You are on page 1of 143

Chapter 1 Electromagnetic Fields

Lecture 1 Maxwell’s equations


1.1 Maxwell’s equations and boundary conditions
Maxwell’s equations describe how the electric field and the magnetic field are generated,
and how they change in space and time. Lights are electromagnetic waves which obey
Maxwell’s equations. Maxwell’s equations have three major equivalent forms.
1) General form of Maxwell’s equations:
1.It is also called Maxwell’s equations in vacuum.
2.It actually applies to all cases, either in vacuum or in a medium. It is thus called the
general form of Maxwell’s equations.
t 
E 
0 

B 
E   
t 
  B  0

 E 
  B  0  J t   0 
  t 
Here  total   free   bound , J total  J free  J bound .
1
Definitions of the auxiliary fields:
D   0E  P 

H  B / 0  M ,  B  0 ( H  M )   Our text was wrong in the definition of
H.
Bound charge and bound current density:
 b    P  E – Electric field
 Refer to an electrodynamics
P  H – Magnetic field
Jb    M   textbook. E.g., Griffiths.
t  D – Electric displacement
B – Magnetic induction
We then come to 2) Maxwell’s equations in matter:
 – Electric charge density
1.It contains no more physics than the general form,
except that the charge density and the current density J – Electric current density
are decomposed into free and bound.  – Permittivity
2.It will be extensively used in our course.  – Permeability
 0 – Permittivity of vacuum
D  f 
  0 – Permeability of vacuum
B 
E   P – Electric polarization
t 
 We then drop the “f” subscripts. M – Magnetization
 B  0
D 
H  Jf  
t 
2
Material equations (Constitutive relations):
D  E

H   1B, ( B  H) 

Note:  and  are of essential importance in our course. They characterize the materials
and are something you can explore all through your life.
1.They are position dependent in an inhomogeneous medium.
2.They are tensors in an anisotropic medium.
3.They depend on the field strength (E and H) in a nonlinear medium.
We then come to 3) Maxwell’s equations in a homogeneous medium:
1.It is valid only in a homogeneous medium.
2.It takes the same general form, with the changes  0   , 0   ,  t   f , J t  J f .
f 
E 
 

B
  E   
t 
  B  0

 E  
  B   J f   
 t  
3
Lecture 2 Energy of the electromagnetic fields
Boundary conditions: Boundary conditions are the relations between the
electromagnetic fields on two sides of the interface that separates the two media.

Gauss’ theorem:    FdV   F  dS


V V

Stokes’ theorem:    F  dS   F  dl
S S

  B  0  n  (B 2  B1 )  0  B2 n  B1n
  D    n  (D  D )    D  D  
 2 1 2n 1n
 B
  E    n  (E2  E1 )  0  E2 t  E1t
  t
 D
  H  J   n  ( H 2  H1 )  K  H 2 t  H1t  K
 t
When  0, K 0, we have
BB22nn  BB11nn n – Unit normal to the surface
D
 D22nn  D D11nn  – Surface charge density

E E22tt  E
E11tt K – Surface current density
H
H22tt  H
H11tt
4
1.2 Poynting’s theorem and conservation laws
Electromagnetic waves carry energy and momentum.
Conservation of the energy of the electromagnetic fields requires:
Energy flowing in Increase of the energy of Work done by the
through the boundary = the electromagnetic fields + electromagnetic fields on the
surface of a volume inside the volume electric charges inside the volume

Considering energy transfer in a unit volume and in a unit time:


U W
  S   S
t t
U
S – Energy flow through a unit area in a unit time
U – Energy density of the electromagnetic fields W

W – Work done by the electromagnetic fields on the


electric charges in a unit volume

Question: What are the expressions for S, U and W ?

5
Work done by the electromagnetic fields on the charges in a unit volume in a unit
time:
W D  

   E  v  B   v  v  E  J  E  E     H   
t  t   D B
  J  E     E  H   E   H
B t t
   E  H   H     E  E     H    H   E     H 
t 

This suggests S  E  H, U e  E  D, and U m  H  B.


D
U e ( E, D)   E  D is a function of the final E and D fields.
0
D
Choosing a straight path of D in the final D direction, then U e ( E, D)   E  D.
0

Assuming linear medium, i.e.,  and  are independent of E and H, then


1
E  D  E   (E)  E  E  D  E  E  D   ( E  D).
2

U U – Energy density of the electromagnetic fields.


   S  J  E
t Linear response assumed. Otherwise
D B
1 U   E  D   H  B.
U  Ue  U m   E  D  B  H 0 0
2
S – Poynting vector. Energy flow through a unit area in a
S  E H
unit time. Instantaneous intensity of light.
Continuity equation. Poynting’s
theorem.
6
Lecture 3 Wave equations
1.3 Complex function formulism
Lights are most often described by sinusoidal time-varying fields:
a(t )  A cos(t   )
Complex representation:
a(t )  Re[ A e i (t  ) ]  Re[ Ae it ], where A  A e i .
~ i t
Let a (t )  Ae  A e
i (  t  )
represent the field, and we always mean its real part.
Complex representations have no problems with linear mathematical operations:
~ ~ ~ ~ ~ ~
E3  E1  E2  Re[ E3 ]  Re[ E1  E2 ]  E3  E1  E2 .
1 ~ 1 ~~
It does not work for the product of fields: U e  ED 
 U e  ED
2 2

The correct way is: U e 


2  2
EE
2
 
1  1 ~ ~* 1 ~ ~ * 
DD  

This is often used in nonlinear optics.

7
Time averaging of two sinusoidal products with the same frequency:
a (t )  A cos(t   ), b(t )  B cos(t   )
~
a~(t )  Aeit  A ei (t  ) , b (t )  Beit  B ei (t   )

a ( t ) b( t ) 
2
 2

a a  b b
1 ~ ~* 1 ~ ~*
  
1 ~ ~ ~ ~* ~* ~ ~* ~*
4
ab  ab  a b  a b 

4

1 ~ ~ * ~* ~ 1
2
 ~ 1
2
~
a b  a b  Re[a~(t )b * (t )]  Re[a~* (t )b (t )]

Examples:
Time-averaged Poynting vector and the energy density:
1 ~ ~
S  Re[E  H * ]
2
1 ~ ~ ~ ~
U  Re[E  D*  B  H * ]
4
Conventions:
We then drop the tidal sign, but we need to be careful in
1)multiplying fields, and
2)multiplying a field with a scalar or tensor which are possibly complex numbers.

8
1.4 Wave equations and monochromatic plane waves
Each of the field vectors of light oscillates in space and time, which satisfies a wave
equation. Assume the medium is isotropic ( and  are scalars), linear ( and  are
independent of fields), and there is no free charge or free current ( =0, J =0).
B   
E    1   
t        E     H  0 
   t 
B  H 
  1  E2

       E    2  0 
H 
D     t 
   2E 
t     H   2  0 
t t  
D  E 
 


1  1 1
     E        E       E ,      E     E   2E 
    

 2E 
 E   2   log      E     E  0
2

t 
  D  0    (E)    E  E   

 2E
 E   2   log      E   E   log    0.
2
The Maxwell’s equations and the
t materials equations are mathematically
H2
symmetric in the exchanges of
2H   2   log      H    H   log    0. E  −H, B  D, and −
t
9
Further suppose the medium is homogeneous ( and  does not change in space under a
translational shift), the wave equations are
 2E
 E   2  0
2

t A partial, linear, second order,


 2H homogeneous differential equation
 H   2  0
2

t

Solution to the wave equations: Plane waves Wavelength

  Aei (t k r ) (e i ( k r t ) is also good.)


Amplitude : A
Phase :   t  k  r
Angular frequecy :   2  2 / 
2  Period
Wave vector : k     
 v
1 c
Phase velocity : v  
 n
Index of refraction : n   /  0  0

10
Vector nature of the fields of electromagnetic waves (in an isotropic medium):
E  u1E0ei (t kr )
H  u 2 H 0ei (t  kr )
  E  0  k  u1  0  
  Transverse waves 
  H  0  k  u2  0 

B kˆ  u1  u 2 
E   
t kE0  H 0  H 0  E0  /  
u1 , u 2 , kˆ are mutually orthogonal, and u1  u 2  kˆ .

E and H are in phase if  and  are real.
In an anisotropic medium ( and  are tensors), only D and B are perpendicular to k.

Time-averaged Poynting vector (intensity of light) and energy density:


1 1  1 1
E0 kˆ   E0  v  kˆ , I   E0  v
2 2 2
S  Re[E  H * ] 
2 2  2 2

U 
1
4
1
4
 2 2 1

Re[E  D*  B  H * ]   E0   H 0   E0
2
2

1 2
Ue  U m   E0
4 11
Lecture 4 Propagation of a laser pulse
1.5 Propagation of a laser pulse; group velocity
Intense ultra-short laser pulses (t ~ 10-15 second, femto-second) are extremely important
in exploring the dynamic structure of atoms and molecules, and their interaction with light.

Bandwidth of a laser pulse:


A laser pulse can be treated as a sum of many plane
waves with different colors. The pulse has a
bandwidth in frequency domain, which satisfies the
uncertainty rule: t  1.
In the language of Fourier transform, at a fixed
point in space
1 
E (t )   E ( ) exp it  d
2 

2
E (t ) dt is the energy in dt.
1 
E ( )   E (t ) exp  it  dt
2  
2
E ( ) d is the energy in d.
0

12
Example: A Gaussian laser pulse

 2 ln 2t 2 
E (t )  E0 exp  2
 i  
0 ,
t
 T 
T is the pulse duration at the fwhm intensity.
1   2 ln 2t 2 
E ( )   0  T 2
E exp   i     
0  dt
t
2 
1   2 ln 2  iT 2       2 T 2      2 

2 
E 0 exp 


T 2



t 
4 ln 2
0




8 ln 2
0
 dt

T

E0  2 ln 2   0  2  
 exp  
  4 ln 2 / T   2 ln 2 / T
2
2
2

The Fourier transform of a Gaussian


2 ln 2 E0  2 ln 2   0  2  pulse is again a Gaussian distribution.
 exp 
  2  All E() are in phase.
This is a transform-limited pulse.

4 ln 2
 is the fwhm bandwidth.
T
T  4 ln 2  2.77 in intensity, t  5.55  2 in amplitude distributi on.
13
Lecture 5 Group velocity
Group velocity: A laser pulse normally has a carrier and an envelope. Group velocity is

the velocity
1 at which the envelope of the laser pulse travels.
E ( x, t ) 
2   E ( ) exp{i[t  k ( ) x ]}d

1

2   E ( ) exp{i[(   )t  (k  k ) x]}d  exp[i( t  k x)]

0 0 0 0

dk 1 d 2k
k  k0  (  0 )  (  0 ) 2  . For small  ,
d 0 2 d 2 0 t =0
vg
1   dk 
  d  exp[i (0t  k0 x )]
E ( x, t )   E ( ) exp  i (   )
0  t  x

vp
2   d 0 
 1 
  dk 
 f x  t   exp[i (0t  k0 x )]
  d   
 0 

 Envelope  Carrier t =t

  dk 
1

Group velocity : vg   
  d   
 0

 0 c
 Phase velocity : v p  
 k0 n(0 )
14
More about group velocity:
  dn 
v g  v p for normal dispersion   0 .
  d 
1
 dk    dn 
vg    v g  v p for anomalousdispersion   0 .
 d    d 
c c 
  .  dn
dn dn v g  c when n    1.
n  n d 
d d 
 d nk
Here v g is not the pulse velocity any more because of higher order .
 d n

Group velocity dispersion:


1
E ( x, t ) 
2   E ( ) exp{i[(   )t  (k  k ) x ]}d

0 0  exp[i (0t  k0 x )]

dk 1 d 2k 1 d 3k
k  k0  (  0 )  (  0 ) 
2
(  0 )3 .
d 0 2 d 2 0
6 d 3
0

d 2k 3 d 2 n
• Group velocity dispersion  broadens or compresses a laser pulse.
d 2 2c 2 d2
d 3k
• distorts a laser pulse.
d 3

15
Lecture 6 Dispersion
1.6 Dispersion and the Sellmeier equations
 non  magnetic
  ( )
n   n( )  . How do we get n ()?
 0 0 0 0

Electric polarization: The electric dipole moment per unit volume induced by an
external electric field.
For an isotropic and linear material, P  D   0E  (   0 )E
n( )  n( )   ( )  D  P  x (t )

Atom = electron cloud + nucleus. How is an atom polarized ?


Restoring force: F   k E x
E
Natural (resonant) frequency: 0  k E me
+
External force: FE  qe E (t )  qe E0 exp(it )
dx
Damping force:   me (does negative work)
dt
dx d 2x
Newton’s second law of motion: qe E0 exp(it )  me0 x   me  me 2
2

dt dt
16
dx d 2x
qe E0 exp(it )  me x   me
2
0  me 2
dt dt
qe E0 / me qe / me
Solution: x(t )  x0 exp(it )  exp(i t )  E (t )
0    i
2 2
0    i
2 2

x0 is frequency dependent  Electric polarization (and thus  and n ) is frequency


dependent  n().

Electric polarization (= dipole moment density):


Nqe2 / me
P(t )  Nqe x(t )  2 E (t ) 
0   2  i
P(t ) Nqe2 / me
  0   0  2
E (t ) 0   2  i

Dispersion equation:

 Nqe2  1 
n ( )   1 
2
 2  n2  1  N
0  0 me  0    i 
2

17
Quantum theory: 0 is the transition frequency.
Nqe2 fj
For a material with several transition frequencies: n ( )  1 
2

 0 me
 2
  2  i j 
Oscillator strength:  f j  1
j 0j

Normal dispersion: n increases with frequency.


Anomalous dispersion: n decreases with frequency.

n  n'in"
 2 2  Re (n)
exp(ikx)  exp  i n' x  n' ' x 
   

n'  Phase velocity


n"  Absorption (or amplification) -Im (n)

18
Sellmeier equation: An empirical relationship between refractive index n and wavelength
 for a particular transparent medium:
Nqe2 fj
Refer to n ( )  1 
2

 0 me

j
2
   i j
2
.
0j

• Sellmeier equations work fine when the wavelength range of interests is far from the
absorption of the material.
dn d 2 n
• Beauty of Sellmeier equations: n( ), , ,  are obtained analytically.
d d
2

• Sellmeier equations are extremely helpful in designing various optics. Examples: 1)


Control the polarization of lasers. 2) Control the phase and pulse duration of ultra-
short laser pulses. 3) Phase-match in nonlinear optical processes.

