classical mechanics fall 2011 chapter 8: two-body central...

13
1 Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force Problems 1. Center of Mass and Relative Coordinates The basic problem is that of two particles interacting via a central force such as the gravitational force or the electric force. For such forces, the potential energy (and force) is a function of the position of one particle relative to the other. First, let us use the positions of the two particles relative to the origin to describe the system. Each position has three components so the total number of generalized coordinates needed is six. The total kinetic energy is T = 1 2 m 1 r 1 2 + 1 2 m 2 r 2 2 . (8.1) The potential energy is U = U r 1 r 2 ( ) . (8.2) Defining the relative position r = r 1 r 2 , (8.3) the potential energy becomes U = U (r ). (8.4) Thus, if we use the relative position vector r (representing 3 generalized coordinates) to describe the system, the potential energy will be a function of r (magnitude of r ) only, which is very convenient. We now need another position vector to have the six coordinates needed for the complete description of the system. The center of mass is a characteristic of the system as a whole and its position is a possible choice for the other coordinate. What makes the CM position the best choice is that the total momentum of the system is O 1 2

Upload: others

Post on 01-Jan-2020

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  1  

Classical Mechanics Fall 2011

Chapter 8: Two-Body Central Force Problems

1.  Center  of  Mass  and  Relative  Coordinates  

The  basic  problem  is  that  of  two  particles  interacting  via  a  central  force  such  as  the  gravitational  force  or  the  electric  force.  For  such  forces,  the  potential  energy    (and  force)  is  a  function  of  the  position  of  one  particle  relative  to  the  other.  First,  let  us  use  the  positions  

of  the  two  particles  relative  to  the  origin  to  describe  the  system.  Each  position  has  three  components  so  the  total  number  of  generalized  coordinates  needed  is  six.  The  total  kinetic  energy  is    

  T = 1

2m1r1

2 + 12m2 r2

2.   (8.1)  

The  potential  energy  is    

  U =U r1 −r2( ).   (8.2)  

Defining  the  relative  position    

  r = r1 −

r2,   (8.3)  

the  potential  energy  becomes  

  U =U(r).   (8.4)  

Thus,  if  we  use  the  relative  position  vector   r  (representing  3  generalized  coordinates)  to  

describe  the  system,  the  potential  energy  will  be  a  function  of  r  (magnitude  of   r )  only,  

which  is  very  convenient.  We  now  need  another  position  vector  to  have  the  six  coordinates  needed  for  the  complete  description  of  the  system.  The  center  of  mass  is  a  characteristic  of  the  system  as  a  whole  and  its  position  is  a  possible  choice  for  the  other  coordinate.  What  makes  the  CM  position  the  best  choice  is  that  the  total  momentum  of  the  system  is  

O  

1  

2  

 

 

 

Page 2: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  2  

conserved  because  there  are  no  external  forces  acting  on  it.  Further,  the  total  momentum  of  the  system  is  equal  to  the  momentum  of  the  CM:  

  m1r1 +m2

r2 = MrCM ,   (8.5)  

where    

  M = m1 +m2   (8.6)  

The  conservation  of  momentum  therefore  implies  that    

  MrCM = constant,   (8.7)  

i.e.,  

  rCM = constant.   (8.8)  

Thus,  the  CM  moves  with  constant  velocity.  We  can  therefore  choose  an  inertial  reference  frame  in  which  the  CM  is  at  rest  to  analyze  the  motion.  This  frame  is  called  the  CM  frame.  In  the  CM  frame,  only  the  relative  motion  of  the  particles  is  important;  the  CM  is  at  rest.  To  see  this  more  quantitatively,  let  us  convert  from   (

r1,r2 )  to   (

rCM ,r )  coordinates.  Recall  that    

 

rCM ≡R = m1

r1 +m2r2

M,   (8.9)  

where  we  have  changed  notation  for  the  position  of  the  CM  for  convenience.  Thus,  

 

R = m1r1 +m2

r2M

.   (8.10)  

Further,    

  r = r1 −

r2.   (8.11)  

Eqs.  (8.10)  and  (8.11)  yield  

 

r1 =R + m2

Mr   (8.12)  

and  

 

r2 =R − m1

Mr .   (8.13)  

The  total  kinetic  energy  is  then  

Page 3: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  3  

 

T = 12m1r1

2 + 12m2 r2

2

= 12MR + 1

2m1m2

Mr 2.   (8.14)  

 

In-­‐class  Problem:  Prove  Eq.  (8.14).  

