coupled  by  three  identical  springs.  

Two  Identical  Masses  Coupled  By  Three  Identical  Springs  

Let  m1 = m2 = m  and   k1 = k2 = k3 = k.The  mass  and  spring-­‐constant  matrices  are    

  M = m 00 m

⎛⎝⎜

⎞⎠⎟

and K = 2k −k−k 2k

⎛⎝⎜

⎞⎠⎟

.   (11.12)  

Eq.  (11.11)  gives  

  2k −mω 2 −k−k 2k −mω 2

= 0,  

which  yields    

  (2k −mω 2 )2 − k2 = 0,  

or,    

  (k −mω 2 )(3k −mω 2 ) = 0.  

The  solutions  are    

  ω1 =km

and ω 2 =3km

.   (11.13)  

These  two  frequencies  are  called  the  normal-­mode  frequencies  or  normal  frequencies  of  the  coupled  oscillator  system.  Let  us  now  substitute  ω1 into  Eq.  (11.10).  We  have    

  k −k−k k

⎛⎝⎜

⎞⎠⎟

a1a2

⎝⎜⎜

⎠⎟⎟= 0,  

which  gives  two  equations,  both  of  which  are  equivalent  to  a1 − a2 = 0,  i.e.,  a1 = a2.  Eq.  (11.8)  tells  us  that    

Page 4: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  4  

  z1 = z2 = aeiω1t = A1e

i(ω1t−δ1 ),  

where  we  have  taken  a1 = a2 = a  and  a = A1eiδ1 .  In  matrix  form,  we  have    

  z = aa

⎛⎝⎜

⎞⎠⎟eiω1t =

A1A1

⎝⎜⎜

⎠⎟⎟ei(ω1t−δ1 ).  

 Taking  the  real  parts  gives  the  solution  for  the  first  normal  mode:    

 x1(t) = A1 cos(ω1t −δ1)x2 (t) = A1 cos(ω1t −δ1). [First normal mode]

  (11.14)  

Clearly,  both  masses  oscillate  precisely  in  phase,  having  the  same  displacement  from  equilibrium  at  all  times.  It  follows  that  the  middle  spring  is  always  relaxed  and  has  no  effect  on  the  motion.  That  is  why  the  first  normal  frequency  is  exactly  that  of  a  single  mass  attached  to  one  spring.  

We  now  substitute  ω 2 into  Eq.  (11.10),  yielding  

          −k −k−k −k

⎛⎝⎜

⎞⎠⎟

a1a2

⎝⎜⎜

⎠⎟⎟= 0,  

which  gives   a1 + a2 = 0,  i.e.,  a1 = −a2.  Thus,  we  obtain    

  z = a−a

⎛⎝⎜

⎞⎠⎟eiω2t =

A2−A2

⎝⎜⎜

⎠⎟⎟ei(ω2t−δ2 ),  

where   a = A2eiδ2 .  Taking  the  real  part  gives  the  displacements  as  a  function  of  time  for  the  

second  normal  mode:  

 x1(t) = A2 cos(ω 2t −δ 2 )x2 (t) = −A2 cos(ω 2t −δ 2 ). [Second normal mode]

  (11.15)  

In  this  mode,  the  masses  oscillate  exactly  out  of  phase.  The  displacements  have  the  same  magnitude  but  are  opposite  in  direction.  For  any  given  displacement,  the  middle  spring  is  stretched  or  compressed  by  an  amount  equal  to  twice  the  magnitude  of  that  displacement.  Along  with  the  force  exerted  by  the  leftmost  or  rightmost  spring,  the  magnitude  of  the  net  force  on  each  mass  is   3k x .  This  explains  the  value  of  the  second  normal  frequency.    The  general  solution  is  a  superposition  of  the  solutions  for  the  first  and  second  modes.  (Remember  that  the  equations  of  motion  are  linear  so  the  principle  of  superposition  is  valid.)  Thus,  the  general  solution  is    