Example: BK7 glass


Coefficient Value
B1 1.03961212

B2 2.31792344×10−1

B3 1.01046945

C1 6.00069867×10−3 μm2

C2 2.00179144×10−2 μm2

C3 1.03560653×102 μm2
19
Chapter 2 Propagation of Laser Beams
Lecture 1 Gaussian laser beams
2.1 Scalar wave equation
Lasers are coherent electromagnetic waves. Control of laser beams is the main topic of
this course. Laser beams usually have a concentrated profile in the transverse direction,
which diverges slowly when propagating. The beam profiles in space can be obtained by
solving the wave equation.
In an isotropic, linear  2E
 E   2   log      E   E   log    0
2
medium with no free charge: t

Suppose within one wavelength the fractional change of  and  are small (slow-varying
refractive index),
Taylor series in dyadic algebra :
 1
  log   λ   log   λλ :  log     1    log   1, 2 2 log   1.
 2
E  2E E E E
Also  E ~ 2 ,  2 ~ 2 ,   E ~ , E ~ .
2

 t   
 2E
We then have  E   2  0.
2

t
20
Assume any Cartesian component of the electric field varies as
 ( x, y, z , t )  E ( x, y, z )e it , then
 2 n 2 (r )
 E  K (r ) E  0, with K (r )    
2 2 2 2
2
.
c
Helmholtz equation

Lens-like (quadratic index) media:


Consider a cylindrically symmetric lens-like medium
 k  k n n n k
n 2 (r )  n02 1  2 r 2 , with 2  2 , k  0 , k 2  2 , 2 r 2  1.
 k  k n0 c c k

Significances of lens-like media:


1) k2 (or n2) describes the inhomogeneity of the
medium. k2 =0 returns to homogeneous medium.
2)Lens-like media simulate the behavior of many n(r)
optics, including lenses and spherical mirrors.
3)Lens-like media are an approximation of many
optical fibers.

21
Solving the Helmholtz equation:  2 E  K 2 (r ) E  0.
2 2  2 2  k2 2 
K (r )  2 n (r )  2 n0 1  r   k 2  kk 2 r 2 .
2

c c  k 
Suppose the light mainly propagates in the z direction, let E ( x, y, z )   ( x, y, z )e ikz
Suppose the solution of the field is cylindrically symmetric, then
1     1 2 2 2 1  2
 
2
r      
r r  r  r 2  2 z 2 r 2 r r z 2
 2 1  2 
 2     
 2   ( x, y, z )e ikz   k 2  kk 2 r 2   ( x, y , z )e ikz  0
 r r r z 
 2 1   
 2     ' '2ik 'kk2 r 2  0 (with  '  )
 r r r   z
 2 1    Suppose ' '  k ' , ' does not change much in one 
 2    2ik 'kk 2 r 2  0  
 r r r   wavelength. Slow - varying amplitude approximat ion. 
  k 2

Let  be in the form   E0 exp 
 i P ( z )  r   , then
  2 q ( z )  Will pick up
later.
k
2
k 1  1  2  1  k
ar 2  b  0 i 
   r 2  2i  k 2 r 2  '2kP'kk2 r 2  0      '  0, and P'   
2

q q q  q   q  k q



22
2.2 Gaussian beams in a homogeneous medium

Complex beam parameter. q is a constant.


2
1 1
    '  0  q( z )  z  q0 0
q q
i i  z
P'    P( z )  i ln1    the integration constant is unimportant.
q z  q0  q0 
  k 2 
   z kr 2  
  E0 exp i  P( z )  r    E0 exp i  i ln1    
  2 q ( z )   
   q0  2 ( z  q )
0 
w02 n
Let q0 be purely imaginary (so that z =0 has the most intensity), and q0  i .

      
      
  z kr 2

  E  ikr 2
  E0 exp  i  i ln 1   0
exp  
   w0 n  
2
w n  
2 z   w0 n  
2
 i  2 z  i 0    1  i w 2 n 2
  z  i  
       
       
 0

 
 
E0   z     r2 ikr 2 
 exp i arctan  2  exp   2 
 z 
2
 w n
 0   2   z 
2
   w0 n   
2

1   2   w0 1   w 2 n   2 z 1   z   
 w0 n     0       
23
Fundamental Gaussian laser beams:

  w0  Beam waist,
w0  r2  kr 2  
E ( x , y , z )  E0 exp 2  exp i kz   ( z )   w( z )  Beam width, spot size
w( z )  w ( z )    2 R ( z ) 
z0  Rayleigh range ( z R )
  z  2    z 2  w02 n
w ( z )  w0 1   2    w0 1    , z0 
2 2 2 b  2 z0  Confocal parameter
  w0 n     z0    R ( z )  Radius of curvature
  w2 n  2   2
  ( z )  Phase delay, Gouy phase
 z 
R ( z )  z 1   0

   z 1   z  
0
 w( z ) 
   z         lim arctan 
z 
 z 
 z   z  
 ( z )  arctan 2   arctan    Beam divergence
 w0 n   z0  w0n

24
Lecture 2 Geometry of Gaussian laser beams

Geometry of the fundamental Gaussian laser beams:

1. Beam waist w0. The minimum spot size. The geometry of the beam is uniquely
determined by the size and location of w0 (suppose  and n are given).
2
 z 
2. Beam width w( z )  w0 1   2  , where the radial intensity drops to 1/e2=0.135.
 w0 n 

Also called spot size. Beam area =  w2(z). Hyperbolically depends on z. The
hyperbola can be thought as the boundary of the beam.
w02 n
3. Rayleigh range z R   , where the beam area is doubled compared to that at the
beam waist, so the intensity is halved. Confocal parameter b=2zR can be thought as the
focal range along the beam direction, where the intensity does not drop too much. The
Rayleigh range is proportional to w02, indicating that smaller beam diffracts faster.
 z 
 ( z )  arctan 2 .
4. Gouy phase  w0 n  Phase delay compared to a plane wave. The total

phase delay from –to  is , among which /2 occurs in the focal range.

25
  w2 n  2 
5. Radius of curvature R( z )  z 1   0  . The radius of curvature of the wave front.
  z  
Refer to a spherical wave at a far point:


exp  ikR   exp  ik x  y  z
2 2 2
 


r 2 
z 

 exp  ikz 1  2  exp  ikz  i
kr 2 
2 R

   
R(z) is infinity at z =0 and z =, where the wave fronts are flat. R(z) z when z is large,
where the beam is close to a spherical wave starting from the focal point.
 w( z )  
5. Beam divergence   lim arctan  . Half apex angle of the hyperbolic beam
z 
 z  w0 n
boundary. Inversely proportional to w0, indicating that smaller beam diffracts faster.

26
2.4 High order Gaussian beam modes in homogeneous medium

If we are not limited to   0 , the Helmholtz equation  2 E  k 2 (r )E  0 in a

homogeneous medium has the solution

w0  2x   2 y   x2  y2  k ( x2  y 2 ) 
El , m ( x , y , z )  E 0 H l  H m   exp 2  i kz   l  m  1 ( z )  
w( z )  w( z )   w( z ) 
 
 w ( z )  2 R ( z ) 
  z  2 
w 2 ( z )  w02 1   2   Hermite polynomials
  w0 n  
  w 2 n  2 
R ( z )  z 1   0  
  z  
 z 
 ( z )  arctan 2 
 w0 n 

27
Hermite-Gaussian laser modes Laguerre-Gaussian laser modes

28
Elliptical Gaussian laser beams:

w0 x w0 y   x 2 y2   kx 2 ky 2  
E ( x , y , z )  E0 exp  2  2   i kz   ( z )   
wx ( z ) wy ( z ) w
  x ( z ) w y ( z )   2 R x ( z ) 2 R y ( z )  
   ( z  z )  2 
wx2 ( z )  w02x 1   x
 
  w 0
2
x n  
 (z  z ) 2 
 y 
wy2 ( z )  w02y 1    
   w 2
0y n  
  w2 n  2 
Rx ( z )  ( z  z x )1   0x
 
   ( z  z )
x  

  w2 n  2  Elliptical Gaussian laser beams
 0y 
R y ( z )  ( z  z y )1     are often seen in the lab.
   ( z  z y )  

1   (z  zx )  1 (z  zy ) 
 ( z )  arctan    arctan  
2  w 2
0x n  2  w 2
0y n 

29
30
Lecture 3 Ray equation
The electromagnetic wave in space is given by
E(r, t )  E0 (r ) exp it   (r ) (r)= const

Suppose the fractional change of n (and thus E0) is small n(r)


in one wavelength, then P
dr
2 P P 2 dr O
Phase :  (r )  O n(r )ds  O  (r )  dr   (r )   n(r ) ds .

P dr
Optical path length :  (r )   n(r )ds,  (r )  n(r ) .
O ds
Ray dr is defined to be along ,and|dr|=ds. Here s is the distance traveled by the wave.

d  dr  d  dr    (r ) 
n ( r )     ( r )        ( r )   n (r )     ( r )   f    f 
1
 f
2

ds  ds  ds  ds    2
2
  ( r ) n 2 ( r )
   n(r ) d  dr 
2 n (r ) 2 n (r ) Compare to  m   (V )
dt  dt 
d  dr  Geometrical optics is a simplification to wave
Ray equation: n (r )  n(r ) optics, essentially like classical mechanics to

ds  
ds  quantum mechanics.
31
Application of the ray equation
Paraxial rays in a medium with n = n(x). Suppose the incident ray is in the x-z plan.
x
d  dr  
n (r )   n ( r ) 
ds  ds  
Paraxial rays along z  ds  dz  z

d  dx  dn d 2 x 1 dn
  n( x )    2 
dz  dz  dx dz n dx dT
 0,
dn
0
dx dx

Example 1: Optical mirages. Inferior mirage


(road, desert)
n  n0 (1  x) (x  1)
d 2 x 1 dn 1 2 dT dn
2
    x  z  bz  c  0, 0
dz n dx 2 dx dx

b, c are determined by initial conditions


(e.g., the injection height and Superior mirage
(sea, airplane)
direction).

32
Example 2: Optical fiber with quadratic index media.
 
n 2 (r )  n02 1   2 r 2 , ( 2 r 2  1,   k 2 k )

Let' s consider rays in the x-z plane, n 2 ( x)  n02 1   2 x 2 . 
d 2 x 1 dn
2
   2 x
dz n dx
 x( z )  a cos z  b sin z  A sin  z  B 

 x ( 0)  a
Ray matrix of a quadratic index medium:
x( z )  a cos z  b sin z   x' ( z )   a sin z  b cos z
 x' (0)  b

 x( z )   cos z  1 sin z  x(0) 
     
  Will pick up
 x' ( z )     sin z cos z  x' (0)  later.

33
*(Reading) Lagrangian optics:
Lagrangian optics is a branch of geometrical optics which deals with the propagation of
rays in inhomogeneous media.
Fermat’s principle (the principle of least time):
2 2
P P  dx   dy  P
  n( x, y, z )ds    n( x, y, z ) 1       dz    n( x, y, z ) 1  x'2  y '2 dz  0
S S
 dz   dz  S

Compare to
Optical Lagrangian : L( x, y, x' , y ' , z )  n( x, y, z ) 1  x'2  y '2 t2
   L(qk , q k , t )dt  0.

t1
d L L d L L
Optical Lagrangian equations :  , 
 dz  x '  x dz y ' y

d L L  
  d nx' 2 n 
dz x' x    1  x ' 2
 y '
2 dz 1  x'2  y '2 x  d  dx  n
L  n( x, y , z ) 1  x '  y ' 
2
  n  
 ds  ds  x
ds
1  x '2  y ' 2  
dz 
d  dr 
  n (r )   n(r ) Optical Lagrangian equations are
ds  ds  equivalent to the ray equation.
34
Lecture 4 Ray matrices
Ray matrices of optical elements:
The propagation of a ray through many optical elements can be described by using ray
matrices.
dr
A ray at position z is uniquely characterized by its height r(z) and its slope r ' ( z )  .
dz
 r ( z) 
We therefore use a column matrix (a vector) to represent the ray: ~ r ( z )   .
 r ' ( z ) 
An optical element transfers an input ray vector into an output ray vector.
  f   f 
2r  f ( r ,
1 1r ' )  f ( 0,0)  
 r  1  r '  r1 ' 
 r  r1'
r2'
  1 0  1 0 r1 r2

r '  g (r , r ' )  g (0,0)   g  r   g  r ' 
2 1 1  r  1  r '  1
  1 0  1 0

In paraxial optics the optical element is represented by a 22 matrix:

 r2   A B  r1  Note that there may be different


      definitions of ray vectors, e.g. (nr', r). The
 r '2   C D  r '1  ray matrices will accordingly be different.

35
Examples of ray matrices.
I). A thin lens with focal length f .
r2  r1   1 0  r 
  r2   1  1 
r  
r '2  r '1  1   r '2    f 1  r ' 
f   1 

II). Refraction

r2  r1 
 r2'
n2 sin r '2  n1 sin r '1  n2 r '2  n1r '1 
r1'
r1 r2
 r2   1 n0  r1 
   0 1  
r '   r ' 
 2  n2  1  n1 n2

36
Ray matrices of some common optical elements and media:

1 d 
 
0 1 

 1 0
 1 
 1
 f 

1 0
 n1 
0 
 n2 

37
Note that there may
 1 0 be different sign
 n2  n1 n1  conventions for R
  and for the z axis
 n2 R n2  after reflection.

