You are on page 1of 381

OPTICAL COATINGS FROM

DESIGN THROUGH
MANUFACTURE

ANGUS MACLEOD

Thin Film Center Inc


2745 E Via Rotonda
Tucson, AZ 85716-5227
www.thinfilmcenter.com
info@thinfilmcenter.com
.
Copyright © Thin Film Center Inc, 1999-2007
All rights reserved
.
Macleod: Optical Coatings Page i Design through Manufacture

Contents
1 INTRODUCTION .....................................................................................................1
2 FUNDAMENTALS...................................................................................................5
2.1 Media and Waves..............................................................................................5
2.2 Interfaces and Films at Normal Incidence ........................................................7
2.3 Interfaces and Films at Oblique Incidence .......................................................11
2.4 Some Complications .........................................................................................15
2.5 Effect of Rear Surface.......................................................................................16
3 THEORETICAL TECHNIQUES ..............................................................................19
3.1 The Quarterwave Rule ......................................................................................19
3.2 Admittance Loci and the Admittance Diagram ................................................22
3.3 Symmetrical Periods .........................................................................................29
4 COLOR IN OPTICAL COATINGS..........................................................................33
4.1 Introduction.......................................................................................................33
4.2 Color .................................................................................................................33
4.3 Coatings for Color and Optimal Color Stimuli.................................................37
4.4 References.........................................................................................................41
5 MATERIALS.............................................................................................................43
5.1 Dielectrics .........................................................................................................43
5.2 Metals................................................................................................................46
5.3 Semiconductors.................................................................................................47
5.4 Metal and Dielectric Films ...............................................................................48
6 COATINGS WITH DIELECTRIC LAYERS ...........................................................51
6.1 Antireflection Coatings.....................................................................................51
6.2 Reflection-Increasing Coating ..........................................................................68
6.2.1 Quarterwave Stack ...................................................................................68
6.2.2 Broadband Reflectors ..............................................................................70
6.3 Edge Filters .......................................................................................................71
6.4 Bandstop or Minus Filters.................................................................................77
6.5 Chapter 6 References ........................................................................................79
7 MULTIPLE CAVITY FILTERS...............................................................................81
7.1 Bandpass Filters ................................................................................................81
7.2 Broad Bandpass Filters .....................................................................................81
7.3 Narrow Bandpass Filters...................................................................................81
7.4 Multiple-Cavity Filters .....................................................................................85
7.5 Higher Performance in Multiple Cavity Filters ................................................87
7.6 Tilted Performance of Narrowband Filters .......................................................94
7.6.1 Fabry-Perot Filters ...................................................................................94
7.6.2 Multiple-Cavity Filters ............................................................................96
7.7 Chapter 7 References ........................................................................................96
8 COATINGS WITH METALLIC LAYERS ..............................................................99
8.1 Induced Transmission Filters............................................................................105
8.2 Chapter 8 References ........................................................................................113
9 OBLIQUE INCIDENCE ...........................................................................................115
9.1 Metal Layers at Oblique Incidence...................................................................116
Macleod: Optical Coatings Page ii Design through Manufacture

9.1.1 Below Critical Angle ...............................................................................116


9.1.2 Beyond Critical Angle .............................................................................118
9.2 Coatings for One Plane of Polarization ............................................................119
9.3 Polarizers ..........................................................................................................119
9.3.1 Brewster-Angle Polarizers.......................................................................119
9.3.2 Plate Polarizer ..........................................................................................121
9.4 Non-Polarizing Coatings...................................................................................121
9.4.1 Reflecting or Beam-Splitting Coatings....................................................121
9.4.2 Edge Filters ..............................................................................................125
9.5 Surface Plasmon Resonances............................................................................134
9.6 Coupled plasmon-waveguide resonator............................................................139
9.7 Ellipsometry......................................................................................................141
9.8 Chapter 9 References ........................................................................................143
10 INHOMOGENEOUS LAYERS AND RUGATE FILTERS ...........................145
10.1 Chapter 10 References ......................................................................................147
11 SHORT-PULSE PHENOMENA IN COATINGS ...........................................149
11.1 Short pulses.......................................................................................................149
11.2 Group Velocity .................................................................................................151
11.3 Group Velocity Dispersion ...............................................................................152
11.4 Chirping ............................................................................................................156
11.5 Optical Coatings - Phase Change......................................................................156
11.6 Group Delay and Group Delay Dispersion.......................................................157
11.7 Chromatic Dispersion .......................................................................................157
11.8 References.........................................................................................................161
12 PERFORMANCE ENVELOPES .....................................................................163
12.1 References.........................................................................................................169
13 COATING MANUFACTURE .........................................................................171
13.1 Thermal Evaporation ........................................................................................171
13.2 Critical Areas ....................................................................................................173
13.3 Dielectric Thin Film Coating Materials............................................................177
13.4 The Energetic Processes ...................................................................................182
13.5 Recent Developments .......................................................................................189
13.6 Chemical Vapor Deposition..............................................................................190
13.7 Other Techniques ..............................................................................................191
13.8 Chapter 13 References ......................................................................................192
14 MICROSTRUCTURE ......................................................................................197
14.1 Origin of Microstructure...................................................................................199
14.1.1 Columnar Microstructure .................................................................199
14.1.2 Crystalline Microstructure................................................................202
14.2 Effects of Microstructure ..................................................................................204
14.2.1 Optical Properties .............................................................................204
14.2.2 Mechanical Properties ......................................................................207
14.2.3 Moisture Adsorption.........................................................................210
14.2.4 Moisture and Mechanical Properties................................................213
14.3 Chapter 14 References ......................................................................................214
15 MONITORING, ERRORS AND TOLERANCES...........................................217
Macleod: Optical Coatings Page iii Design through Manufacture

15.1 Optical Monitoring Techniques ........................................................................218


15.2 The Quartz Crystal Monitor..............................................................................228
15.3 Tolerances .........................................................................................................229
15.4 Multiple Cavity filters.......................................................................................238
15.4.1 Uniformity ........................................................................................242
15.5 Other Possibilities .............................................................................................243
15.6 Chapter 15 References ......................................................................................245
16 OPTICAL CONSTANT DERIVATION..........................................................249
16.1 Chapter 16 References ......................................................................................258
17 REVERSE ENGINEERING.............................................................................259
17.1 Random and Systematic....................................................................................259
17.2 Systematic Problems in General .......................................................................260
17.3 Single layers......................................................................................................261
17.4 Problems of the model in n and k extraction ....................................................265
17.4.1 Measurements...................................................................................265
17.4.2 Single homogeneous dielectric layers ..............................................266
17.4.3 Absorbing Films ...............................................................................269
17.4.4 Inhomogeneous Films ......................................................................270
17.4.5 Incorrect Models...............................................................................271
17.5 Multilayers ........................................................................................................276
17.6 Implications ......................................................................................................281
17.7 Conclusion ........................................................................................................281
17.8 References.........................................................................................................282
18 COATINGS IN SYSTEMS ..............................................................................283
18.1 General..............................................................................................................283
18.2 Reflection from a rejection filter ......................................................................285
18.3 Spectral width of measurement.........................................................................288
18.4 Tilted and cone response...................................................................................289
18.5 Multiple Reflections and Stray Light................................................................296
18.6 Effect of cement................................................................................................299
18.7 Temperature ......................................................................................................300
18.8 Transient response ............................................................................................302
18.9 Contamination Sensitivity.................................................................................304
18.10
Other Effects .....................................................................................................304
18.11
References.........................................................................................................304
19 ABSORPTION AND CONTAMINATION SENSITIVITY ...........................305
19.1 Absorption Of Light..........................................................................................306
19.2 Contamination sensitivity .................................................................................309
20 OPTICAL COATINGS IN COMMUNICATIONS .........................................313
20.1 Introduction.......................................................................................................313
20.2 Wavelength Division Multiplexing ..................................................................313
20.3 Separating the C and L bands ...........................................................................318
20.4 Gain Equalization .............................................................................................319
20.5 Antireflection Coatings.....................................................................................323
Macleod: Optical Coatings Page iv Design through Manufacture

20.6 Dispersion Compensation .................................................................................324


20.7 References.........................................................................................................326
21 ANTIREFLECTION COATINGS ON PLASTICS .........................................329
21.1 Substrate Materials ...........................................................................................329
21.2 Durability ..........................................................................................................329
21.3 Surface Treatment.............................................................................................331
21.4 Optical Coatings ...............................................................................................333
21.5 Chapter 21 References ......................................................................................336
22 BIBLIOGRAPHY.............................................................................................337
23 CHARACTERISTICS OF THIN-FILM DIELECTRIC MATERIALS ..........339
23.1 References.........................................................................................................348
24 A QUICK TOUR OF THE PROGRAM...........................................................353
24.1 A Word about Data Files... ...............................................................................364
24.2 ...about Materials Databases .............................................................................364
24.3 ...and about Conventions...................................................................................364
25 EXERCISES .....................................................................................................367
Macleod: Optical Coatings Page 1 Design through Manufacture

1 INTRODUCTION
This course attempts a systematic, coherent and connected account of optical coatings, their
design, properties and manufacture. It ranges from the fundamentals through to advanced topics
and is intended for practitioners from experienced workers, organizing and extending their
knowledge, to people new to the field. The course starts with the fundamentals because the
structure of everything that follows depends on them. Many of the terms can be defined and
expressions written in different ways that are equally valid, but also incompatible, and
conventions and definitions are, therefore, of great importance. The notes follow a logical
progression that is natural in a written account that may be later used as a reference. The course
that accompanies the notes concentrates on conveying an understanding of the concepts and of
their utility. Thus, although the course material is contained within the notes, the notes are more
extensive and the order of topics in each is slightly different. Further, a course is an interactive
process and not a predictable rehearsal of information in which the instructor plays the active
and the student the passive role. Each course is a unique and dynamic entity that adjusts itself to
the abilities and interests of a particular group. It may follow a roughly similar path each time it
is taught but the details and precise route will vary. A course that does not accommodate these
changes is equivalent to no more than a recorded reading of a book.
Almost all thin-film optical coatings depend, at least partly, on interference for their
operation. Interference effects in single films are responsible for the colors of soap bubbles, of
oxidized metal, or of oil films on water. Light reflected at the inner and outer boundaries of the
film interferes constructively if the path difference is an integral number of wavelengths, or
destructively if it is an odd number of half wavelengths. Since the path difference between the
beams depends on layer thickness and angle of incidence, the colors vary with layer thickness
and with the angle at which they are viewed. In a thin-film optical coating, interference effects of
this type, but usually more complex because more films are involved, are used to produce a wide
range of useful characteristics. In common with soap bubbles and oil films they share a
sensitivity of performance to layer thickness and to angle of incidence.
In an optical coating, the films, together with their support, or substrate, are generally solid.
The particular materials used for the films vary with the application and considerations, such as
the environment in which the coatings will be used, frequently imply that the choice of materials
should be as much for their mechanical as for their optical properties. Only when films can be
protected in use, as by a cemented cover, can optical properties be the sole criterion.
It is possible to construct assemblies of thin films which will reduce the reflectance of a
surface and hence increase the transmittance of a component, or increase the reflectance of a
surface, or which will give high reflectance and low transmittance over part of a region and low
reflectance and high transmittance over the remainder, or which will have different properties for
different planes of polarization and so on. Thin film coatings are often known by names which
describe their function, such as antireflection coatings, beam splitters, polarizers, long-wave-pass
filters, band-stop or minus filters, or which describe their construction, such as quarter-wave
stack, quarter-half-quarter coating and so on. Some of the different types of filter are indicated in
Figure 1-1 to Figure 1-3 together with some of the terms used to describe their most important
features.
Macleod: Optical Coatings Page 2 Design through Manufacture

In a thin-film assembly, or multilayer, the intensity of light reflected at each interface depends
on the refractive indices of the materials on either side and thus the intensities of the various
beams involved in the interference can be adjusted by choosing the refractive indices of the
films. The phases of the beams on the other hand can be adjusted by changing the layer
thicknesses. There are thus two parameters associated with each layer, thickness and refractive
index, which can be chosen to give the required performance, the number in a multilayer being
twice the number of layers. Complete freedom of choice is not possible since suitable coating
materials are limited. There are also reasons related to ease of production which make it
desirable to use layer thicknesses which are related to each other in a simple manner. The total
number of layers which it is possible to use is very large, but the greater the number of layers the
longer the production cycle, and the greater the probability of an unacceptable error, so that it is
better to keep the number of layers in a coating as low as possible.
Thus the number and range of the independent variables associated with a coating, is
effectively limited, and the optimum theoretical performance is, therefore, also limited as well as
there being inevitable drops in performance in manufacture due to constructional variations.
These performance limits can take different forms, such as the spectral width over which the
required performance can be maintained, the minimum bandwidth of a narrow-band filter, the
sharpness of edge of a long-wave-pass filter, the usable angles of incidence and the ripple in the
pass band, and so on. There is always a limit to the width of the spectral region over which
acceptable performance is achieved. The regions outside are often referred to as sidebands and it
is frequently necessary to add subsidiary filters to correct the performance in the sidebands.
A film in an optical coating is said to be thin when interference effects can be detected in the
light which it transmits or reflects, and thick when they can not. Of course, whether or not
interference effects can be detected, depends as much on the source of illumination and the
receiver which is used, as on the films themselves. Even without changing the wavelength, the
same film can be made to appear thick or thin, depending entirely on illumination and detection
conditions. There is, however, seldom any confusion. In normal coatings, the films will be thin
while the substrates will be thick.

Figure 1-1. An edge filter of idealized design defining the principal terms used for describing the optical
characteristic. Large bands of unwanted transmittance, like that on the left, are usually described as
sidebands.
Macleod: Optical Coatings Page 3 Design through Manufacture

Figure 1-2. A broad-band-pass filter of idealized design defining the principal terms used for describing the
optical characteristic.

Figure 1-3. A Narrow-band-pass filter of idealized design. Here the peak and center wavelengths coincide.
The small region of unwanted transmittance on the left is too small to be called a sideband and the term leak
is used instead.
Macleod: Optical Coatings Page 5 Design through Manufacture

2 FUNDAMENTALS
2.1 Media and Waves
The discussion of optical thin films is largely the discussion of interference effects caused by
partial reflections at boundaries between films and recombination of the resulting beams in
which their phase differences are significant. So that we can talk unambiguously about these
effects we need a framework of conventions and definitions.
A wave is simply a disturbance that propagates in such a way that a later disturbance can be
related to an earlier one primarily as a spatial displacement. The rate of change of the
displacement is the velocity of the wave. In general, because of an effect called dispersion, the
shape of the disturbance changes somewhat as it propagates and this complicates matters. It
becomes difficult, for example, to assign a definite velocity to the wave. We are concerned with
optical waves, which are propagating electromagnetic disturbances. The effects we are dealing
with are linear. Linearity means that when we want the resultant disturbance of a number of
optical waves we simply add the individual disturbances, the addition involving the addition of
the electric fields and the magnetic fields of the waves separately. It also means that we can
break any electromagnetic wave down into any set we like of fundamental components that are
simply added together to make up the original wave. The linear nature of the effects means that
we can carry out our calculations on the components separately and then add them to get the
final result. Since we have complete freedom of choice in respect of the components we use, we
can choose them to make the calculations as simple as possible. We therefore choose plane,
linearly polarized, monochromatic waves. In such waves the electric and magnetic fields vary as
a sine or cosine function of both time and distance along the direction of propagation. They have
no dependence on the spatial coordinates normal to the propagation direction. The electric
vector, magnetic vector and direction of propagation form, in that order, a right-handed set (that
is where they obey the corkscrew rule). The linear polarization means that the electric and
magnetic vectors have a constant direction although since they vary cyclically, for half the time
they point in exactly the opposite direction.
All electromagnetic waves, whatever their character, travel at the same velocity in free space.
This is not true when a wave propagates through a medium other than free space. There are two
important constants of optical media, the refractive index, n, and the characteristic optical
admittance, y. Both are defined with respect to a propagating harmonic wave.

velocity of the wave in free space


n= (2.1)
velocity of the wave in the medium

amplitude of magnetic field


y= (2.2)
amplitude of electric field

The high frequencies of the optical region permit a simplification:


Macleod: Optical Coatings Page 6 Design through Manufacture

y=n Y (2.3)

where Y (1/377 siemens) is the admittance of free space. By changing the units of characteristic
admittance from siemens to free space units, we can use the same number for both n and y, a
useful simplification that should not be allowed to disguise the fact that n and y are quite
different quantities. At microwave frequencies this simplification, which is due to the absence of
direct magnetic influence on the electrons, is no longer possible.
We write the expressions for the harmonic waves in the complex form because this affords the
simplest technique for the manipulation of phase shifts. The scalar form of the expression for
such a wave, propagating along the positive direction of the z-axis, is:

E exp[i (ω t − 2π nz / λ )] (2.4)

where ω is the angular frequency of the wave and λ is the wavelength. We are able to use the
complex form because in linear systems we are adding or subtracting waves only and in such
operations involving complex numbers the real and imaginary parts are kept completely
separate. Nonlinear effects, however, mix the real and imaginary parts and then we can no longer
use the complex form without complications.
The linearity of the effects implies that ω is always constant for a given wave. (ω can be
altered only by non-linear effects). If the velocity varies therefore so must the wavelength. But
we traditionally use the wavelength to characterize the wave and so to make to make the
wavelength constant whatever the wave velocity, we define λ as the wavelength in free space.
The n then appears in the numerator of the second term because the actual wavelength is λ/n. We
usually combine n with z when we call the product, nz, the optical path, as distinct from the
physical path z. Note that the order of the two terms in the phase factor could be reversed
without changing the direction of the wave. Either form is completely valid but in order to avoid
ambiguities one form must be chosen as a convention. The usual form adopted in thin film work
is that shown. E is the complex amplitude and can also be written as |E|exp(iϕ) where ϕ is the
relative phase of the wave. The part, exp[i (ω t − 2π n z / λ )] , of the wave expression is known as
the phase factor. It is convenient to arrange that all the beams that are to be combined in an
interference calculation should have the same phase factor, and then relative phases involve the
complex amplitudes only. E can actually represent either magnetic or electric field amplitude
but, since in the optical region only the electric vector produces significant effects on a medium,
the amplitude of the wave is normally considered to be the electric amplitude. H, the magnetic
amplitude can be derived by relationship (2.2), H = y E .
Absorption in a medium can be included by permitting n, and y, to become complex. Then n is
replaced by (n-ik) where k is known as the extinction coefficient, and y by (n-ik)Y, or (n-ik) free
space units. (n-ik) is known usually as the complex index of refraction. In thin film optics we
mainly deal either with materials where k is very small compared with n, the dielectric materials,
or where k is large compared with n, the metallic materials. Dielectric materials are transparent
while metals are opaque, unless very thin.
Macleod: Optical Coatings Page 7 Design through Manufacture