Page 5: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  5  

 x1(t) = A1 cos(ω1t −δ1)+ A2 cos(ω 2t −δ 2 )x2 (t) = A1 cos(ω1t −δ1)− A2 cos(ω 2t −δ 2 ). [General solution]

  (11.16)  

There  must  be  four  unknown  constants  because  two  second-­‐order  differential  equations  were  solved.  In  matrix  form,  the  general  solution  is    

  x(t) = A111

⎛⎝⎜

⎞⎠⎟cos(ω1t −δ1)+ A2

1−1

⎛⎝⎜

⎞⎠⎟cos(ω 2t −δ 2 ).   (11.17)  

We  can  rewrite  Eq.  (11.17)  in  the  following  form,  which  is  sometimes  more  convenient:  

  x(t) = (B1 cosω1t +C1 sinω1t)11

⎛⎝⎜

⎞⎠⎟+ (B2 cosω 2t +C2 sinω 2t)

1−1

⎛⎝⎜

⎞⎠⎟.   (11.18)  

The  general  motion  is  therefore  a  combination  of  oscillations  at  both  normal  frequencies.  Note  that  there  are  two  normal  frequencies  because  the  system  consists  of  two  oscillators.  Solving  associated  2 × 2  determinant  equation  gives  a  quadratic  equation  in  ω 2 ,  which  has  exactly  two  solutions.  

In-­class  Problem:  Taylor,  Problem  11.7.  

 

2.  Normal  Coordinates  

Let  us  add  Eqs.  (11.1)  and  (11.2)  with  the  masses  equal  (=  m)  and  the  spring  constants  all  the  same  (=  k).  We  obtain  

  m(x1 + x2 ) = −k(x1 + x2 ).   (11.19)  

Let  us  define  the  normal  coordinate    

  ξ1 = x1 + x2.   (11.20)  

Substituting  ξ1  in  Eq.  (11.19)  gives  

  mξ1 = −kξ1.   (11.21)  

Eq.  (11.21)  is  the  equation  of  motion  for  a  single  harmonic  oscillator  with  angular  frequency  ω1 = k /m .  We  see  that  rewriting  the  equations  of  motion  in  terms  of  the  normal  coordinate  causes  them  to  become  uncoupled  and  the  resulting  motion  is  oscillation  at  a  single  frequency  equal  to  the  normal  frequency  ω1.  

We  now  subtract  add  Eq.  (11.2)  from  Eq.  (11.1)  giving    

  m(x1 − x2 ) = −3k(x1 − x2 ).   (11.22)  

Page 6: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  6  

Defining  a  second  normal  coordinate  

  ξ2 = x1 − x2   (11.23)  

and  substituting  it  in  Eq.  (11.22)  yields  

  mξ2 = −3kξ2.   (11.24)  

The  second  normal  coordinate  gives  oscillation  at  the  second  normal  frequency  ω 2 = 3k /m  only.  We  can  initiate  oscillations  with  a  single  frequency  using  initial  conditions.  If  both  masses  are  given  the  same  initial  displacement  and  released  from  rest,  then  the  normal  coordinate  ξ1  will  initially  be  non-­‐zero  whereas  the  normal  coordinate  ξ2  will  initially  be  zero  and  will  remain  zero  for  all  time.  Hence,  the  system  will  oscillate  with  normal  frequency  ω1 .    

In-­class  Question:  How  would  you  generate  oscillations  at  a  single  frequency  ω 2 ?  

Note  that  if  we  write  the  equations  of  motion  for  the  normal  coordinates  in  matrix  form,  we  obtain  

  Mξ = − ′K ξ,   (11.25)  

where    

  ξ =ξ1ξ2

⎝⎜⎜

⎠⎟⎟,  

  M = m 00 m

⎛⎝⎜

⎞⎠⎟

and ′K = k 00 3k

⎛⎝⎜

⎞⎠⎟

.  