1 0
2 
 1
R 

    k 
 cos k 2 l  k
sin  2 l  
  k 
  k2  k 
 
  k 2 sin  k 2 l   k2  
cos l 
 k  k  k 
     

38
Ray matrices of combined optical elements:
Example: Thin lens combination. Two lens separated by d.
 A B   1
1 0 1 d  1 0
      1 
1  0 1   1
 C D   f 2    f1 
 d 
 1 d 
f1
 
 1 1 d d 

 f   1  d
 1 f 2 f1 f 2 f 2 

Significance of the ray matrix of a lens-like media:


    k2  
 cos k2 l  k
sin l  
 A B   k 
      k2  k 

 C D    k2 sin k2 l   k2  
cos l 
 k  k  k 
     

1) Free propagation, lenses, and spherical mirrors (cases 1, 2, and 5) are equivalent to a
specific lens-like medium. E.g, a lens corresponds to l0, and k2=k/fl.
2) Single refraction (cases 3 and 4) may be equivalent to a combination of two
successive lens-like media.
39
Lecture 5 The ABCD law
2.3 Fundamental Gaussian beams in a lens-like medium
Fundamental Gaussian beams in a lens-like medium: E ( x, y, z )   ( x, y, z )e  ikz
2
  k   1   1  k2 i
  E0 exp  i  P( z )  r 2  ,     '  0, and P'   .
  2 q( z )   q q k q
2
 1   1  k2  dx  1 k  k 
    '  0,   2  arctan x     2 tan  2 z  c    tan  z  c .
q q k  x 1  q( z ) k  k 
1
Considering   tan  z1  c  , then at z2  z1  l ,
q( z1 )
1 tan  z1  c   tan l
  tan  z2  c    tan  z1  c   l   
q( z2 ) 1  tan  z1  c  tan l
1
cosl    sin l
q( z1 ) cosl  q( z1 )   1 sin l Aq( z1 )  B
  q( z2 )  
1   sin l  q ( z )  cos l Cq( z1 )  D
 1 sin l   cosl 1
q( z1 )
Aq1  B
 q2 
Cq1  D

40
Physical significance of the complex beam parameter q(z):
1 1 
1) Define   i for a beam in a lens-like medium.
q ( z ) R ( z ) n0 w 2 ( z )
  kr 2    r2   kr 2 
Then   exp i     exp  2  exp  i 
  2 q ( z )   w ( z)   2R( z ) 

• The real part of 1/q(z) specifies the radius of curvature of the wave front R(z).
• The imaginary part of 1/q(z) specifies the spot size of the beam w(z).

nw02
2) In a homogeneous medium, q( z )  q0  z  i z

• The real part of q(z) specifies the location of the beam waist z=0.
• The imaginary part of q(z) specifies the beam waist w0.

41
Another derivation for the transformation of the complex beam parameter:
Ray equation in a lens-like medium. Consider rays in an r-z plane.
 k  k
n 2 ( r )  n02 1  2 r 2 , ( with 2 r 2  1).
 k  k
d  dr  ds  dz d
 dr  dn d 2 r k2 r" k 2
 n ( r )    n ( r )   n ( r )    2
 r  0   0
ds  ds  dz  dz  dr dz k r k
2
 1   1  k2  
   '  0. q ( z ) transforms in space 
 r / r'   r / r'  k  
  r
2 using the same rule as that for  Aq1  B
 1   1  k2
    '  0
 r' 
  q2 
 Cq1  D
q q k  

 r2   A B  r1   r  A r / r '  1  B 
          
 r '2   C D  r '1   r ' 2 C  r / r '  2  D 

The ABCD law: The transformation of the complex beam parameter q(z) through an
optical system can be described by:
Aq1  B 1 C  D / q1 A B
q2  or  , where   is the ray matrix of the optical system.
Cq1  D q2 A  B / q1  C D 

42
Applications of the ABCD law:
Importance of the ABCD law: The law enables us to trace the Gaussian beam parameter
q(z) through a sequence of optical elements. The beam spot size w(z), the beam radius
R(z), the beam waist w0 and its location can be obtained at any point along the beam.
 1 0
I) Thin lens: M   1 
 1
 f 
Aq1  B 1 1 1
q2      w2  w1
Cq1  D q2 q1 f  
 1 1 1 Compare to the lens equation
1 1   R   1 1 1
 
 i R1 f
q ( z ) R ( z ) nw2 ( z )   2 so si f

II) Focusing a Gaussian laser beam:


A Gaussian laser beam with waist size w01 is incident at its waist on a thin lens with focal
length f. Please find the location and size of the waist of the output beam.

w03=?

l =?
43
Planes 1, 2, 3 are defined in the figure. We want to find w03 and l.
Strategy: We find q2 first. The real part of q2 then specifies the location l of the output
beam waist. The imaginary part of q2 specifies the beam waist w03.
1 1    
 i   i 2  
q1 R1 nw01 2
nw01  1 1  
   i
1 1 1  q2 f nw01 2

  
q2 q1 f  
nw03 
2
q2  i  l
 
 f
 l  2
 f   1   f  

f 1 i 
2    nw2 
nw03
2
1  nw  01 
i l    01 
 
 1    
2
w01
 i f  w 
f nw01
2 1   2 
 
03
2 2
 nw 01    nw 01 
 1   
  f  
Please note that to find w03 and l it is not necessary to calculate q3 (as our textbook does).
They are already embedded in q2. This problem also applies to the focusing of a laser
beam far from its beam waist, since q1 is again purely imaginary.
44
II) Eigen mode of Gaussian beams in a quadratic-index medium:
Let us try to find a Gaussian laser beam whose beam parameter q(z) does not change in a
quadratic-index medium specified by k2.
Aq  B  ( D  A)  ( D  A) 2  4CB 
q  Cq  ( D  A) q  B  0  q 
2

Cq  D 2C 
 
 k2  k  k2   
 cos l  sin  l  
 A B    k  k  k  
      2   
 
C D  k2  k2   k2  
 sin  l  cos l  
 k k k  
     
1 k  R  
 i 2  
q k  1 1
 2 k  4  1  4
1 1   w      
 i k k
 2  n
 0 2n
q R n0 w 2  

In the eigen mode the spot size w is independent of z. This is caused by the focusing
nature of the medium, which cancels the diffraction nature of the beam.

45
Chapter 3 Polarization of Light Waves
Lecture 1 Polarization
3.1 The concept of polarization
Introduction:
1)Since F = qE, polarization determines force direction. The generation, propagation and
control of lasers thus crucially depend on their state of polarization.
2)The electric field E is defined as the polarization state of the light.
3)In an anisotropic material the index of refraction depends on the polarization state of the
light, which is used to manipulate light.
3.2 Polarization of monochromatic plane waves
In a monochromatic plane wave propagating in the z direction, the x and y components of
the electric field oscillate independently.
E( z , t )  Re[ A exp[i (t  kz )]] Ex

 Re[( A ei x ˆi  A e y ˆj) exp[i (t  kz )]]


i
x y

E x ( z , t )  Ax cos(t  kz   x )
E y ( z , t )  Ay cos(t  kz   y ) Ey

46
Ey

Ex

Ey
E

Ex

47
The trajectory of the end point of the electric field vector, at a fixed point as time goes, is
E x  Ax cos(t  kz   x )  y
  (eliminating t  kz )
E y  Ay cos(t  kz   y )  y' Ay E x'
2 2
 Ex   E y   E  E  a
      2 x  y  cos   sin 2 
    b 
 Ax   Ay   Ax  Ay  x
Ax
  y  x,     .

Light is thus generally elliptically polarized.


A complete description of the elliptical polarization needs
1)Orientation angle 
2)Ellipticity (shape) a/b, and
3)Handedness (sense of revolution, can be combined to show the sign of ellipticity).
They can be obtained by rotating the ellipse to its normal coordinates:
 E  2  E y '  2
 x '      1
 a   b 

 2 2 Ax2  Ay2   Ax2  Ay2   4 Ax2 Ay2 cos 2 
2

a , b 
 2
 2 Ax Ay
 tan 2  cos 
 Ax
2
 A 2
y
 48
Handedness (sense of revolution) of elliptical polarization:
Looking at the approaching light, if the E vector revolves counterclockwise, the
polarization is right-handed, with sin <0. If the E vector revolves clockwise, the
polarization is left-handed, with sin >0. Some books use the opposite definition.

= - -3/4 -/2 -/4 0/4 /2 3/4 

Linear polarization:
   y   x  0, 
Ey Ay

Ex Ax

Circular polarization:
 = -/2 /2
  y x  
2
Ax  Ay
49
Lecture 2 Jones vector
3.4 Jones vector representation
A plane wave can be efficiently and uniquely described by a Jones vector in terms of its
complex amplitudes on the x and y axes:
 Ax ei x 
J  
 A e i y 
 y 
If we are only interested in the polarization state of the wave, we use the normalized
Jones vector, with J+J=1:
 cos  A
J ( ,  )   i ,   arctan y (auxiliary angle),    y   x .
 e sin  Ax

Examples: Some relations:


1)Linearly polarized light:
ˆ  L
xˆ   yˆ  0, R ˆ 0
 cos   cos  1   0
  and  , especially xˆ   , yˆ   . ˆ  1  xˆ  iyˆ  , Lˆ  1  xˆ  iyˆ 
 sin    sin   0 1  R
2 2
2)Right- and left-handed circularly polarized light:
xˆ 
2

1 ˆ ˆ

R  L , yˆ  
i ˆ ˆ
2
RL 
ˆ  1  1  and Lˆ  1 1 .
R
2   i  2  i 
50
Examples of Jones vectors:

= - -3/4 -/2 -/4 0/4 /2 3/4 

1  1  1  1  1  1  1  1  1 1
J          
5   2  5   2 (1  i )  5   2i  5  2 (1  i )  5  2 
1  1  1 1 1  1  1  1 
       
5  2 (1  i )  5  2i  5  2 (1  i )  5   2 

Light intensity:
 Ax ei x 
J   
i y ,

I J J Ae  * i x
x Ae * i y
y 
 Ax ei x 

 A ei y 
  Ax
2
 Ay
2

 Ay e   y 
Jones vectors are important when applied with Jones calculus, which enables us to
track the polarization state and the intensity of a plane wave when traversing an
arbitrary sequence of optical elements.
51
(*Reading) What is the orientation angle  of the ellipse ax  by  cxy  1 ?
2 2

How long are the principle axes?


Solution 1: Suppose the ellipse is erect after rotating the x, y axes by an angle , that is
 x'  x cos   y sin   x  x' cos   y ' sin 
 ,  , and Ax'2  By '2  1.
 y '   x sin   y cos   y  x' sin   y ' cos 

ax 2  by 2  cxy  a ( x' cos   y ' sin  ) 2  b( x' sin   y ' cos  ) 2  c( x' cos   y ' sin  )( x' sin   y ' cos  )
 x'2 (a cos2   b sin 2   c sin  cos  )  y '2 (a sin 2   b cos2   c sin  cos  )

 A  a cos2   b sin 2   c sin  cos 
 x' y ' (a sin 2  b sin 2  c cos 2 )  1 
   B  a sin   b cos   c sin  cos 
2 2

Ax'  By'  1
2 2
  c
 a sin 2  b sin 2  c cos 2  0  tan 2 
 a b

52
 
(Let rm be the principle axes,  m   or   ) 
2 
1 1 c 
rm2   2  a  (b  a ) sin 2  m  sin 2 m  
a cos  m  b sin  m  c sin  m cos m
2 2
rm 2 
c 
tan 2 m  
a b 
1 c 2 sin 2  m  c 
 a   sin 2    a  tan  m 
2  tan 2 m 
m
rm2 2 

c 2 tan  m  ( a  b)  ( a  b)  c 
2 2
tan 2 m    tan  m  
a  b 1  tan 2  m c 
1/ 2
1 ( a  b)  ( a  b)  c 2 2 ( a  b)  ( a  b) 2  c 2 
  rm   
rm2 2  2 

ax 2  by
by 2  cxy
cxy  Ax
Ax''2By
2 2 2 2
If
If ax By''2 ,, then
then
c c
 tan 2 m  tan 2 m 
a b a b
( a (ba)b)(a (ba)
2 b) 22  c 2
c
A, BA, B 
2 2

53
2 2
 E   Ey   E  E y 
For the case of  x      2 x   cos   sin 2  ,
   
 Ax   Ay   Ax  Ay 
2 cos 

c Ax Ay 2 Ax Ay
tan 2 m    2 cos 
ab 1 1 A  A 2

2
 2 x y
Ax Ay
1 / 2
 ( a  b )  ( a  b) 2  c 2 
1 / 2  2 2  
 1 1  1 1   2 cos    / 2 sin 2  
E x ' 0 , E y '0     2  2   2  2     
2  A Ay   Ax Ay  
   Ax Ay  x   
 
1 / 2 1 / 2

 A2  A2  A2  A2 2  4 A2 A2 cos 2   
2

   
x y x y x y

 2 Ax Ay sin    2
  
2 2 2 2
   Ax  Ay 
2
Ax2  Ay2  4 Ax2 Ay2 cos 2  
1/ 2
 A2  A2  A 2
A 
2 2
 4 A A cos  
2 2 2


x y x y x y

 2 
 

54
Solution 2: Let us use a little linear algebra.
 a c / 2  x  
ax 2  by 2  cxy  1   x y      1
 c / 2 b  y  
 x'   cos  sin   x   x 
       R  
 y '    sin  cos   y   y 
 A 0  x'  
Ax'  By'  1   x' y ' 
2 2
   1 
 0 B  y '  
A 0  a c / 2  1  cos  sin   a c / 2  cos   sin  
   R  R     
 0 B   c / 2 b    sin  cos   c / 2 b  sin  cos  
 c 
 a cos   b sin   c sin  cos  (cos 2   sin 2  )  (a  b) sin  cos  
2 2

 2 
c
 (cos   sin  )  (a  b) sin  cos 
2 2
a cos   b sin   c sin  cos  
2 2

2 

 A  a cos 2   b sin 2   c sin  cos 

  B  a sin 2   b cos 2   c sin  cos 
c c
 (cos 2   sin 2  )  (a  b) sin  cos   0  tan 2 
2 a b

55
Solution 3: The problem is: given the condition ax  by  cxy  1 for x and y, what is
2 2

the maximum of x2 + y2? This can be solved by the Lagrange multiplier method.
f ( x, y,  )  x 2  y 2   (ax 2  by 2  cxy  1)
f 
 2 x   (2ax  cy )  0 
x  y 2by  cx 2b tan   c 2 tan  c
  tan       tan 2 
f  x 2ax  cy 2a  c tan  1  tan 2  a b
 2 y   (2by  cx)  0
y 

Solution 4: In polar coordinates, x  r cos  , y  r sin  . The curve is now

1
ax 2  by 2  cxy  1  r 2  .
a cos 2   b sin 2   c sin  cos 
dr
d
0
d 1
 2 0
d  r   d
d
 a cos 2   b sin 2   c sin  cos    0 
  

c
 a sin 2  b sin 2  c cos 2  0  tan 2  .
a b
1  A  a cos 2   b sin 2   c sin  cos 
rm 
2

a cos 2   b sin 2   c sin  cos   B  a sin 2   b cos 2   c sin  cos 

56
Lecture 3 Diagonalizing the tensor of a quadratic surface
Question: What is the orientation angle  of the ellipse y
ax 2  by 2  cxy  1  Ax'2  By '2 ? y'
x'
How long are the semiaxes?
A-1/2
B -1/2

Solution 5: 
x
Diagonalizing the tensor of a quadratic surface

Surppose a quadratic surface is expressed in the x - y - z and x' - y' - z ' coordinate systems as
 S11 S12 S13  x   S '11 S '12 S '13  x ' 
     
x y z   S21 S22 S23  y    x ' y ' z '   S '21 S '22 S '23  y '   1
S S32 S33  z   S' S '32 S '33  z ' 
 31  31
That is,
Sij xi x j  xi Sij x j  xi ' Sij ' x j '  r T Sr  r 'T S' r '  1 (sum over repeated indices).