The irradiance (sometimes called intensity) of the plane wave, that is the mean rate of flow of
energy carried by unit area of the wave, is given by

I = 0.5 Re(EH *) (2.5)

This rather curious form is necessary because multiplication of electric and magnetic fields is a
nonlinear operation. This expression is valid only when the complex form of the wave is used.
2
Usually it is sufficient to simplify it as I ∝ n E where we are using the symbol I rather than the
more usual E for irradiance to avoid confusion with the electric field. The expression permits us
to derive the relationship between absorption coefficient and extinction coefficient as

α = 4π k / λ . (2.6)

2.2 Interfaces and Films at Normal Incidence


We know, by experience, that the effect on an incident wave, of an interface between two
media, is to split the wave into two parts. One part is transmitted into the emergent medium
while the other is reflected in the incident medium. For the moment we consider normal
incidence only. The various waves can be described by expressions of the form:

Incident:Ei exp[i (ω t − 2π n0 z / λ )] (2.7)


Transmitted:Et exp[i (ω t − 2π n1 z / λ )] (2.8)
Reflected:Er exp[i (ω t + 2π n0 z / λ )] (2.9)

where the various E's are complex of the form E exp[i ϕ ] so that they contain the relative
phases. Comparison of the relative phases is difficult because one of the phase factors has a
different form from the two others. We therefore fix the coordinate system so that the interface is
the origin. Then z is zero at the interface. There the phase factors reduce to exp[i ω t ] and the
relative amplitudes and phases can be compared by comparing the complex amplitudes, E. We
deal only with linear effects here, and so all we need are the appropriate ratios. To avoid
ambiguities we must first define positive directions for the electric fields. The usual convention
is shown in Figure 2-1. We choose the simplest possible convention for the electric fields but,
because the vectors form a right-handed set, the implied convention for the magnetic vectors is
less simple. Figure 2-1 also contains the notation for the indices of refraction and the admittances
of the two media.
Macleod: Optical Coatings Page 8 Design through Manufacture

Incident Reflected
n0 y0
E E
H H

E
H
z
n1 y1
Transmitted
Figure 2-1. Convention for positive directions of E and H for incident, reflected, and transmitted waves.
Given this convention, we define the amplitude reflection coefficient as

ρ = (Er / Ei ) (2.10)

and the amplitude transmission coefficient as

τ = (Et / Ei ) (2.11)

The corresponding ratios of irradiances are known as reflectance and transmittance


respectively and given by

R =| ρ |2 (2.12)

T = {Re ( y1 ) Re ( y0 )} | τ |2 = (n1 / n0 ) | τ |2 (2.13)

It can readily be derived that

y0 − y1
ρ= (2.14)
y0 + y1

2 y0
τ= . (2.15)
y0 + y1

Note that in order for reflectance, R, to have a meaning there must be no absorption in the
incident medium. In other words y0 must be real. No such restriction applies to ρ. [The reason is
a curious coupling between incident and reflected irradiances in an absorbing medium due to
what is called the mixed Poynting vector.]
The boundary conditions at the interface are that the total components of electric and
magnetic fields parallel to the interface, what are known as the tangential components, are
continuous across it. We can consider these total tangential fields as characterizing the interface.
This is the approach taken in calculating the properties not just of a single interface but of any
Macleod: Optical Coatings Page 9 Design through Manufacture

arbitrary optical coating. The ratio of the total tangential magnetic field to the total tangential
electric field is an admittance that we call the surface admittance.

Hq
Admittance of surface q: Yq = (2.16)
Eq

where Hq and Eq are the tangential components at surface q. This concept of the admittance of an
interface or surface is a particularly powerful one in optical thin film work.
The expressions, (2.12) and (2.14), for a single interface, are derived from the boundary
conditions that the total electric and magnetic fields are continuous across a boundary. In this
respect, y1 can be considered as the admittance of the emergent medium or as the ratio of the
ratio of the total magnetic field to the total electric field at the interface. In the first case it is the
characteristic admittance of a medium, in the second case it is the admittance of the interface. If
the optical admittance presented by the surface to the incident medium is Y, then the amplitude
reflection coefficient of the system will be given by
y0 − Y
ρ= (2.17)
y0 + Y

and the reflectance, as before, and provided y0 is real, by

R =| ρ |2 . (2.18)

It does not matter how complex the system of layers and interfaces might be, if we know the
admittance of the final, front, interface we can immediately calculate, using (2.17) and (2.18),
the reflection characteristics of the system. The calculation of the reflecting properties of the
multilayer then reduces to the change produced by the multilayer in the surface admittance of the
system. The multilayer can be considered as an admittance transformer. In fact, it is extremely
productive to consider first the admittance of the substrate and its transformation by each layer in
turn. Transmittance, unfortunately, cannot be dealt with as simply, but in dielectric systems it is
just the complement of reflectance.
We find that when we need to include ideas of transmittance through an absorbing system of
layers that it is better to keep the tangential components of field separate until the end and only
then to calculate the final surface admittance. Since the effects are linear we can normalize these
fields to simplify the expressions and we usually make the tangential electric field component
unity at the final interface. In calculating the transfer of the tangential fields we start at the rear
of the thin-film system and work through each of the layers to the front. The transfer through
each film involves a linear relationship between the fields in which the coefficients depend on
the parameters of the thin film. This linear relationship can be expressed in terms of a matrix
multiplication in which the film properties are expressed as a square, four-element matrix,
defined by the layer phase thickness and the optical admittance of the material. The calculation
of a multilayer consisting of q layer can then be written:
Macleod: Optical Coatings Page 10 Design through Manufacture

⎧ ⎡ i sin δ j ⎤ ⎫
⎡ B ⎤ ⎪ q ⎢ cos δ j ⎪⎡ 1 ⎤
⎢ ⎥ = ⎨∏ ⎢ y j ⎥⎥ ⎬ ⎢ ⎥ (2.19)
⎣C ⎦ ⎪ j =1 ⎢ ⎥ ⎪ ⎣ ysub ⎦
⎩ ⎣iy j sin δ j cos δ j ⎦ ⎭

where

B and C are the normalized tangential E and H fields at the front interface

δ j = 2π (n j − ik j )d j / λ = phase thickness of layer j

d j = geometrical thickness of layer j

y j = (n j − ik j ) = optical constants of layer j

ysub = (nsub − ik sub ) = optical constants of substrate

Note that the form of the square matrix is such that it contains information only concerning the
film and not the surrounding environment. The matrix is therefore called the characteristic
matrix of the layer j.
Let y0 be the optical admittance of the incident medium. This will be the medium in which all
measurements will be assumed to be made. We therefore assume that y0 is real. Then the final
admittance is given by:

C
Y= (2.20)
B

and the reflectance is then

( y0 B − C )( y0 B − C )∗
R= (2.21)
( y0 B + C )( y0 B + C )∗

Continuing the restriction that y0 should be real, the phase shift on reflection is given by

⎡ iy0 ⎛⎜ CB ∗ − BC ∗ ⎞⎟ ⎤
ϕ = arctan ⎢ 2⎝ ∗ ∗

⎥ (2.22)
⎢⎣ y0 BB − CC ⎥⎦

and the transmittance by


Macleod: Optical Coatings Page 11 Design through Manufacture

4 y0 Re( ysub )
T= (2.23)
( y0 B + C )( y0 B + C )∗

Then, if we denote absorptance by A, then

A = 1− R − T (2.24)

and this allows us to calculate A if we know T and R.

2.3 Interfaces and Films at Oblique Incidence


The situation is a little more complicated if the incidence is oblique. The directions of the
beams in the different media are determined by Snell’s Law

n0 sin ϑ0 = n1 sin ϑ1 (2.25)

Incident Reflected

Interface

Transmitted

Figure 2-2. Refraction tilts the transmitted beam with respect to the incident beam and so, although both
beams are drawn with the same cross sectional area, they subtend different areas at the boundary.
A simple interface is shown in Figure 2-2. It is tempting to define the reflectance, R, as the
ratio of the irradiance of the reflected beam to that of the incident beam, and transmittance, T, as
the ratio of the irradiances of transmitted and incident beams but the drawing makes it clear that
equal areas normal to the beams do not subtend equal areas at the interface. This means that
these simple definitions of R and T do not add to unity when there is no absorption. So that
(2.24) is valid, we therefore define reflectance and transmittance to be the ratios of the
components of irradiance, or intensity, normal to the surfaces of the films. This formulation is
equivalent to considering components of the electric and magnetic fields of the waves which are
parallel to the interfaces, that is the tangential components. Amplitude reflection coefficient, ρ,
amplitude transmission coefficient, τ, reflectance, R, transmittance, T, and absorptance, A, are
then defined as:

reflected amplitude (tangential component)


ρ= (2.26)
incident amplitude (tangential component)
Macleod: Optical Coatings Page 12 Design through Manufacture

transmitted amplitude (tangential component)


τ= (2.27)
incident amplitude (tangential component)

Irradiance reflected normally


R= (2.28)
Irradiance incident normally

Irradiance transmitted normally


T= (2.29)
Irradiance incident normally

A = 1− R − T (2.30)

In general, at oblique incidence the components of electric and magnetic field parallel to the
interface may involve a movement out of the plane containing the ray direction and the field
direction. This implies a change in the polarization state of the reflected and transmitted beams.
Fortunately there are two states where the components remain in the same plane, one with
electric vector parallel to the plane of incidence, known as p-polarized light, and one with
electric vector perpendicular to the plane of incidence, known as s-polarized light. Only in these
two cases will the polarization state of the incident beam be preserved in the reflected and
transmitted beams. But we can use these as fundamental components in which we can express
any arbitrary polarization. After any arbitrary polarization is split into a combination of these
two components they can be solved separately, and the final result then given by the combination
of the two solutions. We retain the definition of surface optical admittance as the ratio of the
components of magnetic and electric field parallel to the interfaces but this is different for the
two modes of polarization. To complete the arrangement we replace the characteristic admittance
of the media with the ratio of the components of magnetic and electric field parallel to the
interface, here too, we have one value for p-polarization and a different one for s-polarization.
For each of our thin films we must recast the phase thickness as the phase shift experienced in a
traversal of the film with no shift in lateral position. The phase thickness has the same value for
both polarizations. It is proportional to the cosine of the angle of propagation in the film and so
is less than at normal incidence.
The principal difference between oblique incidence and normal incidence is, therefore, the
existence of two solutions, one for p-polarization and the other for s-polarization. We do need to
redefine our sign convention as shown in Figure 2-3. It is consistent with the earlier one for
normal incidence, but it should be noted that it is not always that used elsewhere. It does,
however, avoid an ambiguity which is found in an alternative scheme and is, therefore, to be
preferred. The main point about a sign convention is that, once it is chosen, it must be adhered
to.
The results are:
s-polarization:

Optical admittance = ηs = y cos ϑ = (n − ik ) cos ϑ (2.31)


Macleod: Optical Coatings Page 13 Design through Manufacture

p-polarization:

Optical admittance = η p = y / cos ϑ = (n − ik ) / cos ϑ (2.32)

these admittances being measured in free space units.

and the phase thickness of the layers becomes for each plane of polarization,

δ = (2π / λ )(n − ik )d cos ϑ (2.33)

ϑ is given by Snell's Law:

n0 sin ϑ0 = n1 sin ϑ1 = … = (n − ik ) sin ϑ (2.34)

In the general case, y, and n, are complex, implying that ϑ is also complex. We will return to this
point in a moment.

Incident p Reflected

s s p
ϑ0 ϑ0

Transmitted ϑ1
p
s

z
Figure 2-3. The sign convention for the positive directions of electric field for oblique incidence. p and s
indicate the two planes of polarization.
It is important to note that the admittances of the incident and exit media must also be
converted according to (2.31) and (2.32). The arrangement for the calculation of parameters at
oblique incidence then becomes at a single surface:

(η0 − η1 )
ρ=
(η0 + η1 )
(2.35)
(η − η )(η − η )∗
R= 0 1 0 1 ∗
(η0 + η1 )(η0 + η1 )

where, as before, η1 must be real.


Macleod: Optical Coatings Page 14 Design through Manufacture

2η0
τ=
(η0 + η1 )
(2.36)
4η0 Re(η1 )
T=
(η0 + η1 )(η0 + η1 )∗

and

R +T =1 (2.37)

Note that the amplitude transmittance and reflection components are not strictly the Fresnel
components. In fact, all except τp do coincide with the Fresnel expressions.
For a multilayer coating we need to replace not only the admittances by the tilted form but
also the phase thicknesses.

⎧ ⎡ i sin δ j ⎤ ⎫
⎡ B ⎤ ⎪ q ⎢ cos δ j ⎪⎡ 1 ⎤
⎢ ⎥ = ⎨∏ ⎢ η j ⎥⎥ ⎬ ⎢ ⎥ (2.38)
⎣C ⎦ ⎪ j =1 ⎢ ⎣ηsub ⎦
⎩ ⎣iη j sin δ j cos δ j ⎥⎦ ⎪⎭

The advantages of this method should now be clear. The expressions are all exactly as the
earlier expression for normal incidence, but use the appropriate s- or p-admittances and the
corrected value of phase thickness (2.34). Then

(η0 B − C )(η0 B − C )∗
R= (2.39)
(η0 B + C )(η0 B + C )∗

the phase shift on reflection is given by

⎡ iη0 ( CB∗ − BC ∗ ) ⎤
ϕ = arctan ⎢ 2 ∗ ⎥ (2.40)
⎢⎣ η0 BB − CC ⎥⎦

the transmittance by

4η 0 Re(η sub )
T= (2.41)
(η 0 B + C )(η 0 B + C )∗

and the phase shift on transmission by

⎡ i (η 0 B + C − η 0 B∗ − C ∗ ) ⎤
ζ = arctan ⎢ ⎥ (2.42)
⎢⎣ η 0 B + C + η 0 B + C ⎥⎦
∗ ∗
Macleod: Optical Coatings Page 15 Design through Manufacture

The expressions (2.31) to (2.33) are correct, but complex angles of incidence are difficult to
visualize and to handle. We therefore use the following method which has a better physical
basis, and can also be derived by setting cos ϑ = √ (1 − sin 2 ϑ ) . This avoids difficulties over signs
and quadrants.

(n j − ik j ) cos ϑ j = (n 2j − k 2j − n02 sin 2 ϑ0 − 2in j k j )1/2 (2.43)

the correct solution being in the fourth quadrant. Then from (2.31), (2.32) and (2.43),

η js = (n 2j − k 2j − n02 sin 2 ϑ0 − 2in j k j )1/2 (2.44)

and

η jp = (n j − ik j ) 2 / η js (2.45)

The corresponding expression for δ is

2π d
δ= (n 2j − k 2j − n02 sin 2 ϑ0 − 2in j k j )1/2 (2.46)
λ

2.4 Some Complications


When a beam of light is subject to reflections at surfaces in series that are separated by large
distances the absolute phase of the beam is lost. However the relationship between the p- and the
s-polarizations is preserved. Relative retardation, ∆, is defined as the difference between the
phases of the two components. The retardation is changed when the beam is reflected or
transmitted through a system. Unfortunately the normal thin film convention for positive
directions of the electric vector has the p-direction, the s-direction and the direction of
propagation as a right-handed set in both incident and transmitted beams but left-handed in the
reflected beam. This makes it very difficult to keep track of retardation when multiple surfaces
are involved. We therefore always define relative retardation with respect to a right-handed set
of axes. This makes it additive but it complicates the definition slightly.

∆ = ϕ p − ϕ s in transmission (2.47)

but

∆ = ϕ p − ϕ s ± 180 in reflection (2.48)

If we have successive transmissions and reflections the resultant ∆ is simply the sum of the
individual ones:

∆ = ∆ a + ∆b + ∆ c + ∆ d + … (2.49)
Macleod: Optical Coatings Page 16 Design through Manufacture

At oblique incidence, the admittances are replaced with the appropriate tilted admittances, η.
This applies to the incident medium and substrate as well as to the layers. The layers,
additionally, have their phase thicknesses changed. The incident medium, then, has split
admittances, one for each plane of polarization and this leads to difficulties in the visualization
of performance. It is much simpler if the admittance remains at the normal incidence value. This
can be arranged by introducing a set of normalized admittances such that the admittance of the
incident medium remains constant. The new definition of what will be called the modified
admittances, is then
(n − ik ) cos ϑ0
ηp = (2.50)
cos ϑ

and

(n − ik ) cos ϑ
ηs = (2.51)
cos ϑ0

The value of δ remains as before. The normalization process involves an extra cosϑ0. The
admittances are already normalized by dividing by Y, the admittance of free space. This further
operation simply changes the normalization factor. Reflectances, transmittances, phase shifts and
so on, remain unchanged. The modified admittances are used in the admittance diagram,
described in greater detail later.
δ, the phase thickness of a layer is given by (2πnd/λ). It is convenient to think in terms of a
reference wavelength to which the thicknesses of the various layers is referred. Thus instead of
specifying that nd is so many nanometres we say that nd is a certain fraction of λ0, the reference
wavelength. (nd/λ0) is therefore the important quantity. For a quarterwave layer, nd/λ0 is 0.25
and so on. We therefore modify the expression for δ so that it becomes

⎛ nd ⎞ ⎛ λ0 ⎞
δ = 2π ⋅ ⎜ ⎟⋅⎜ ⎟ (2.52)
⎝ λ0 ⎠ ⎝ λ ⎠

The quantity (λ0/λ) is dimensionless and is usually given the symbol g. δ is proportional to g.
For a quarterwave layer δ becomes (π / 2 ) g . It is much easier to think in terms of g rather than
wavelength and so it is much used in coating design. We will make considerable use of it
especially in connection with the admittance diagram. (nd/λ0) is also a dimensionless quantity,
the optical thickness in terms of λ0. Frequently designs are expressed in terms of this quantity
rather than either d or nd because it immediately conveys a qualitative idea of the interference
properties.

