Anda di halaman 1dari 57

PC421 Winter 2013 Chapter 3.

Optical Waveguides Slide 1 of 57

Chapter 3.
Optical Waveguides

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 2 of 57

Most of the photonic devices that we will study in PC421 incorporate optical
waveguides - structures that confine light in one or two dimensions while allowing
for propagation in the remaining dimension(s). This geometry permits light to be
split, re-routed, and recombined. In this chapter, we will explore the relation between
a waveguides geometry (size and refractive index) and the properties of the optical
fields that it guides.
Most of the chapter will be concerned with slab waveguides, in which confinement
exists in only one dimension. These are relatively easy to analyze - although this still
require numerical techniques (Matlab alert!) Towards the end of the chapter, we will
encounter the more common channel waveguides, in which 2D confinement
occurs. Some approximate solutions will be presented.
Much of the material in this chapter is not in Numais text; we draw it from a variety of
sources.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 3 of 57


Outline

Table of Contents
1

Symmetric Slab Waveguides


Ray Analysis
Wave Analysis
Graphical Solution for the Characteristic Equations
Numerical Solution for the Characteristic Equations
Mode Cutoff
Low- and High-Frequency Limits
Normalized Parameters
Normalization Constant
Mode Orthogonality
Confinement Factor
Confinement Limits - Mode Field Diameter
TM Modes

Asymmetric Slab Waveguides


TE Modes
Cutoff Condition

Bending Loss

Waveguiding in a Lossy or Gain Medium

Multilayer Waveguides

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 4 of 57


Outline

Channel Waveguides
Marcatili Method
Effective Index Method (EIM)
Examples

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 5 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Ray Analysis TN 3.1 3.2


In your earlier optics courses, you learned about Snells law that relates the incident
and transmitted angles of rays at a boundary between media of different refractive
indices n1 and n2 :


n1
n1 sin 1 = n2 sin 2 , or 2 = sin1
sin 1 .
(3.1)
n2
Suppose that we are incident from medium 1, which has the higher refractive index:
n1 > n2 . For the inverse sine function above to have a real solution, we require that
the incident angle be less than a critical angle
c = sin1 (n2 /n1 ).

(3.2)

For larger incident angles, we have a condition of total internal reflection (TIR), in
which 100% of the the rays power reflects from the boundary.
Now, consider the case of a high-index dielectric core of width h sandwiched
between low-index dielectrics (for historical reasons, these are called the cladding
and substrate; in general they can both be thought of as cladding layers) (Fig. 1).
We now have two parallel boundaries. If the ray angle is greater than c at both
boundaries, then the ray will be trapped within the high-n material; it propagates by
zigzagging between the two boundaries. This is the concept of a dielectric
waveguide, on which all photonic devices in PC421 are based. For the remainder of
this chapter, we will explore the properties of this guided wave.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 6 of 57


Symmetric Slab Waveguides

Cladding Layer
Film
Substrate
Fig. 1:

(a)
Fig. 2:

Cross section of a 2D optical waveguide. TN Fig. 3.3.

(b)

(c)

Propagation modes: (a) radiation mode; (b) substrate radiation mode; (c) guided mode. TN Fig. 3.5.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 7 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Ray Analysis cont

TN 3.1 3.2

We can identify three types of propagation modes, depending on whether or not TIR
exists at one or more interfaces (it is assumed that nf > ns > nc , such that if TIR
exists at the film-substrate interface, it also exists at the film-cover interface). These
modes are shown in Fig. 2. While a few exotic devices make use of substrate
radiation modes, we will only be concerned with guided modes in PC421.
Although there are a continuum of possible ray angles that produce TIR, there are
actually only a discrete number of possible guided modes in any waveguide possibly just a single one.
Consider Figs. 3 and 4. A ray propagating at an angle f to the surface normal will
have a wavenumber k0 nf in the core material. The component of this wavenumber in
the direction of the waveguide axis (z ) is called the propagation constant :
= k0 nf sin f .

(3.3)

Now, lets keep track of the accumulated phase shifts as the ray completes one full
round trip between the two interfaces. These are equal to k0 nf h cos f (propagation
from substrate boundary to cover boundary), 2c (the phase shift upon TIR at the
film-cover interface), k0 nf h cos f (propagation from cover boundary to substrate
boundary), and 2s (the phase shift upon TIR at the film-substrate interface).

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 8 of 57


Symmetric Slab Waveguides

Fig. 3:

Coordinate system for a guided mode. TN Fig. 3.7.

Fig. 4:

Definition of the propagation constant. TN Fig. 3.8.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 9 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Ray Analysis cont

TN 3.1 3.2

To obtain a lightwave propagating through the optical waveguide without decay, the
total phase shift in the round trip must be a multiple of 2:
2k0 nf h cos f 2c 2s = 2m (m = 0, 1, 2, ).

(3.4)

Now, it would seem that there would be an infinite combination of (f , s , c ) values


that would solve this equation. However, the reflection angles s ,c depend on f ,
according to the Fresnel equations. Therefore, f uniquely specifies s ,c , and the
preceding equation has a finite number of discrete solutions, indexed by the mode
number m. The phase shifts upon reflection can be found on p. 44 of the text.
Importantly, they depend on the polarization of the ray. Therefore, so do the
allowed values of f .
Once the allowed ray angles (which are denoted as m ) are found, Eq. (3.3) and Fig.
4 can be used to find the allowed propagation constants m , transverse wave
numbers, and effective indices:
m = k0 nf sin m ,

m = k0 nf cos m ,

Nm =

m
k0

(for the latter, another common notation is neff ). It is trivial to show that
nf > N > (nc , ns ) in a TIR waveguide.