Thus,  the  normal  coordinates  diagonalize  the  M  and  K    matrices.  (In  this  simple  case,  the  M  matrix  was  diagonal  from  the  start.)  This  happens  because  the  normal  coordinates  can  be  written  in  terms  of  the  eigenvectors  for  the  system.  In  fact,  notice  that  we  can  write  the  normal  coordinates  of  the  oscillators  in  matrix  form  as    

  ξ = x1a1 + x2a2 = x111

⎛⎝⎜

⎞⎠⎟+ x2

1−1

⎛⎝⎜

⎞⎠⎟=

x1 + x2x1 − x2

⎝⎜⎜

⎠⎟⎟=

ξ1ξ2

⎝⎜⎜

⎠⎟⎟.   (11.26)  

The  normal  coordinates  can  be  expanded  in  a  vector  space  in  which  the  eigenvectors  a1  and  a2  that  solve  Eq.  (11.10)  play  the  role  of  unit  vectors,  i.e.,  they  are  the  basis  vectors  that  span  the  space.  The  displacements  are  the  “components”  or  coefficients  of  the  basis  vectors.  We  can  also  write  the  displacements  in  terms  the  same  basis  vectors  with  the  normal  coordinates  being  the  coefficients:  

Page 7: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  7  

  x = 12 ξ1a1 + 1

2 ξ2a2 =12

ξ1 + ξ2ξ1 −ξ2

⎝⎜⎜

⎠⎟⎟.   (11.27)  

Note  that  in  this  case,  ξ1  is  the  coefficient  of  a1  and  so  is  associated  with  the  first  normal  mode  only.  (ai    and  ξi can  be  defined  so  that  that  the  inelegant  factor  of   12  goes  away.)  

In-­class  problem:  Taylor,  Problem  11.11.  

3.  Two  Weakly  Coupled  Oscillators  

Let  us  assume  that  in  our  two-­‐coupled-­‐oscillator  system,  the  masses  are  identical  (m)  and  the  rightmost  and  left  most  springs  are  identical  (spring  constant  k).  We  further  assume  that  the  middle  spring  has  a  spring  constant  k2  that  is  different  from  that  of  the  other  springs.    The  spring  constant  matrix  is    

  K =k + k2 −k2−k2 k + k2

⎝⎜⎜

⎠⎟⎟  

and    

  K −ω 2M =k + k2 −mω

2 −k2−k2 k + k2 −mω

2

⎝⎜⎜

⎠⎟⎟.  

Setting  the  determinant  equal  to  zero  and  solving  yields  

  ω1 =km

and ω 2 =k + 2k2

m.   (11.28)  

We  see  that  we  obtain  the  same  normal  frequencies  for  the  three  identical  springs  case  if  k2 = k .  As  before,  for  the  first  normal  mode,  the  masses  oscillate  in  phase;  for  the  second,  they  oscillate  out  of  phase.  

Let  us  consider  the  case  of  a  weak  middle  spring,  i.e.,   k2 k .  In  this  case,  the  second  normal  frequency  is  very  close  to  the  first.  It  is  useful  to  observe  that  superposing  two  sinusoidal  functions  of  the  same  amplitude  but  different  frequencies  gives  

  ′A cosω1t + ′A cosω 2t = 2 ′A cos ω1 −ω 2

2⎛⎝⎜

⎞⎠⎟ t

⎡⎣⎢

⎤⎦⎥cos ω1 +ω 2

2⎛⎝⎜

⎞⎠⎟ t

⎡⎣⎢

⎤⎦⎥.   (11.29)  

Thus,  if  we  use  initial  conditions  to  give   A1 = A2  and  δ1 = δ 2  (e.g.,  cart  2  is  given  an  initial  displacement  from  equilibrium  and  released  from  rest;  cart  1  is  at  rest  at  equilibrium)  in  the  general  solution  similar  to  Eq.  (11.16),  we  obtain  the  solution  

Page 8: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  8  

 x1(t) = Acosεt cosωtx2 (t) = Asinεt sinωt,

  (11.30)  

where    

  ε = ω1 −ω 2

2 and ω = ω1 +ω 2

2.   (11.31)  