57
Suppose R is the othogonalcoordinate tranformation that relates the two coordinate
systems, r '  Rr , R 1  R T , then
r T Sr  ( R 1r ' )T S( R 1r ' )  r 'T (R 1 )T S(R 1r ' )  r 'T ( RSR 1 )r '  r 'T S' r '  1
 S'  RSR 1.  S that specifies the quadratic surface is a tensor.
Suppose the othogonal coordinate tranformation R diagonizes S,
S'  RSR 1  SR 1  R 1S' 
 R111 R121 R131   R111 R121 R131  S1 ' 0 0
 1 1 1
  1 1 1
  
S R21 R22 R23    R21 R22 R23  0 S 2 ' 0  
 R 1 1
R32 1 
R33  1 1 1 
0 S3 ' 
 31   R31 R32 R33  0
 R111   R111   R121   R121   R131   R131 
 1   1   1   1   1   1 
S R21   S1 '  R21 , S R22   S2 '  R22 , S R23   S3 '  R23 .
 R 1   R 1   R 1   R 1   R 1   R 1 
 31   31   32   32   33   33 
Also since RR 1  1, we have
 R111   1   R121   0   R131   0 
 1     1     1   
R  R21    0 , R  R22    1 , R  R23    0 .
 R 1   0   R 1   0   R 1   1 
 31     32     33   
58
Therefore we conclude:
If the orthogonal transformation R diagonizes S into S',
 S1 ' 0 0
 
RSR 1  S'   0 S2 ' 0 
0 0 S3 ' 

then
1) The diagonal elements of S' are the eigenvalues of S.
2) The row vectors of R, or the column vectors of R-1, are the eigenvectors of S.
3) The new coordinate axes lie along the eigenvectors of S.

 a c / 2  x   a c / 2
ax 2  by 2  cxy  1   x y      1  S   
 c / 2 b  y  c / 2 b 

Strategy :
1) Diagonizi ng S. That is, find the eigenvecto rs and eigenvalue s of S.
2) The eigenvecto rs are the principle axes of the curve, which show the orientatio n angle  .
3) The eigenvalue s are the coefficien ts in Ax '2  By '2  1. The semiaxes are then simply
1 A and 1 B.

59
a c/2 ( a  b)  ( a  b) 2  c 2
 0  1, 2  .
c/2 b 2
a   c / 2  x  ( a  b )  ( a  b) 2  c 2 c
    0  x y 0
 c/2 b    y  2 2
 c 

r1, 2   
2 
 ( a  b)  ( a  b)  c 
2

  ( a  b)  ( a  b ) 2  c 2 
1 / 2

 A  1  ( a  b)  ( a  b)  c
2 2
1
 rm1   
 2 A  2 

1 / 2
  ( a  b)  ( a  b) 2  c 2 
 ( a  b)  ( a  b)  c
2 2
1
  B  2   rm 2   
 2 B  2 

 tan 2  22 x1 y1 2  2 x1 y1

c
 x1  y1 (  y1 y2 )  y12 a  b


60
Chapter 4 Electromagnetic Propagation in
Anisotropic Media
Lecture 1 Light propagation in anisotropic media
Introduction:
1) The optical properties (e.g., refractive indices and absorption) of an anisotropic
medium remarkably depend on the propagation direction and the polarization state of
the incident light, as well as the external forces (electric, acoustic and mechanic)
exerted on the materials.
2) Anisotropic media exhibit many interesting and important phenomena, including
birefringence, double refraction, conic refraction, optical rotation, Faraday effect, and
electro-optical phenomena. These effects are employed to design and fabricate
various optical devices, including polarizers, filters, beam splitters, rotaters, and many
electro-optical devices. A significant portion of laser technology, and nonlinear optics
deal with the generation and control of light using the optical properties of anisotropic
crystals.
3) We therefore conclude that studying the propagation of light in anisotropic media is
inevitable in understanding the physics of lasers and their application in the lab.

61
4.1 The dielectric tensor of an anisotropic medium
Crystals are made up of regular periodical arrays of molecules. In an anisotropic crystal,
the polarization induced by an electric field is in general not in the electric field direction.
This can be seen if we consider that the electrons are anisotropically bonded in the crystal.

Pi   0  ij E j (sum assumed,  ij is the susceptibility.)


Di   0 Ei  Pi   0 ( ij   ij ) E j   ij E j .
 ij is called the dielectric tensor.

We currently make the following assumptions on the media:


1)Homogeneous. The medium is identical if translated internally.
2)Nonabsorption.  is real. Complex permittivity tensors will be considered later when
we study the optical activity of crystals.
3)Linear.  does not depend on the strength of the external electric field. Nonlinear
polarization is the basis of nonlinear optics, which we learn only a little in this course.
4)Nonmagnetic. Our results can be easily extended to magnetically isotropic
media.

62
The dielectric tensor ij is symmetric in a nonabsorption medium.
B D 
E   ,H  
t t     ( E  H)  E  D   H  B 
  ( E  H)  H  (  E)  E  (  H) 
 S  ED   H  B  U  U  U  E  D
e m e
  E  E
i ij j


1 1 1    ij   ji
U e  E  D  U e   ij ( Ei E j  Ei E j )  ( ji   ij ) Ei E j 
   
2 2 2 
This is an intrinsic symmetry. It comes from the nature of the thermodynamic
requirement that Ue is a state function of E, which takes all Ei as independent variables.
This does not require the symmetry of the crystal, or the linearity of the response.
This can be generalized as follows.
dU  X i dYi  e.g., F  ds, E  dD  
  2  2
Yi  aij X j   d   aij X j dX i , 
 X i X j X j X i
U  U ( X i , X j ,)   ( X i , X j ,)  U  X iYi 
 aij  a ji
 Dx    x 0 0  E x   nx2 0 0  E x 
The real and symmetric dielectric        
D  0  y 0  E y    0  0 n y 0  E y 
2
tensor ij can then be diagonalized in  y    0 0 n 2  E 
 D   0 0   E 
the principle dielectric axis system:  z   z  z   z  z 

Relative permittivity 63
4.2 Plane wave propagation in anisotropic media
In an anisotropic medium, the phase velocity of light depends on its polarization state and
its propagation direction. For a given propagation direction, there exist in general two
eigenwaves, each has its own eigen refractive index (or equivalently eigen phase
velocity ) and eigen polarization. All light traveling in that direction can be decomposed
onto the two eigenwaves.
Question: For a given wave normal direction s in the crystal, what are the eigen refractive
indices and eigen polarizations?
This is answered by solving the Maxwell’s equations with an anisotropic dielectric tensor.
Suppose the phase of all the electromagnetic fields (E, H, D, and B) varies in the form:
  n  k in 
exp i  t  k  r    expi  t  s  r , where s  . Then    s,  i
  c  k c t
B n
E   H sE E
t c D
D n

H   D   sH
t c k (s)
H (B) 
Relations between the directions of the vectors:
1) D, H, and s are mutually perpendicular. S=E×H (t)
2) D, E, s, and E×H (energy flow) lie in the same plane.
3) The Poynting vector S=E×H is generally not along s.
64
Wave normal direction s and energy transfer direction t
k
Wave normal direction s  is the direction cosine of the wave vector.
k
n
Wave vector k  s is perpendicular to the wavefronts.
c
S EH S
Energy transfer direction is t   , with the velocity .
S EH U
c
Phase velocity v p  s is the velocity for phase transfer.
n
kc k
Refractive index n   , where  is the wavelength in vacuum.
 2
E
D

k (s)
H (B) 
S=E×H (t)

65
Lecture 2 Eigenwave equation
We continue to search for the eigenwaves propagating along a given direction in a crystal.
B n 
E   H s  E
  D  n  0s  (s  E)    n  0s  s   E  0
t c  2 2
 
D n
 H   D   sH 
t c 

This is the eigenwave equation for determining the eigen refractive indices (eigen values)
and polarization (eigen states) of a plane wave propagating in a prescribed direction s. We
now realize it in the matrix form.
   n  s  s  E  0
2

  E  n  0  E  s(s  E)  0  (ni / n  1) Ei   si s j E j  0
0 2 2 2

s  (s  E)  s(s  E)  E j

 
  si s j  (1  ni2 / n 2 ) ij E j  0 
j Fun math tricks:
 s x2  (1  nx2 / n 2 )  E x   0  sz sy 
sx s y sx sz  
   s   sz 0  sx 
 s y sx s y  (1  n y / n )
2 2 2
s y sz  E y   0  s
 y sx 0 
 2 
 s z s x s z s y s 2
z  (1  n 2
z / n )  E z  s  s  ss  1

66
Solving the eigenwave equation:
For a nontrivial solution of E, the determinant of the matrix must be 0.

 sx2  (1  n x2 / n 2 ) sx s y s x sz 
 
det  s y sx s y  (1  n y / n )
2 2 2
s y sz 0
 s s s s s 2
 (1  n 2
/ n 2 
)
 z x z y z z

sx2 s 2y sz2 1
This can be simplified into 2    .
n  nx n  n y n  nz n
2 2 2 2 2 2

This is called the Fresnel’ s equation of wave normals. It is a quadratic equation of n 2 ,


which in general supports two refractive indices for a given direction ( s x , s y , sz ).
s x2 s 2y sz2
Another form of the same equation is    0.
1 1 1 1 1 1
  
n 2 n x2 n 2 n 2y n 2 nz2
The corresponding electric field E of each eigenstate is then found to be given by
sx sy sz
Ex : E y : Ez  2 : :
n  n x2 n 2  n 2y n 2  nz2

67
One easier way to solve the eigenwave equation is given below:
   n  s  s  E  0
2

  E  n  0  E  s(s  E)  0   i Ei  n  0  Ei  si (s  E)
0 2 2

s  (s  E)  s(s  E)  E
n 2 si (s  E) n 2 si2 s x2 s y2 s z2 1
 Ei  2  eq.1  s  E   2 2 (s  E )    
n  ni2 i  x , y , z n  ni n 2  nx2 n 2  n y2 n 2  nz2 n 2
sx sy sz
eq.1  E x : E y : E z  2 : : .
n  nx2 n 2  n y2 n 2  nz2

Question: For a given wave normal direction s in the crystal, what are the eigen refractive
indices and eigen polarizations?
Answer:
1)In general two refractive indices, n1 and n2, are given by solving the Fresnel’s equation
of wave normals.
s x2 s y2 s z2 1
  
n 2  nx2 n 2  n y2 n 2  nz2 n 2
2)The polarization direction of E of each eigen state is given by
s sy sz
Ex : E y : Ez  2 x 2 : 2 :
n  nx n  n y2 n 2  nz2
68
4.4 Phase velocity, group velocity, and energy velocity
 n( )
The phase velocity of a plane wave is v p  s, with k  .
k c
Therefore in general each of the two eigenwaves in a crystal has its own phase velocity.
Group velocity is the velocity of energy flow of a laser pulse in a dispersive medium.
We now generalize the concept of group velocity, where the light pulse moves in a bundle
of directions centered at k0. We decompose the light in the k-space (momentum space).
E(r, t )   A(k ) exp i (k )t  k  r  d k 
 k

 (k )   (k 0 )   k (k ) k  (k  k 0 )   

  
0

E(r, t )  exp i   (k 0 )t  k 0  r    A(k ) exp i  k (k ) k t  r  (k  k 0 ) d k


 k

 
0

 f  k (k ) k t  r exp i   (k 0 )t  k 0  r  
0

    
 Group velocity v g   k (k )   , , .

 xk k z k z 

In k-space, (k) = const is called a wave normal surface. Group velocity is the gradient
of the wave normal surface in k-space, and is therefore always perpendicular to the wave
normal surface.

69
S E H
The energy velocity is defined as v e   .
U 1  E  D  H  B
2
For a non-absorptive medium, vg= ve. A neat proof is given in our textbook.

group velocity
kz
(k) = const
normal surface
k-space

ky
kx

70
Lecture 3 Wave normal surface in the k-space
We now express the Fresnel’s equation of wave normals in terms of k=(kx, ky, kz), that is,
in the k-space. This is not mathematically new compared to what we have done by using
n and s to express k. However, the k-space gives us more convenience.
  2    2 2 2  
 2

  n  0 s  s  E  0   2   0 k  k  E  0    k i k j
    k  2 ni  ij  E j  0 
 c  j   c  
 2 2 
 2 nx  k y2  k z2 kxk y kxkz 
c  E x 
 2 2  
 k ykx n y  k 2
x  k 2
z k ykz  E y   0 
c2
  E z 
 2 2
 kzkx kzk y 2
nz  k x2  k y2 
 c 
2 2 
 2 nx  k y2  k z2 kxk y kxkz 
c 
 2 2 
det k ykx n y  k 2
x  k 2
z k ykz   0.
c2
 
 2 2 2
 kzkx kzk y 2
nz  k x  k y 
2

 c 
71
This determinate can be simplified into
 k x2 k y2 2   2 1       2 4

n n
k
  1 1 1 1 1
 2 2  2 z 2  k x  k y  k z  k x  2  2   k y  2  2   k z  2  2  2  4  0.