2.5 Effect of Rear Surface


Most coating design work, at least in the early stages, is involved with just one surface of the
substrate. The substrate is considered sufficiently thick for the rear surface of the substrate to
have no effect. This assumption is, of course, essential at the beginning of the design process.
Macleod: Optical Coatings Page 17 Design through Manufacture

Trying to include the coupling of the properties of the coatings on front and rear surface would
make the design of each single coating almost impossibly complicated. Eventually, however, a
system has to be constructed and coating are then combined by placing them on opposite sides of
a substrate. Further, many substrates can be placed close to each other - a multiple element lens
for example. Thus we must have some way of dealing with such combinations.
We make the assumption that any substrate is sufficiently thick for all interference effects to
be washed out either by the variation of thickness over the aperture of the system, or the range of
angles of incidence, or the range of wavelengths included in the measurement. [If interference
effects are to be included then the substrate is treated simply as a thin film and the complete
system as just one coating.] When there are no interference effects the summation of the beams
is said to be incoherent. It is fairly easy to show that the incoherent summation is equivalent to
the integration of the interference effects over a range of angles, of wavelength and of thickness
sufficiently wide to smooth out all fringes.. In other words, there is no conflict between the
coherent case, that is the interference case, and the incoherent.

Figure 2-4. The various beams


and quantities involved in the
Ra incoherent multiple beam
a summation for a substrate. The
two substrate surfaces are a and
Ta R' b. These surfaces may be coated
a
and so although the
T int transmittance is identical in
R either direction, the reflectances
b are not necessarily equal. Back
b
reflectance is indicated by a
Tb R' prime.
b

The arrangement of the various beams is shown in Figure 2-4. It is assumed that the irradiance
of the incident beam is unity and that a steady state exists. Then the sum making up the
transmitted irradiance is

T = TaTintTb + TaTint RbTint Ra′TintTb + TaTint ( RbTint Ra′Tint ) Tb + ...


2
(2.53)

TaTintTb
i.e. T= (2.54)
(1 − Rb Ra′Tint2 )
and that giving the reflected irradiance,

R = Ra + TaTint RbTintTa + TaTint RbTint Ra′Tint RbTintTa + ... (2.55)


Macleod: Optical Coatings Page 18 Design through Manufacture

Ta2Tint2 Rb
i.e. R = Ra + (2.56)
(1 − Tint2 Ra′ Rb )
(2.54) and (2.56) can be used in an iterative solution for the properties of any number of coated
or uncoated substrates in series where the surfaces are sufficiently parallel for an infinite number
of combined beams but without interference effects.
In some cases the reflected beams may walk gradually out of the system so that not all will be
included in the sum. In an extreme case the surfaces could be so far from parallel that no
reflected beam from any surface except the very front one would enter the aperture of the
reflectance receiver and no reflected beam whatsoever would enter the transmittance receiver.
Then the expressions are rather simpler:

T = TaTintTb (2.57)

and R = Ra (2.58)

(2.54) and (2.56) represent the behavior of parallel-sided substrates; (2.57) and (2.58) represent
wedged substrates. For a transmitting system, (2.57) represents the transmittance of directly
imaged light, in other words useful light in an image while (2.58) represents the total possible
light including any stray light. The maximum level of stray light therefore is represented by the
difference:

TaTintTb
Tstray = −T T T (2.59)
(1 − Rb Ra′Tint2 ) a int b
TaTb Rb Ra′Tint3
i.e. Tstray = (2.60)
(1 − Rb Ra′Tint2 )
If there is no absorption then the internal transmittance will be unity and the sum of
reflectance and transmittance for each surface will be unity. Then (2.54) can be written as

1
T= (2.61)
1 1
+ −1
Ta Tb

Note that for small equal values of T for both surfaces the (-1) term can be neglected and the
overall transmittance is simply one half of the single surface transmittance. This is an important
result for the overall rejection of a filter made up of separate components. These components
should either be arranged to be wedged or the internal transmittance should be arranged to be
less than unity. An absorption filter that makes up the assembly could usefully be placed
between the reflecting components.
Macleod: Optical Coatings Page 19 Design through Manufacture

3 THEORETICAL TECHNIQUES
3.1 The Quarterwave Rule
Thereare some simple relationships which involve dielectric, i.e. lossless, layers of optical
thicknesses which are an integral multiple of a quarter wavelength. For such layers

δ = (2π / λ ).q.(λ / 4) = q(π / 2) (3.1)

and the appropriate matrices become of the form

⎡ i⎤
⎢ 0 ±
η⎥ (3.2)
⎢ ⎥
⎢⎣ ±iη 0 ⎥⎦

for an odd number of quarter waves, or

⎡ ±1 0 ⎤
⎢ 0 ±1⎥ (3.3)
⎣ ⎦

for an even number of quarter waves, that is for an integral number of half waves.
It is then easy to show that an odd number of quarter waves of admittance ηf on a substrate of
admittance ηsub, acts as a single surface of admittance ηf2/ηsub, while the halfwaves act as
though they did not exist. Then the admittance of the combination of halfwave layer and
substrate is simply ηsub.

For a single film on a substrate, these special cases mark the extrema of reflectance and
transmittance. In between these values, the reflectance or transmittance behaves in a sinusoidal
manner as shown in Figure 3-1. The quarterwave rule can be expressed diagrammatically as
shown in
Figure 3-2.
Multiple application of the quarterwave rule leads to the following expression for the
equivalent admittance of a series of dielectric quarterwaves at normal incidence:

y12 y32 y52


Y= (3.4)
y22 y42 y42 ysub

where ysub is in the denominator as shown for an odd number of layers or numerator for an even
number. For oblique incidence the y values are replaced by the appropriate values of η.
Macleod: Optical Coatings Page 20 Design through Manufacture

Medium 1.00, Layer 2.35, Substrate 1.52


40
35
30
Reflectance (%) 25
20
15
10
5
0
0 1 2 3 4 5
Quarterwave Optical Thicknesses
Figure 3-1. The reflectance of a single layer as a function of its thickness in quarterwaves.

Incident Incident
y0 y0
medium medium
Quarterwave
yf Surface
layer
≡ admittance
yf2/ysub
Substrate ysub

Figure 3-2. Diagram illustrating the quarterwave rule where the film-substrate combination can be replaced
by a single equivalent surface.

Since we are dealing with dielectric materials where the optical constants are all real, the
reflectance of the multilayer is then given by the expression:

( y0 − Y ) 2
R= (3.5)
( y0 + Y ) 2

The quarterwave rule applies equally well to absorbing as transparent substrates, but, with the
former, ysub is complex, and then (3.4) must be modified.
Because of the simplicity of quarterwave layers, they are frequently used in designs, and a
convenient shorthand notation represents high index quarterwaves by H, low index quarterwaves
by L and intermediate index quarterwaves by M, or by any other convenient capital letter.
Designs can then be written as, for example,

Glass | HLHLHHLHLHLHLHLHHLHLH | Glass

which could also be written as


Macleod: Optical Coatings Page 21 Design through Manufacture

Glass | ( HL) 2 H ( HL)5 2 H ( LH ) 2 | Glass

or as Glass | ( HLHLH ) 2 L( HLHLH ) 2 | Glass

Note that HH and LL represent halfwaves. Halfwave layers do not alter the reflectance of a
substrate and hence act as though they were missing. For this reason they are sometimes known
as “absentee” layers. To calculate the reflectance, then, of a mixture of quarter and halfwave
layers, the halfwave layers should first be eliminated and then the reflectance of the remaining
quarterwaves calculated using the quarterwave rule. A typical design of a single-cavity
narrowband filter might be:

Glass | HLHLHLHHLHLHLH | Glass

where we are assuming that a cover slip has been cemented over the assembly so that there is
glass (or cement) on both sides. Now we can use the “absentee” property of halfwave layers to
eliminate the central halfwave or spacer layer:

Glass | HLHLHL[ HH ]LHLHLH | Glass

to give: Glass | HLHLHLLHLHLH | Glass .

But this new equivalent design can be written in turn as:

Glass | HLHLHLLHLHLH | Glass

when the new LL layer that has appeared can also be eliminated to give:

Glass | HLHLHHLHLH | Glass .

Thus we see that the entire assembly of layers can be eliminated in turn until no further layers
are left and the reflectance of the final assembly is zero and hence the transmittance unity. If the
incident medium is air rather than glass then the design becomes:

Air | HLHLHLHHLHLHLH | Glass

Elimination of the halfwaves follows the identical scheme which we can write as:

Air H ⎡⎢ L ⎡ H ⎡ L ⎡⎣ H ⎡⎣ L [ HH ] L ⎤⎦ H ⎤⎦ L ⎤ H ⎤ L ⎤⎥ H Glass
⎣ ⎣ ⎣ ⎦ ⎦ ⎦

leaving Air | Glass.


Macleod: Optical Coatings Page 22 Design through Manufacture

The reflectance of this will be given by

( yair − y glass ) 2
R= (3.6)
( yair + yglass ) 2

and clearly the transmittance can be increased by adding an antireflecting quarterwave to the
assembly. If we assume values of 2.35 for yH and 1.35 for yL, then a single quarterwave of low
index will reduce the reflectance to around 1% (use the quarterwave rule to calculate it
accurately) and hence raise the transmittance to 99%. The extra L layer can conveniently be
placed at the left hand side of the design next to the air incident medium.
We can take the earlier two-cavity design:

Glass | HLHLHHLHLHLHLHLHHLHLH | Glass

and carry out the same procedure to determine the transmittance at the reference wavelength.

Glass ⎡ H ⎡ L ⎡⎣ H ⎡⎣ L [ HH ] L ⎤⎦ H ⎤⎦ L ⎤ H ⎤ L ⎡ H ⎡ L ⎡⎣ H ⎡⎣ L [ HH ] L ⎤⎦ H ⎤⎦ L ⎤ H ⎤ Glass
⎣ ⎣ ⎦ ⎦ ⎣ ⎣ ⎦ ⎦

equivalent to

Glass | L | Glass

This does not give maximum transmittance and an additional L layer is needed to transform
this into

Glass | LL | Glass

when the LL layer is an “absentee” and can be eliminated. The full design is then

Glass | LHLHLHHLHLHLHLHLHHLHLH | Glass

3.2 Admittance Loci and the Admittance Diagram


In this technique, we plot, in an admittance diagram, that is the complex plane, the locus of
the optical admittance of a multilayer measured at a notional surface, parallel to the substrate,
which is swept through each layer, starting at the substrate and ending at the front surface. It is
as if we plotted the admittance of the multilayer during the entire deposition process. For each
dielectric layer, the locus is an arc of a circle centered on the real axis. For metallic layers the
locus is a more complicated spiral. In this explanation we consider dielectric layers.
We measure Y in free space units and assume that we have one single layer of admittance y on
a substrate of admittance α + i β . Then
Macleod: Optical Coatings Page 23 Design through Manufacture
⎡ i sin δ ⎤
⎡ B ⎤ ⎢ cos δ ⎡ 1 ⎤
⎢C ⎥ = ⎢ y ⎥⎢ ⎥ (3.7)
⎣ ⎦ iy sin δ cos δ ⎥ ⎣α + i β ⎦
⎣⎢ ⎦⎥

where δ = 2π y d / λ

d = layer geometrical thickness

y = film admittance (real since the film is dielectric)

and Y =C/B

Y will in general be complex of the form a + ib and we can then write:

α cos δ + i ( y sin δ + β cos δ )


Y = a + ib = (3.8)
[cos δ − ( β / y ) sin δ ] + i (α sin δ ) / y

We separate real and imaginary parts

a[cos δ − ( β / y ) sin δ ] − b(α sin δ ) / y = α cos δ (3.9)

b[cos δ − ( β / y ) sin δ ] − a(α sin δ ) / y = ( y sin δ + β cos δ ) (3.10)

and eliminate δ to give the equation of the admittance locus:

a 2 + b 2 − a( y 2 + α 2 + β 2) / α + y 2 = 0 (3.11)

i.e. a circle with center ([ y 2 + α 2 + β 2 ] / 2α , 0) passing through the point (α , β ) . For the loci of
constant δ we assume that β is zero and then eliminate α giving

⎛ 1 ⎞
a 2 + b 2 + by ⎜ − tan δ ⎟ − y 2 = 0 (3.12)
⎝ tan δ ⎠

i.e. a circle with center (0, − 0.5 y{[1/ tan δ ] − tan δ }) , that is, on the imaginary axis, passing
through the point ( y, 0) .
Macleod: Optical Coatings Page 24 Design through Manufacture

Figure 3-3. Admittance locus of a dielectric layer.


The circle is traced out clockwise. If the starting point is on the real axis, then a semicircle
corresponds to a quarterwave. Drawing the circle becomes easy if we note that, if one point of
intersection with the real axis is given by α, then the other point of intersection will be y2/α.

Figure 3-4. Isophase and isoreflectance circles.


In a similar way we can plot isoreflectance contours on the same diagram. These are circles
which are also centered on the real axis. Their centers are given by (y0[1+R]/[1-R], 0) and their
radii by 2y0(√R)/(1-R), where y0 is the admittance of the incident medium.
Isophase contours, that is the phase change on reflection, can also be added. The most
important are the boundaries between the various quadrants. These are the real axis and the circle
with center at the origin and passing through the point (y,0). The scheme is shown in Figure 3-4.
Metal layers, or other absorbing layers, normally require the assistance of a computer. As
y = n − ik tends to −ik , i.e. as n tends to zero, the locus tends to a circle centered on the real
axis passing through the point (0, −k ) and traced out clockwise. Metal layers shall be briefly
considered at the end of this section.
An advantage of the admittance diagram beyond the obvious visualization of performance, is
that contours of electric field amplitude, of importance in damage performance assessment, are
very simple in form and can readily be added when all the layers are dielectric. These are
illustrated in Figure 3-5 and are lines perpendicular to the real axis and with field increasing
towards the origin.
Macleod: Optical Coatings Page 25 Design through Manufacture

Figure 3-5. Lines of constant electric field amplitude. The figures are in volts/metre if the transmitted
irradiance is 1watt/sq metre.
We continue to work with admittances in free space units. Then we can write for the net
irradiance passing into the multilayer,

(1/ 2) Y Re( BC ∗ ) = (1/ 2) Y Re(YBB∗ ) = (1/ 2) YBB∗ Re(Y ) (3.13)

where Y = admittance of free space = 1/ 377 siemens

BB∗ = E 2 = (electric field amplitude) 2

Now, when the layers are dielectric the net irradiance actually entering the assembly is the
same as that which emerges. Let unit irradiance emerge from the multilayer system. Then
1
∗ 2 1 27.46
E = ( BB ) = 1
= 1
(3.14)
⎡⎣(1/ 2 ) Y Re (Y ) ⎤⎦ 2 ⎡⎣ Re (Y ) ⎤⎦ 2

If the incident irradiance is unity, then


1
27.46T 2
E= 1
(3.15)
⎡⎣ Re (Y ) ⎤⎦ 2

where T is the transmittance. E is in volts/metre provided that Y is in free space units (i.e. if Y is
numerically equal to refractive index).
We can illustrate some of the features of the admittance diagram with a particularly simple
example. This is the V-coat antireflection coating, which consists of two layers, a high index
layer next to the substrate and a low index layer outermost. The admittance diagram consists of
two circles, one representing the high-index layer and with its starting point the substrate and the
other the low-index layer with its termination point as the incident medium. Provided these two
circles intersect then a solution is possible. Since there will usually be two points of intersection,
two solutions are normal. This is illustrated in Figure 3-6.
Macleod: Optical Coatings Page 26 Design through Manufacture

Figure 3-6. This diagram shows the two circles in the admittance plot for the V-coat. The two solutions are
given by the two points of intersection of the circles. If there are no points of intersection then there is no
possible solution.
The solution usually adopted involves the thinner high-index layer and thicker low-index
because this involves less total optical thickness and so the coating is likely to be less sensitive
both to wavelength and to angle of incidence effects. We adopt as an example a combination of
high index 2.25, low index 1.45, substrate 1.52 and incident medium 1.00. Then the optical
thicknesses of the two layers must be 0.0714λ0 for the high-index layer and 0.3275λ0 for the
low-index. The admittance diagram of this solution is shown in. Figure 3-7

Figure 3-7. This diagram shows the admittance diagram for the particular solution we are adopting for the V-
coat. The thin high-index layer is followed by the thicker low-index.
Macleod: Optical Coatings Page 27 Design through Manufacture

Figure 3-8. The performance of the V-coat of Figure 2.15 plotted as a function of g (=λo/λ).
From the admittance diagram we can immediately see that the electric field in the high-index
layer falls from substrate to the junction with the low index. In the low-index layer the field falls
further until the point of intersection of the locus with the real axis. That point represents the
minimum field. From there on to the front of the assembly the field rises and reaches a turning
value at the front surface where the value is at its greatest. An accurate plot of electric field,
where the assumption of an incident irradiance of 1 watt/sq metre has been made, is shown in
3.15 and has exactly the form expected.
The substrate is normally considered infinite in extent with no returning wave and therefore
no standing wave. In this particular case that we have just considered, we are dealing with a
coating that is a perfect antireflection coating and so there is no reflected beam in the incident
medium and, therefore, no standing wave there either. In many cases, however, there is a
reflected beam and therefore a standing wave in the incident medium. We can deal with this by
adding a thick layer to the front surface of the coating consisting of incident medium material.
This, of course, has no effect on the reflectance or transmittance of the assembly but the addition
of the appropriate admittance circle to the admittance diagram permits the ready calculation of
the variation of electric field amplitude in the region in front of the coating.
Macleod: Optical Coatings Page 28 Design through Manufacture

Figure 3-9. The variation of electric field amplitude through the V-coat. The incident beam is assumed to
have an irradiance of 1 watt/sq metre.
The admittance diagram can also be used in the determination of suitable points for the
insertion of halfwave performance flattening layers and for buffer layers. Basically the ideas
behind these two types of layer are similar. The halfwave layer being an “absentee” layer can be
inserted anywhere in the system and will not affect the performance at λ0. However, judicial
choice of index can result in improved (or worsened) performance at other wavelengths. The
admittance diagram assists in the correct choice of insertion point and of halfwave index. A
buffer layer is a layer of quite definite index the thickness of which can be varied without
altering performance λ0. It is not quite as useful as the halfwave and is not nearly so often used.
The insertion point must be a point of intersection of the locus with the real axis and the
admittance of the buffer layer inserted must correspond to the admittance at the point of
intersection. Clearly, then, the thickness of the inserted layer can be quite arbitrary because its
admittance locus has collapsed to a single point. Again the admittance diagram shows clearly
where buffer layers can be inserted.
The admittance of a metal layer is a little more complicated than a dielectric. For a lossless
metal in which the refractive index, and hence optical admittance, is purely imaginary, and given
by -ik, the loci are a set of circles with centers on the real axis and passing through the points ik
and -ik, which are on the imaginary axis. Figure 3-10 shows the typical form. The circles are like
the dielectric ones, traced out clockwise so that they start on ik and end on -ik.
Macleod: Optical Coatings Page 29 Design through Manufacture

Figure 3-10. Typical form of ideal metal admittance loci. The loci are described clockwise and are a set of
circles centered on the real axis and passing through the points (0,k) and (0,-k).