Defining  the  reduced  mass  

  µ = m1m2

m1 +m2

,   (8.15)  

we  have    

  T == 1

2MR + 1

2µr 2 .   (8.16)  

The  Lagrangian  for  the  two-­‐body  system  is  therefore  

  L == 1

2MR + 1

2µr 2 −U(r).   (8.17)  

The  Lagrange  equation  for  the  R  coordinate  (really  three  equations  –  one  for    for  each  component)  is    

 

ddt

MR( ) = 0,   (8.18)  

i.e.,  

  MR = constant.   (8.19)  

Thus,  the  momentum  of  the  CM  is  conserved  as  predicted  above.  This  also  means  that   R  is  

constant.  Shifting  to  the  CM  frame  (where   R =  0),  the  Lagrangian  becomes  

    L = 1

2µr 2 −U(r).   (8.20)  

Eq.  (8.20)  looks  like  the  Lagrangian  for  a  single  particle  with  mass  equal  to  the  reduced  mass.  Thus,  with  the  change  in  coordinates,  we  have  transformed  the  two-­‐body  problem  to  what  is  effectively  a  one-­‐body  problem.  This  greatly  simplifies  the  analysis.  Note  that  if   m1 m2,  µ ≈ m2.  In  this  case,  particle  1  hardly  moves  and  the  problem  is,  to  good  approximation,  a  one-­‐body  problem  in  which  particle  2  moves  relative  to  a  fixed  force  center  located  at  the  position  of  particle  1.  

 

Page 4: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  4  

2.  Angular  Momentum  

The  total  angular  momentum  of  the  system  is    

  L = r1 ×m1r1 +r2 ×m2

r2.   (8.21)  

Substituting  for   r1  and  

r2  using  Eqs.  (8.12)  and  (8.13),  along  with   R = 0  in  the  CM  frame,  

we  obtain  

  L = r × µr .   (8.22)  

In-­‐class  Problem:  Prove  Eq.  (8.22).  

Thus,  the  total  angular  momentum  is  that  of  a  single  body  with  mass   µ  at  position   r

relative  to  the  origin,  which  is  consistent  with  the  effective  one-­‐body  Lagrangian  we  obtained  above  in  the  CM  frame.  

Now,  for  an  isolated  two-­‐particle  system  in  which  the  particles  interact  via  a  central  force,  the  net  external  torque  on  the  system  is  zero  and  so  the  total  angular  momentum  is  constant.  The  angular  momentum  vector  therefore  has  a  constant  direction,  which  is  perpendicular  to  the  vectors  

r  and   r according  to  Eq.  (8.22).    Thus,  the  relative  position  

and  relative  velocity  vectors  are  always  in  the  same  plane.  It  follows  that  the  motion  must  lie  in  one  plane,  i.e.,  the  motion  is  two-­‐dimensional.  

 

3.  Two-­‐Dimensional  Equations  of  Motion  

Since  the  motion  is  confined  to  a  plane  and  the  potential  energy  is  only  a  function  of  the  relative  position  r,  we  will  use  2D  polar  coordinates  to  describe  the  motion.  The  radial  coordinate  r    is  equivalent  to  the  magnitude  of  the  relative  position  vector.  The  Lagrangian  is    

  L = 1

2µ( r2 + r2 φ 2 )−U(r).   (8.23)  

The  coordinate  φ  is  ignorable  ( L  is  independent  of  it),  therefore    

 

∂L∂ φ

= µr2 φ = lz = l = constant.   (8.24)  

Thus,  angular  momentum  is  conserved.    