(3.5)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 10 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis


In this section, we will analyze the waveguiding properties of a dielectric slab
waveguide using Maxwells equations. This structure provides the foundation for all
other waveguides that utilize total internal reflection (TIR).
Figure 5 shows a simplified view of a slab waveguide. It consists of a high-index core
of thickness h surrounded on both sides by identical semi-infinite lower-index
claddings of index ncl . We will assume that the waveguide is infinite in extent in the
y - and z directions, with propagation along z . This implies that the dependence of
the E and H fields on y is negligible (/y = 0).

Fig. 5:

A symmetric slab waveguide

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 11 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


In the last chapter (Eq. 2.20), we derived from Maxwells equation the wave
equation,
2 E
2 E
(3.6)
2 E = 0 0 r 2 = 2 .
t
t
2
2
For time-harmonic fields, /t , so we can rewrite the wave equation as
(2 + 2 )E = 0.

(3.7)

By placing the origin (x = 0) at the center of the waveguide, we will exploit the
symmetry of the structure to find even and odd mode solutions. The refractive
indices are ncl for |x | h /2 (the cladding), and nf for |x | < h /2 (the core).

The guided modes can have two polarizations, TE (transverse electric) and TM
(transverse magnetic). In the former, the E field is parallel to the core-cladding
boundary, while in the latter, the H field is parallel to this boundary.
A waveguide mode - also called a normal mode or eigenmode - is defined as a
wave solution to Maxwells equations with all boundary conditions satisfied, for which
the transverse spatial profile of the fields and the polarization remain unchanged
during propagation. As we have seen, each mode is characterized by a propagation
constant m . It is also characterized by a particular spatial distribution of the EM
field; this latter information is not provided by the ray method of waveguide analysis.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 12 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


. As
For TE polarization, the electric field has only a y component: E = Ey (x , y , z )y
we have said, the mode propagates along z and has no y -dependence. Therefore, it
can be written

E = E y ( x ) e i z y
(3.8)
i t
is suppressed here and for most of PC421). We
(the time dependence e
automatically know the H field as well, using Maxwells equations for a
time-harmonic field - specifically, E = i H, which gives us
!
Ey
1

(3.9)
z e i z ,
i Ey x +
H=
i 0
x
where we assume that = 0 everywhere. Even and odd modes are sketched in the
figure below.

Fig. 6: Electric field profiles of the first three guided modes


in a waveguide. This particular guide is asymmetric
(nc , ns ), so the field profiles are asymmetric as well. TN
Fig. 3-12.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 13 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


TE Even Modes
We can see from eq. (3.7) that the Laplacian (i.e. 2nd derivative) of E is proportional
to -E in each medium. Therefore, E must be oscillatory or exponentially
growing/decaying (depending on the sign of the proportionality constant). For a
wave to be guided, it must have a field solution that is oscillatory inside the core
and exponentially decaying in the cladding. For the even modes, the oscillatory
portion must be a cosine function. Using our slab geometry, we can write
(
C0 e (|x |h /2) |x | h /2,
(3.10)
Ey (x ) =
C1 cos x
|x | h /2,
By substituting this solution into eq (3.7), we see that the variables , , and must
satisfy
2
2 + 2 = 2 0 0 nf2 = 2 nf2 = k02 nf2
(3.11)
c
2
(3.12)
2 + 2 = 2 0 0 ncl 2 = 2 ncl 2 = k02 ncl 2 ;
c
notice that the first of these equations agrees with the ray-analysis result, Eq. (3.5).
C0 and C1 are constants. Their values are very important! The ratio of C0 to C1 can
only take a discrete set of values while still satisfying the boundary conditions on the
field components.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 14 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


The first boundary condition is continuity of tangential E field, which in the present
case means that Ey is continuous at the core-cladding interface. This tells us that
 
h
C0 = C1 cos
(3.13)
2
(by symmetry, we only need to care about either - but not both - of the core-cladding
boundaries). The second boundary condition is continuity of tangential H field,
which in the present case means that Hz is continuous at the boundary. This tells us
that
 

,
(3.14)
C0 =
C1 sin
i 0
i 0
2
which simplifies a bit to

 

C0 =
C1 sin
.
0
0
2

(3.15)

We now have two equations in the two unknowns C0 and C1 . These can be written
"
#
#"
# "
1
cos 2h
C0
0

.
(3.16)
=
sin 2h
0
C1
0
0

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 15 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


This matrix equation has a solution only if the determinant of the 2 X 2 matrix is
equal to zero. Forcing this condition, we obtain the characteristic equation
= tan


h
.
2

(3.17)

Because the tangent function is periodic, there may exist multiple solutions (that is,
pairs of [, ]) to this equation. Each represents a guided TE mode with even
symmetry. The first two of these even modes - m = 0 and 2 - are illustrated in Fig. 6.

By the way, we now know the ratio of C0 to C1 , but we dont know either of
their absolute values. Knowing the ratio allows us to eliminate one of these
constants. The remaining constant is proportional to the amplitude of the field
that exists in the waveguide, and is irrelevant to this discussion.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 16 of 57


Symmetric Slab Waveguides

Symmetric Slab WGs - Wave Analysis cont


TE Odd Modes
For the odd modes, we can write

C0 e (x h /2)

C1 sin x
Ey (x ) =

C0 e (x +h /2)

x h /2,
|x | h /2,
x h /2.