Figure 3-11. Sketch of the admittance loci and isophase thickness contours of a typical high-performance
metal layer where n is non zero.
Real metallic layers depart somewhat from this ideal model but if the metal is of high
performance, i.e. if the ratio k/n is high, then the loci are similar to the perfect case. It is as if the
diagram were rotated slightly about the origin so that the points where all circles intersect are
(n, − k ) and (− n, k ) respectively, although the circles can never reach the point (− n, k ) since
admittance loci are constrained to the first and second quadrants of the complex plane. The
direction of the loci is now better described as terminating on ( n, − k ) , although most are still
described in a clockwise direction, Figure 3-11.

3.3 Symmetrical Periods


Many of the most powerful analytical design techniques are based on the concept of
symmetrical periods. It can be readily shown that any combination of thin films which is
symmetrical, that is one in which the sequence of layers is unchanged when they are listed in
reverse order, can be represented by a single equivalent film having an optical admittance and
Macleod: Optical Coatings Page 30 Design through Manufacture

phase thickness which can be calculated from the parameters of the individual layers. If we
represent the product of the characteristic matrices of the layers of the symmetrical period by a
single matrix, M, then we find that M11= M22, and that the determinant of the matrix is unity. The
product matrix, M, can therefore be put in the form:

⎡ i sin γ ⎤
⎡ M 11 M 12 ⎤ ⎢ cos γ ⎥
⎢M = E (3.16)
⎣ 21 M 22 ⎥⎦ ⎢ ⎥
⎣iE sin γ cos γ ⎦

1
where E = ( M 21 / M 12 ) 2 (3.17)

and γ = arccos( M 11 ) (3.18)

E is known as the equivalent admittance and γ as the equivalent phase thickness. Analytical
expressions can be derived for E and γ but, even for a very small number of layers, they are
somewhat complicated. For three layers of the form ABA, we have:

M 11 = cos 2δ a cos δ b − 0.5[( yb / ya ) + ( ya / yb )]sin 2δ a sin δ b (3.19)

M 12 = (i / ya ){sin 2δ a cos δ b + 0.5[( yb / ya ) + ( ya / yb )]cos 2δ a sin δ b


(3.20)
+0.5[( ya / yb ) − ( yb / ya )]sin δ b }

M 21 = (iya ){sin 2δ a cos δ b + 0.5[( yb / ya ) + ( ya / yb )]cos 2δ a sin δ b


(3.21)
−0.5[( ya / yb ) − ( yb / ya )]sin δ b }

If straightforward evaluation is all that is required, then it is simpler to write a computer


program in terms of the multiplication of the basic matrices. The equivalence is valid for
dielectric or metallic layers, but is most often applied to combinations of layers which are
entirely dielectric. When the equivalent parameters are calculated for an assembly of dielectric
layers, it is found that, in some regions, the equivalent phase thickness and the equivalent
admittance are imaginary and others where they are both real. There are no regions where they
are not either both real or both imaginary. Regions where both are imaginary may be thought of
as representing a pseudo-metallic behavior (only at optical frequencies) and are therefore
associated with stop bands. Here the reflectance increases towards unity as the number of similar
periods used in series in the multilayer increases. The regions in which the parameters are real,
represent dielectric behavior and are potential pass bands. The parameters γ and E are
multivalued and we have complete freedom of choice in choosing which particular value to use.
We find it most convenient to choose γ as that value nearest the total phase thickness of the
symmetrical combination and we also choose the positive value for E. In the pass regions, the
equivalent phase thickness of the period is found to be approximately equal to the total phase
thickness of the layers of the period. If now q periods are placed in series, then, regardless of the
actual number, the multilayer will act as a single layer with the equivalent admittance and with
Macleod: Optical Coatings Page 31 Design through Manufacture

phase thickness q times that for a single period. To transform the regions of real admittance into
pass bands, all that is required is a coating at either end to match the equivalent layer into the
substrate and into the surrounding medium.
We shall see that an arrangement such as (0.5HL0.5H) or (0.5LH0.5L), where two eighth
wave layers surround a single quarterwave, is of particular importance in the design of edge
filters. We illustrate the equivalent parameters of such an arrangement with H representing a
quarterwave of high index material such as titanium dioxide with nH=2.35 and nL=1.45. Results
are given in Table 2.1. Note that γ/π rather than γ is listed.
The technique is also used in cases where a layer with an unobtainable admittance is required
in a design. It is sometimes possible to design a replacement consisting of a symmetrical period.
In this case the thicknesses of the layers in the symmetrical period are chosen such that the
equivalent admittance is the quantity required but at the same time the equivalent phase
thickness is a quarterwave, i.e. γ =π/2. From (3.18) to (3.20) with some algebraic manipulation,
we eventually derive the relationships:

( yb2 + ya2 )( E 2 − ya2 )


cos 2δ a = (3.22)
( yb2 − ya2 )( E 2 + ya2 )

2 ya yb 1
and tan δ b = ⋅ (3.23)
( yb + ya ) tan 2δ a
2 2

with optical thicknesses na d a / λ0 and nb db / λ0 given by δ a / 2π and δ b / 2π respectively.


Note that the form of (3.21) makes it possible to achieve only equivalent admittances that lie
within the range ya to yb.
Table 3-1. E and γ for nH=2.35 and nL=1.45
(0.5H L 0.5H) (0.5L H 0.5L)
g E γ/π E γ/π
0.00 1.84594 0.00000 1.84594 0.00000
0.05 1.84451 0.05147 1.84737 0.05147
0.10 1.84017 0.10295 1.85173 0.10295
0.15 1.83280 0.15448 1.85917 0.15448
0.20 1.82218 0.20607 1.87001 0.20607
0.25 1.80798 0.25774 1.88470 0.25774
0.30 1.78972 0.30954 1.90393 0.30954
0.35 1.76676 0.36149 1.92867 0.36149
0.40 1.73818 0.41365 1.96038 0.41365
0.45 1.70275 0.46610 2.00117 0.46610
0.50 1.65869 0.51893 2.05433 0.51893
Macleod: Optical Coatings Page 32 Design through Manufacture

(0.5H L 0.5H) (0.5L H 0.5L)


g E γ/π E γ/π
0.55 1.60342 0.57229 2.12514 0.57229
0.60 1.53305 0.62642 2.22269 0.62642
0.65 1.44136 0.68173 2.36409 0.68173
0.70 1.31753 0.73895 2.58628 0.73895
0.75 1.14019 0.79975 2.98853 0.79975
0.80 0.85451 0.86902 3.98764 0.86902
0.85
0.90
0.95
1.00 Both E and γ are imaginary in this region
1.05
1.10
1.15
1.20 6.46271 1.13098 0.52726 1.13098
1.25 4.84347 1.20025 0.70352 1.20025
1.30 4.19156 1.26105 0.81294 1.26105
1.35 3.83146 1.31827 0.88935 1.31827
1.40 3.60229 1.37358 0.94593 1.37358
1.45 3.44420 1.42771 0.98934 1.42771
1.50 3.32944 1.48107 1.02345 1.48107
1.55 3.24328 1.53390 1.05063 1.53390
1.60 3.17717 1.58635 1.07250 1.58635
1.65 3.12578 1.63851 1.09013 1.63851
1.70 3.08567 1.69046 1.10430 1.69046
1.75 3.05451 1.74226 1.11556 1.74226
1.80 3.03070 1.79393 1.12433 1.79393
1.85 3.01314 1.84552 1.13088 1.84552
1.90 3.00108 1.89705 1.13543 1.89705
1.95 2.99402 1.94853 1.13810 1.94853
Equivalent
2.00 layer is ½-
wave
Macleod: Optical Coatings Page 33 Design through Manufacture

4 COLOR IN OPTICAL COATINGS


4.1 Introduction
Anyone who works with optical coatings knows that they can present exceedingly attractive
colors. These colors originate in interference effects that enhance reflectance or transmittance in
certain parts of the visible spectrum and inhibit it in others. Although colors occur with both
transmitted and reflected light it has long been observed that the most vivid effects are usually to
be found in reflection. In the same way that coatings can be designed to have desired spectral
properties they can also be designed to present desired colors. This is a little more complicated
than the usual design processes because of the subjective nature of color itself. Color is strictly a
human response to a luminous stimulus. The response varies with the individual observer. Since
the coating is not luminous in itself, a source of light is also required before the colors can be
observed. We avoid the obvious difficulties inherent in the variations between individual
observers and light sources by using theoretical standards that represent more or less well the
average properties, and permit us to remove the subjective nature of the problem. The design
goals can then be presented in unambiguous terms but there remain some complications. In order
to observe the color there must be an acceptable level of reflected or transmitted light. This, in
turn, has a major influence on what can be achieved. A further requirement is that the coatings
must be sufficiently simple for large-scale production at reasonable cost.
The coatings that are considered here are ones that derive their colors primarily from
interference effects rather than any intrinsic colors of the materials themselves. Although the thin
films may be deposited over rough substrates, the thin-film itself is essentially a specular effect
and this is the thrust of the treatment here.
Colors in thin films have attracted many workers over the centuries. Isaac Newton spent a
great deal of time studying and describing the colors of soap bubbles and his account of them is
still the most complete and graphic that exists. Newton struggled to describe the colors, and
actually succeeded rather well, but it was a further 100 years before, with the work of Thomas
Young, color vision began to be understood, and another 130 before agreed standard ways of
defining colors objectively were established. There are clear advantages in an objective system
of color definition yet one admits missing the excitement and enthusiasm inherent in Newton’s
descriptions.
Newton understood very well that the colors in thin films vary with angle of inclination. This
effect, inherent in interference, gives a depth to the colors that is lacking in normal pigments and
is one of the particularly attractive features of decorative interference optical coatings.
Most decorative coatings are used in reflection with just a few occasions when colors in
transmission are required. Coatings that must perform only in reflection do not have to transmit
and therefore can include heavily absorbing layers, especially metals, giving additional
possibilities in their design.

4.2 Color
Color[1] is a subjective, human, response to the spectral quality of light. It depends on the
output of receivers in the retina that have three different responses called red, green and blue. We
Macleod: Optical Coatings Page 34 Design through Manufacture

can measure the color of any light by comparing the response with those for known standard
colors. The accepted method involves mixing different amounts of light from three standard
sources of different color and suitable power, known as primary sources, until the mixture gives
the same color sensation as the sample, a process known as Color Matching. The relative
amounts of the three primaries is then a measure of the color. But individuals differ in their
responses and we need an objective measurement. A Standard Observer was therefore created
from measurements on many individuals. The human eye is very complicated in its response and
there are, in consequence, more than one Standard Observer but that most frequently used, and
used here, was defined by the International Committee on Lighting (CIE) in 1931.

Response
2.0

1.5

xbar
1.0 ybar
zbar
0.5

0.0
300 400 500 600 700 800
Wavelength (nm)

Figure 4-1. The three Color Matching Functions of the CIE 1931 Standard Colorimetric Observer. The
primary sources were defined so that y ( λ ) corresponds exactly to the photopic relative response curve of
the eye. [There exists also a CIE 1964 Supplementary Standard Observer with similar matching functions.]
The Standard Colorimetric Observer is defined as three Color Matching Functions that were
derived by determining the amount of each standard primary required to match narrow spectral
lines. Narrow spectral lines are considered to represent the most intense colors. The concept of
purity is used to describe how closely the intensity of a color resembles that of a spectral line.
Measurements showed that with any real set of primary sources it was necessary sometimes to
subtract an amount of a given primary to match a spectral line. To avoid this, the three standard
primary sources are defined as having a spectral purity beyond that of any spectral line. They
exist, therefore, only theoretically. They are roughly red, green and blue in color and are known
as X, Y and Z respectively.
Thin film coatings, like most objects, do not emit light on their own. The color stimulus is
derived from light that is either reflected or transmitted. The spectral quality of the illuminant
source is therefore important and Standard Sources, some shown in Figure 4-2, are defined.
Macleod: Optical Coatings Page 35 Design through Manufacture

Relative output
300

200
Illuminant D 65
Illuminant A
Equal energy
100

0
300 400 500 600 700 800 900
Wavelength (nm)

Figure 4-2. Since the coatings must be illuminated for their colors to be observed standard sources must form
part of the color definition. Here are three standards. Illuminant D 65 represents daylight but not direct
sunlight. Illuminant A represents an incandescent light source. Equal energy is not a standard but is
frequently used in thin film coating calculations.
The assessment of color assumes the coating illuminated by a standard source and the
reflected of transmitted light collected by a detector with response equal to each of the three
color matching functions in turn. Three quantities, X, Y and Z, are then derived by multiplying
each detector output by 100 and then dividing by a reference. The reference consists of a
measurement of the light source, in the absence of the coating, by the detector with the y ( λ )
response. The three derived quantities, X, Y and Z, are then known as the Tristimulus Values and
they are manipulated in various ways to yield a final set of color coordinates.

∫λ S ( λ ) R ( λ ) x ( λ ) d λ
X = 100 ⋅ (4.1)
∫ S (λ ) y (λ ) dλ
λ

∫λ S ( λ ) R ( λ ) y ( λ ) d λ
Y = 100 ⋅ (4.2)
∫ S (λ ) y (λ ) dλ
λ

∫λ S ( λ ) R ( λ ) z ( λ ) d λ
Z = 100 ⋅ (4.3)
∫ S (λ ) y (λ ) dλ
λ

Note the 100 factor is not required if the coating response, R(λ), is in percent.
Macleod: Optical Coatings Page 36 Design through Manufacture

Since the y ( λ ) color matching function has been chosen to match the photopic sensitivity of
the eye Y is a measure of how well the coating reflects luminous light. Y is known as the
Luminosity Factor, or, sometimes, Luminous Reflectance or Transmittance. Note that the
luminosity factor does depend to some extent on the spectral distribution of the source.
The Chromaticity Coordinates, x, y and z, are normalized values of X, Y and Z such that their
sum is unity.

X Y Z
x= , y= and z = (4.4)
X +Y + Z X +Y + Z X +Y + Z

This arrangement permits a two-dimensional plot of the chromaticity coordinates. It is customary


to use x and y for the plot. The simplicity of this plot is a great advantage. For a complete
description of the color, it is only necessary to add a value for Y.
Unfortunately the chromaticity coordinates are not necessarily proportional to the degree of
the color sensation. It is often important to have some idea of the changes that can be tolerated in
the coordinates before an unacceptable shift in the perceived color occurs. Attempts have been
made to construct color diagrams where the distance between two color specifications is
proportional to their perceived difference. Two such approximately correct spaces recommended
by the CIE are designated as L*a*b* and L*u*v*.
If Xn, Yn and Zn are the tristimulus values of the standard illuminant with Yn=100, then
13
∗ ⎛Y ⎞
L = 116 ⎜ ⎟ − 16 (4.5)
⎝ Yn ⎠

⎡⎛ X ⎞1 3 ⎛ Y ⎞1 3 ⎤ ⎡⎛ Y ⎞1 3 ⎛ Z ⎞1 3 ⎤
a = 500 ⎢⎜

⎟ − ⎜ ⎟ ⎥ and b = 200 ⎢⎜ ⎟ − ⎜ ⎟ ⎥

(4.6)
⎢⎣⎝ X n ⎠ ⎝ Yn ⎠ ⎥⎦ ⎢⎣⎝ Yn ⎠ ⎝ Z n ⎠ ⎥⎦

and if

4X 4Xn
u′ = and un′ =
X + 15Y + 3Z X n + 15Yn + 3Z n

9Y 9Yn
v′ = and vn′ =
X + 15Y + 3Z X n + 15Yn + 3Z n

then

u ∗ = 13L∗ ( u ′ − un′ ) and v∗ = 13L∗ ( v′ − vn′ ) (4.7)


Macleod: Optical Coatings Page 37 Design through Manufacture

4.3 Coatings for Color and Optimal Color Stimuli


Optimal Color Stimuli[1] are object color stimuli with greatest luminous reflectance or
transmittance for a given chromaticity. The characteristics used for the calculations are very
simple but they can be shown to give best performance. Optimal color stimuli are the limit of
what is achievable with an optical coating. It is no use trying to do better. All the stimuli have
the form of the curve of Figure 4-3. The boundaries of the different zones are completely sharp.
Tables giving their location in wavelength exist[1]. The two sets of optimal color stimuli in
Figure 4-4 have a luminance factor, Y, of 50.These characteristics can be plotted in a
chromaticity diagram and can be used to assess the quality of any coating that is produced by
demonstrating how closely it approaches the maximum possible performance for its luminosity.