For  weak  coupling,  the  average  frequency  ω  is  very  close  to  both  normal  frequencies  and  the  difference   ε  is  very  small.    Eq.  (11.30)  underlies  the  phenomenon  of  beats,  which  represent  a  slow  periodic  variation  of  the  amplitude  when  two  waves  with  frequencies  that  are  nearly  the  same  are  superposed.  Here,  the  beats  are  due  to  the  superposition  of  normal-­‐mode  oscillations  (corresponding  to  the  two  normal  coordinates)  that  have  nearly  equal  frequencies.  The  period  of  the  amplitude  variation  is  2π / ε .  From  Eq.  (11.30),  we  see  that  when  the  time  varying  amplitude  of  mass  1  is  maximum  ( cosεt = 1),  the  amplitude  of  mass  2  is  minimum  and  vice-­‐versa.  Thus,  the  energy  of  the  system  is  continuously  being  transferred  from  one  oscillator  to  the  other.  

In-­class  Problem:  Taylor,  Problem  11.13  (a)    

         

 

 

 

 

 

 

 

4.  The  Double  Pendulum:  Lagrangian  Approach  

The  double  pendulum  system  (depicted  below)  is  quite  a  bit  more  complicated  than  the  cart-­‐spring  system  that  we  have  analyzed.  Therefore,  we  will  use  the  Lagrangian  formulation  to  obtain  the  equations  of  motion.  

Weakly  coupled  oscillators  

x1  

x2  

Page 9: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  9  

           

The  two  generalized  coordinates  are  φ1  and  φ2 .  We  take  the  zero  of  gravitational  potential  energy  to  be  the  fixed  support  to  which  the  upper  pendulum  is  attached.  The  potential  energy  of  each  mass  is  given  by    

  U1 = −m1gL1 cosφ1 and U2 = −m2g(L1 cosφ1 + L2 cosφ2 ).  

Hence,  the  total  potential  energy  is  

  U = −(m1 +m2 )gL1 cosφ1 −m2gL2 cosφ2.   (11.32)  

 The  kinetic  energy  of  mass  1  is  given  by    

  T1 =12 m1L1

2 φ12.  

Note  that  the  velocity  of  mass  2  relative  to  an  inertial  frame  (e.g.,  one  attached  to  the  fixed  support)  is  the  vector  sum  of  the  velocity  of  mass  2  relative  to  mass  1  and  the  velocity  of  mass  1.  Thus,  the  square  of  the  velocity  of  mass  2  is  calculated  as  

  v22 = v2 ⋅

v2 = (v1 +v21) ⋅(

v1 +v21) = v1

2 + v212 + 2v1 ⋅

v21 = L12 φ1

2 + L22 φ2

2 + 2L1L2 φ1 φ2 cos(φ1 −φ2 ),  

and  T2 = 12 m2v2

2.  Thus,  the  total  kinetic  energy  is    

  T = 12 (m1 +m2 )L1

2 φ12 +m2L1L2 φ1 φ2 cos(φ1 −φ2 )+ 1

2 m2L22 φ2

2.   (11.33)  

To  simply  things,  let  us  consider  small  oscillations.  Thus,  the  generalized  coordinates  and  their  time  derivatives  will  be  small.  We  will  retain  terms  up  to  second  order  in  φi  and   φi .  Thus,   cosφi ≈1− 1

2φi2 .  Further,  the  terms   φi

2  and  φiφ j  in  the  kinetic  energy  are  already  second-­‐order  and  so   φ1 φ2 cos(φ1 −φ2 ) ≈ φ1 φ2.  Then  in  the  small-­‐oscillation  approximation,  the  potential  energy  is,  after  dropping  constant  terms,  

  U = 12 (m1 +m2 )gL1φ1

2 + 12 m2gL2φ2

2.   (11.34)  

The  kinetic  energy  becomes  

  T = 12 (m1 +m2 )L1

2 φ12 +m2L1L2 φ1 φ2 + 1

2 m2L22 φ2

2.   (11.35)  

Note  that  in  the  small-­‐oscillation  approximation,  the  kinetic  energy  is  a  homogeneous  quadratic  function  of  the  generalized  velocities,  i.e.,  every  term  is  a  product  of  two  