The  Lagrange  equation  for  the  radial  coordinate  is    

  ddt

∂L∂ r

= ∂L∂r,   (8.25)  

Page 5: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  5  

which  gives  

  µr = µr φ 2 − dU(r)

dr.   (8.26)  

The  first  term  in  Eq.  (8.26)  is  the  centrifugal  force.  Since  the  angular  momentum  is  a  constant,  it  is  useful  to  substitute  for   φ  in  Eq.  (8.26)  using  Eq.  (8.24):  

  µr = l2

µr3− dU(r)

dr.   (8.27)  

Note  that  the  centrifugal  force  is  now  a  function  of  r  only.  (The  other  parameters  are  constants).  We  can  define  a  fictitious  centrifugal  potential  energy  Ucf (r) such  that  

  Fcf (r) = −dUcf (r)dr

.   (8.28)  

Using  Fcf (r) = l2 µr3 and  integrating  Eq.  (8.28),  we  obtain  

  Ucf =l2

2µr2,   (8.29)  

where  we  have  taken  Ucf (∞) = 0.  Let  us  now  define  an  effective  potential  energy  Ueff (r)  such  that    

  Ueff (r) =Ucf (r)+U(r) =l2

2µr2+U(r).   (8.30)  

With  this  definition,  Eq.  (8.27)  becomes  

  µr = −

dUeff

dr.   (8.31)  

Eq.  (8.31)  is  equivalent  to  the  equation  of  motion  of  a  single  particle  moving  with  a  potential  energy  Ueff .  

Conservation  of  Energy  

Multiplying  both  sides  of  Eq.  (8.31)  by   r  gives  

 

ddt

12µ r2⎛

⎝⎜⎞⎠⎟ = −

dUeff (r)dt

,   (8.32)  

i.e.,  

Page 6: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  6  

 

ddt

12µ r2 +Ueff (r)

⎛⎝⎜

⎞⎠⎟ = 0.   (8.33)  

Integrating  yields  

 

12µ r2 +Ueff (r) = E = constant.   (8.34)  

If  the  actual  potential  energy  U(r)derives  from  an  attractive  inverse-­‐square  force  (F(r)∝−1 r2 )  such  as  gravity,  then    

  U(r) = −αr

, (Attractive inverse square force)   (8.35)  

where  α is  a  constant.  In  this  case,    

  Ueff (r) =l2

2µr2− αr.   (8.36)  

For  large  r,  the  1/r  term  dominates;  therefore,  Ueff (r) is  negative.  For  small  r,  the  1/r2  (centrifugal)  term  dominates  and  Ueff (r) is  positive.  It  follows  that  Ueff (r)  must  have  a  minimum.  A  sketch  of  Ueff (r)  is  shown  in  the  figure.  When  E =Ueff ,   r = 0 ,  i.e.,  the  radial  velocity  is  instantaneously  zero.  The  values  of  r  at  which  this  occurs  are  the  radial  turning  points.  At  these  r  values,  the  equivalent  one  body  is  neither  moving  toward  or  away  from  

the  force  center.  If  the  total  energy  E  >  0,  then  there  is  only  one  such  r  value  (rmin)  where  radial  motion  toward  the  force  center  ceases.  Clearly  for   r > rmin ,  E >Ueff (r) ;  therefore,  the  body  can  move  to  infinity.  The  motion  is  unbounded.  For  E  <  0,  in  general  the  radial  motion  is  bounded  between   rmin  and   rmax .    An  example  of  such  bounded  motion  is  an  elliptical  orbit  of  a  planet  in  our  solar  system.  When  the  planet  is  closest  to  the  sun,   r = rmin  and  when  the  planet  is  farthest  from  the  sun,   r = rmax .  Note  that  a  bounded  orbit  does  not  necessarily  mean  that  the  orbit  is  closed.  

Question:  At  what  value  of  E  would  a  planet  move  in  a  circular  orbit?  