R or T
100%

0%
Wavelength
Figure 4-3. Optimum color stimuli are simple rectangular functions in transmittance or reflectance. The
boundaries of the transmittance or reflectance zones have been calculated and are published in tabular
form[1].

Reflectance (%) Reflectance (%)


100 100
80 80
60 60
40 40
20 20
0 0
400 500 600 700 400 500 600 700
Wavelength (nm) Wavelength (nm)

Figure 4-4. The two sets of curves shown give an optimum Luminance Factor, Y, of 50. The curves on the
left represent coatings that reflect a band of wavelengths and absorb others and on the right that absorb a
band of wavelengths and reflect others.
Macleod: Optical Coatings Page 38 Design through Manufacture

y
0.9
520
530
0.8 540 Illuminant D65
510
550
0.7 Y=20
560

0.6 Y=50 570


500 580
0.5 Y=90
590
600
0.4
620
650700
0.3
490

0.2

0.1 480
White Point
470
0.0 450
360

-0.1
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
x

Figure 4-5. We can plot the contours of the optimum color stimuli in a chromaticity diagram. Y, the
luminosity, takes values of 100 (the white point), 90, 50, 20 and zero (the outer spectrum line).

Imaginary

Titanium with oxide


100

80
Reflectance (%)

y0 ~yf2/y0 Real 60

40

20
Metal optical constants
0
should be in the yellow 400 500 600 700
region Wavelength (nm)

Figure 4-6. Quite simple coatings that give good color are oxide layers on metals that act as efficient
antireflection coatings. Titanium is a good example of a metal that has such performance.
Some metals lie on or very near that admittance circles of their oxide that passes through the
admittance of the incident medium and so gives zero, or near zero, reflectance. The very deep
single minimum in the visible region leads to brilliant residual colors of high saturation.
Titanium, shown in Figure 4-6, is a good example. Blue heat-treated steel is another.
Figure 4-7 shows the chromaticity coordinates of the oxide-coated titanium of Figure 4-6
plotted over the optimal stimuli contours. The Luminance Factor, Y, for the coating is 27
implying a very reasonable performance for a particularly simple coating. The curves of Figure
4-7 show that the maximum luminosity factor for the given chromaticity coordinates is around
50.
Macleod: Optical Coatings Page 39 Design through Manufacture

y
0.9
520
530
0.8 540
Illuminant D65
510
550
0.7 Y=20
560

0.6 Y=50 570


500 580
0.5 Y=90
590
600
0.4
620
650700
0.3
490
0.2

0.1 480

470 Ti with oxide


0.0 450
360

-0.1
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
x

Figure 4-7. The oxide-coated titanium chromaticity coordinates plotted over the optimum color stimulus
curves show that it has reasonable performance for such a simple coating. The luminance factor value of 27
indicates that it falls a little short of the possible value of 50 for the achieved chromaticity coordinates.
If no transmittance is required of the coating, then a simple two-layer combination of a
dielectric and metal can act as an antireflection coating for any high-performance metal, for
example, aluminum, to give intensely colorful effects. This design dates back to at least the
1940’s[2-4] and has appeared in many guises since then. The arrangement is shown in the right
hand part of Figure 4-8. The reflector is adjusted so that the two beams are of equal amplitude.
The phase matching layer is then adjusted in thickness to assure destructive interference. In
Figure 4-8 the order of the phase matching layer was arranged to be rather higher than first to
narrow the low reflectance fringe and make it a better fit with the appropriate optimal color
stimulus.
The performance curve to the right of Figure 4-8 was obtained for a coating consisting of a
magnesium fluoride phase matching layer and chromium reflector deposited over aluminum.
Such interference coatings show the usual movement of the characteristic to longer wavelengths
as the angle of incidence increases[5-7]. In Figure 4-9 the locus of the chromaticity coordinates
of the antireflected aluminum as the angle of incidence changes from 0° to 45° show a gradual
change from magenta through red, orange and yellow to green. The Luminance Factor is 44 at
normal incidence, which, for such a simple coating, compares well with the maximum of just
over 50 that can be assessed from the diagram. This implies a quite well-designed coating. Here
the color change is enhanced by using a low index dielectric film (MgF2 here). This general two-
layer approach is used in many applications.
Macleod: Optical Coatings Page 40 Design through Manufacture

Al/MgF2/Cr. 0° Magenta. 45° green.


100

80

Reflectance (%)
Reflector 0°
60
Dielectric phase
40
matching layer
20
45°
Metal layer 0
400 450 500 550 600 650 700
Wavelength (nm)
Figure 4-8. The two-layer coating on the right can serve as an antireflection coating for any high performance
metal. The reflector is adjusted to equalize the beams and the phase matching layer to assure phase
opposition. A typical performance at normal incidence (magenta) and at oblique incidence (green) is shown
on the right. This particular coating[7] consists of magnesium fluoride (phase matching) and chromium
(reflector) and the phase matching layers was made quite thick to ensure a narrow fringe and good quality
color.

y
0.9
520
530
0.8 540
510
550
0.7 Y=20
560

0.6 Y=50 570


500
580
0.5 Y=90
590
45°
0.4 600

620
650700
0.3 0°
490

0.2

0.1 480

470
0.0 450
360

-0.1
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
x

Figure 4-9. The plot of chromaticity coordinates of the coating of Figure 4-8 show the variation of color as the
coating is tilted from normal incidence, where it appears magenta in color, to 45º where it becomes green. The
luminance factor at normal incidence is 44. Since the maximum possible value can be seen to be just over 50, the
coating can be considered to be well designed.
Finally, A dichroic coating is shown in Figure 4-10. The chromaticity coordinates with
designs of 11 and 21 layers are shown and they are almost identical. The corresponding
luminance factors are 46.87 and 50.85, the latter being essentially optimal. The enormous
increase in the number of layers, while it clearly makes a difference to the spectral profile of the
coatings does little for the color performance of the coating.
Macleod: Optical Coatings Page 41 Design through Manufacture

y
0.9 520
Transmittance (%) 530
0.8 540
100 510
550
0.7 Y=20
80 560

0.6 Y=50 570


60
500 580
0.5 Y=90
40 590
11-layer 21-layer
600
20 Y=46.87 Y=50.85 0.4
620
650700
0 0.3
490
400 500 600 700
0.2
Wavelength (nm)
0.1 480

470
0.0 450
360

-0.1
-0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
x
Figure 4-10. Optimal color stimuli applied to a dielectric coating. The 11-layer coating gives a performance
that is virtually indistinguishable in color terms from that of the 21-layer. The increase in luminance factor
from 46.9 to 50.9 represents a rather small gain for an almost doubling of the number of layers.

4.4 References
1. Wyszecki, Günter and W S Styles, Color science. 2nd ed. 1982, New York: John Wiley
& Sons.
2. Hadley, L N and D M Dennison, Reflection and transmission interference filters. Part I.
Theory. Journal of the Optical Society of America, 1947. 37(6): p. 451-465.
3. Turner, A F, Some current developments in multilayer optical films. Journal de Physique
et le Radium, 1950. 11: p. 443-460.
4. Mohler, N M and J R Loofbourow, Optical filters. American Journal of Physics, 1952.
20: p. 499-515 and 579-588.
5. Phillips, Roger, Mike Nofi, and Robert Slusser. Color effects from thin film designs. in
Eighth International Conference on Vacuum Web Coating. 1994. Las Vegas, Nevada:
Bakish Materials Corporation, POB 148, Englewood, NJ 07631, USA. p. 270-284.
6. Phillips, Roger W, Robert Slusser, and Becky Holton. Angle dependent metamerism in
multi-layer thin films. in Ninth International Conference on Vacuum Web Processing.
1995. Tucson, Arizona: Bakish Materials Corporation, POB 148, Englewood, NJ 07631,
USA. p. 127-138.
7. Phillips, Roger W and Anton F Bleikolm, Optical coatings for document security.
Applied Optics, 1996. 35(28): p. 5529-5534.
Macleod: Optical Coatings Page 43 Design through Manufacture

5 MATERIALS
5.1 Dielectrics
An atom consists of a massive nucleus that is positively charged, surrounded by electrons that
are negatively charged. The entire assembly is electrically neutral. A very few, sometimes only
one, of the outer electrons, the valence electrons, are much less tightly bound to the atom than
the others. Thus the atom can be thought of as a massive core consisting of the nucleus and the
tightly bound electrons, all with a net positive charge, together with a much more mobile
electron, or electrons, with negative charge, attached to it. They behave rather like two
oppositely charged masses attached by a spring. When an electromagnetic wave is incident on an
atom or molecule, the electric field of the wave pushes the valence electron and the positive core
apart. Because the electron is many orders of magnitude lighter than the core we can consider the
electron as moving and the core remaining effectively stationary. The electric field of the wave is
oscillating and so the electron oscillates in position at the same frequency. The wave also has a
magnetic field but magnetic fields can react with charged particles only when they are moving
and for the action to be significant the speed of the charged particle must be a good fraction of
the speed of light. The high frequency of the light wave implies that the electrons never attain
any significant speed before they are reversed and so magnetic effects are virtually zero. This is
the origin of the simplification in optics that the refractive index and characteristic admittance
are proportional. This simplification is not possible in the lower-frequency microwave region.
The vibrating electron-atom system extracts energy from the electromagnetic field to induce
and sustain the vibration. A positive and a negative charge that are slightly separated is called a
dipole. An oscillating dipole radiates an electromagnetic wave, just like an antenna. The
frequency of the wave is identical to the oscillation frequency of the dipole. Thus we have a
mechanism that extracts electromagnetic energy and another that emits it continuously, as long
as the incident wave is present.
Molecules are made up of two or more atoms that are bound together. They may behave in a
quite similar way to an atom, appearing to consist of a relatively mobile electron or electrons
surrounding a positively charged mass. Sometimes the more mobile electrons are attached more
strongly to one or more atoms that make up a substructure of the molecule. In those cases the
molecule behaves more like two large oppositely charged masses and the molecule is known as
ionic. Whichever is the more appropriate model, atoms and molecules in an electromagnetic field
behave also like a dipole and extract and return energy from the field.
In most cases, virtually all the energy that is received by the dipole is reemitted.
When we have just one atom or molecule present the emission from the dipole in any
direction is proportional to the cosine of the angle between the direction and the axis of the
dipole. This effectively redirects a small amount of light in a direction completely different from
the original and such a process is known as scattering. This particular type of scattering is
known as dipole scattering. The scattering is not much changed if we introduce just a few more
atoms or molecules. They are so far apart that they act quite independently and the scattered light
is just stronger according to their number. The blue color of the illumination from the sky during
the day is because it consists of light scattered by the few molecules that make up the outer
Macleod: Optical Coatings Page 44 Design through Manufacture

atmosphere. When, however, the molecules become much more numerous, so numerous that
their outputs are closely related, a curious thing happens. The resultant electric and magnetic
fields cancel out except for a net wave that propagates in exactly the same direction as the
primary wave. The independent scattering is usually called incoherent scattering. The
cancellation of other light and the enhancement of the forward-propagating scattered wave is
called coherent scattering. Gases at atmospheric pressure, liquids and solids have atoms and
molecules sufficiently closely packed to give coherent scattering.
However, the transfer of energy to the dipoles and its return all takes time and the scattered
field lags behind the primary field. To obtain the resultant we must add primary and scattered
beams and again this must take account of magnitude and direction. When we do all that we find
that the resultant wave while it has the same frequency as the primary wave moves more slowly
through the medium. The denser the packing of the atoms or molecules and the greater the
induced dipole, the slower is the speed of the resultant wave.
Refractive index is the ratio of the speed of the primary wave to the speed of the resultant
wave. Thus, the greater the density of the material and the greater the dipole interaction, the
greater is the refractive index.
The dipole interaction becomes greater if the bond or spring that holds the electron to the
atom or molecule is weaker. Higher refractive index materials tend to have more weakly bound
electrons.
There are some other effects. The electron and nucleus attached by a spring act as a little
mechanical resonator. The nearer the frequency of the wave is to the natural frequency of the
resonator, the greater is the interaction. Imagine that we have a wave that is at a frequency that is
rather lower than the natural frequency of the resonator. As the frequency of the wave is
increased, so the interaction increases and, with it, the refractive index. When the frequency gets
quite close to the natural frequency of the resonator, losses in the system become much more
important and high absorption results so that the material is no longer transparent. This
absorption persists right through the resonant frequency, which is usually in the ultraviolet
region. Once the frequency of the wave becomes sufficiently high the interaction with the
resonator becomes quite small and the refractive index is close to unity. This happens beyond the
ultraviolet in the x-ray region.
In the other direction, when molecules and/or dense materials like solids and liquids are
involved, there are resonances that involve the heavier molecular components. These, because of
the mass of the components, correspond to frequencies in the infrared. The lowest frequencies
relate to the ionic materials. As the frequency of the wave is reduced towards the resonance the
absorption rises and terminates the transparent region. The full theory of the effect shows that in
both resonances the refractive index first falls then rises rapidly and then falls again as the
wavelength increases through the resonance. The gradual fall of refractive index with increasing
wavelength in the region of transparency is called normal dispersion. The rapid rise with
increasing wavelength in the center of the resonance is known as anomalous dispersion. It is
actually not anomalous but well understood.
Although materials do exhibit regions of usable transparency beyond the resonances that
bound the transparent region of normal dispersion, they are either in the soft x-ray region or in
Macleod: Optical Coatings Page 45 Design through Manufacture

the far infrared and therefore beyond the normal wavelength range in which the bulk of optical
applications are situated.
We have seen that the refractive index is larger the weaker the bond between the electrons and
the remainder of the atoms or molecules. However, the weaker the bond the weaker the effective
spring in the resonator consisting of the electron and its opposite number and therefore the lower
the resonant frequency. To reduce the wavelength corresponding to the shortwave boundary of
the region of transparency means a strengthening of the bond and therefore a reduction in the
refractive index of the material. It is impossible to find good high-index materials in the
ultraviolet, and the further into the ultraviolet we go the more difficult it is to find even slightly
elevated indices.

n, k

Tr anspar ent r egion

n
Nor mal disper sion

k k

Wavelengt h
Figure 5-1. Schematic diagram of the variation of refractive index and extinction coefficient in a typical
transparent optical material. The variation of refractive index in the transparent region is known as normal
dispersion.
The boundary of the region of transparency in the infrared is also dominated by resonances
but here we want the frequency to be reduced if the region is to be broadened. Reduction in the
resonant frequency implies a weakening of the bonds and an increase in the mass of the resonant
units. Ionic materials made up of heavy atoms give best results. The weak ionic bonds imply that
the materials are soluble and the very heavy atoms for longest wavelength resonators imply that
the metallic ions are likely to be poisonous.
The materials we have been discussing so far are ones that are insulators at low or zero
frequencies. Such materials are known as dielectrics and the term is used in optics to denote
materials of high transparency in the region of normal dispersion.
Dielectric materials include the optical glasses, many crystalline materials and plastic and
organic materials for optics.
There are two other major classes of materials that are much used in optics. Like the
dielectrics they are classified by terms borrowed from low frequency electrical behavior. They
are metals and semiconductors.
Macleod: Optical Coatings Page 46 Design through Manufacture

5.2 Metals
Metals conduct electricity at low or zero frequency and this means that they have current
carriers that are free to move through their bulk. Dielectrics are characterized by electrons that
are bound and not free to move. Metals are characterized by free electrons.
A single metal atom does not have a free electron. The outer valence electrons are bound to
the remainder of the atom by strong bonds. When metal atoms are combined into a metallic
solid, however, a strange thing happens. The various metal atoms share their valence electrons
with each other so that they become free to move through the solid, and to conduct electricity.
The electrons remain free even at the very high frequencies of optical waves.
Free electrons are affected by the electric fields of electromagnetic waves and they oscillate
but they do not form oscillating dipoles because they are not bound into resonant units. The
electrons simply absorb the light and do not return it to the primary wave. Light, therefore,
cannot penetrate far into a metal. The extinction coefficient, k, is very large. As the wavelength
increases, so does the extinction coefficient. In fact it is usually roughly proportional to λ.
A perfect metal with completely free electrons cannot support the propagation of a normal
electromagnetic wave. The wave that is incident on the surface of the metal is completely
reflected. The electric and magnetic fields do actually penetrate into the metal but they decay
very rapidly and exponentially and the electric and magnetic fields have a phase difference of
90° and so they can carry no energy. The term evanescent wave is used to describe the
electromagnetic behavior within the metal. The refractive index, n, is zero because the
oscillations of the electrons within the metal are all exactly in phase regardless of the depth into
the metal.
Real metals do not quite achieve this ideal behavior. There are some bound electron effects
like the dielectric materials and the free electrons offer some resistance. The extinction
coefficient, k, is large but the electron oscillations gradually change in phase with depth so that
the refractive index n is not zero, although it can be quite small. A certain proportion of the
energy does actually enter the metal where it is absorbed. We classify the metals that approach
ideal behavior as good and those that do not as poor.
At shorter wavelengths and higher frequencies, the electrons do not respond as freely to the
oscillating electric field and their effects diminish. The extinction coefficient becomes small and
the refractive index rises to unity. In the x-ray region metals are essentially transparent, rather
like air or a vacuum. The frequency where the change over from metallic to dielectric behavior
occurs is called the plasma frequency, Figure 5-2. The plasma frequency is usually in the
ultraviolet region of the spectrum.
Strictly speaking, the plasma frequency is slightly lower than the frequency at which the n and
k curves cross. The greater the relaxation time, that is the time it takes for a current to decay to
1/e of its value when an electric field is switched off, the more free the electrons and the closer
the plasma frequency to the cross-over frequency.
Macleod: Optical Coatings Page 47 Design through Manufacture

n, k

k
n

Plasma
fr equency

Wavelengt h
Figure 5-2. Variation of the optical constants of a typical metal. For wavelengths that are less that that
corresponding to the plasma frequency the ideal metal is reasonably transparent. For longer wavelengths it
reflects and absorbs strongly.
The normal incidence reflectance of a metal surface with refractive index n and extinction
coefficient k in a dielectric medium of index n0 is given by

R=
b n − ng + k
0
2 2
(5.1)
b n + ng + k
0
2 2

Because k is high, the absorption coefficient is also very high and the depth to which the light
penetrates is very small in a metal. Thus even quite thin metals appear just as bulk metals and a
quite thin layer on the surface of a transparent substrate is enough to make the substrate into a
reflector with the reflectance of bulk metal. This is the principle behind metal front-surface
mirrors and the metalizing of plastics to give them a metallic appearance.