Page 10: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  10  

generalized  velocities  multiplied  by  a  constant.  Likewise,  the  potential  energy  is  a  homogeneous  quadratic  function  of  the  generalized  coordinates.    Applying  Lagrange’s  equations  gives  

 

ddt

∂L∂ φ1

= ∂L∂φ1

, i.e.,

(m1 +m2 )L12φ1 +m2L1L2

φ2 = −(m1 +m2 )gL1φ1.   (11.36)  

Also,    

 

ddt

∂L∂ φ1

= ∂L∂φ1

, i.e.,

m2L1L2φ1 +m2L2

2 φ2 = −m2gL2φ2.   (11.37)  

Note  that  because  T  and  U  are  homogeneous  quadratic  functions,  the  equations  of  motion  are  linear  and  so  can  be  solved  using  matrix  methods  and  other  techniques.  As  usual,  we  can  rewrite  the  equations  of  motion  in  matrix  form:  

  Mφ = −Kφ,   (11.38)  

where    

  M =(m1 +m2 )L1

2 m2L1L2m2L1L2 m2L2

2

⎝⎜⎜

⎠⎟⎟,   (11.39)  

  K =(m1 +m2 )gL1 0

0 m2gL2

⎝⎜⎜

⎠⎟⎟,   (11.40)  

and  

  φ =φ1φ2

⎝⎜⎜

⎠⎟⎟.   (11.41)  

Let  us  take  m1 = m2 = m  and   L1 = L2 = L.  This  gives  

  M = mL2 2 11 1

⎛⎝⎜

⎞⎠⎟

and K =2mgL 0

0 mgL

⎝⎜

⎠⎟ = mL2 2g / L 0

0 g / L

⎝⎜

⎠⎟ .   (11.42)  

For  non-­‐trivial  solutions  to  the  equations  of  motion  to  exist,  we  must  have  det(K −ω 2M) = 0 .  This  gives    

Page 11: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  11  

  mL22(ω 0

2 −ω 2 ) −ω 2

−ω 2 ω 02 −ω 2

= 0,  

where  ω 0 = g / L  is  the  frequency  for  a  simple  pendulum.  Solving  the  resulting  equation  gives  the  normal  frequencies  

  ω12 = (2 − 2)ω 0

2 and ω 22 = (2 + 2)ω 0

2,   (11.43)  

In-­class  Problem:  Fill  in  the  details  leading  to  Eq.  (11.43).  

To  find  the  eigenvector  matrix  a  corresponding  to  ω1 ,  we  set   (K −ω12M)a = 0 ,  which  gives  

 

mL2ω 02 2( 2 −1) 2 − 2

2 − 2 2 −1

⎝⎜⎜

⎠⎟⎟

a1a2

⎝⎜⎜

⎠⎟⎟= mL2ω 0

22 2 −1( ) 2 1− 2( )2 1− 2( ) 2 −1

⎜⎜⎜

⎟⎟⎟

a1a2

⎝⎜⎜

⎠⎟⎟

= mL2ω 02 2 −1( ) 2 − 2

− 2 1

⎝⎜⎜

⎠⎟⎟

a1a2

⎝⎜⎜

⎠⎟⎟= 0.

 

This  gives  the  equations  

  2a1 − 2a2 = 0 and − 2a1 + a2 = 0,  

which  are  the  same  equation  apart  from  a  constant  factor  of  − 2 .  Thus,   a2 = 2a1.  The  solution  for  the  first  normal  mode  is  therefore   φ = aeiω1t ,  or,    

 φ1

φ2

⎝⎜⎜

⎠⎟⎟= A1

12

⎝⎜

⎠⎟ cos ω1t −δ1( ). [First mode]   (11.44)  

The  two  masses  vibrate  in  phase  but  the  amplitude  of  the  lower  pendulum  is   2  times  that  of  the  upper  one.  For  the  second  normal  mode,  we  set   (K −ω 2

2M)a = 0 .  One  finds,  using  the  same  procedure  as  above,  that  

 φ1

φ2

⎝⎜⎜

⎠⎟⎟= A2

1− 2

⎝⎜

⎠⎟ cos ω 2t −δ 2( ). [Second mode]   (11.45)  

In  this  case,  the  pendulums  oscillate  out  of  phase,  with  the  lower  pendulum  again  having  an  amplitude  that  is   2  times  the  upper  one.  