Ueff  

r  

 

rmin   rmax  

rmin   E>0  

E<0  

Page 7: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  7  

In-­‐class  Problem:  Taylor,  Problem  8.13  

4.  Equation  of  the  Orbit  

For  ease  of  discussion  and  visualization,  let  us  assume  that  µ ≈ m1 so  that  the  motion  of  particle  1  is  nearly  identical  to  that  of  the  equivalent  one  body.  Particle  2  is  nearly  stationary  (the  CM  is  very  close  to  its  position)  and  particle  1  orbits  particle  2.    

While  much  can  be  learned  by  examining  the  radial  equation,  to  find  the  trajectory  of  the  orbiting  particle,  we  need  to  consider  the  variation  of  the  angle  φ  as  well.  (Recall  that  the  motion  is  two-­‐dimensional,  so  only  the  two  polar  coordinates  are  needed  to  describe  the  orbit.)  The  trajectory  is   r(φ) ,  i.e.,  r  as  a  function  of  φ .  To  determine  the  trajectory,  we  go  back  to  the  radial  equation  of  motion:  

  µr = l2

µr3+ F(r),   (8.37)  

where  

  F(r) = − dU(r)dr

  (8.38)  

is  the  central  force  by  which  the  particles  interact.  We  will  substitute  

  u = 1r   (8.39)  

into  Eq.  (8.37)  and  also  rewrite  d dt  using  the  chain  rule:  

 

ddt

= ddφ

dφdt

= φ ddφ

= lµr2

ddφ

= lu2

µddφ.   (8.40)  

Thus,  

  r = dr

dt= lu

2

µddφ

1u

⎛⎝⎜

⎞⎠⎟ = − l

µdudφ.   (8.41)  

  r = d r

dt= lu

2

µddφ

− lµdudφ

⎛⎝⎜

⎞⎠⎟= − l

2u2

µ2d 2udφ 2

.   (8.42)  

Substituting  these  results  in  Eq.  (8.37)  gives  

  − l2u2

µd 2udφ 2

= l2u3

µ+ F(r),   (8.43)  

i.e.,  

Page 8: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  8  

  ′′u (φ) = −u(φ)− µl2u2

F(u).   (8.44)  

Solving  this  equation  for  a  given  central  force  F  will  give  the  trajectory   r(φ)  by  using  the  relation  u = 1/ r .  

5.  Orbits  for  an  Inverse  Square  Force  

For  an  attractive  inverse-­‐square  force,  

  F(r) = − αr2

= −αu2,   (8.45)  

where  α is  a  proportionality  constant.  Substituting  for  F(r)  in  Eq.  (8.44)  gives  

  ′′u (φ)+ u(φ) = αµl2.   (8.46)  

This  is  a  second-­‐order  linear  inhomogeneous  differential  equation  similar  to  that  of  a  forced  harmonic  oscillator  with  no  damping  and  a  constant  applied  force.  The  solution  to  the  homogenous  equation  is  sinusoidal  (like  the  harmonic  oscillator  with  ω = 1)  and  the  particular  solution  is  just  a  constant  up =αµ l2 .  Thus,  the  complete  solution  to  Eq.  (8.46)  is  

  u(φ) = Acos(φ −δ )+ αµl2,   (8.47)  

where  A  and  δ  are  constants  of  integration.  We  can  make  δ = 0  by  an  appropriate  choice  of  the  direction  corresponding  to  φ = 0 .  Recalling  that  r    =  1/u,  Eq.  (8.47)  becomes  

  r(φ) = c1+ ecosφ

,   (8.48)  

where   e = Al2 /αµ  is  a  positive  dimensionless  constant  and    

  c = l2

αµ   (8.49)  

is  also  a  positive  quantity  that  has  the  dimensions  of  length.  Eq.  (8.48)  is  the  equation  of  the  orbit  for  an  attractive  inverse-­‐square  central  force.  