5.3 Semiconductors
Semiconductors also get their name from their electrical behavior at low or zero frequency.
They are poor conductors showing that they have some free electrons but not nearly as many as
metals and as the temperature is reduced their conductivity falls.
If we imagine a dielectric material where the binding of the electrons to their molecular units
is very weak we have a material that has high refractive index but that has an absorption onset at
quite long wavelengths. Also because the binding of the electrons is exceedingly weak, they can
readily be detached by thermal effects from their molecules or atoms to become free. Thus
semiconductors appear like very high-index dielectrics with reasonable transparency in the
infrared. Because of the tendency of some electrons to be free, the residual absorption in
semiconductors is greater than in pure dielectrics. Impurities, particularly metallic impurities,
encourage the appearance of free electrons and purity is important if optical properties are to be
good.
A typical semiconductor that is much used in infrared optics is germanium. Germanium is
transparent at wavelengths longer than 1.6µm and has a refractive index of around 4.0. Gallium
Macleod: Optical Coatings Page 48 Design through Manufacture

aluminum arsenide alloys, have transparency from around 800nm and refractive indices around
3.2.
Semiconductors can be thought of as slightly lossy dielectrics with their regions of
transparency in the infared.

5.4 Metal and Dielectric Films


In optical coatings it is usual to divide the materials into two main classes, metals and
dielectrics. Semiconductors in their region of transparency are usually considered to be similar to
dielectrics.
The ideal properties of metals and dielectrics are listed in Table 5.1. Real materials depart
more or less from these ideals. For example, dielectrics have finite k and metals finite n. In each
case losses are associated with the departures from perfection but those in real metals are
significantly higher than those in real dielectrics.
Dielectric layers exhibit interference effects with properties that cycle as wavelength or
thickness changes. The phase thickness δ=2πnd/λ is the most important quantity. This is the
change in phase suffered by a wave on one traversal of the film. As λ increases, δ becomes
smaller, since n is changing only slightly, and so the properties of the film weaken. Metals
ideally do not exhibit any cyclic properties. The reflectance simply increases or decreases in the
same sense as β and/or k. Since β is substantially constant while k increases with λ, the
properties of the metallic film become stronger with wavelength. Ideally, a dielectric layer
should be most useful in a coating with high transmittance over an extended spectral region
while a metal would be most appropriate in a coating with high reflectance over a wide region. If
many layers are combined into a single coating then a dielectric assembly would lend itself to
reflect a band of shorter wavelengths and transmit a wide range of longer wavelengths while a
mixture of dielectric and metallic layers would be more suitable for a coating to transmit a
shorter band of wavelengths and reflect all longer ones. These ideal properties are summarized in
Figure 5-3
Macleod: Optical Coatings Page 49 Design through Manufacture

Table 5.1
Dielectrics Metals
k=0 y = − ik Y
y = nY k ∝λ
n independent of λ 2π kd
β= = constant
2π nd 1 λ
δ= ∝
y ∝λ
λ λ
y = constant
Become weaker, i.e. T Become stronger, i.e. R
increases, as λ increases increases, as λ increases
progressive wave evanescent wave
Eei bω t −κ z g
-α z
Ee 2 eiω t
Real dielectrics show lower Real metals show higher
losses losses
Summary of ideal properties of metals and dielectrics.

Figure 5-3. Broad characteristics of single layer (on left) and multilayer coatings (on tight) of ideal materials.
Macleod: Optical Coatings Page 51 Design through Manufacture

6 COATINGS WITH DIELECTRIC LAYERS


We now discuss specific types of optical coatings with reference to the techniques we have
established so far. In this chapter we deal principally with coatings having broad characteristics.
A separate chapter deals with narrowband filters.

6.1 Antireflection Coatings


An important application of thin film optics, and probably the largest in commercial terms, is
the reduction of surface reflectance.
The simplest form of antireflection coating is the quarterwave layer. The quarterwave rule
tells us that the reflectance of a quarterwave of admittance yf on a substrate of admittance ysub in
a medium of admittance y0 is

( y0 − y 2f / ysub )2
R= (6.1)
( y0 + y 2f / ysub )2

Zero reflectance will result if y 2f = y0 ysub . This is of course true only at those wavelengths for
which the layer is an odd number of quarterwaves. At other wavelengths the reflectance is higher
and at those for which the antireflection coating is a whole number of halfwaves thick and hence
an absentee layer, we know, immediately, that the reflectance is exactly that of the uncoated
substrate. Elsewhere we can use the methods developed earlier to calculate the performance. A
direct application of the matrix approach yields the following expression for the reflectance of a
single surface of admittance ysub coated with a layer of admittance yf and phase thickness δ in a
medium of admittance y0.

( y0 − ysub )2 cos2 δ + [( y0 ysub / y f ) − y f ]2 sin 2 δ


R= (6.2)
( y0 + ysub )2 cos2 δ + [( y0 ysub / y f ) + y f ]2 sin 2 δ

Probably the commonest application of an antireflection coating is in the reduction of the


surface reflection of crown glass of admittance 1.52. For a perfect antireflection coating in air
(admittance 1.0) we require a thin film material of admittance 1.52 , i.e. 1.233. No sufficiently
robust material of admittance as low as this exists. Magnesium fluoride, which has an admittance
of around 1.38 over the visible region, has, over the years, been found to be best in this
application. When deposited on a hot substrate at a temperature of 300°C, it is extremely
durable. The minimum reflectance of a quarterwave of magnesium fluoride on crown glass is
approximately 1.25% per surface. Figure 6-1 shows a curve of reflectance over the visible region
for such a coating. The higher reflectances in the red and the blue gives the coating its
characteristic magenta appearance in reflected light which has led to the use of the term
"bloomed" when describing an optical component treated with a reflectance reducing coating.
Magnesium fluoride will reduce to zero the reflectance of a substrate of admittance 1.382 , i.e.
1.90.
Macleod: Optical Coatings Page 52 Design through Manufacture

Design1: Reflectance
5

Reflectance (%)
3

0
400 450 500 550 600 650 700
Wavelength (nm)
Figure 6-1. The reflectance of a single-layer antireflection coating for the visible region.

Figure 6-2. The W-coat. An antireflection coating for the visible region in which the single quarterwave
coating has been broadened by the addition of a high-index (1.9) halfwave layer between the quarterwave
and the substrate. The design is now: Air | LHH | Substrate with yL=1.38, yH=1.9.

Figure 6-3. The admittance locus of the W-coat of Figure 6-2. The halfwave layer being an absentee does not
alter the reflectance at λ0(510nm).
Macleod: Optical Coatings Page 53 Design through Manufacture

Figure 6-4. The admittance locus of the W-coat of Figure 6-2. At 650nm, the high-index layer, now less than
a halfwave, moves the starting point of the following low-index layer away from the real axis in such a way
as to ensure that its termination point is near the point (1,0) where the reflectance is zero.
The characteristic of the single quarterwave layer can be much improved by adding a high-
index halfwave layer between the magnesium fluoride and the substrate, which broadens the
characteristic without altering the reflectance at the reference wavelength, where, of course, the
halfwave layer is an absentee. The admittance diagram of Figure 6-4 is of assistance in assessing
the action of the halfwave layer. This type of coating is often called a W-coat because of the
shape of the characteristic curve.
For very high-index substrates, such as silicon with admittance around 3.5 and germanium
with admittance 4.0, film materials with indices higher than magnesium fluoride are required,
and zinc sulfide, with admittance 2.3, and zinc selenide, with admittance around 2.5, are suitable
single-layer antireflection coatings, especially if good transmittance is essential beyond 10µm.
For shorter wavelengths, oxide materials such as zirconium dioxide, which are very much
tougher than either zinc selenide or sulfide, can be used.
The single layer coating is ideal only if low reflectance is required at one single wavelength.
High performance over a wider region, or improved performance at a single wavelength when
materials suitable for the single quarterwave are not available, demands additional layers. High-
index substrates present simpler design problems.
The vector diagram for a coating consisting of two quarterwaves chosen such that all vectors
are of the same length and negative in direction is shown in an earlier section. The required
indices are given by

y13 = y02 ysub (6.3)

y23 = y0 ysub
2
(6.4)

This gives zero reflectance at g=2/3 and 4/3 and a reflectance due to just one of the vectors at
g=1. This is a considerable improvement over the bare substrate. Low reflectance over a range of
wavelengths greater than 2:1 is obtained.
Macleod: Optical Coatings Page 54 Design through Manufacture

Figure 6-5. A two layer antireflection coating for germanium (y=4.0) with ideal values of admittance,
y1=1.59 and y2=2.52.

For germanium, n = 4.0, the ideal materials have indices n = 1.59 and n = 2.50, giving a
reflectance of 5.6% at λ0, compared with 36% for an untreated surface. Zinc selenide or zinc
sulfide as the high index material and cryolite as the low index, or combinations of oxides such
as zirconium oxide and aluminum oxide, are suitable, although with indices which are not quite
ideal. Figure 6-5 shows a typical performance of the ideal coating for the near infrared.
This vector design technique can be extended to greater numbers of layers, and hence greater
numbers of zeros, or to slightly different arrangements of two layers, giving zeros which are
closer together or further apart.
For low admittance substrates, the same theoretical approach is valid, but, as for the single
layer coating, the low indices which are required, cannot be obtained. For zero reflectance at just
one single wavelength, a two-layer arrangement, involving a high admittance layer next to the
substrate followed by a low admittance film, known often as a V-coat, is frequently used. The
admittance diagram helps in understanding the V-coat and an earlier section shoed the
admittance diagram for such a coating designed for glass (n=1.52) in an incident medium of air.
The vector diagram can also be used in understanding the coating. Analytical techniques yield
the following solutions for the optical thicknesses of the two layers:

RS UV
1

n1d1 1 LM ( y − y )( y 2 − y y ) y 2 OP 2
⋅ arctan ± M 2 3
=
MN ( y1 y3 − y0 y22 )( y0 y3 − y12 ) PPQ
0 2 0 3 1
(6.5)
λ 2π T W

R LM ( y − y )( y y − y ) y OP U
1
2 2
n2 d2 1 2
⋅ arctan S± M
T MN ( y y − y y )( y − y y ) PQVW
= 3 0
P
0 3 1 2
(6.6)
λ 2π 2
1 3 0
2
2
2
2 0 3
Macleod: Optical Coatings Page 55 Design through Manufacture

The values of optical thicknesses found from these equations must be correctly paired, and the
admittance diagram is helpful in this. A high-index thickness less than a quarterwave should be
paired with a low-index thickness greater than a quarterwave and vice versa. Titanium oxide and
silicon oxide are often used for such coatings for the visible region.
The limiting case of the V-coat is a coating consisting of two quarterwaves. The admittance
plot for this coating is two circles which just touch at the point of intersection with the real axis
to the right both of substrate and medium admittance. The relationship between the admittances
of the two layers must be given by:

( y1 / y2 )2 = y0 / y3 (6.7)

This quarter-quarter coating is one of a class of two layer coatings sometimes called
Muchmore coatings after the author of the first paper describing them[1].
For broader bandwidths, the most usual form of coating is the well-known quarter-half-quarter
coating. This coating was probably discovered by trial and error[2] but the most useful analytical
design method was devised by Thelen[3]. We simply take a Muchmore quarter-quarter coating,
and add a high admittance half-wave flattening layer between the two quarterwaves. This gives a
coating which consists of a quarterwave of intermediate admittance of around 1.7 next to the
substrate, followed by a halfwave of high (around 2.1) admittance, and ending with an outer
quarterwave of low admittance, usually magnesium fluoride. At the reference wavelength, the
halfwave is an absentee layer leaving the other two layers to act as an antireflection coating,
similar to the V-coat, for the reference wavelength. At wavelengths other than the reference, the
halfwave layer broadens the characteristic so that, instead of the single narrow zero which would
be obtained from the two quarterwaves on their own, a broad characteristic, covering the better
part of the visible region, is obtained.

Figure 6-6. The admittance diagram of a typical quarter-half-quarter coating. The design is Air | LHHM |
Glass where yH=2.3, yL=1.38, yM=1.7 and yglass=1.52.
Macleod: Optical Coatings Page 56 Design through Manufacture

Figure 6-7. The characteristic of the quarter-half-quarter coating of Figure 6-6.


We can also mention a coating which consists of the V-coat, with the low-index layer split
into two where the locus crosses the real axis. A halfwave layer of high admittance inserted here,
will act to flatten the characteristic, in much the same way as in the quarter-half-quarter coating.
This is a design which was originally published by Vermeulen[4] who arrived at it in a
completely different way. Arriving at the correct layer thicknesses is very much simpler,
however, when we look on it as a flattened V-coat. In this form, the performance is reasonably
good over the visible region but can be made rather better by a process of refinement in which
the layers are gradually relaxed by a computing process to values that give optimum
performance. Such a characteristic curve is shown in Figure 6-9. Many of the most modern
antireflection coatings use only two materials which makes the process simpler and more
reliable. This four layer coating is a good example of such coatings.

Figure 6-8. The admittance locus of the flattened V-coat, or Vermeulen, coating. The design is Air | LHH
0.2833L 0.2351H | Glass with yH=2.2, yL=1.38 and yglass=1.52. The admittance locus is plotted for the
reference wavelength, λ0 of 510nm.
Macleod: Optical Coatings Page 57 Design through Manufacture

Figure 6-9. The characteristic of the flattened V-coat of Figure 6-8.

Figure 6-10. The characteristic of the flattened V-coat after refinement. The design is now Air | 0.994L
2.129H 0.356L 0.238H | Glass.
The antireflection coatings for the visible region frequently require quarterwave layers of
intermediate indices, around 1.7. Such indices are not easily obtainable and those materials
which do have indices in that region are often very sensitive to preparation conditions so that
their indices are not particularly stable. Mixtures of materials have been, and are, used. An
alternative solution involves symmetrical periods. A symmetrical period, that is a combination of
layers which is completely symmetrical about its mid-point, can be replaced by a single
equivalent layer with optical thickness and refractive admittance which can be calculated
analytically. It is possible to work backwards, and to choose a symmetrical combination which is
a replacement for a single layer. The technique has already been described in the section on
symmetrical periods. We choose two readily obtainable indices, and the technique allows us to
replace any quarterwave layer with admittance between those of the two layers, with a
symmetrical period of our chosen materials. Halfwave layers can be treated as two quarterwaves
in series. This technique has the advantage that any antireflection coating consisting of
quarterwave or halfwave layers, can be replaced by an assembly of just two layer materials, the
highest and lowest in the coating design, for example. The disadvantage is that fractional
Macleod: Optical Coatings Page 58 Design through Manufacture

thicknesses, sometimes difficult to reproduce in manufacture, are required. An example of such


an approach where the quarter-half-quarter coating has been replaced by a four-layer equivalent
is shown in Figure 6-11.

Figure 6-11. The characteristic of a four-layer antireflection coating based on the quarter-half-quarter coating
of Figure 6-7 in which the intermediate index M has been replaced by a three-layer symmetrical period. The
design is now Air | L 2.286H 0.39L 0.286H | Glass, where yH=2.3, yL=1.38 and yglass=1.52. Note the
similarity between this design and the Vermeulen one. Refinement of this four-layer coating would give
virtually the same result as Figure 6-10.
Buffer layers[5] can sometimes be of use in the design of antireflection coatings. A halfwave
layer is an absentee layer at the reference wavelength whatever the admittance of the structure
within which it is inserted. It can have any dielectric characteristic admittance whatsoever and
still it remains an absentee. Its admittance is therefore the design parameter that can be varied to
achieve a desired flattening, or even sharpening, effect. A buffer layer can be inserted only at
certain points within a multilayer and it must have a prescribed admittance but the performance
of the coating at the reference wavelength is unaffected by changes in the thickness of the buffer
layer. Consider the quarter-quarter coating:

Air| LM | Glass

with L a quarterwave of admittance 1.38, M a quarterwave of admittance 1.70 and where air has
admittance 1.00 and glass 1.52. The admittance diagram at λ0, which we assume in what follows
to be 510nm, is shown in Figure 6-12.
Macleod: Optical Coatings Page 59 Design through Manufacture

Figure 6-12. The admittance diagram of the quarter-quarter antireflection coating at the reference
wavelength. The termination point of the layer next to the substrate is 1.9. A layer of admittance 1.9 and any
thickness added at this point will have a locus consisting of a point and will not affect the structure of the
admittance diagram in any way at this wavelength.
A layer added to a medium that has identical admittance will have an admittance locus
consisting of a point which is completely independent of the layer thickness. This is the principle
of the buffer layer. The intersection of any locus with the real axis is therefore a potential point
for insertion of a buffer layer. It is sufficient that the buffer layer have characteristic admittance
equal to the admittance of the point of intersection with the real axis. In Figure 6-12 there is a
point of intersection with the real axis at 1.9. A buffer layer of admittance 1.9 can therefore be
inserted at that point without disturbing the admittance diagram at that wavelength. We therefore
add a buffer layer, arbitrarily a quarterwave thick, to obtain the design

Air| LHM | Glass

where H indicates a quarterwave of admittance 1.9. The performance of the new design is
compared with the old in Figure 6-13 where again 510nm is assumed as the reference
wavelength. The improvement is modest but positive. The admittance diagram at wavelength
600nm is shown in Figure 6-14. The buffer layer now acts as a bridge between the adjacent
layers in much the same way as a halfwave flattening layer.
Macleod: Optical Coatings Page 60 Design through Manufacture

Figure 6-13. The performance curves of the two designs. The buffer layer has flattened the quarter-quarter
design but not to the same extent as the halfwave layer in the quarter-half-quarter coating.