(3.18)

Using the same procedure of matching the tangential field components (Ey and Hz )
at x = h /2, we obtain the characteristic equation
= cot


h
.
2

(3.19)

This is very similar, but not identical, to the characteristic equation for the TE even
modes.
Before the advent of modern computing, the characteristic equations were solved
graphically. We will now show how this is done. Then, we will introduce an algorithm
for numerical solution.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 17 of 57


Symmetric Slab Waveguides

Graphical Solution for the Characteristic Equations


The propagation constant of the guided modes can be found by the following
procedure. First, multiply the characteristic equations by h /2:

 

2h tan 2h ,
TE even modes,
h

 
(3.20)
=

h cot h , TE odd modes.


2
2

Now, subtract Eqn. (3.12) from Eqn. (3.11) to eliminate , and multiply by (h /2)2 .
This results in
!2
 2 
 2

h
h
k0 h 2 2
h
(3.21)
= 2 0 0 (nf2 ncl 2 )
+
=
(nf ncl 2 ),
2
2
2
2

where k0 = /c . Now we will plot the curves of Eqns. (3.20) and (3.21) on the same
graph, with normalized coordinates X = h /2 and Y = h /2; that is, we plot
(
X tan X , TE even modes,
Y=
and X 2 + Y 2 = R 2 .
(3.22)
X cot X , TE odd modes
q
k h
nf2 ncl 2 .
Note that the latter equation appears as a circular arc of radius R = 02
Every point at which the lines intersect represents a guided TE mode, alternating
even and odd and labeled, in turn, TE0 , TE1 , TE2 , and so on, the index indicating the
number of zeros in the field profile.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 18 of 57


Symmetric Slab Waveguides

h/2

0
0

4
h/2

Fig. 7: A graphical solution to determine and for TE modes. This particular waveguide supports five TE modes, shown
by the intersection of the black line with the blue (even) and red (odd) lines.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 19 of 57


Symmetric Slab Waveguides

Numerical Solution for the Characteristic Equations


From Eqns. (3.11) and (3.12), we can write
q
p
= k0 nf2 N 2 , = k0 N 2 ncl 2 .

(3.23)

Therefore, finding the discrete Xm or Ym from the graphical method leads easily to
the effective indices Nm and thus the propagation constants m .
Also, as we shall soon see, the effective index is bounded below by ncl and above by
nf . Therefore, the easiest numerical solution proceeds as follows:
1

Create a vector of trial N that extends from nf to ncl with the desired resolution
(usually 104 ). Call this N.
Use the above equations to create corresponding vectors and . These will
both be functions of N.
Using Eqns. (3.17) or (3.19), create
  vectors equal to  
h
h
and + cot
(3.24)
tan
2
2
Use any appropriate root finder to locate the values of N for which either of the
above equations are equal to zero. These values of N are the effective indices
of the guided modes, Nm . Note that the functions in step 3 have poles. It may
take a while for you to figure out how to deal with these poles when solving for
the roots.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 20 of 57


Symmetric Slab Waveguides

Mode Cutoff
A mode is said to be cut off when the wavelength is sufficiently long (i.e. is
sufficiently low) that the mode is no longer supported by the waveguide. From the
graph, it is clear that the cutoff condition for the TEm mode occurs at R = m/2.
q
That is,
k0 h

nf2 ncl 2 = m , m = 0, 1, 2, .
(3.25)
2
2
The implication is that the fundamental (TE0 ) mode has no cutoff frequency; it
exists for any values of h , , nf , and ncl , as long as the WG is symmetric.
In most applications, we desire single-mode operation. That is, we require that our
waveguide can support exactly one TE mode. In this case, we must have
q

k0 h
(3.26)
nf2 ncl 2 < .
2
2
This is accomplished by an appropriate choice of the operating frequency , the
core thickness h , and the material parameter nf2 ncl 2 ; for a fixed , single-mode
operation is forced by having a small index difference, a small h , or both.
For example, if we must operate at a wavelength 0 , with a core thickness h , then
single-mode operation requires that
 2
0
.
(3.27)
nf2 ncl 2 <
2h

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 21 of 57


Symmetric Slab Waveguides

Mode Cutoff cont


In the case that nf is just barely greater than ncl (as in glass waveguides, for
example), we use the fact that

nf2 ncl 2 = (nf ncl )(nf + ncl ) 2nf n.

(3.28)

The core/cladding index difference n required for single-mode operation under


these weakly-guiding conditions is therefore
n <

20
8nf h 2

(3.29)

Note that all of these equations (indeed, this entire section) makes the simplification
that nf and ncl are independent of frequency. Of course, this is not true, as n = n()
or n(). The dispersion of the material indices will be important for some concepts
later in the course.
For all TIR waveguides, the effective indices get smaller as the mode number
increases. That is,
nf > N0 > N1 > N2 > > ncl .
(3.30)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 22 of 57


Symmetric Slab Waveguides

Mode Cutoff cont


Below cutoff, there exist radiation modes. They are not discrete as are the guided
modes; rather, they form a continuum. With radiation modes, N < ncl . This means
that is imaginary. Since the field in the cladding has the form exp(x ), an
imaginary implies that the field is oscillatory in the cladding, rather than
exponentially decaying. This allows power to flow out of the guide during
propagation. That is, the modes radiate. We wont study radiation modes in PC421,
although we will briefly mention them in the next chapter when we discuss
multimode interference devices.
In your fiber optics course, you may have heard of cladding modes. These are
modes which are guided by TIR at the cladding/air or cladding/jacket interface (that
is, the cladding + core forms an effective waveguide core, with the environment as its
cladding). We will not consider these modes any further, as we are dealing with
integrated waveguides, for which cladding modes dont exist unless our waveguide is
specifically designed to support these modes.
Cladding modes in optical fiber are very useful for sensor applications, however. A
small change in the refractive index of the environment around the fiber can lead to a
noticeable change in the effective index of the cladding modes.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 23 of 57