In  Eq.  (8.48),  note  that  if  e  <  1,  there  is  an  upper  bound  and  a  lower  bound  on  the  values  of  r  because  −1≤ cosφ ≤1 .  We  have    

  rmin =c

1+ e and rmax =

c1− e

.   (8.50)  

Page 9: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  9  

Thus,  the  orbit  is  bounded  for  e  <  1.  For    e  =  0,  there  is  only  one  value  of  r,  i.e.,  r  =  c.  This  corresponds  to  a  circular  orbit.  For   e ≥1 ,  there  is  a  lower  bound  on  r  but  no  upper  bound  (note  that  r  cannot  be  negative).  Thus,  in  this  case,  the  orbit  is  unbounded.  The  values  of  rmin  and  rmax  can  also  be  found  by  using  energy  considerations.  Recall  that  radial  turning  points  occur  when   r = 0 .  Thus,  

  l2

2µr2+U(r) = E.  

Taking  U(r) = −α r for  the  potential  energy  corresponding  to  the  inverse  square  force  and  substituting  it  in  the  equation  above,  we  obtain  the  quadratic  equation  

          2µEr2 + 2µαr − l2 = 0,  

for  which  the  solutions  are    

  r± =α2E

−1± 1+ 2El2 µα 2( )⎡⎣⎢

⎤⎦⎥.   (8.51)  

We  see  that  if  −µα 2 2l2 < E < 0 ,  there  are  two  solutions,  corresponding  to  rmin  ( r+ )  and  rmax  ( r− ).  These  are  bounded  orbits  for  which  the  eccentricity  e  <  1.  For  E    >  0,  there  is  only  one  positive  solution,  corresponding  to  rmin.  Thus,  these  orbits  are  unbounded  and  the  eccentricity  e  >  1.  The  explicit  relationship  between  E  and  e  can  be  shown  by  adding  rmin  and  rmax  from  Eq.  (8.51)  and  equating  the  result  to  that  obtained  by  adding  rmin  and  rmax  from  Eq.  (8.50).  The  result  is    

  e = 1+ 2cEα

= 1+ 2El2

µα 2 .   (8.52)  

In-­‐class  Problem:  From  Eq.  (8.51),  show  that  the  radius  of  the  circular  orbit  for  a  given  l,  α ,  and  µ  is  r  =  c.    

The  Orbits  

The  general  equation  for  a  conic  section  is    

  r(φ) = λ1+ ecosφ

,   (8.53)  

The  parameter  e  is  called  the  eccentricity  and  it  determines  the  particular  conic  section.  The  parameter  λ  is  called  the  semi-­‐latus  rectum.  We  see  that  equation  for  a  conic  section  has  exactly  the  same  form  as  the  equation  of  the  orbit  for  an  inverse-­‐square  force,  with  λ = c.  It  follows  that  the  orbits  are  conic  sections.  The  table  below  gives  the  conic  section  for  a  given  value  of  e  and  the  corresponding  sign  of  the  total  energy  E.  

 

Page 10: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  10  

Eccentricity     Energy   Conic  section  0   <0,  minimum   circle  

0 < e <1   <0   ellipse  1   0   parabola  >1   >0   hyperbola  

 

Elliptical  Orbits  

To  show  that  e  <  1  in  Eq.  (8.48)  corresponds  to  an  ellipse,  one  can  convert  from  polar  coordinates  to  Cartesian  coordinates.  The  result  is    

  (x + d)2

a2+ y

2

b2= 1,   (8.54)  

where    

  a = c1− e2 , b = c

1− e2, and d = ea.   (8.55)  

Eqn.  (8.54)  is  the  equation  of  an  ellipse  with  semimajor  axis  a  and  semiminor  axis  b.  The  origin  is  at  O,  the  position  of  the  body  being  orbited  (e.g.,  the  sun),  which  is  also  the  

position  of  one  focus  of  the  ellipse.  The  position  of  the  orbiting  body  (e.g.,  a  planet)  when  it  is  closest  to  the  body  being  orbited  is  called  the  perigee.  If  the  body  being  orbited  is  the  sun,  it  is  called  the  perihelion.  At  this  point,  the  distance  between  the  two  bodies  is  rmin.  The  position  at  which  the  orbiting  body  is  farthest  away  is  called  the  apogee  (or  aphelion  in  the  case  of  the  sun).  The  distance  between  the  bodies  is  then  rmax.  It  is  easy  to  see  that    

  rmin = a(1− e) and rmax = a(1+ e).   (8.56)  