2 2 2 2 2

  n y nz  n 
2 2
 y z nx nz nx n y   nx nz   x n y  c c
This is an (k) = constant wave normal surface. The surface is composed by all the wave
normals k that have the same frequency . Note that the relative permittivities also depends
on frequency. For future use, we simplify the equation for some special cases.
1) If nx  n y  nz , then
k x2 k y2 k z2
   1.
 2
 2
 2
k 2  2 nx2 k 2  2 n y2 k 2  2 nz2
c c c
This is the general form of Fresnel’ s equation of wave normals.
2) If k y  0, then
 k x2  k z2  2  k x2 k z2  2 
  2  2  2  2   0
 n2 c  nz nx c 
 y

3) If nx  n y  no , nz  ne , then
 k x2  k y2 k z2  2  k 2  2 
  2  2  2  2   0.
 n2 no c  no c 
 e
72
Wave normal surface in the k-space:
The wave normal surface (k) = constant contains the end points of all the k vectors
for a given frequency . Here are the general characteristics of the wave normal surface.
1)The wave normal surface consists of two shells, with only four points in common.
2)The two lines that go through the origin and the four common points are called the
optic axes. When light is propagating in the direction of one of the two optic axes, there
is only one possible k value, and thus only one refractive index.
3)Any other light propagating direction intersects the two shells of the wave normal
surface at two different points, giving two possible k values, and thus two refractive
indices for the two eigenwaves.
4)For each eigenwave, the energy flow (group velocity or energy velocity) of the light is
perpendicular to the wave normal surface at its k point.
Once wave vector k is known, the polarization of the corresponding E field of the
eigenwave is given by
kx ky kz
Ex : E y : Ez  2 : :
k  nx2 2 / c 2 k 2  n y2 2 / c 2 k 2  nz2 2 / c 2

73
74
75
76
Spatial relations between the fields of the two eigenwaves:
s  (sx , s y , sz )
 sx sy s z  E1

e 2 , , D1
 n  n2 n2  n2 n2  n2  
 x y z 
S2(t2)
 nx2 s x n y2 s y n 2
s z  

d 2 , 2 , 2 z
 n 2e  s k (s)
n n n n n n 
2 2 2 H1 (B1) 
 x y z 
D2 
d1  d 2   n12e1  s   d 2  n12e1  d 2  n22e 2  d1  E2 S1(t1)

e1  d 2  e 2  d1  H2 (B2)
 d1  d 2  e1  d 2  e 2  d1  0, if n1  n2 .
d1  s d 2  s 1
e1  e 2  2
 2
 2 2
0
n1 n2 n1 n2

Orthogonality of the two eigenwaves:


E1  D 2  E 2  D1  0  U e  (E1  E 2 )  (D1  D 2 )  U1  U 2
s  (E1  H 2 )  s  (E 2  H1 )  0  s  S  s   (E1  E 2 )  (H1  H 2 )  s  (S1  S 2 )

77
Lecture 4 The index ellipsoid

4.3 The index ellipsoid


When light propagates in a crystal, the D vectors of the two eigenwaves and the wave
normal direction s form a mutually perpendicular triad. It is therefore more convenient to
present the field vectors in the D space.
The electric energy density Ue is given by
1  Dx2 D y Dz2  1  Dx2 Dy Dz2 
2 2
1
Ue  E  D       .
2 2   x  y  z  2 0  nx2 n y2 nz2 
2
Dx2 D y Dz2
The constant energy density surface in D space is then n 2  n 2  n 2  2 0U e .
x y z

x2 y2 z2
If we denote D / 2 0U e  r  ( x, y, z ) then we have 2  2  2  1.
nx n y nz

This surface is called the index ellipsoid (or the optical indicatrix).

78
The role of the index ellipsoid:
For a given arbitrary wave normal direction s, the index s
ellipsoid can be used to
1)Find the indices of refraction of the two eigenwaves. D2
2)Find the corresponding directions of the D vectors of n2
the two eigen waves. n1
D1
The prescription is as follows:
1)Draw a plane that is through the origin and is
perpendicular to s. This plane intersects the index
ellipsoid surface with a particular intersection ellipse.
2)The lengths of the two semiaxes of the intersection
ellipse, n1 and n2, are the two indices of refraction of
the eigenwaves.
3)The two axes of the intersection ellipse are each
parallel to the D vectors of the eigenwaves.

79
Proof of the prescription of using the index ellipsoid:
 x2 y 2 z 2
 2  2  2 1
The intersection ellipse is given by  nx n y nz
 xs  ys  zs  0
 x y z

The lengths (squared) of the semiaxes of the ellipse is given by the extrema of
r 2  x 2  y 2  z 2 subject to the above two auxiliary conditions. This can be solved by
the Lagrange undetermined multipliers method, construct
 x2 y2 z 2 
F ( x, y, z , 1 , 2 )  x  y  z  1  xsx  ys y  zsz   2  2  2  2  1
2 2 2
n 
 x n y nz 
F F F F F
The extrema of r 2  x 2  y 2  z 2 is then given by solving      0.
x y z 1 2
  2 1 xi si 2 xi2  
  i
       
2
x 2 
0 r 2 
F F F s  x  
i 2 n i  
   0  xi  1 i  2 2 i  0   
x y z 2 ni   s x  1si  2 si xi   0  1  
2
si xi 
  i i 2 ni2  2
2
ni2 
 i  i

2   r 2 
   r 2  2  s x x s y y sz z 
 
2 sx x s y y sz z    xi 1  2   r si  2  2  2   0
1  2r  n 2  n 2  n 2    ni   nx ny nz 
  x y z   80
 rm2  2  s x xm s y ym s z z m  
xmi 1  2   rm si  
 2  2  0 (Here m meams extrema.)
n  n2 ny nz  
 i   x

If we according to the prescripti on assume the extrema rm2  n 2 , 
D Di  i Ei 
and assume rm // D, that is rm  n , then xmi  n n 
|D| |D| | D| 

 s x x E x s y  y E y s z  z E z 
n si 
2
  
 n2 ny2
nz  n 2 si 0 ni2 (s  E)
2

 i Ei   x
 
n 2 / ni2  1 n 2  ni2
 sx sy sz
E :
 x y z E : E  : :
n 2 si (s  E)  n 2  nx2 n 2  n y2 n 2  nz2
Ei  2 
n  ni2 s
 x 
2
s 2
y s z2 1
 
 n2  n2 n2  n2 n2  n2 n2
 x y z

That is, if we assume the extrema of the intersection ellipse be rm  n and rm // D , then the
resultant E fields will be on the right polarizations, and the resultant n1 and n2 will satisfy
the Fresnel’s equation of wave normals.

81
Comparison of the wave normal surface method and the index ellipsoid method
We have studied the wave normal surface method and the index ellipsoid method on
light propagating in anisotropic crystals. Let us compare them here.
1)The wave normal surface shows us the allowed two wave vectors, and thus the
refractive indices, at a given light propagation direction. The index ellipsoid does
similar things, plus it also shows us the allowed direction of the D vectors.
2)The index ellipsoid displays the refractive indices and the directions of the D vectors
in a convenient visual way. It also involves easier mathematics. Because of this
simplicity, it is often first considered in solving problems.
3)However, the knowledge in the wave normal surface is much more profound. For
example, it displays the optic axes. It shows the wave vector variation in the k-space. It
shows the energy transfer direction. It has more mathematical base, and therefore
allows for deeper theoretical derivations.
4)We tentatively summarize that for solving problems of light propagating in
anisotropic crystals, we may first try the index ellipsoid method. However, we need to
keep in mind that we have a backup more powerful wave normal surface method. This
is especially true when we are directly dealing with the wave vectors rather than
merely the refractive indices in the problems.

82
Lecture 5 Light propagation in uniaxial crystals
4.5 Classification of anisotropic media (crystals)
Crystals are optically classified into 3 groups, namely the isotropic (or cubic) , the
uniaxial, and the biaxial crystals, according to the number of independent elements of
their dielectric tensors in the principle axis systems.
Our text shows many examples of isotropic and anisotropic crystals commonly used in
making optical devices.

Isotropic Uniaxial Biaxial


 0 0   x 0 0  x 0 0
     
Dielectric tensor 0  0 0 x 0 0 y 0
0 0   0 0  z  0 0  z 
   
Positive crystal Usually choose
when z>x, x<y<z
negative crystal
when z<x.
Examples NaCl, diamond Quartz (positive) Mica
Calcite (negative) Topaz
BBO (Beta-Barium
Borate, negative)
83
4.6 Light propagation in uniaxial crystals z
The index ellipsoid:
The equation of the index ellipsoid of a uniaxial crystal is s

x2  y2 z 2
 2 1 De n )

no2 ne e
y
no   x /  0   y /  0 , ordinary refractive index no
Do
ne   z /  0 , extraordinary refractive index
ne  no : positive uniaxial crystal  prolate spheroid
ne  no : negative uniaxial crystal  oblate spheroid
The index ellipsoid is rotationally symmetric around the z-axis. Let s be in the y-z plane
with a polar angle . The two polarization directions of the D vectors are: Do is parallel
to the x-axis, De is in the y-z plane and is perpendicular to s.
The corresponding refractive indices are:
no  no ,

  
 ne ( ) cos
2



ne ( ) sin 
2
 cos 2  sin 2  
 1  ne ( )   
1 / 2

 .
 2 2 2 2
 no ne  no ne 

When s is on the z direction, ne(0°) = no. Therefore the z-axis is the optic axis.
84
The wave normal surface: kz  / c
If nx  n y  no , nz  ne , then the wave normal s
De()
surface is reduced to ne)
 no
 k x2  k y2 k z2  2  k 2  2  k y  / c
  2  2  2  2   0.
 n2 no c  no c  Do n o  ne
 e

The refractive indices are given by solving the


two factors: Positive uniaxial crystal
no  no ,
 1 / 2
  cos 2  sin 2   kz  / c
ne ( )   2
 2
 .
  no ne  De() s
Notes to uniaxial crystals: n
 neo)
1)At a given propagation direction, there are in general k y  / c
two eigen refractive indices, each has its own eigen Do n o n e
polarization direction. This is called birefringence.
2)The E field of the o-ray is always polarized
perpendicular to the plane that contains s and the optic Negative uniaxial crystal
axis, while the e-ray is polarized parallel to that plane.

85
Lecture 6 Double refraction
4.7 Double refraction at a boundary
Up to now we discussed the light propagation in an anisotropic crystal when it is already
inside the crystal, but how is the light refracted into the crystal?
We recall that the boundary condition requires
1)The wave vectors of the incident, reflected, and refracted light ( k 0 , k r , and k )t lie in
the plane of incidence.
2)The tangential components of all three wave vectors on the boundary interface is the
same: k0 sin  0  k r sin  r  kt sin  t  const.
We know that in the crystal the length of kt depends on the angle of refraction t , as well
as the orientation of the optic axes. Each t supports two kt because of the double shell
structure of the wave normal surface. k 0 sin  0
Also because of the double shell structure of  k1 sin 1
k0  k 2 sin  2
the wave normal surface, in general there are 0
two refraction angles, 1 and 2, both satisfy boundary
k1 sin 1  k 2 sin  2  k0 sin  0  const , or

n1 sin 1  n2 sin  2  n0 sin  0
 k1
k2
wave normal
surfaces
86
The refraction at a boundary can be explained on the intersection between the wave
normal surface and the plane of incidence.
At the boundary, the k0 beam is uniquely decomposed into the reflected beam and the two
eigenstates of the refracted k1 and k2 beams, according to the boundary conditions. Each
refracted beam then propagates separately. This is called double refraction.
Please note that in general the polarization of the k1 and k2 beams are not orthogonal since
they are in different directions. The incident light still can be uniquely decomposed into the
reflected beam and the two refracted beams according to the boundary conditions.

Internal double reflection:


When light is internally reflected from the surface k0 sin  0
of an anisotropic material, double reflection may  k1 sin 1
occur. It can be discussed similarly on the k0  k 2 sin  2
intersection between the wave normal surface and 0
boundary
the plane of incidence.

 k1
k2
wave normal
surfaces
87
Double refraction at the boundary of a uniaxial crystal: Examples
The directions of the D vectors are shown. There exist an ordinary wave with the
refractive index no, and an extraordinary wave whose refractive index ne depends on its
direction of propagation.

optic
axis optic
optic
Positive axis ki ki axis ki
uniaxial
crystal
ko
ko ko
ke
ke ke

optic
axis
optic optic
Negative axis ki ki axis ki
uniaxial
crystal ke ke
ke
ko ko ko

88
o e

89
Deviation angle between the energy flow of the o-ray and e-rays
Suppose the light is incident normally on the surface of a uniaxial crystal. Then the wave
vectors koand ke are in the same direction. Let us see how much the energy flow of the e-
ray is deviated away from that of the o-ray. z
Do  ( Dox ,0,0)  E o  ( Dox /  x ,0,0) te
s (to)
t o is in the E o - s plane and is perpendicular to E o  t o  s.