In-­class  Problem:  Taylor,  Problem  11.16.  

 

Page 12: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  12  

5.  Lagrangian  Approach:  General  Case  

We  consider  a  holonomic  system  with  n  degrees  of  freedom,  so  that  there  are  n  generalized  coordinates   q1,q2,....,qn .  If  there  are  N  particles,  the  Cartesian  positions  

rα  are  functions  of  

the  generalized  coordinates:  

  rα = rα (q1,q2,....,qn ). α=1, 2, ...., N   (11.46)  

To  find  the  velocities,  we  take  the  time  derivative  of  the  positions  and  use  the  chain  rule.  For  example,  the  time  derivative  of  the   x -­‐coordinate  of  particle  α  is  given  by  

  xα = ∂xα

∂qiqi

i=1

n

∑ ,  

and    

  xα2 = ∂xα

∂qi

∂xα∂qjqi qj

j=1

n

∑i=1

n

∑ = Aijα qi qj

ij∑ .  

Similar  equations  can  be  written  down  for   yα2  and   zα

2 .  We  have  assumed  in  the  above  that  the  generalized  coordinates  are  not  explicit  functions  of  time,  i.e.,  ∂qi / ∂t = 0.The  total  kinetic  energy  is  given  by  

T = 1

2α∑ mαrα2 .  Using  the  expression  above,  we  can  therefore  

write  the  kinetic  energy  of  the  system  as    

  T = 1

2 Aij qi qjij∑ ,   (11.47)  

where   Aij  is  a  function  of  the  generalized  coordinates.  To  obtain  an  expression  for  the  potential  energy,  we  will  perform  a  Taylor  expansion  about  the  equilibrium  values  of  the  generalized  coordinates,  which  we  assume  are  zero:    

  qi = 0 (equilibrium) i = 1,2,...,n  

Hence,  we  have  

  U(q1,q2,..,qn ) =U(0)+∂U∂qi

⎛⎝⎜

⎞⎠⎟i

∑0

qi +∂2U

∂qi ∂qj

⎝⎜⎞

⎠⎟ 0qiqj

j∑

i∑ + ...   (11.48)  

U(0)  is  a  constant,  which  we  will  set  to  zero.  Since  the  derivatives  are  evaluated  at  the  equilibrium  position,  the  first  derivatives   (∂U / ∂qi )0=  0.  Further,  we  will  assume  that  the  oscillations  are  small  and  drop  terms  that  are  higher  than  those  quadratic  in  the  generalized  coordinates.  Hence,  the  potential  energy  becomes  

  U = 12 Kijqiqj

ij∑ ,   (11.49)  

Page 13: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  13  

where    

  Kij =∂2U

∂qi ∂qj

⎝⎜⎞

⎠⎟ 0.  

Assuming  that  the  equilibrium  positions  correspond  to  a  minimum  in  U  (we  are  looking  for  oscillations  after  all!)  ,  then  the  Kij  are  positive  numbers.  Also,  Kij  =  Kji  since  the  order  of  differentiation  is  immaterial.  