The  significance  of  the  eccentricity  e  can  be  seen  by  computing  the  ratio  of  the  semimajor  and  semiminor  axes  using  Eq.  (8.55):  

  ba= 1− e2 , or, e = 1− b

2

a2 .   (8.57)  

b  

a  

c    

d   O  

rmin  

Page 11: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  11  

Clearly,  as   e→1,  b  becomes  very  small  compared  to  a  and  the  ellipse  becomes  very  elongated.  As  e  approaches  zero,  the  ellipse  becomes  more  circular  and  is  precisely  a  circle  when  e  =  0.  The  fact  that  the  planets  describe  elliptical  orbits  about  the  sun  is,  of  course,  the  content  of  Kepler’s  first  law.  

Kepler’s  second  law,  which  states  that  the  line  joining  a  planet  and  the  sun  sweeps  out  equal  areas  in  equal  times,  is  the  result  of  the  conservation  of  angular  momentum  (you  proved  it  in  chapter  3).  The  rate  of  sweeping  out  area  is  dA / dt = l / 2µ,  where  l  is  the  constant  angular  momentum  and   µ  is  the  reduced  mass.  One  can  use  this  result  and  the  fact  that  the  area  of  an  ellipse  is  πab to  derive  Kepler’s  third  law,  which  relates  the  orbital  period  to  the  semimajor  axis  length:  

  τ 2 = 4π2µ

αa3.   (8.58)  

The  quantity  α is  the  proportionality  constant  in  the  generic  attractive  inverse-­‐square  force  law  F(r) = −α r2 .  For  the  gravitational  force,  α = Gm1m2 ,  so  that    

  τ 2 = 4π 2µGm1m2

a3.   (8.59)  

If  the  mass  of  the  sun  is  m1  =  Ms    and  the  mass  of  a  planet  is  m2,  then  the  equation  above  becomes,  to  an  excellent  approximation,  

  τ 2 = 4π 2

GMs

a3.   (8.60)  

Finally,  one  can  calculate  the  total  energy  of  an  object  in  an  elliptical  orbit  in  terms  of  the  orbit  parameters.  Using  Eq.  (8.51),  we  find  that  

  E = − α2a,   (8.61)  

where  2a  is  the  length  of  the  major  axis.  For  the  gravitational  force,  the  energy  of  an  object  in  an  elliptical  orbit  is  therefore  

  E = −Gm1m2

2a.   (8.62)  

In-­‐class  Problem:  Taylor,  problem  8.18.  

Unbounded  Orbits:  Parabolas  and  Hyperbolas  

As  seen  in  the  table  above,  when   e ≥1 ,  the  orbit  is  unbounded.  This  condition  corresponds  to  E ≥ 0 ,  i.e.,  if  the  total  energy  is  zero  or  positive,  the  orbit  is  unbounded.  For  e  =  1  and  E  =  0,  the  orbit  is  a  parabola.  Note  that  in  the  orbit  equation  Eq.  (8.48),  the  denominator  approaches  zero  and  r  approaches  infinity  when  φ → ±π .  One  can  easily  show  that  when  e  

Page 12: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  12  

=  1  is  inserted  into  the  orbit  equation  and  the  equation  is  converted  to  Cartesian  coordinates,  the  result  is   y2 = c2 − 2cx ,  which  is  the  equation  of  a  parabola.  

If  e  >  1  (and  E  >  0),  we  see  from  the  orbit  equation  that   r→∞ as  φ → ±φmax ,  where  φmax  satisfies  

  ecosφmax = −1. (hyperbolic orbit)   (8.63)  

Thus,  the  orbit  is  confined  to  the  range  of  φ  values  −φmax ≤φ ≤φmax.  Converting  the  orbit  equation  to  Cartesian  coordinates  yields  the  more  readily  recognized  equation  for  a  hyperbola:  

  (x −δ )2

γ 2 − y2

β 2 = 1,   (8.64)  

where  the  parameters  δ , β, and γ  depend  on  semi-­‐latus  rectum  c  and  the  eccentricity  e.  