Symmetric Slab Waveguides

Low- and High-Frequency Limits


In the low-frequency
limit, all modes m > 0 can be cut off. At the cut-off point, we
q

nf2 ncl 2 = m/2 and = 0, as can be seen from the graphical


solution. At this point, h /2 = m/2, and from eq. (3.11),

= 0 0 ncl = ncl /c .
(3.31)

have (k0 h /2)

Therefore, the propagation constant of the mode approaches that which would
be encountered in the bulk cladding material. This is to be expected; at cutoff,
the mode power is no longer localized at the core - nearly all of it propagates in the
cladding.
In the high-frequency
limit, R  . The graphical solution shows that, for the TEm
 
h
(m+1)
mode, 2h 2 and 2 . Therefore, from eq. (3.11),

= 0 0 nf = nf /c .

(3.32)

Here, virtually all of the power is localized to the core, so the mode propagation
constant is identical to that of the bulk core material.
These two extreme cases, and the transition between them, are illustrated on the
following slide.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 24 of 57


Symmetric Slab Waveguides

Fig. 8: Dispersion curves of the guided TEm modes. They are confined to the region between the light lines of the cladding
and core material. In the low-frequency limit, they approach the former and in the the high-frequency limit, they approach the
latter. The light lines should actually be slightly curved because the core and cladding material is dispersive. The region above
the cores light line is forbidden, as such modes would not have an oscillatory profile in the core. The region below the
claddings light line represents the continuum of radiation modes. The cutoff frequencies of modes 1 and 2 are also shown.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 25 of 57


Symmetric Slab Waveguides

Normalized Parameters
Earlier, we saw that the propagation constant is related to the effective index as

N=

.
k0

(3.33)

Furthermore, we have just seen that nf > N > ncl ; that is, the effective index is
bounded by the cladding and core material indices. This is a property of all TIR
waveguides.
Another widely-used parameter is the normalized propagation constant b . It takes
a value between zero and one:

b=

N 2 ncl 2
.
nf2 ncl 2

This parameter is plotted on the next slide, as a function of the normalized


frequency V , where
q
V = k0 h nf2 ncl 2 .

(3.34)

(3.35)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 26 of 57


Symmetric Slab Waveguides

Fig. 9: (a) Effective index of the TEm modes vs. frequency. The cut-off frequencies of the TE1 and TE2 modes are
indicated. Normalized propagation parameter b vs. normalized frequency V for the TEm modes.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 27 of 57


Symmetric Slab Waveguides

Normalization Constant
Many slides ago, we learned that the ratio of C0 to C1 must take certain values in
order to satisfy the boundary conditions on the fields at the core-cladding interface.
The absolute value of these constants does not affect the mode condition. However,
it is sometimes helpful to set the constants such that the modes are normalized.
That is, so that their total guided power is unity.
Because the mode propagates in the z -direction, we can write this power using the
Z
Poynting vector:
1
(E H )
zdx = 1.
P = Re
(3.36)
2

As mentioned previously, for the TE modes we have Hx = Ey , and thus


Z
Z

1
2
P=
Ey Hx dx =
Ey dx .
2
20

(3.37)

Carrying out the integral (which must be separated into its core and cladding
portions), we find that for both even and odd modes,
v
t
4

.
C1 =
(3.38)
h + 2

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 28 of 57


Symmetric Slab Waveguides

Mode Orthogonality
The mode solutions of a waveguide possess the property of orthogonality. We
wont prove it here, but it was probably covered in an earlier math class - for a
related example, review your PC321 notes where it was shown that the solutions to
the Schrdinger equation are also orthogonal. Mathematically, orthogonality of the
modes is expressed as
"
1
(Em Hn ) z dx dy = mn ,
(3.39)
2

where m and n are the indices of any two modes and mn is the Kronecker delta
function (equal to 1 if m = n and 0 otherwise).

Physically, the result of mode orthogonality is that we can consider all guided modes
to evolve along the propagation direction independently from each other. If we
couple light into only the nth mode of a waveguide, only the nth mode will emerge
from the other end. On the other hand, an arbitrary input can be decomposed into a
linear combination of the orthogonal modes, propagated independently, and then
re-composed at the output. Since the modes have different m , they will de-phase
during propagation, and the re-composed output field profile will no longer resemble
the input profile. In the next chapter, we will see that many devices operate on the
principle of modal interference, or modal coupling (wherein a z -dependent geometry
produces a controlled coupling among modes).