Going  back  to  the  kinetic  energy,  recall  that  the  Aij  parameters  are  functions  of  the  generalized  coordinates.  We  can  perform  a  Taylor  expansion  about  the  equilibrium  positions  for  these  functions  as  we  did  for  the  potential  energy.  The  first  term  in  the  expansion  is  a  constant  [ Aij (0) ]  and  the  other  terms  are  first  and  higher  orders  in  the  generalized  coordinates.  Since  we  are  assuming  that  the  qi  values  are  small,  the  associated  generalized  velocities   qi  will  also  be  small.  Thus,  the  product   qi qj  in  the  kinetic  energy  is  already  second-­‐order  in  the  small  generalized  velocities.  Since  we  are  keeping  terms  in  the  kinetic  energy  only  up  to  second  order,  we  must  keep  only  the  first  (constant)  term   Aij (0)in  the  Taylor  expansion  of  Aij.  Thus,  the  kinetic  energy  becomes  

  T = 1

2 Mij qi qjij∑ ,   (11.50)  

where  Mij = Aij (0)  is  a  positive  constant.  Note  that  in  the  small-­‐oscillation  approximation,  the  kinetic  energy  is  a  homogenous  quadratic  function  of  the  generalized  velocities  and  the  potential  energy  is  a  homogeneous  quadratic  function  of  the  generalized  coordinate,  which  guarantees  that  the  equations  of  motion  will  be  linear  and  solvable.  

Equations  of  Motion  

Lagrange’s  equations  for  the  general  case  are  

 

ddt

∂L∂ qi

= ∂L∂qi

. i = 1,2,...,n   (11.51)  

Before  proceeding,  we  note  that    

 

T = 12 Mij qi qj

ij∑ = 1

2 Mii qi2

i∑ + 1

2 Mij qi qji< j∑ + M ji qj qi

i< j∑⎡

⎣⎢

⎦⎥

= 12 Mii qi

2

i∑ + Mij qi qj

i< j∑ ,

  (11.52)  

where  the  last  equality  follows  because  Mij = M ji .  The  expression  for  U  in  Eq.  (11.49)  can  be  rewritten  in  a  similar  way.  Using  Eq.  (11.52),  we  obtain  

Page 14: Classical Mechanics Fall 2011 Chapter 11: Coupled ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · Classical Mechanics Fall 2011 Chapter 11: Coupled Oscillators and Normal

  14  

 

∂L∂ qi

= ∂T∂ qi

= Miiqi + Miji< j∑ qj = Mij

j∑ qj .  

Thus,    

 

ddt

∂L∂ qi

= Mijj∑ qj   (11.53)  

Using  similar  manipulations,  we  find  that    

 

∂L∂qi

= − ∂U∂qi

= − Kijj∑ qj .   (11.54)  

Thus,  Lagrange’s  equations  give  

 

Mijj∑ qj = − Kij

j∑ qj , i = 1,2,....,n   (11.55)  

or,  in  matrix  form,  

  Mq = −Kq,   (11.56)  

where  q  is  a  1× n column  matrix  containing  the  generalized  coordinates.  If  we  take  the  trial  solution  to  be  the  complex  column  matrix   z = aeiωt  with  q = Rez ,  the  matrix  form  of  the  equation  of  motion  becomes  the  eigenvalue  equation  

  (K −ω 2M)a = 0,   (11.57)  

for  which  solutions  exist  if    

  det(K −ω 2M) = 0.   (11.58)  

The  n  normal-­‐mode  frequencies  satisfy  Eq.  (11.58),  which  is  called  the  secular  equation  (or  characteristic  equation).  Substituting  each  normal  frequency  into  Eq.  (11.57)  gives  the  column  matrix  a  (the  eigenvector)  which  represents  the  solution  for  that  normal  mode.  The  general  solution  to  the  equation  of  motion  is  a  linear  superposition  of  all  the  normal-­‐mode  solutions.  If  we  take  the  eigenvectors  to  be  column  matrices  consisting  of  the  numerical  coefficients  from  solving  Eq.  (11.57),  then  the  ith  normal  coordinate  can  be  obtained  from    

  ξi = q ⋅ai ,   (11.59)  

where  ai  is  the  eigenvector  corresponding  to  the  ith  normal  mode.  In  Eq.  (11.59),  q  is  a  1× n  row  matrix  and  ai  is  a  n ×1  column  matrix.  Eq.  (11.59)  tells  us  that  the  ith  normal  coordinate  is  the  ith  projection  (“component”)  of  the  expansion  of  the  solutions  q  in  vector  space  consisting  of  the  orthogonal  eigenvectors  a.