De  (0, De cos , De sin  )  E e  (0, De cos /  x , De sin  /  z ) De n ) 
E
e
 y
t is in the E - s plane and is perpendicular to E .
e e e e
no
Suppose t e has a polar angle  . Do(Eo)
2
D sin  /  z  no 
tan  e    tan 
De cos /  x  ne 
 n  2 
The angle between t e and t o is then         arctan  tan  
o

 ne  
2 tan   ne 
If ne is very close to no , we have     1, and the maximum deviation
1  tan 2   no 
occurs at   45 with   (ne  no ) / no .
Example:
For KDP (KH2PO4) crystal, no=1.50737, ne=1.46685. At =45°, the angle between the e-
ray and the o-ray is =1.56°. 90
Lecture 7 Light propagation in biaxial crystals z
4.8 Light propagation in biaxial crystals s
The index ellipsoid:
The equation of the index ellipsoid of a biaxial crystal is n()=ny 
x2 y2 z 2 D2
   1 (suppose nx  n y  nz ) x
nx2 n y2 nz2 ny
D1
Generally for a wave normal direction s (,) there exist two
allowed vectors D1 and D2 with their specific refractive indices
n1 (,) and n2 (,).
If s is in the x-z plane, we have n1()= ny always, and n x<n2 ()< nz. Since n x<ny< nz,
there exists a special angle  which makes n2 ()= n1 ()= ny. That is, the intersection
ellipse is a circle and there is no difference between the two refractive indices. This
special direction  is called the optic axis of the crystal.
For a biaxial crystal, there are two such optic axes, both are on the x-z plane. They are
located symmetrically on each side of the z-axis. Thus comes the name biaxial crystal.
It is not difficult to find that the polar angle of the two optic axes is given by
 1 1   1  n y2  nx2
tan     2  2   2  12    nz .
n  n  n n
2 2
 x ny   y nz  nx z y

91
When an unpolarized light is propagating along the optic axis of a biaxial crystal, the
allowed D vector is in any direction perpendicular to the s vector. This results in the
allowed E vectors locating in one plane. The ray direction t is then found to form a cone
on one side of the light propagation direction s. Any plane perpendicular to s intersects
the cone with a circle. This phenomenon is called conical refraction.
Conical refraction is predicted by William Hamilton in 1832, and was confirmed
experimentally by Humphrey Lloyd the next year. Both were Irish scientists.
s  ( s x ,0, s z )
D  s  Dx s x  Dz s z  0  E x x s x  E z  z s z  0  E  ( x s x ,0,  z s z )
t is in the plane of E, s, D, and t  E.
We discuss the details of conical refraction later using the wave normal surfaces.

t1
s
t2
D1 t3
E1
D2
E2 y

E3(D3)
92
Wave normal surfaces:
We already know that the wave normal surface of a biaxial crystal consists of two
complex shells, with only four points in common. Let us consider the intersections
between the wave normal surface and the three coordinate planes. If we set ky=0, then
 k x2  k z2  2  k x2 k z2  2 
  2  2  2  2   0
 n 2
c  nz nx c 
 y

This consists of a circle with the radius ny/c, and an ellipse with semiaxes nz/c and
nx/c. The intersections with the other two coordinates planes are similar, each has a circle
and an ellipse. The optic axes lie in the x-z plane.

k y  / c kz  / c kz  / c
nz ny ny

nx nx nx
k x  / c k x  / c k y  / c
ny nz ny nz nx nz

93
Lecture 8 Conical refraction
Conical refraction:
The group velocity of light is the gradient of the wave normal surface in the k-space, and
is thus perpendicular to the wave normal surface. However, on the optic axis, the two
shells of the wave normal surface degenerate into a point, which is a singular point where
the gradient is not well defined. The energy flow of light at this point is governed by the
nature of the singularity. We then need to examine the shape of the wave normal surface
at that point. The wave vector at the singular point is given by

k 0  ( k0 x , k0 y , k0 z ), with

 n n 2y  n x2 kz  / c
k0 x  n y sin   z
c c n n
2
z
2
x
ny
k0 y  0 nx k0

 n n n
2 2 kx  / c
k0 z  n y cos   x z y
ny nz
c c n n
2
z
2
x

n n 2y  n x2
tan   z
nx nz2  n 2y

94
We are interested in the neighborhood of this singular point. We therefore do a Taylor
expansion of the wave normal surface at this point.
k x  k 0 x  x

Let k y  k0 y  y , substitute it into the equation of the wave normal surface
k  k  z
 z 0z

 k x2 k y2 2   2 1  2 1  2 1   2  4
  
k z
 n2n2 n2n2 n2n2  x
  n 2

1
n 2 
1
 k  k y  k z  k x     k y     k z   
2 2 2
 n2 n2   n 2
1
n 2 
c 2
 4  0,
c
 y z x z x y    y z   x z   x y 
and keep up to the second power of x, y, and z, we have
(I confirmed it by Mathematica.)

4(k0 x x  k0 z z )( nx2 k0 x x  nz2 k0 z z )  y 2 (n 2y  nx2 )(n 2y  nz2 )  0. kz  / c


ny
This is a cone with its vertex at k0. The cone is 2  2
nx 1
symmetric about the y =0 plane. 
kx  / c
ny nz

95
The cone intersects the y=0 plane by two lines, which make an angle 2 given by
2    2  1 
k0 x 
tan 1     tan  
k0 z 
2 2  (n 2y  n x2 )(nz2  n 2y )
nk n
tan  2   x2 0 x   x2 tan    tan 2   n 2 2
nz
nz k0 z nz  x

n n y  nx 
2 2

tan   z 
n x nz2  n 2y 
We then rotate the x-z axes by an angle of /2−+ to the x'-z' axes. The cone will be
erect in the new coordinates ( I confirmed it by Mathematica):
kz  / c x'
1 x'2 ny z'
z' 
2
y' 
2
2
(1  tan  )
2
tan 2  2
nx

The energy flow is everywhere perpendicular to k0 kx  / c
this cone surface, which form a light cone of ny nz
z '2  (1  tan 2  ) y '2  x'2 tan 2 
for conical refraction.
96
The light cone due to conical refraction z '2 (1  tan 2  ) y '2  x'2 tan 2 
has the following characters:
1)The cone contains the optic axis as z '   x' tan  . kz  / c x'
2)All other lights on the cone are above the optic axis ny z' 2
2
(if the z axis is upward). nx
3)The cone has an apex angle of 2 in the y'=0 plane. 
4)The cone intersects any plane that is perpendicular k0 kx  / c
to the optic axis with a circle. ny nz

Proof of 4): We rotate x'-z' axes by an angle of – to


the x"-z" axes. The cone will be
z"2  x" z" tan 2   y"2  0
It is a circle at any x"=const.

97
Lecture 9 Phase-matching in second harmonic generation
Second harmonic generation
When an intense laser beam passes through a crystal, the molecules of the crystal can be
nonlinearly polarized, which causes the crystal radiate at the doubled frequency of the
incident light. This is called second harmonic generation.
Suppose the incident light has a frequency of . It passes through a crystal with thickness
L in the x direction. Suppose the E-field of the incident light (which is called the
fundamental beam) is E ( , x, t )  E0 ( ) exp i k ( ) x  t .
We suppose the conversion to second harmonic is small so that the amplitude of the
fundamental beam is almost constant. We take care of the polarization issue later.
The second order nonlinear polarization of the material is
P ( 2 ) (2 , x, t )   ( 2 ) (2 ;  ,  ) E ( , x, t ) E ( , x, t ).
Here (2) is the second order nonlinear optical susceptibility .
According to the theory of radiation, the complex field for
the second harmonic that is generated by the crystal inside
dx but observed at L is dx

e(2 , x, L, t ) dx  P ( 2) (2 , x, t ) dx  exp ik (2 )( L  x) x


e(2 , x, L, t ) dx  aE ( , x, t ) E ( , x, t ) exp ik (2 )( L  x) dx L
Here a is just a constant.
98
The total field of the second harmonic at the output surface of the crystal is then
E (2 , L, t )  a  E ( , x, t ) E ( , x, t ) exp ik (2 )( L  x) dx
L

 aE ( ) exp  i 2t   exp i 2k ( ) x  ik (2 )( L  x) dx


L
2
0
0

sin  kL / 2  L
 aLE 02 ( ) expi 2k ( )  k (2 )  exp  i 2t 
kL / 2  2
4
Here k  k (2 )  2k ( )   n(2 )  n( ) is the phase mismatch.

(I believe it should be called the wave-vector mismatch, and kL should be called the
phase mismatch. This mistake was made by somebody many years ago.)
The intensity of the produced second harmonic is
2 2 sin 2  kL / 2
I (2 )  E (2 , L, t )  a L I ( ) 2 2
.
 kL / 2 2

For efficient second harmonic generation:

1)Conversion efficiency I ( 2 )
 I ( ).
I ( )
2)Phase-matching condition k  0, that is n(2 )  n( ).
This requirement is critical due to the narrow shape of the sinc function.
99
Phase-matching in BBO crystal
BBO (Beta-Barium Borate, β-BaB2O4) is a negative
uniaxial optical crystal, i.e., ne< no.
x2  y2 z2
The index ellipsoid is 2  1
no ( ) ne2 ( )
with the z-axis as the optic axis.
Let  be the direction of the light wave, the refractive
index of the e light, ne(, ), is then
1

cos  sin 
2 2
1  cos  sin  
2 2 2
   ne ( ,  )   2  2 
no2 ( ) ne2 ( ) ne2 ( ,  ) n
 o ( ) ne ( ) 

Phase-matching condition requires that n(2 )  n( ) . This is impossible if both the
fundamental and the second harmonic are o-rays, or both are e-rays, because of the
dispersion of the material. However, since ne< no, and normally no,e (2 )  no ,e ( ) , we
can let  be o-ray and 2 be e-ray, and hopefully to accomplish the phase-matching
condition. That is o + o  e, which is called type-I phase-matching. For contrast, type-
II phase matching refers to o + e  o or e.
100
To achieve type-I phase-matching in BBO, the phase-matching angle m, where k=0,
is given by
ne (2 , m )  no ( ) 
1 no 2 ( )  no 2 ( 2 )
 
 cos2  m sin 2  m  2    m  arcsin n  2 (2 )  n  2 (2 ) .
ne (2 , m )   2  2   e o
n
 o ( 2 ) ne ( 2 )  

This is not hard to calculate because we have the analytic Sellmeier equations for the
refractive indices as a function of the wavelength or frequency. The Sellmeier
equations of BBO are (in m):
no2() = 2.7359+0.01878/( 2-0.01822)-0.01354  2
 
m degree

Phase-matching angle
80
n () = 2.3753+0.01224/( -0.01667)-0.01516 
e
2 2 2 in BBO crystal
60

We then find that the BBO crystal can achieve 40

the phase-matching condition for input


wavelengths down to a little more than 400 nm.


20

 m
0.4 0.6 0.8 1

101
Angular sensitivity of phase-matching:
4
k ( ,  )   ne (2 ,  )  no ( ) I (2)
1.0

k k
k ( ,  )  k ( ,  m )    ...   . 0.5
  m   m

k 2 3  1 1 
 no ( )  2  2  sin 2 m
  m  n
 o ( 2 ) ne ( 2 ) 
2.783 kL
sin  kL / 2  1
2
k
  kL  2.783    L  2.783 
 kL / 2 2 2   m
1
2.783  3  1 1  
2    no ( )  2  2  sin 2 m
L    no ( 2 ) ne ( 2 )  

This is only 0.11°for a BBO crystal of 1-mm long with an input beam at 600 nm.
Therefore the phase-matching condition is extremely angular-sensitive.
In practice the experimental angular width of phase matching can be larger, because 1)
The angle is measured outside the crystal. 2) The laser has a bandwidth. 3) Intensity
saturation. 4) A Gaussian beam has a divergence in its propagation directions.
Similarly we can also calculate the wavelength sensitivity for phase-matching. It is
found to be about 1 nm for 1-mm thick BBO crystal, which is again quite sensitive.
102

Angular and wavelength sensitivity
of phase-matching in BBO crystal

m, ,  degree

kL=0
40.7
40.6 

40.5 kL=±2.783


40.4
40.3
 m
0.599 0.6 0.601 0.602

103
Lecture 10 Optical activity
4.9 Optical activity
Optical activity refers to the phenomena that when a linearly polarized light is passing
through a medium, the polarization plane is rotated. It is thus also called optical rotation.
When this happens, the medium is said to be optically active. The rotation of the
polarization plane is proportional to the path length of the light, and can be measured by
degree/centimeter. Common optically active media include quartz, sugar and syrup. It can
be used to measure blood sugar concentration in diabetic people.
Optical activity occurs in a chiral molecule, where the molecule does not overlap with its
mirror image, and the electron cloud takes some kind of helical shape. Left- and right-
circularly polarized light is different in polarizing a chiral molecule, which causes different
refractive indices. This is called circular birefringence (or circular double refraction).
Most amino acids (the building blocks of life) are left-handed, which makes optical
activity prominent in nature.
If when observed facing the light the
rotation of the polarization plane is
counterclockwise, the substance is
dextrorotatory (right-handed). When
the rotation is clockwise, it is called
levorotatory (left-handed).
104
Explaining optical activity using circular birefringence:
Suppose light is linearly polarized on the x-axis, and is propagating along the z-axis.
Suppose the refractive indices for circularly polarized light is nr and nl.

At z  0 : E(0)  E0 xˆ exp it  


2

E0 ˆ ˆ

R  L exp it 

E0 ˆ  nr   n 
At distance z : E( z )   R expi z   Lˆ exp  i l z  exp it 
2   c   c 


E0 
  xˆ  iyˆ  exp  i nr z    xˆ  iyˆ  exp  i nl z  exp it 
2   c   c 
E    n   n     n   n   
 0 xˆ exp  i r z   exp  i l z   iyˆ exp  i l z   exp  i r z   exp it 
2    c   c    c   c  
E    (nl  nr ) z  (n  nr )z    (nl  nr ) z  (n  nr )z  
 0 xˆ 2 cos exp  i l   iyˆ  2 sin i exp  i l   exp it 
2   2c  2c   2c  2c  
  (nl  nr ) z  (nl  nr ) z   (nl  nr )z  
 E0 xˆ cos  yˆ sin  exp i
  t  
 2c 2c   2c 
 (nl  nr ) 
The specific rotatory power is   (nl  nr ),
2c 
which is very sensitive in measuring circular birefringence.