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 29 of 57


Symmetric Slab Waveguides

Confinement Factor
The optical confinement factor is the fraction of the total power residing in the core.
This is a very important concept in active photonic devices as it is generally only the
core material that provides gain, while the cladding is transparent. is calculated as:

2
R
R
1
E (x ) dx
Re(E H )
z dx
2 core
core
= R
.
(3.40)
= 1R
2

Re(E H )
z dx
E (x ) dx
2 total
total
Substituting the expressions for E and H from the beginning of this chapter, we find
1

!
that

2 cos2 (h /2)

h
i
= 1 +
even TE modes

h 1 + sin(h )
h
(3.41)
1

!
2

2 sin (h /2)

h
i
odd TE modes.
= 1 +

h 1 sin(h )
h
We see that the confinement factor is a function of the core width h , the material
indices, and the wavelength. It is implicitly a function of the mode number m, since
and are m-dependent. For a mode that is very far from cutoff (such as the
fundamental mode in a highly multimode waveguide), is very large and
approaches 1. On the other hand, close to cutoff, rapidly becomes small and is
highly reduced; here, much of the guided mode extends into the cladding.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 30 of 57


Symmetric Slab Waveguides

Confinement Limits - Mode Field Diameter


For many applications, we wish for the guided mode(s) to have as small a spatial
extent as possible. There are many reasons for this, most of which are beyond the
scope of this course. However, it should be obvious that if the guided fields are
narrow, we can have a denser packing of devices on our optical chips.
Because the modes exhibit an exponential decay in the cladding, they technically
have an infinite extent. However, we can quantify the mode field diameter as being
the core width plus the 1/e distance on either side. That is,
MFD = h +

1
2
1
or MFD = h +
+ ,

1
2

(3.42)

for symmetric and asymmetric guides, respectively (the latter case will be discussed
in a moment). These equations are valid for any mode (as long as you use the
appropriate ), but are commonly applied only to the fundamental mode.
When the core width h is several , the MFD scales with h . However, this is not the
case for core widths that are on the order of or smaller. Here, there is a rapid
increase in MFD, as shown on the next 2 slides. The reasons for this increase are
twofold. First, the portion of the field in the core is subject to the diffraction limit:
MFD > /(2nf ). Second, as the core width decreases, the WG becomes weaker,
increasing .

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 31 of 57


Symmetric Slab Waveguides

10
9
8

MFD/

7
6
5
4
3
2
1
0
0

h/

Fig. 10:

Mode field diameter (blue) vs. core thickness (both normalized by wavelength). Here, nf = 1.55 and ncl = 1.54.
The green line is the limit of MFD=h. The exact position of the minimum MFD is a function of the core-cladding index
difference, but it generally lies around h / = 1.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 32 of 57


Symmetric Slab Waveguides

Fig. 11:

As the core width (yellow shaded area) decreases, the spatial extent of the mode decreases as well, up to a limit.
Further decrease results in a rapid expansion of the mode, with a loss of confinement within the core.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 33 of 57


Symmetric Slab Waveguides

TM Modes
Recall from the beginning of this chapter that for TE polarization, we had
!
E y
1
i z
+

z .
E = Ey (x )e y and H =
i Ey x
i i
x

(3.43)

. By
For TM polarization, the magnetic field has only a y component: H = Hy (x )e i z y
Ampres equation,
E
(3.44)
H= ,
t
we can write
!
Hy
1

E=
i Hy x
z .
(3.45)
i
x
The guided modes are found as before, by matching the tangential fields (in this
case, Hy and Ez ) at the core-cladding interface. The resulting characteristic
equations are
 2  
 
h
ncl
h
h
=
tan
even TM modes
(3.46)
2
nf
2
2






h
ncl 2 h
h
=
cot
odd TM modes
2
nf
2
2

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 34 of 57


Symmetric Slab Waveguides

TM Modes and Birefringence


These equations look identical to those for the TE modes, other than the (ncl /nf )2
factor. In fact, in the TE case, we had ratios of cl /f = 0 /0 , which cancelled out.
The graphical and numerical methods for finding the TM effective indices are
identical to those used in the TE case, as long as the (ncl /nf )2 factor is included.
Because ncl < nf by definition (otherwise we wouldnt have internal reflection), it can
be seen by comparing the characteristic equations that we will always have, for a
given mode number m,
N TE > N TM .
(3.47)
m

We can therefore define the waveguides birefringence,


TE
TM
Bm = Nm
Nm
.

(3.48)

This indicates that if we input a signal with random polarization into the waveguide,
the TE and TM components will drift out of phase as the signal propagates.
Although a few devices exploit this fact, most photonic devices - particularly passive
ones - are desired to operate in a polarization-insensitive fashion. For these
devices, we need to design our waveguides with B as low as possible.
The figure on the next slide shows the confinement factors of the first few modes for
both TE and TM polarizations; note that there is a slight difference between them,
and that the difference reduces to zero at the low- and high-frequency limits. The TE
mode always has slightly better confinement within the core.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 35 of 57


Symmetric Slab Waveguides

Fig. 12:

Optical confinement factors for the TEm modes (solid) and TMm modes (dashed).

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 36 of 57


Asymmetric Slab Waveguides

Asymmetric Slab Waveguides - TE Modes


It is not always the case that the dielectric constant in the upper and lower cladding
is equal. For example, if we deposit a thin layer of high-index cover material onto a
low-index substrate, then this forms a waveguide in which the upper cladding is air.
From a technological standpoint, this is the simplest type of slab waveguide. It can
be analyzed using ray analysis, but we will only use wave analysis in PC421.
Consider the waveguide shown in Fig. 13, which has different cladding indices
labeled ns and nc . Due to the asymmetry, the field solutions will not be even or
odd modes. It no longer simplifies the situation to place the origin in the middle of
the core. Therefore, we place it at one boundary. The general solution for TE modes
is of the form

C1 e 1 x
x 0,

cos(
sin(
C
x
)
+
C
x
)
0
x h,
Ey (x ) =
(3.49)
2
3

C4 e 2 (x h )
x h.
Because Ey satisfies the wave equation in each of the three regions, we have
21 + 2 = 2 0 0 ns2 = (k0 ns )
2 + 2 = 2 0 0 nf2 = (k0 nf )

2
2

22 + 2 = 2 0 0 nc2 = (k0 nc ) .