Figure 6-14. The admittance diagram for the buffer layer coating at a wavelength of 600nm. The design is
Air|LHM|Glass where L, H and M are quarterwaves of admittances 1.38, 1.9 and 1.7 respectively. Air has
admittance 1.00 and glass 1.52.
Although the buffer layers here is responsible for an improvement in performance that is
modest compared with the quarter-half-quarter coating, nevertheless the design can form a useful
starting design for refinement. A refined performance is shown in Figure 6-15. The design is
now

Air|0.978 L1676
. H 0.649 M | Glass
Macleod: Optical Coatings Page 61 Design through Manufacture

Figure 6-15. The performance of an antireflection coating of design Air | 0.978L 1.676H 0.649M | Glass
where the admittances have the same values as in Figure 6-14. The reference wavelength is 510nm. This
design was derived by refining the three quarterwave design of Figure 6-13.
Antireflection coatings are sometimes required for two separated spectral regions. There is no
consistent technique that will always work but simply a series of solutions of particular cases.
The most straightforward is two zeros of reflectance at two wavelengths having a three to one
ratio. Here, all that is required is a layer that is a quarterwave of the correct admittance to give
zero reflectance at the longer of the two wavelengths. Since the layer will be three quarterwaves
at the other wavelength it will give zero reflectance there also, provided there is no dispersion.
Figure 6-16 shows the performance of a single-layer antireflection coating for glass at 0.5µm and
1.5µm consisting of a single layer of MgF2.
A halfwave high-index layer can be used to flatten the performance in the 500nm region. It
should be inserted such that the existing three quarter wavelength layer is cut into a halfwave
next to the substrate and a quarterwave outermost, that is a design of Air | LHHLL | Glass. The
LL layer can contribute to the flattening effect. However, although this new HH layer will be an
absentee layer at 510nm, it will be far from absentee at three times that wavelength. To prevent
deterioration of the performance at 1530nm the layer must be made a halfwave there which
implies three halfwaves at 510nm, or a design of Air | LHHHHHHLL | Glass. The performance
of this design is shown in Figure 6-17. The characteristic curve looks much as might be
expected, except that there is a deep additional minimum at near 800nm. This could therefore be
a useful design for that wavelength coupled with the visible region. To understand the source of
this additional unexpected minimum we turn to an admittance diagram for that wavelength.
Macleod: Optical Coatings Page 62 Design through Manufacture

Figure 6-16. Single layer of MgF2 (n=1.38) on glass (n=1.52) λ0=510nm.

Figure 6-17. The performance of design Air | LHHHHHHLL | Glass where L has index 1.38 and H has index
1.95 (to avoid overemphasis of the flattening effect) and λ0=510nm.
Macleod: Optical Coatings Page 63 Design through Manufacture

Figure 6-18. The admittance diagram of the coating of Figure 6-17 where the wavelength is 800nm. The low
index layers are the smaller arcs and the large circle corresponds to the high-index HHHHHH layer.
The reason is now clear. The high index layer sweeps round in a large circle that bridges the
gap between the end of the first low index layer and the correct starting point for the second.
Because the high-index layer is quite thick the circle is described almost twice. This implies that
the correction could involve a thinner high-index layer. Of course any change in the thickness of
the high-index layer will affect the performance at 1530nm but if we now neglect that
wavelength and concentrate on the visible and one minimum in the near infrared, we can reduce
the high index layer thickness to two halfwaves instead of three. This will preserve a flattening
in the visible, and, since the longwave minimum now requires a high-index locus a little less
than a complete circle, the minimum will be pushed to a longer wavelength. The performance of
such a coating, with design Air | LHHHHLL | Glass is shown in Figure 6-19 and the admittance
diagram for the longwave minimum at near 1100nm in Figure 6-20.
Macleod: Optical Coatings Page 64 Design through Manufacture

Figure 6-19. The performance of design Air | LHHHHLL | Glass with other parameters as in Figure 6-17.

Figure 6-20. The admittance diagram of the coating of Figure 6-19 at a wavelength of 1100nm corresponding
to the minimum in the near infrared.
The concept of the buffer layer can sometimes be useful in the design of antireflection
coatings for two wavelength regions. The layer is designed to be a buffer in the shorter
wavelength region so that it takes part in the achievement of the longer wavelength performance.
We can take as an example, the design of a coating for the visible region and for the narrow
region around 1100nm using materials of indices 1.38 and 2.4. We have to arrange for one of the
layers to be a buffer and we will take it to be of index 2.4. A possible starting design is given in
Table 6-1.
Macleod: Optical Coatings Page 65 Design through Manufacture

Table 6-1
Starting design with buffer layer
Layer Index Thickness
0 1.000000 Medium
1 1.380000 0.250000
2 2.400000 0.088189
3 1.380000 0.064763
4 2.400000 0.500000
5 1.380000 0.064763
6 2.400000 0.088189
7 1.520000 Substrate
λ0=510nm

The admittance locus is shown in Figure 6-21. Layers 7 and 6 are designed to move from the
substrate at 1.52 to the real axis at 2.4 where layer 5, the buffer layer is situated. This layer is
simply a point in the admittance diagram. Layers 4 and 3 are exactly equivalent to layers 7 and 6,
but in reverse order, to return the admittance to the starting point. Finally a quarterwave of 1.38
completes the coating. The thickness of the buffer layer is arbitrarily set as a half wavelength.
The performance of this coating is shown in Figure 6-22. It has the general form of an
antireflection coating but the reflectance is not particularly low.

Figure 6-21. The admittance locus of the six-layer starting design. The buffer layer is represented by a single
point at 2.4 on the real axis.
Macleod: Optical Coatings Page 66 Design through Manufacture

Figure 6-22. The performance of the starting design with admittance locus in Figure 6-21. Although the
average reflectance is not particularly low, the bandwidth of the design is large and so it is very suitable as a
starting design for refinement.
Nevertheless, this does form a very useful starting design for subsequent refinement. Using a
specification of zero reflectance over the regions 400nm to 700nm and 1000nm to 1200nm we
arrive finally at the design in Table 6-2.

Table 6-2
Final design
Layer Index Thickness
0 1.000000 Medium
1 1.380000 0.304693
2 2.400000 0.126933
3 1.380000 0.061316
4 2.400000 0.711801
5 1.380000 0.067280
6 2.400000 0.091886
7 1.520000 Substrate
λ0=510nm

The performance of this design is shown in Figure 6-23 and the admittance diagrams for
510nm and for 1100nm in Figure 6-24 and Figure 6-25. The buffer layer is now slightly
displaced at 510nm so that it exercises a small effect, but it is still largely a buffer. At 1100nm,
however, it takes a very large part in assuring low reflectance. This six-layer design is quite an
efficient one in terms of performance for a given number of layers.
For starting designs to have good potential for refinement there must be plenty of
opportunities for adjustment. This implies buffer layers, halfwave layers and especially those
Macleod: Optical Coatings Page 67 Design through Manufacture

cusp-like features at the junction between two layers where there is virtually a reversal of
direction.

Figure 6-23. The performance of the six-layer antireflection coating after refinement.

Figure 6-24. The admittance locus at 510nm of the coating of Figure 6-23. The small circle on the right is the
buffer layer, now not quite a buffer although its influence is very small.
Macleod: Optical Coatings Page 68 Design through Manufacture

Figure 6-25. Admittance locus of Figure 6-23 at 1100nm corresponding to the long wavelength minimum.
The buffer layer corresponds to the large partial circle on the right and is clearly exerting great influence on
the performance in contrast to its slight effect in Figure 6-24.

6.2 Reflection-Increasing Coating


6.2.1 Quarterwave Stack
A single quarterwave layer changes the admittance of a substrate from ysub to y2/ysub. Thus, a
succession of quarterwaves will give an admittance of the form:

y12 y32 y52…


Y= 2 2 2 (6.8)
y2 y4 y6 … ysub

with ysub in the denominator, as shown, or numerator, depending on an odd or even number of
layers, respectively. Thus if there are alternate high and low index layers then the admittance can
be made either very high or very low, in each case leading to high reflectance. Under normal
circumstances, the highest reflectance is obtained with high index layers next to the substrate and
outermost. If there are x+1 high index layers and x low index, then the reflectance is given by

LM y − y b g OP
2 x +2
H
2

R=M
y y P
0 2x

MM y + y b g PP (6.9)
L sub
2 x +2
H

N y y Q
0 2x
L sub
Macleod: Optical Coatings Page 69 Design through Manufacture

With lossless layers, the possible reflectance is limited by the number of layers which can be
deposited. The higher this number, the higher the reflectance. In practice, the ultimate
reflectance which can be achieved depends on the residual losses, absorption and scattering. For
good quality reflectors for the visible region the total residual losses could be assumed to be of
the order of 0.05 per cent. If the values of the extinction coefficients for both high and low index
layers are known then a useful formula which relates the total absorption loss of a quarterwave
stack to the intrinsic losses of the materials from which it is constructed, is:

A=
b
2π n0 k H + kL g final layer of high index (6.10)
cn 2
H −n 2
L h
and A=
c
2π nH2 k L + nL2 k H h final layer of low index (6.11)
c
n0 n − n 2
H
2
L h
A typical quarterwave stack characteristic is shown in Figure 6-26.
Several features of this characteristic should be noted. The high reflectance associated with
the wavelength or frequency for which the layers are quarterwaves, also exists at those
wavelengths or frequencies for which the layers are odd multiples of a quarterwave. Thus, there
is high reflectance not just for g=1, but for g=3, g=5, g=7 and so on. Each high reflectance zone
is limited in extent. There are sharp transitions between zones of high reflectance and of high
transmittance. The width of the high reflectance zone is given by 2∆g, where g = λ0 / λ and

∆g =
2
⋅ arcsin
RS n H − nL UV (6.12)
π Tn H + nL W
The principal features of the quarterwave stack are thus the regions of high reflectance and
low transmittance around frequencies or wavelengths for which the layer optical thicknesses are
odd integral numbers of quarterwaves, separated by regions of relatively high transmittance with
appreciable ripple. By itself, the quarterwave stack is used as an efficient reflector with very low
losses, but, in addition, the quarterwave stack is used as a sub-unit of many different types of
optical coating.
Macleod: Optical Coatings Page 70 Design through Manufacture

Figure 6-26. A typical quarterwave stack characteristic. The design of this coating is Air|(HL)5 H|Glass with
yH=2.35, yL=1.35 and yglass=1.52. The characteristic is plotted in terms of g (=λ0/λ).

6.2.2 Broadband Reflectors


The limited width of the high reflectance zone of the quarterwave stack is often a
disadvantage and it is useful to be able to produce coatings having a wider reflectance zone.
There are three principal techniques for achieving this. The first consists simply of depositing a
number of quarterwave stacks on top of each other with different reference wavelengths. The
main problem here is to avoid the occurrence of transmission peaks in the high reflectance zone.
Unless special precautions are taken, these always occur when two overlapping quarterwave
stacks are deposited on top of each other, because the assembly acts in much the same way as the
Fabry Perot filter, to be described later. The cure is to insert a single coupling layer which has a
thickness of one quarterwave at a wavelength between the centers of the two stacks. A typical
design would be H(LH) L"(H'L') H' where H(LH) is centered on λ1, (H'L') H' on λ2, and L" on
(λ1 + λ 2 ) / 2 .
The second technique consists of depositing a stack where the thicknesses of the layers are
staggered in some way, the effect being similar to a quarterwave stack where the reference
wavelength varies throughout. Various thickness progressions such as geometric or arithmetic
have been tried with very similar results.
A third approach uses a computer to refine an existing design which has a performance only a
little short of what is required.
All these techniques yield much broader high reflectance zones than the basic quarterwave
stack but there is a number of important disadvantages. The first is that the phase of the
reflectance usually varies fairly rapidly with wavelength and this means that it can be dangerous
to use broadband reflectors in interferometers. Ciddor[6] has designed special coatings
specifically for interferometric use. The second difficulty is an increased sensitivity to absorption
losses. For some wavelengths, the layers which contribute to the high reflectance are buried
beneath many other layers, and the light must penetrate through these upper layers before being
reflected. Any transmission losses in the upper layers reduce the reflectance below the level
which would be achieved by a simple quarterwave stack. Since, in the visible region, the
Macleod: Optical Coatings Page 71 Design through Manufacture

absorption losses are usually greater at the shortwave end, it is marginally better to have the
thinnest layers on top of the stack, so that the longer wavelengths are those which must penetrate
the coating.

Figure 6-27. A broadband reflector for the visible region of design: Air | (HL)5 H 1.2L (1.4H 1.4L)5 1.4H |
Glass with yH=2.35, yL=1.35 and yglass=1.52.

Figure 6-28. The broadband reflector of Figure 6-27 shown with a greatly enlarged vertical scale.

6.3 Edge Filters


Edge filters transmit wavelengths on one side, and stop wavelengths on the other side, of a
critical wavelength, known as the edge wavelength. Filters in which the transmission band is
situated at wavelengths longer than the edge are known as long-wave-pass filters, and those
where it is shorter, as short-wave-pass. Closely allied to the edge filter is the dichroic beam
splitter where the rejected band is not merely stopped but also reflected, so that two spectral
regions are separated.
Macleod: Optical Coatings Page 72 Design through Manufacture

The simplest type of edge filter is an absorption filter and there exists a wide range of colored
glasses and dyed gelatin filters which can be used as long-wave-pass filters for the near
ultraviolet, visible and near infrared. Then, most materials have sharp onsets of absorption due to
electronic transitions as the wavelengths become shorter, semiconductors in the infrared and
visible, and dielectrics in the ultraviolet. Short-wave-pass edges also exist, especially associated
with lattice vibrations in ionic materials, but they are usually less sharp and there is a more
limited choice than for long-wave-pass edges. If a natural edge filter exists, it is often the best
solution. If not, then thin-film interference coatings are employed.
The basic edge filter is the quarterwave stack. Since it is characterized by regions of high
reflectance separated by regions of transmittance, it can be used either as long-wave-pass or
short-wave-pass by suitable choice of the position of the stop bands. The principal deficiency of
the simple quarterwave stack is the prominent ripple in the pass bands. Many design techniques
for reducing this ripple exist. The most powerful and straightforward depend on the properties of
symmetrical periods.
We look for a symmetrical period which can be used as the basis for a quarterwave stack. The
simplest is a three-layer period, consisting of a quarterwave of one material surrounded by two
eighthwave layers of the alternate material. That is either (0.5H L 0.5H) or (0.5L H 0.5L). The
quarterwave stack can then be considered as

0.5 HLHLHLHLH LHL0.5 H

or 0.5LHLHLHLHL HLH 0.5L

slightly, but not too, different from the original quarterwave stack.
The equivalent index and thickness of such a basic period is given in an earlier section. Over a
wider range of g than is listed there, the pattern repeats itself. To change from a simple period to
q periods, we have only to multiply the phase thickness by q. Thus the design of an edge filter
reduces to the design of matching coatings between the basic slab, which is equivalent to the q
periods, and the bounding media.
The ripple in edge filters can be seen in this model as due to the mismatching of the
equivalent single layer representing the stack, and the surrounding media. The amplitude of the
ripple can be estimated by considering the envelopes. The extreme values will be given when the
equivalent layer is either a quarterwave thick or a halfwave. The unmatched system will
therefore have reflectance envelopes given by

Ly −E / y
R=M 0
2
sub
OP for the quarter wave thicknesses
2

(6.13)
Ny +E / y
0
2
sub Q
and R=M
Ly − y0 sub
OP for the half wave thicknesses
2

(6.14)
Ny + y0 sub Q
Macleod: Optical Coatings Page 73 Design through Manufacture

Figure 6-29. The design of an edge filter using symmetrical periods reduces to the task of finding two
matching layers as shown. These must match the substrate to the slab of material of admittance E, equivalent
to the stack, and match the slab of admittance E to the incident medium. In their simplest form these will be
quarterwaves at the appropriate value of g or λ of admittance Eysub and yo E respectively.
Since E is a function of g, so too are the envelopes. For quarterwave matching layers, ya
between incident medium and multilayer and yb between multilayer and substrate we have the
modified envelope equations:

L y − y y / ( E y ) OP E is a quarter wave
R=M 0
2
a
2
b
2
sub
2

(6.15)
N y + y y / (E y )Q
0
2
a
2
b
2
sub

and R=M
L y − y y / y OP E is a half wave
0
2
a sub
2
b
2

(6.16)
Ny + y y / y Q
0
2
a sub
2
b

These expressions strictly apply only to the region around the value of g where the matching
layers are quarterwaves but they give a good qualitative indication of performance over a wider
region.
In some cases, the multiple set of periods by itself, is a sufficiently good match for the
substrate and surrounding medium, often the case in the visible region. If a matching assembly is
required, a quarterwave of the correct index at the appropriate wavelength will usually be
sufficient. There are advanced techniques for eliminating ripple in the few cases where
quarterwave layers cannot achieve the desired performance. These usually involve constructing
the matching modules from symmetrical periods to create the same kind of rapid variation that is
seen in the value of E near the edge of the high reflectance zone.
Straightforward edge filters are shown in Figure 6-30 and Figure 6-31.
Macleod: Optical Coatings Page 74 Design through Manufacture

Figure 6-30. A longwave pass filter for the visible region of design
Air | (0.5H L 0.5H)7 | Glass, where yH=2.35, yL=1.35, yglass=1.52 and yair=1.00. λ0=450nm. Here the
symmetrical period by itself is sufficiently well matched to the surrounding media.