105
Theory on optical activity:
A helical molecule can be polarized along its axis by a circulating current driven by an
circular electric field produced by a time-varying magnetic field: p  E   H
The constitutive equation of a medium driven by a plane wave is then revised to
D  E  i 0 Gs  E, because H
    E, and   is

Gs=G is called the gyration vector. The length G of the gyration vector describes the
rotation power of the medium. It varies with the direction of the wave, and can be
expressed by a gyration tensor g of the medium as G  s  gs  g ij si s j .

Now our original eigenwave equation   n 2 0s  s  E  0 is revised to


  i Gs  n  s  s  E  0
0
2
0

After realizing it into the matrix form, we have


 2 G G 
 x
s  (1  n 2
x / n 2
) s s
x y  i 2
s z s x s z  i 2
s y 
 n n  E x 
 s s i G s G     0  sz sy 
s y  (1  n y / n )
2 2 2
s y sz  i 2 sx  E y   0  
 y x n2 z n  s   sz 0  sx 
 G G  E z   s
 y sx 0 
 sz sx  i 2 s y sz s y  i 2 sx s z  (1  nz / n ) 
2 2 2
s  s  ss  1
 n n 

106
For a nontrivial solution of E, the determinant of the matrix must be 0, which is
simplified to
s x2

s y2

s z2

1


G 2 nx2 s x2  n 2y s y2  nz2 s z2  (*1)
  
n 2  nx2 n 2  n y2 n 2  nz2 n 2 n 2 n 2  nx2 n 2  n y2 n 2  nz2 
The corresponding wave normal surface is
6
2  
2
kx
2
ky 2
kz
G   
nx2 k x2  n y2 k y2  nz2 k z2 
   1  c
2 2 2 2 2 2 2 2  2 2  2  2 2  2  2 2 
k  2 nx k  2 n y k  2 nz
2 2 2
k  k  2 nx  k  2 n y  k  2 nz 
c c c  c  c  c 
I believe this surface is formidable for anyone attempting to draw using Mathematica.
2 2 2
 2

Fortunately equation 1 can be simplified to n  n1 n  n2  G , where n1 and n2
2

are the solutions of the equation with G=0, i.e., when there is no optical activity.
On the optic axis, we have n1  n2  n , the eigen refractive indices are then
G
n 2  n 2  G, and n  n  when G is small.
2n
G
These are two circularly polarize waves, with a rotary power of   .
n
107
The eigen polarization state in an optically active medium:
If we solve equation 1 for the two general eigen refractive indices, and then substitute into
the eigenwave equation, we may have the eigen polarization states for light in an optically
active medium. Surely it is too complicated. Our textbook did it in a simple way. We can
rewrite the eigenwave equation into that for the D vector, and express it in the (D1, D2, s)
coordinate system, thus reducing one dimension. The corresponding eigen polarization,
when G is small, and in terms of D, is
D2
  2 2 
 1  1 1  1 1 1   G  
 2  2    2 2  

 2  n2 n2  
J    1 2  4  n1 n2   n1 n2 
 D1
 iG 
  2 2 
 n1 n2 

We conclude that
1)The eigen polarization states are generally two orthogonal elliptically polarized
waves, oriented along the D1 and D2 axes.
2)In an isotropic medium, or when the light is propagating in the optic axis of an
anisotropic medium, the eigen polarizations are left- and right- circularly polarized
light.
3)In an anisotropic medium, and when the light is not close to the optic axis, because G
is usually much smaller than  n2light
n,12 the 2
is almost linearly polarized.
108
4.10 Faraday rotation
Faraday rotation refers to the phenomena that when a linearly polarized light is passing
through a medium placed in an external magnetic field, which is along the light
propagation direction, the polarization plane is rotated.
The origin of the effect is as follows. The electrons are moving in a molecule, driven by
the electric field of the light. The external magnetic field will displace the electrons
laterally due to the Lorentz force qv×B. The induced dipole momentum then includes a
term proportional to E×B. The constitutive equation is then
D  E  i 0B  E
Hereis the megnetogyration coefficient, B is the gyration vector. The discusses
followed should be similar to optical activity, with a specific rotation given by   VB ,
where V is called the Verdet constant, can be measured by deg/Gauss·mm.
One distinct difference between optical activity and Faraday effect is the reversibility of
the effect when the light is going back. Optical activity is reversible, while in Faraday
effect when the light is reversed the rotation is doubled. This is ready to explain by the
two constitutive equations:
 k
D  E  i 0 Gs  E, optical activity B
D  E  i 0B  E, Faraday effect k
B
k
d B
109
Optical isolator:
An optical isolator, or optical diode, is an optical device that allows the transmission of
light in only one direction. It is used to prevent unwanted light feedback into a laser
cavity. The operation of the device depends on the non-reversibility of Faraday effect.
The optical isolator in the figure consists of three
parts. An input polarizer allows only vertically
polarized light to pass. A Faraday rotator rotates
the polarization by 45°. An output polarizer
(analyser) just let the 45° polarized light to pass
by. If light is feed back from somewhere later in
the path, the Faraday rotator will turn it into
horizontally polarized light, which is then
absorbed (or deflected) by the first polarizer.

110
Chapter 5 Jones Calculus and Its Application
to Birefringent Optical Systems
Lecture 1 Wave plates
Wave plates (retardation plates) are optical elements used to transform the polarization
states of light. They are made from one or more pieces of birefringent crystals.

•Let us consider a plate made of a uniaxial crystal with a thickness of l. Usually the plate
is cut so that its optic axis lies in the plane of the plate surface.
•For a normally incident light, the polarization directions of the two eigenwaves both lie
in the surface of the plate, and are mutually orthogonal. One polarization direction
coincides with the optic axis, with a refractive index ne. The other is perpendicular to the
optic axis, with a refractive index no.
•The polarization direction with the larger refractive index is called the slow axis, and the
polarization direction with the smaller refractive index is called the fast axis, regardless of
whether it is an ordinary light or an extraordinary light. The refractive indices are then
designated as nf and ns, respectively.

111
In the frame of the slow and fast axes, suppose the input light is linearly polarized with the
field
 
Ein  Es sˆ  E f fˆ eit , where Es and E f are both real.
At the output surface, the light field is changed into
  nsl  ˆ  n f l  ˆ  it
Eout   Es exp  i s  E f exp  i f  e .
  c   c  
The phase difference (retardation or retardance) between the f and s polarization is then

l
Γ   ns  n f 
c
When Γ  (2m  1) , the plate is a half - wave plate.
When   (2m  1 / 2) , the plate is a quarter - wave plate.

f f f
l
s s
s

112
Half-wave plates and quarter-wave plates
1) A half-wave plate converts a linearly polarized light into another linearly polarized
light, mirrored by the fast or slow axis.
2) A quarter-wave plate converts a linearly polarized light into a circularly polarized
light, when the input polarization is 45° to the fast and slow axes. At other azimuth
angles it converts a linearly polarized light into an elliptically polarized light oriented
along the fast or slow axes.
f f f f
s s s s

Half-wave plate Quarter-wave plate

f f f f
s s s s

Quarter-wave plate Quarter-wave plate

113
Zero-order and multiple-order wave plates
2
A half-wave plate with a retardation of    ns  n f l  (2m  1) with m  0

is called a zero-order half-wave plate. The thickness of the plate is
 500 nm
l   25m (typically for crystal quartz).
2( ns  n f ) 2  0.01
This thickness is apparently not easy to fabricate and not easy to handle.

We can use a thickness of l  (2m  1) , which is a multiple-order wave
plate. 2( ns  n f )
   2
The wavelength sensitivity of a half-wave plate is l  n  n  
f   l  2m  1.
  
s
 
Therefore a multiple-order wave plate has a limited bandwidth.
One technique to solve this problem is to make a compound zero-order wave plate
from two plates, with their optic axes intercrossed. The difference in thickness
between the two plates determines the overall retardation.  Zero-order wave plates
thus have broad bandwidths.
2
  ns  n f  l1  l2 

114
Reading: Achromatic wave plates: y (e)
2
( , l1 , l2 )   ns1 ( )  n f 1 ( ) l1   n f 2 ( )  ns 2 ( ) l2  y (o)

MgF2
Estimation of l1 and l2 (e.g., QWP):
In the wavelength range considered, the phase x (o)
retardation should be as close as possible to /2 Crystal Quartz x (e)
(minimum rms, or similar criteria).
2
2 1000 nm  
g (l1 , l2 )   ( , l1 , l2 )   d
1  700 nm  2

g (l1 , l2 ) 
 0
l1 
g (l1 , l2 )   l1=603 m, l2=477 m.
 0
l2 

Question: Why do we need two materials?


( , l1 , l2 ) 2   dns1 dn f 1   dn f 2 dns 2  

  2 (ns1  n f 1 )l1  (n f 2  ns 2 )l2     
l1     l2 
    d d   d d  
For a normal compound zero-order wave plate (one material), l1-l2 is fixed, thus
 /  is fixed. For achromatic wave plates (two materials), l1 and l2 can be chosen
to minimize  /  , which greatly expands the bandwidth.
115
Lecture 2 Jones matrix
5.1 Jones matrix formulation
While it is not difficult to track the polarization of light passing through an individual
wave plate or polarizer using junior algebra, when there is a combination of several such
optical elements, and a certain goal is aimed, the algebra involved can be complicated.
Jones calculus is created to study the transformation of polarization using linear algebra,
where the polarization of light is represented by a Jones vector, and the function of an
optical element is represented by a 2×2 matrix.
A fixed lab coordinate system (instead of the
principle axes of the crystal) is normally used. The
azimuth angle of the retardation plate is defined as
the angle between the slow axis and the lab x axis, or
between the fast axis and the lab y axis. The light is
propagating in the –z direction.

116
Wave plates:
Let us first derive the 2×2 matrix for a wave plate. Suppose the input light has an arbitrary
polarization state (Vx ,V y )T. At the surface of the plate this light need to be decomposed into
the two eigenwaves, which is a transformation to the s-f coordinate system.
 Vs   cos sin Vx   Vx   cos sin 
       
   R( ) , with R( )   .
V
 f   sin cos  V
 y  V
 y   sin cos  

The polarization state after the plate, in the s-f coordinates, is then
 V 's   exp  insl / c  0  Vs  i  e  i / 2 0  Vs   Vs 
      e     W0  
V '    0 exp  in  l / c  V   0 i / 2 
e V f  V 
 f  f  f    f
l
is the phase retardation, and    ns  n f 
l 1
Here Γ   ns  n f  is the average
c 2 c
 e  i / 2 0 
phase change. W0  e  i
 is the Jones matrix for the wave plate expressed
i / 2 
in  0 e 

its own principle s-f coordinate system.


In
 Vthe x-y coordinate
  sinsystem
  Vs 'the
 polarization
 Vs '  state after the plate is
x '  cos
       
V '  sin cos V '   R( )V ' 
 y   f   f 
117
The overall effect of the retardation plate is then
 Vx '  V   cos sin  i  e
 i / 2
0 
   R(  )W0 R( ) x , with R( )   , W0  e  .
i / 2 
V y '  V y    sin cos   0 e 

The e-ifactor can be dropped if we are not dealing with interference. Therefore a
retardation plate is characterized by its phase retardation  and azimuth angle  , and is
represented in the lab frame by
W  R ( )W0 R ( )
Note that the transformation is unitary: W+W=1. It does not change the inner product
between two Jones vectors. This is because a wave plate does not absorb light.
Linear polarizers (analyzers):
An ideal linear polarizer with its transmission axis on the x axis is
1 0
P0  e i  , with  as the absolute phase change.
0 0
For a linear polarizer oriented at an azimuth angle , the Jones matrix is
P  R( ) P0 R ( )
1 0 0 0
Particularly Px   , Py   .
0 0 0 1
118
Combination of wave plates and polarizers:
For a series of wave plates and polarizers, we need to just multiply the Jones matrix
of individual element in sequence.
Wave plates examples:
Half-wave plate:
A half-wave plate with its slow axis oriented at =45°. The input light is linearly
polarized in the vertical direction:
1 1  1 e i / 2 0  1  1 1  0  i 
W  R( )W0 R( )    
i / 2 
    
2 1 1  0 e  2   1 1   i 0 
 0  i  0    i  1
V '         i  is a horizontal ly polarized light.
  i 0  1   0  0

For a general azimuth angle , a half-wave plate will rotate a horizontally or vertically
polarized light by 2.
A half-wave plate will change a left-handed circularly polarized light into a right-handed
circularly polarized light, and vice versa, regardless of the azimuth angle of the plate.

119
Quarter-wave plate:
A quarter wave plate with its slow axis oriented at =45°. The input light is linearly
polarized in the vertical direction:
W  R (  )W0 R ( )
1 1  1 e i / 4 0  1  1 1
   
i / 4 
 
2  1 1  0 e  2   1 1
1  1  i Our textbook
   is wrong here.
2  i 1 
1  1  i  0  1   i   i  1
V'        
2   i 1  1  2  1  2  i 
is a left - handed circularly polarized light.

If the input light is horizontally polarized, it


will be changed into a right-handed
circularly polarized light.

120
5.2 Intensity transmission
For a Jones vector J   E x , E y  , the corresponding electric field is E   E x xˆ  E y yˆ. e it .
The absolute intensity of the light is then
1 2 1 2 2
I   E x xˆ  E y yˆ v  v( E x  E y ).
2 2
For convenience, if we only care the relative intensity of light in one medium (e.g., air),
the factor (1/2)v is a constant and can be dropped. We therefore define the intensity of
light as
2 2
I  E x  E y  E   E  J   J.
 Ex '   Ex 
For a Jones matrix transformation    M   , the intensity transmittance is then
 Ey '  Ey 
2 2
Ex '  E y '
T 2 2
.
Ex  E y

121
Intensity transmission examples:
A birefringent plate sandwiched between two parallel polarizers:
Suppose the transmission axes of the polarizers are both vertical. The slow axis of the
birefringent plate is oriented at 45° from the x axis. The plate introduces a phase retardation
d

Γ  ns  n f  
 2 ns  n f 
d
c 
The Jones matrix is
1 1  1 e i / 2 0  1  1 1  cos  / 2  i sin   / 2 
W  R( )W0 R( )    
i / 2 
    
2 1 1  0 e  2  1 1    i sin   / 2  cos   / 2  
Let the incident light be unpolarized, with unit intensity. After the first polarizer, the
1  0
Jones vector is  . The electric field of the final transmitted light is
2 1
 0 0  cos  / 2  i sin   / 2   1  0  1  0 

E'   

 
 
 
  
 
 0 1   i sin   / 2 cos  / 2  2  1  2  cos  / 2 

The intensity transmittance is


 1
T  E '  E '  cos2  cos2   ns  n f  .
1  d
2 2 2  
122
A birefringent plate sandwiched between two crossed polarizers:
As above, but let the transmission axis of the final polarizer be horizontal. The electric
field of the final transmitted light is
 1 0  cos  / 2   i sin   / 2   1  0   i  sin   / 2 
E '         
 0 0   i sin   / 2 cos  / 2  2  1  2 0 

The intensity transmittance is


 1
T  E '  E '  sin 2  sin 2   ns  n f  .
1  d
2 2 2  

This is complementary to the case of parallel polarizers.