(3.50)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 37 of 57


Asymmetric Slab Waveguides

Asymmetric Slab Waveguides - TE Modes cont


Here, we have four unknowns (C1 , C2 , C3 , and C4 ). The four equations we use to
solve these modes are the continuity of tangential fields Ey and Hz at both of of the
core-cladding boundaries. This can be set up as a 4 4 determinant equation, but it
can also be solved step-by-step as shown here.
At x = 0, continuity of Ey simply gives us C1 = C2 , while continuity of Hz results in
C3 = (1 /)C1 .

Fig. 13:

An asymmetric dielectric slab waveguide.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 38 of 57


Asymmetric Slab Waveguides

Asymmetric Slab Waveguides - TE Modes cont


We can re-write the general solution as

C1 eh 1 x

C1 cos(x ) +
Ey (x ) =

C e 2 (x h )
4

sin(x )

x 0,
0 x h,
x h,

(3.51)

where we can identify C1 as the Ey amplitude at the x h= 0 interface. Next, we


i apply
1
continuity of Ey at x = h , which requires that C4 = C1 cos(h ) + sin(h ) ,
allowing us to update the solution as
x
e 1
x 0,

i
h

cos(x ) + 1 sin(x )
0
x h,
Ey (x ) = C1
(3.52)

i
h

cos(h ) + 1 sin(h ) e 2 (x h ) x h .

Finally, we impose continuity of Hz at x = h , which requires that




1
sin(h ) + 1 cos(h ) = 2 cos (h ) +
(3.53)
sin(h ) .

Dividing both sides by cos(h ) and rearranging, we are left with the characteristic
equation
1 + 2
.
tan(h ) = 
(3.54)

1 1 2 2

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 39 of 57


Asymmetric Slab Waveguides

Asymmetric Slab Waveguides - TE Modes cont


The preceding equation can be inverted, to remind us that there may be multiple
solutions - the TEm modes:
 
 
2
1
+ tan1
+ m, m = 0, 1, 2, .
h = tan1
(3.55)

If we require that the modes be normalized to unit power, then


v
t
40
.

C1 =
h + 1 + 1
1

(3.56)

Without proof, we provide the characteristic equations for TM modes:


 2

n
n2
nf2 1 + nf2 2
s
c
tan(h ) =
nf4
2 n 2 n 2 1 2

(3.57)

s c

h = tan1

nf2 1
ns2

+ tan1

nf2 2
nc2

+ m, m = 0, 1, 2, .

(3.58)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 40 of 57


Asymmetric Slab Waveguides

Asymmetric Slab Waveguides - Cutoff Condition


For the wave to be guided, nf must be greater than both ns and nc . Lets assume that
ns > nc . When reducing until = k0 ns , 1 will vanish before 2 does. Physically,
we have reached a point where TIR only occurs at one of the core-cladding
boundaries (the one with nc ; the mode is leaky at the other boundary). At this point,
 
2
h = tan1
(3.59)
+ m.

The tan1 term represents a shift (to the right) of the curves that we found in the
graphical solution (Fig. 7). As a result, for an asymmetric waveguide, there exists
a cutoff frequency even for the TE0 mode.
Dispersion curves for the guided modes of an asymmetric waveguide are shown on
the next slide. Note that the curves are confined between the light lines of the core
and the higher-index cladding.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 41 of 57


Asymmetric Slab Waveguides

Fig. 14: Dispersion curves for modes of an asymmetric dielectric waveguide. The light lines should actually be curved
because the core and cladding material is dispersive.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 42 of 57


Bending Loss

Bending Loss
A straight waveguide made from lossless materials will in theory exhibit no
propagation loss - this is why the numerical mode solving technique produced
real-valued or N . Whenever a waveguide is curved, we have what is known as
bending loss. There are two ways to describe this phenomenon qualitatively. Fig.
15 uses the ray picture, in which a ray that bounces along the core-cladding interface
in a curved waveguide will eventually hit the interface at an angle that does not result
in TIR; therefore, it radiates. A weakly-guided WG will suffer more from bending loss
because the range of allowable TIR angles is much smaller.

Fig. 15: Explanation of bending loss by ray analysis (blue/green) and wave analysis (red). R is the radius of curvature of the
bent waveguide. SOK Fig. 2-32.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 43 of 57


Bending Loss

Bending Loss cont


A wave analysis can also be used to explain bending loss. In a straight waveguide,
the phasefronts of the mode all propagate with a velocity c /N , which by definition is
between c /nf and c /ncl . In a curved waveguide, the portions of the mode that lie
toward the inside edge of the curve will move slower, while the portions that lie
toward the outside edge must move faster; this increase and decrease depends on
the distance from the core. Because the mode extends an infinite distance from the
core, there must be a portion of the mode that is being asked to propagate
faster than the speed of light in the cladding. As this is impossible, that portion of
the mode radiates. Again, for weakly-guided WGs, the evanescent tail extends
farther into the cladding, where it is more susceptible to bending losses.
In fact, when a waveguide is curved, its propagation constant and mode profile are
perturbed as well. Even when the waveguide is symmetric, the mode will shift toward
the outside edge of the curve. We will return to this concept in the next chapter.
Solving the eigenproblem for a curved waveguide generally involves solving for the
straight waveguide but in a linearly distorted spatial domain. This wont be attempted
in PC421.
In the next chapter, we will see that curved waveguides are often required. In this
case, we must balance the problem of bending loss (for small-radius bends) with that
of propagation loss and device size (for large-radius bends).