6.  Orbital  Transfers  

To  send  probes  to  other  planets  and  to  asteroids,  one  needs  to  be  able  to  change  orbits.  The  probe,  while  on  Earth,  is  in  a  nearly  circular  orbit  about  the  sun.  To  send  it  to  Mars,  say,  you  will  need  to  transfer  the  probe  to  an  elliptical  orbit  (with  the  sun  at  one  focus)  that  intersects  Mars.  Before  the  probe  crashes  into  Mars,  you  will  have  to  transfer  it  to  an  orbit  around  Mars.  How  are  such  transfers  done?  The  spacecraft  fires  rockets  over  a  time  interval  to  change  its  speed  and  trajectory.  We  will  concentrate  on  the  simplest  case,  which  is  when  a  tangential  boost  is  delivered  at  the  perigee  or  the  apogee.    

When  a  tangential  boost  is  given  at  the  perigee  or  the  apogee,  since  there  is  no  radial  component,  the  radial  speed   r  will  remain  zero.  Therefore,  the  new  orbit  will  have  its  perigee  or  apogee  at  that  same  point.  In  other  words,  the  orientation  of  the  major  axis  does  not  change.  

Let  us  assume  the  space  probe  is  at  the  perigee.  Let  the  speed  just  before  firing  be   v1  and  that  just  after  firing  be   v2 .  The  thrust  factor  is  defined  as  

  λ = v2v1.   (8.65)  

Page 13: Classical Mechanics Fall 2011 Chapter 8: Two-Body Central ...faculty.chas.uni.edu/~Shand/Classical_Mechanics... · ! 1! Classical Mechanics Fall 2011 Chapter 8: Two-Body Central Force

  13  

Since  the  radial  distance  does  not  change  (assuming  the  firing  is  instantaneous),  then  the  angular  momenta  before  and  after  have  the  same  relationship  as  the  speeds,  i.e.,   l2 = λl1 .  It  follows  that    

  c2 = λ 2c1.   (8.66)  

Since  rmin  is  the  same  before  and  after,  we  have    

  c11+ e1

= c21+ e2

.   (8.67)  

 The  new  eccentricity  can  now  be  calculated  using  Eq.  (8.67).  One  finds  

  e2 = λ 2e1 + (λ2 −1).   (8.68)  

If  λ >1  (forward  thrust),  we  see  that   e2 > e1 .  The  new  orbit  is  more  elongated,  i.e.,  it  lies  entirely  outside  the  old  one  (see  diagram  on  previous  page).  The  major  axis  will  be  longer  and  hence  the  probe  has  greater  total  energy  in  the  new  orbit  (but  it  is  still  negative!).  If  λ <1  (backward  thrust),   e2 < e1.  The  orbit  is  less  elongated  and  lies  entirely  inside  the  old  one.  If  λ  is  smaller  still,  e2  becomes  negative.  This  means  the  new  orbit  has  its  apogee  at  the  firing  point  instead  of  the  perigee.  Of  course,  the  eccentricity  is  not  really  negative;  φ = 0  now  corresponds  to  the  apogee  and  φ = π  corresponds  to  the  perigee,  which  is  the  reverse  of  the  usual  scenario  and  so   e2  becomes  “negative.”    In  the  typical  transfer  orbit  problem,  one  is  asked  to  solve  for  λ .  You  can  do  this  by  determining   e1  and   e2  and  then  using  Eq.  (8.68)  to  solve  for  λ .  It  is  useful  to  note  that  transfer  orbit  problems  can  also  be  solved  by  using  the  conservation  of  mechanical  energy  and  Eq.  (8.62)  for  elliptical  orbits.  

In-­‐class  Problem:  Taylor,  problem  8.33.