In both cases the transmission is a sinusoidal function of wave number 1/ .

123
Chapter 7 Electro-optics
Lecture 1 Linear electro-optic effect
7.1 The electro-optic effect
We have seen that light propagating in an anisotropic medium can be decomposed into its
normal modes, or eigenwaves, which can be determined by the index ellipsoid:
x2 y2 z 2
   1.
nx2 n y2 nz2
Here nx2   x /  0   11 /  0 . If we define the impermeability tensor ij  ( ) ij /  0 , the
1

index ellipsoid in a general case will be  ij xi x j  1.


In certain crystals, the application of an external electric field will redistribute the charges
in the molecules, thus results in a change in the size and orientation of the index ellipsoid.
This is called the electro-optic effect.
In the presence of an applied electric field E, the index ellipsoid is changed to
ij (E) xi x j  1

ij (E)  ij (0)  rijk Ek  sijkl Ek El  

Here rijk are the linear (Pockels) electro-optic coefficients, and sijkl are the quadratic
(Kerr) electro-optic coefficients.
124
From thermodynamic considerations the linear electro-optic coefficients rijk and the
quadratic electro-optic coefficients sijkl have the following permutation symmetries:
rijk  rjik
sij kl  s ji kl  sij lk  s ji lk

We therefore introduce the contracted indices to replace the paired interchangeable


1  (11), 2  (22), 3  (33), 4  (23)  (32), 5  (13)  (31), 6  (12)  (21).
indices:
Examples : r(12) k  r( 21) k  r6 k , r( 22) k  r2 k , s(13)( 22)  s(31)( 22)  s52

The permutation symmetry reduces the number of the independent elements of rijk from 27
to 18, and that of sijkl from 81 to 36.
A linear electro-optic effect is also called the Pockels effect, in which the change in the
refractive index n  E . Pockels effect is a second order nonlinear optical phenomenon.
It only occurs in crystals that do not possess the inversion symmetry.
For a crystal with the inversion symmetry, we can equivalently inverse the coordinates.
 
In the orriginal coordinates rijk  ij 
Ek 
 'ij  'ij 
In the inversed coordinates r 'ijk     rijk  0.
E 'k  ( Ek ) 
Inverse symmetry  r 'ijk  rijk ,  'ij   ij 

 125
7.2 The linear electro-optic effect
In an external electric field E the index ellipsoid is deformed into
 ij (0)   ij  xi x j  1, or
1  1  2 1  1   2 1  1   2  1   1   1
 2    2   x   2    2   y   2    2   z  2 2  yz  2 2  xz  2 2  xy  1.
 nx  n 1   n y  n  2   nz  n 3   n 4  n 5  n 6

For the linear electro-optic effect, the change in the coefficients is given by
 r11 r12 r13 
 
 r21 r22 r23 
 Ex 
 1  3   1    r31 r32 r33  
 2    rij E j , or   2 
 n 

 r
 E y 
 n i j 1  1, 2 , 3, 4 , 5, 6  41 r42 r43  
  Ez
 r51 r52 r53  
r r62 r63 
 61
Here rij is the linear electro-optic (or Pockels) coefficient, and r is called the electro-
optic tensor. Depending on the symmetry of the crystal, many of the elements of the
electro-optic tensor are 0, and some of them have the same or opposite values.
Table 7.1 in the textbook lists the point groups of crystals. We have 7 crystal systems,
which lead to 32 point groups. Table 7.2 lists the non-vanishing linear electro-optic
coefficients of all the point groups. Table 7.3 gives the value of the linear optic-optic
coefficients of some crystals. 126
From Poon and Kim, Engineering Optics with MATLAB.

127
Example 1:
KDP (KH2PO4) crystal (negative uniaxial, 42m point group).
The E-field is in an arbitrary direction.
0 0 0 0 
   
0 0 0  0 
 E 
  1   0 0 0  x   0 
  2    E y    
 n   r 0 0    r41E x 
 1, 2 , 3, 4 , 5, 6  41
Ez  
0 r 0   r E 
 41
  41 y

0 0 r  r E 
 63   63 z 
x2 y2 z2
   2r41E x yz  2r41E y xz  2r63 E z xy  1.
no2 no2 ne2
The index ellipsoid is deformed and the xyz axes are no longer the principle axes.

i) For the KDP crystal, if the E field is in the z direction, we have


x2 y2 z2
   2r63 E z xy  1.
no2 no2 ne2

The index ellipsoid is tilted in the x-y plane.

128
It seems that we can simplify the problem by choosing a new
principle axis system. We rotate the old coordinate system in y
y' x'
the x-y plane counterclockwise by 45° and construct the new
x′,y′,z′ system, that is
x
1
x  x ' y '  , y  1  x ' y '  , z '  z
2 2
The index ellipsoid in the new coordinate system is
1  2  1  2 z '2
 2  r63E z  x '   2  r63 E z  y '  2  1.
 no   no  ne

x '2 y '2 z '2 n3  1 


If we write it as 2  2  2  1, and considering n     2 ,
n x ' n y ' nz ' 2 n 
 1 3
n
 x'  no  no r63E z
2

 1
then n y '  no  no3r63E z
 2
nz '  ne

129
ii) For the KDP crystal, if the applied field is in the x direction, then
x2 y2 z 2 z
2
 2  2  2r41 E x yz  1.
no no ne z'
The index ellipsoid is tilted in the y-z plane. We rotate the nz'
old coordinate system in the y-z plane counterclockwise by
angle , and construct the new x′,y′,z′ system. Suppose the  y
index ellipsoid in the new coordinate system is
 2no2 ne2 r41E x ny' y'
 tan 2   n 2  n 2
 o e

n x '  no

x' 2
y' 2
z' 2  1 1 1 1  1 1 
2 
   1, then    2  2   2  2    2r41E x  
2
n x2' n 2y ' nz2'  n y ' 2  no ne
2
 no ne  
  
  2 
1
    1 1 1  1 1 
  2  2    2r41E x  
2
 nz2' 2  no2 ne2  no ne  
  
 no5ne2 r412 E x2 If ax 2  by 2  cxy  Ax '2  By ' 2 , then
n y '  no  2(n 2  n 2 ) ,
 o e
tan 2 
c
For small E x , this leads to  5 2 2 2 ab
n '  n  ne no r41E x .
 z e
2(no2  ne2 ) ( a  b)  ( a  b) 2  c 2
 A, B 
2 130
Example 2:
LiNbO3 crystal (negative uniaxial, 3m point group).
Suppose the E-field is in the z direction.
 0  r22 r13   r13 
   
 0 r22 r13   r13 
 0 
  1    0 0 r33    r33 
  2    0     E z 
 n    0 r51 0    0 
 1, 2 , 3, 4 , 5, 6 
Ez
 r 0 0    0 
 51  
 r 0 0  0
 22
 1   1   1 
 2  r13 E z  x 2   2  r13 E z  y 2   2  r33 E z  z 2  1
 no   no   ne 
x2 y2 z 2
If we write the new index ellipsoid as 2  2  2  1, then the new coefficien ts are
nx n y nz
 1 1 1 3
 n2  2
 r E
13 z  n x, y  no  no r13 E z
 x , y no 2

 1  1  r E  n  n  1 n3r E
 n
 z
2
ne
2 33 z z e
2
e 33 z

131
Lecture 2 Electro-optic modulation
7. 3 Electro-optic modulation
In the example of the KDP crystal, if the external E-field is in the z direction, and the light
is propagating in the z direction, the birefringence is
n y '  nx '  no3 r63 E z
Suppose the thickness of the plate is d, and the voltage applied is V=Ezd. The phase
retardation is
2
  n y '  nx'  d  2 no3r63V .
 
Since the retardation is proportional to the applied voltage, we can consequently convert
the polarization state of the incident light into a desired polarization by choosing the
appropriate voltage. This is called electro-optic retardation.
V 
For practical use, we write the retardation as    , where V  3 is called the
V 2no r63
half-wave voltage, that is the voltage for obtaining a phase retardation of .

Example:
For KDP at =0.55 m, no=1.50737, r63=10.6×10-12 m/V, we have V=7.5 kV.

132
Electro-optic amplitude modulation
A typical arrangement of an electro-optic amplitude modulator is shown in the figure. A
KDP crystal is placed between two crossed polarizers. Also in the light path there is a
wave plate that introduces a fixed retardation of /4. The input light is polarized in the x-
direction. A voltage of V  Vm sin mt is applied in the z direction of the crystal. This is
called longitudinal electro-optic modulation since the applied electric field is parallel to
the direction of light propagation. The total phase retardation is
 V 
    m sin mt   m sin mt.
4 V 4
The transmission of the modulator is
I out   1
 sin 2   sin 2   m sin mt   1  sin  m sin mt  .
I in 4  2
I out 1
For m  1,  1  m sin mt .
I in 2
Therefore a small sinusoidal voltage will cause a sinusoidal modulation of the transmitted
light intensity.

133
134
135
Transverse electro-optic modulation:
We now consider the case where the applied electric field is perpendicular to the
direction of light propagation, which is called the transverse electro-optic
modulation.
As shown in the figure, in a KDP crystal, the input light is polarized in the x′-z plane at
45° to each2of the axes. The 2 retardation
 1at the output
 2plane
 is then no3 r63 V
transverse  ( nz  nx ' )l   ne  no  no r63 E z l 
3
 ne  no  l   l
   2    d
Here d is the thickness along which the electric field is applied.
2no3 r63V
For comparisio n, longitudinal 

Compare to longitudinal electro-optic modulation, the advantages of transverse electro-
optic modulation are
1) The retardation can be increased by using a longer crystal, or multiple crystals.
2) The filed electrodes do not interfere with the incident light beam.

136
Yariv, Quantum Electronics, 3rd edition.

137
Phase modulation of light
In the case of KDP, if the external field is in the z direction, and the polarization of the
input light is only in the x' direction, then the applied electric field will change the phase of
the light, instead of its polarization. Suppose the field of the input light is Ein  A cos t ,
the applied electric field is Em sin mt , then the field of the output light is
 2  1 3  
Eout  A cos t   no  no r63 Em sin mt d  (omit the constant phase)
   2  
  3 
 A cos t   sin m t     no r63 Em d , phase modulation index 
  
 A[ J 0 ( ) cos t  J1 ( ) cos(  m )t  J1 ( ) cos(  m )t
 J 2 ( ) cos(  2m )t  J 2 ( ) cos(  2m )t  ]
We see that lights at side bands are generated

138
139
Reading: Lecture 3 Quadratic electro-optic effect
7. 5 Quadratic electro-optic effect
The quadratic electro-optic effect is a third order nonlinear effect, where the change in the
refractive index is proportional to the square of the applied fields. Unlike the linear
electro-optic effect, quadratic electro-optic effect occurs in crystals with any symmetry.
Using contracted indices, the index ellipsoid in the presence of quadratic electro-optic
effect is

Table 7.4 lists the non-vanishing quadratic electro-optic coefficients of all the point
groups. Table 7.5 gives the values of the quadratic optic-optic coefficients of some
crystals. 140
Example 1:
Kerr effect in an isotropic medium. Let the z axis be along the applied electric field.
 s11 s12 s12  0 
  
 s12 s11 s12  0 
  1    s12 s12 s11  E 2 
  2     
 n    ( s11  s12 ) / 2  0 
 1, 2 , 3, 4 , 5 , 6 
 ( s11  s12 ) / 2  
  0 
 ( s11  s12 ) / 2  0 

 1 2 2  1 2 2  1 2 2
 2  s12 E  x   2  s12 E  y   2  s11 E  z  1.
n  n  n 
 1 3
 n  n  n s E 2
x y
2 2
z 2
 o
2
12
If we write it as   1, then 
no2 ne2 n  n  1 n 3 s E 2


e
2
11

1
The birefringence is ne  no  n 3  s12  s11  E 2   n3 s44 E 2  KE 2
2
Here K is called the Kerr constant of the substance.
141
Example 2:
Kerr effect in BaTiO3 (m3m). The applied electric field is ( E , E ,0) / 2 .
 E / 2 
2
 s11 s12 s12
  2
 s12 s11 s12  E / 2 
  1    s12 s12 s11  0 
  2    
 n    s44  0 
 1, 2 , 3, 4 , 5, 6 
 s44  
  0 
 s44  E 2 

 1 1 1 2 2  1 1 1   1 
 2  s11 E 2
 s12 E  x   2  s11 E 2  s12 E 2  y 2   2  s12 E 2  z 2  2s44 E 2 xy  1.
n 2 2  n 2 2  n 
x '2 y '2 z '2
Rotating by 45 in the xy plane gives 2  2  2  1, and
nx ' n y ' nz '
1 1 1
 2  2   s11  s12  E  s44 E
2 2

 nx ' n 2

1 1 1
 2  2   s11  s12  E  s44 E
2 2

 ny' n 2
1 1
 2  2  s12 E 2

 nz ' n 142
1
For s44 E 2  , we have
n2
 1 3 1 2
n
 x'  n  n  s
 2 11 12 s  E 2
 s 44 E 
 2
 1 3 1 2
n
 y'  n  n  s 
 2 11 12 s  E 2
 s 44 E 
 2
 1 3
nz '  n  n s12 E
2

 2

143

You might also like