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 44 of 57


Bending Loss

Bending Loss cont


Bending loss is a highly nonlinear function of the bend radius, in that as the bend
radius is decreased, the bending loss remains low up until a critical radius, Rc , at
which point it increases dramatically. The loss coefficient (in units of inverse length,
as with absorption) is given by
(3.60)
B = Ce R /Rc ,
where C is a constant. As B is large when R < Rc , it is imperative for designers of
photonic devices to know the critical radius for a particular waveguide, and to ensure
that all curves exceed this radius.

Rc depends strongly on the index difference between the core and cladding, as well
as the core thickness. For weak WGs (such as doped glass), it can be on the order
of several millimeters, while for strong WGs (such as semiconductor ribs), it can be a
few tens of microns.
0

loss [dB]

10

15

20
0

Fig. 16:

0.5

1.5
R/Rc

2.5

Bending loss for a curve of constant angle, for various bending radii around the critical radius. Note the log scale.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 45 of 57


Waveguiding in a Lossy or Gain Medium

Waveguiding in a Lossy or Gain Medium


When the permittivity of the waveguide materials are complex,
r = rr + i ir ,

(3.61)

(ir

(i.e. the material is lossy


> 0) or it has gain (ir < 0)), the propagation constant
r
i
becomes k = k + ik - or, using a Taylor expansion,
!
q
p
ir
g
r
i
r
k = 0 (r + i r ) 0 r 1 + i r = k r i ,
(3.62)
2r
2
p
where k r = 0 rr = k0 n is the wave vector in the absence of gain or loss, and
g = k r (ri /rr ) is the gain coefficient (which is negative, representing loss, if ri > 0).
With semiconductor lasers, it is a common situation to have gain in the core and a
slight bit of loss in the cladding. In that case, it can be shown that
g

(0) i + i (1 ) .
(3.63)
2
2

In other words, the real part of the propagation constant is unchanged by the
presence of gain or loss, while the imaginary part depends on how much of the
mode overlaps the lossy and gainy regions. It is most important to note that a mode
propagating along a waveguide with a gain coefficient g in the core only experiences
a gain coefficient of g .

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 46 of 57


Multilayer Waveguides

Multilayer Waveguides TN 3.2


The methods that we have seen for finding the guided modes of a slab waveguide
can be extended to the analysis of a multilayer waveguide - one which consists of
an arbitrary number of layers. The details of the analysis are beyond the scope of
PC421, but they can be found on pp. 50-51 of Numai. Essentially, we express
tangential field continuity at the boundary between the i th and (i 1)th layers as a
2 2 matrix, then we take the product of each of these matrices to define a matrix
for the waveguide as a whole, from which a characteristic equation is derived.
This is only important to us because active (and some passive) photonic devices
necessarily require multiple layers.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 47 of 57


Channel Waveguides

Channel Waveguides TN 3.3


Nearly all useful photonic devices use channel waveguides rather than slab
waveguides. Here, light is confined in two transverse dimensions while it propagates
in the third dimension (which is z , by convention).
If you found the numerical procedure required to solve slab guides difficult, then Im
sorry to tell you that the situation here is much worse. Whereas the numerical
solutions for slab guides were exact, there is no equivalent procedure to exactly
solve for the modes and propagation constants of even the simplest channel
waveguides other than cylindrical fibers. The main problem is that there is no such
thing as a purely TE or TM mode; all channel modes must be of mixed polarization
(a cylindrical fiber supports pure radially or azimuthally polarized modes, but these
are in fact equal mixtures of TE and TM). The polarization mixing increases as the
strength of the waveguiding increases and as the aspect ratio becomes more
symmetric (that is, square waveguides have a higher degree of polarization mixing
than do rectangular waveguides).
In the next few slides, we will examine two approximate methods for solving channel
waveguides. The first method is somewhat intuitive, but it is limited in scope. The
second method is an extension of the slab technique that we have already covered.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 48 of 57


Channel Waveguides

Marcatilis Method TN 3.3


Consider the channel waveguide shown in Fig. 17. We assume that w > h . The
refractive indices in the corners are not shown because they are irrelevant in
Marcatilis method. The rectangular shape of the core ensures that certain
polarization components are more dominant than others. The two types of mode are
called:
HEpq modes Hx and Ey are the dominant components,
EHpq modes Ex and Hy are the dominant components.
The indices p and q indicate the number of nulls in the field in the x and
y direction.

Fig. 17:

Geometry used in Marcatilis method.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 49 of 57


Channel Waveguides

Marcatilis Method cont

TN 3.3

To solve for the HEpq modes, we write equations for the Hx field component in the 5
regions of the waveguide, and assume that Hx = 0 in the corner regions. Along both
directions, the field is oscillatory in the middle portion of the guide and
evanescently decaying in the outer portions. In other words, we assume that:

C1 cos(x x + x ) cos(y y + y ) region 1,

C2 e 2 (x w ) cos(y y + y )
region 2,

3 (y h )
cos(
C
x
+

)
e
region 3,
Hx = e i z
(3.64)
3
x
x

C4 e 4 x cos(y y + y )

region 4,

C cos( x + )e 5 y
region 5,
x

where

x2

2y

+ = (k0 n1 ) ,

x2

22

+ 2 = (k0 n2 )2 , and so on.

Then we get to apply boundary conditions!


Then we get to use Maxwells curl equations to find all of the other field components!
Then we do it all over again starting with Ex to find the EHpq modes!
Lets not attempt this any further.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 50 of 57


Channel Waveguides

Effective Index Method (EIM) TN 3.3


In the effective index method, we convert the 2D problem into a succession of 1D
problems (which weve already learned how to solve). Here, instead of finding HE
and EH modes, we seek quasi-TE and quasi-TM modes. That is, we assume that
the aspect ratio is large enough - and that the guiding is weak enough - such that the
polarization is primarily of one type.
The geometry is shown in Fig. 18. It is identical to that of Marcatilis method, except
that we must specify the refractive indices in the corners, since we do not assume
that the field is zero in these regions.

Fig. 18:

Geometry used in the Effective Index Method.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 51 of 57


Channel Waveguides

Effective Index Method (EIM) cont

TN 3.3

To find the guided modes, as a first step we separate the 9-section waveguide into
columns, each of which can be solved individually as a slab guide to find its own
effective index. Then, in a second step, we treat these three effective indices as the
core and claddings of a slab waveguide, which is solved to give the overall effective
index. This is best explained using the figure on the next slide.
It is important to realize that you must flip your polarization between the two
steps. This is because a polarization that is parallel to the boundaries in step 1 is
perpendicular to the boundaries in step 2, and vice versa.
The EIM isnt used just for rectangular guides. It is also quite useful for finding the
effective indices of rib or ridge guides, which we will encounter later on. In fact, it can
even be used to analyze waveguides that have a smoothly varying index distribution.
Instead of treating them as a 3 - by - 3 problem as in the figure on the previous slide,
we turn it into an N - by - N problem. The slab methods we learned earlier in the
chapter can be extended to the case where there are multiple layers of core.
The EIM can be used to find effective indices, but it is poor at determining mode
profiles. In class, we will see an example to illustrate why this must be the case.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 52 of 57


Channel Waveguides

Fig. 19:

Effective index method. (top) slab waveguides corresponding to the left, middle, and right columns of the channel
guide. (bottom) the corresponding horizontal slab waveguide.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 53 of 57


Channel Waveguides

Examples
Realistically, all waveguide designers use computational techniques such as the
finite difference method, the finite element method, or the imaginary-distance
beam propagation method. Here, the waveguide cross-section is discretized into a
grid containing thousands of small elements, and variations of Maxwells equations
are solved in each element, pieced together by boundary conditions. The following
figures show some mode intensity profiles obtained using these methods. For all
figures, nf = 1.55, ncl = 1.50, = 1.55 , with core dimensions of 5 10 m.

Fig. 20:

Intensity profile of EH00 mode.

Fig. 21:

Intensity profile of EH10 mode.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 54 of 57


Channel Waveguides

Fig. 22:

Intensity profile of EH20 mode.

Fig. 24:

Intensity profile of EH11 mode.

Fig. 23:

Intensity profile of EH01 mode.

Fig. 25:

Intensity profile of EH30 mode.

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 55 of 57


Channel Waveguides

Examples cont
In actual fact, relatively few channel waveguides have rectangular cross-sections.
Several other geometries are shown in the following figures (in the figures, n1 = n2 ).
The mode is confined to the shaded regions).
Fig. 26 is a strip-loaded waveguide. Transverse confinement is produced in a slab
geometry, while lateral confinement is achieved by a narrow, low-index strip above
the core. Even though the core layer has no transverse structure, the evanescent tail
of the mode sees enough of the strip that lateral confinement is achieved. This
should be obvious if you analyze this WG using the EIM.
Fig. 27 is a rib waveguide. Mmmmnn, ribs... ... Sorry about that...its like a ridge
waveguide except that the strip is made from the core material (we etch away the
core material outside of the core region). The confinement isnt as strong as with the
ridge or channel WGs, but there is less propagation loss, as the mode doesnt
interact very strongly with material boundaries (which produce scattering losses).
Fig. 28 is a ridge waveguide. It is similar to a channel waveguide but with a lower
cladding index (usually air) above and beside the core. This configuration is used
when we want very strong confinement (small core, tight bending radius, etc.)

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 56 of 57


Channel Waveguides

Fig. 27:
Fig. 26:

Rib waveguide. Source: JML

Strip-loaded waveguide. Source: JML

Fig. 28:

Ridge waveguide. Source: JML

Fig. 29:

Diffused waveguide. Source: JML

PC421 Winter 2013 Chapter 3. Optical Waveguides Slide 57 of 57


Channel Waveguides

Examples cont
Fig. 29 is a diffused waveguide. A dopant material is diffused into the substrate
through a patterned mask, resulting in a graded-index profile n(x , y ) rather than a
discrete core/cladding boundary. Examples include Ti:LiNbO3 (titanium-diffused
lithium niobate) and ion-exchanged glass (in which network modifiers such as
sodium are replaced by other ions such as silver).
Obviously, the methods described in this chapter can not be used to analyze
diffused waveguides since there is an index gradient. Slab diffused WGs (where
n = n(x )) can be solved using analytical methods for some index profiles, but in
general they must be solved numerically.
Later in this course, we will look at semiconductor lasers and light-emitting diodes
that incorporate optical waveguides. We will see that their geometries are much
more complicated than those shown on the previous slide. This is because we must
confine electric current and charge carriers in addition to the optical mode.

Anda mungkin juga menyukai