Figure 6-31. A shortwave pass filter for the infrared region of design:
Air | (0.5L H 0.5L)7 0.8L | Germanium, using zinc sulfide as low index material and germanium as high.
yH=4.00, yL=2.30, yGe=4.00 and yair=1.00. λ0=4.0µm. Here the symmetrical period by itself is sufficiently well
matched to the air but requires matching to the substrate.
The dispersion of the equivalent admittance near the edge of the pass region is a great
problem. This makes it very difficult to design simple matching assemblies. Ideally the matching
assembly should have dispersion similar to that of the basic symmetrical period. A reasonable
approximation to this is obtained by using a symmetrical period of the same kind but with the
reference wavelength shifted. This is known as the method of shifted periods and was devised by
Thelen. We can illustrate the method in the design of a long wave pass edge filter for the visible
region using materials of indices 2.35 and 1.45. The equivalent admittance of the system
( 0.5 HL0.5 H ) is shown in Table 6-3.
Macleod: Optical Coatings Page 75 Design through Manufacture

Table 6-3
E and γ for nH=2.35 and nL=1.45

(0.5H L 0.5H)
g E γ/π
0.00 1.84594 0.00000
0.05 1.84451 0.05147
0.10 1.84017 0.10295
0.15 1.83280 0.15448
0.20 1.82218 0.20607
0.25 1.80798 0.25774
0.30 1.78972 0.30954
0.35 1.76676 0.36149
0.40 1.73818 0.41365
0.45 1.70275 0.46610
0.50 1.65869 0.51893
0.55 1.60342 0.57229
0.60 1.53305 0.62642
0.65 1.44136 0.68173
0.70 1.31753 0.73895
0.75 1.14019 0.79975
0.80 0.85451 0.86902
0.85
0.90
0.95
1.00 E and γ are imaginary

We assume that the incident medium will be air and that the substrate will be glass of index
1.52. We take the design point as somewhere near g=0.75. The value of admittance at this value
of g is 1.14019. this is quite a reasonable match to air but we will need an admittance of 1.317 to
match it to the substrate. The value of equivalent admittance corresponding to g=0.7 is very
close to this figure at 1.31753. The value of γ / π is 0.73895, very close to 3/4. Two periods will
have a value of γ / π of twice this quantity, that is 1.4779. This is close to 1.5 which is three
quarterwaves and therefore should be a reasonable matching coating between the symmetrical
period and the substrate. We simply have to shift it by the correct amount. It has to be at a
shorter wavelength and therefore we multiply its thicknesses by the factor 0.7/0.75. This leads to
the design:

Air|( 0.5 HL0.5 H )7 ( 0.467 H 0.933L0.467 H )2 | Glass

The performance is shown in Figure 6-32. Compare this with Figure 6-30. The performance of
each filter is shown in Figure 6-33 with an expanded vertical scale. Rather more careful choice
of matching assemblies using a table with a smaller vertical interval would achieve still further
improvements in performance.
Macleod: Optical Coatings Page 76 Design through Manufacture

Figure 6-32. The performance of an edge filter designed using the method of shifted periods. The design is
Air | (0.5 HL 0.5 H) 7 (0.467H0.933L 0.467H) 2 |Glass
where H represents a quarterwave of index 2.35, L of 1.45, air has index 1.00 and glass 1.52.

Figure 6-33. A comparison of the performance in the pass region of the filters of Figure 6-32, i.e. using the
method of shifted periods, and of Figure 6-30, i.e. using the simple approach.
Two common types of edge filter, probably more correctly described as dichroic beam
splitters, are heat-reflecting filters and cold mirrors. The heat-reflecting filter is a short-wave-
pass filter that reflects the near infrared and transmits the visible region, while the cold mirror
reflects the visible and transmits the near infrared. Both are usually based on quarterwave stacks,
the heat-reflecting filter usually being deposited over a heat-absorbing glass and the cold mirror
usually consisting of two quarterwave stacks after the method of Figure 6-27.
Macleod: Optical Coatings Page 77 Design through Manufacture

6.4 Bandstop or Minus Filters


A bandstop filter is one which transmits on either side of a spectral region which is rejected
(the stopband). Bandstop filters, (sometimes also known as minus filters), are essentially
quarterwave stacks, which are designed to have low ripple on either side of the high reflectance
zone, instead of on just one side, as in edge filters. The elimination of the ripple is, however, a
more difficult task than in the edge filter.

Figure 6-34. A bandstop, or minus, filter. The design is


Glass | (0.5LB 0.5L) 2 (0.5LH 0.5L ) 6 (0.5LB 0.5L) 2 |Glass
with yL=1.56, yB=1.91, yH=2.34 and yglass=1.56. After Thelen [7]
useful general analytical technique for eliminating ripple on both sides of the stop band has been
devised by Thelen[7]. This makes use of the concept of equivalent admittance together with a
dummy medium. The filter is designed to have minimum ripple in the dummy medium, which is
then matched to the actual surrounding media. The thickness of the dummy medium can then be
permitted to shrink to zero without perturbing the properties of the filter. The dummy medium
can have any appropriate value of index which simplifies the design of coating. It is a technique
which has been found useful in the design of many types of thin-film coating. In this case the
dummy medium is chosen to have identical admittance with the outermost material, A, of the
basic symmetrical period, ( 0.5 AB 0.5 A) . If the equivalent admittance at g, on the longwave side
of the stopband of such a symmetrical period is given by E, then it can readily be shown that the
equivalent admittance at (2-g), on the shortwave side, will be E ′ = y A2 / E . Since reflectance and
transmittance are unaffected when all admittances are multiplied by a constant factor, or when
they are replaced by their reciprocals, or both, a matching coating consisting of similar
symmetrical periods of the same λ0 and with the same outermost layer material, which are
designed for wavelengths longer than the stopband, must also be correct for wavelengths shorter
than the stopband. The innermost materials will not be the same as the materials for the basic
stack. The remainder of the design task, is the matching of the dummy medium, to the substrate
at one end, and to the surrounding medium at the other. A design of this type is illustrated in
Figure 6-34.
Macleod: Optical Coatings Page 78 Design through Manufacture

Figure 6-35. A minus filter created by refining a basic stack. The thicknesses of the layers in the basic stack
were deliberately detuned to reduce the width of the rejection zone from that corresponding to a quarterwave
stack. The design is given in the table.
This design has three materials and therefore presents a more difficult production problem that
a two-material design. It is possible to construct quite acceptable minus filters with two
materials. Analytical techniques involve constructing the basic minus filter from a series of
symmetrical periods and then grading the outermost periods by gradually reducing the amount of
high index material with respect to the low, so as to make a multilayer matching assembly,
which, although not consisting of strict quarterwaves, nevertheless can be very efficient.
Unfortunately, because the individual periods are a little way from quarterwave equivalents,
quite a number of them have to be used for the match. An alternative that can be slightly more
economical in terms of total number of layers is straightforward refinement of a starting
quarterwave stack. The width of the high reflectance zone can be varied by adjusting the relative
thicknesses of the high and low index layers. It is a maximum when the layers are quarterwaves.
An example of such a design is shown in Figure 6-35. For this 44-layer coating, the refinement
process was a gradual one starting with only the outermost layers being available for refinement
at the start. After the first cycle of refinement, the next to outermost layers were also freed and
the process was repeated. In this way, eventually the outer eleven layers on either side were
permitted to take part resulting in the performance curve shown. After refinement one of the
layers was so thin that it was eliminated, leaving the 42 layers that are listed in the design table.
Macleod: Optical Coatings Page 79 Design through Manufacture

Table 6-4
Minus filter design
Layer Index Thickness Layer Index Thickness
0 1.00 Medium ------- ------- ----------
1 1.45 0.5780 28 2.45 0.1016
2 2.45 0.0321 29 1.45 0.3882
3 1.45 0.4593 30 2.45 0.1016
4 2.45 0.0852 31 1.45 0.3932
5 1.45 0.4142 32 2.45 0.0902
6 2.45 0.0580 33 1.45 0.4288
7 1.45 0.4244 34 2.45 0.0636
8 2.45 0.0969 35 1.45 0.4123
9 1.45 0.3983 36 2.45 0.0733
10 2.45 0.1016 37 1.45 0.4527
11 1.45 0.3882 38 2.45 0.0200
12 2.45 0.1016 39 1.45 0.4580
13 1.45 0.3882 40 2.45 0.0362
14 2.45 0.1016 41 1.45 0.3470
------- ------- ---------- 42 1.52 Substrate
λ0=700nm

6.5 Chapter 6 References


[1] R. B. Muchmore, “Optimum band width for two layer anti-reflection films,” Journal of
the Optical Society of America, vol. 38, pp. 20-26, 1948.
[2] L. B. Lockhart and P. King, “Three-layered reflection-reducing coatings,” Journal of the
Optical Society of America, vol. 37, pp. 689-694, 1947.
[3] J. T. Cox, G. Hass, and A. Thelen, “Triple-layer antireflection coating on glass for the
visible and near infrared,” Journal of the Optical Society of America, vol. 52, pp. 965-
969, 1962.
[4] A. J. Vermeulen, “Some phenomena connected with the optical monitoring of thin-film
deposition and their application to optical coatings,” Optica Acta, vol. 18, pp. 531-538,
1971.
[5] J. Mouchart, “Thin film optical coatings. 5: Buffer layer theory,” Applied Optics, vol.
17, pp. 72-75, 1978.
[6] P. E. Ciddor, “Minimization of the apparent curvature of multilayer reflecting surfaces,”
Applied Optics, vol. 7, pp. 2328-2329, 1968.
[7] A. Thelen, “Design of optical minus filters,” Journal of the Optical Society of America,
vol. 61, pp. 365-369, 1971
Macleod: Optical Coatings Page 81 Design through Manufacture

7 MULTIPLE CAVITY FILTERS


7.1 Bandpass Filters
In this type of filter, a given spectral band is passed while the regions on either side are
stopped. The class can be divided loosely into broad-band-pass and narrow-band-pass filters, the
classification depending on the design rather than on the actual width of the pass band.

7.2 Broad Bandpass Filters


These are normally made simply by placing two edge filters in series. The simplest
arrangement is to have each edge filter on a separate substrate surface. If, however, the entire
filter must be on the same single substrate surface, then the most powerful design technique
involves the equivalent index - symmetrical period approach. In the pass band, the two
quarterwave stacks, consisting of symmetrical periods must be matched to each other, to the
incident medium and to the substrate. Quarterwave matching is usually sufficient.

7.3 Narrow Bandpass Filters


The simplest type of narrow-band filter is based on the Fabry-Perot etalon, which consists of
two equally-reflecting surfaces spaced a distance d apart. The transmittance characteristic
consists of a series of narrow bands, equally spaced in frequency, and given by

TaTb 1
TF = ⋅ (7.1)
[1 − Ra Rb ] 1 + F sin [{(ϕ a + ϕ b ) / 2} − δ ]
2 2

4 Ra Rb
where F=
[1 − Ra Rb ]2

2π n d cos ϑ
δ= the phase thickness of the layer
λ
Ra , Rb are the reflectances of the two etalon plates
Ta , Tb are their transmittances
and ϕ a , ϕ b are the phase shifts on reflection
Usually the reflectors will be well matched and if we assume also that the phase shifts on
reflection are zero or 180°, then the transmittance becomes

T2 1
T= ⋅ (7.2)
[1 − R] 1 + F sin 2 δ
2
Macleod: Optical Coatings Page 82 Design through Manufacture

4R
with F=
(1 − R) 2

which gives the well known series of narrow transmission bands centered on δ=mπ where m is
the order number. The width of the pass band measured at half the peak transmittance, the
halfwidth, is given by
(1 − R)
∆λH = (7.3)
mπ √ R
where λp is the peak wavelength. The finesse of the etalon is defined as the ratio of the interval
between fringes to the fringe halfwidth and is given by

π√R
F = (7.4)
(1 − R)
and the resolving power by
λp
PR = = mF (7.5)
∆λH

A thin-film Fabry-Perot filter, in its simplest form, consists of a spacer, which is always a
dielectric and of optical thickness equal to an integral number of halfwaves at the required
wavelength, surrounded by two thin metal layers, which act as the reflectors. This type of filter is
known as a metal-dielectric. The spacer is seldom greater than three halfwaves thick. There is a
significant phase shift on reflection from the metal layers which displaces the peak wavelength
of the filter from the position predicted from spacer optical thickness only. Absorption in the
metal layers is the principal factor which limits the performance. From (7.2) peak transmittance
is given by

T2 1
TP = 2
= (7.6)
(1 − R) (1 + A / T ) 2
where A is the absorptance of each reflector. In the visible region, the most suitable metal is
silver which has a residual absorption loss of around 4% which implies that the reflectance must
be lower than 92% if the peak transmittance is to be greater than 25%. Typically, first-order
metal-dielectric filters can have bandwidths of 10-20nm in the visible with peak transmittances
in the region of 20-40%. First-order filters of metal-dielectric construction have no long-wave
sidebands (the dispersion of the silver acts to maintain the high reflectance into the infrared and
so retain the rejection). Second order filters can give narrower pass bands with higher peak
transmittance at the expense of the appearance of the first order peak at a longer wavelength.
Thus the first order metal dielectric filter simply requires the addition of a long-wave-pass filter
on the short-wave side of the peak to suppress the higher-order peaks for it to be a useful narrow-
band filter.
A typical commercial metal-dielectric narrow-band filter will consist of a cemented assembly
of the thin-film element and an absorption long-wave-pass filter. The long-wave-pass filter will
Macleod: Optical Coatings Page 83 Design through Manufacture

usually act also as a cover for the metal-dielectric component, protecting it against the effects of
moisture and abrasion. Such filters are fairly straightforward to produce and are, therefore,
usually the cheapest which are available for the visible and near infrared.
In the ultraviolet, silver has poor performance and aluminum is normally used as the metal.
Cements for attaching cover slips are not available beyond around 300nm, and an outer dielectric
layer is usually added for protection which is, however, far from being as effective as a cemented
cover, and ultraviolet filters must be treated with rather greater care than is necessary for visible
filters.

Design1: Transmittance
100
Transmittance (%)

80

60

40

20

0
300 500 700 900 1100 1300 1500 1700
Wavelength (nm)
Figure 7-1. A metal-dielectric Fabry-Perot filter. The design is of the form metal-dielectric-metal deposited
on glass and with a glass cover slip. The metal is silver, with a geometrical thickness of 42nm and the
dielectric, cryolite, with a geometrical thickness of 180nm. Dispersion of the optical constants is included in
the calculation.
The metal-dielectric filter is not able to provide very narrow halfwidths, and, for higher
performance, the metal layers must be replaced with dielectric reflectors, which, as we have
seen, are capable of very low losses. The use of quarterwave stacks in place of the metal layers
permits the achievement of halfwidths in the visible region of 0.5nm with peak transmittance, for
the basic element, in excess of 50%. The most popular materials for these all-dielectric narrow-
band filters in the visible and near infrared, are zinc sulfide and cryolite, which, although soft
and easily damaged by abrasion, have very high optical performance and are very easy to
deposit, and, like the metal-dielectric filters, are normally protected by a cemented cover slip or
sideband-suppression filter. Very recently, because of the need for the greatest possible stability
and because of the presence of slight residual birefringent peak splitting in narrowband filters
deposited by straightforward thermal evaporation (see the section on microstructural effects)
there has been a move to the construction of filters for applications in communication systems
from oxide materials such as silica and titania deposited by one of the energetic processes. The
amorphous structure of these materials removes the form birefringence and their increased
packing density, because of the energetic process, improves the stability.
The quarterwave stack has high reflectance over a limited region only and so the rejection
zone of the all-dielectric filter is limited to the same extent. The sidebands of transmission,
which exist outside the high-reflectance zone of the reflectors, must usually be suppressed, and
the normal method is to place a first-order metal-dielectric filter in series, which removes the
Macleod: Optical Coatings Page 84 Design through Manufacture

long-wave sidebands and much of the short wave, together with a long-wave-pass absorption
filter to eliminate the remaining short-wave bands.

FP1: Transmittance
100

80
Transmittance (%)

60

40

20

0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-2. An all-dielectric Fabry-Perot filter of design
Air | ( HL)5 HH ( LH )5 | Glass , of zinc sulfide, yH=2.35, and cryolite, yL=1.35. yglass=1.52 and λ0 is
1000nm.

FP1: Transmittance
100

80
Transmittance (%)

60

40

20

0
500 750 1000 1250 1500
Wavelength (nm)
Figure 7-3. The Fabry-Perot filter of Figure 7-2 showing the performance over a wider region. Note the high
transmittance in the longwave and shortwave sidebands where the basic quarterwave stack reflectors are no
longer effective.
An expression for the halfwidth of an all-dielectric Fabry-Perot filter can be derived from
(7.3) with the appropriate values of R for the all-dielectric reflectors. However, the phase shift
associated with the reflection from a quarterwave stack changes with wavelength, and the rate of
change narrows the pass band of an all-dielectric filter. Including a correction for this effect, the
halfwidth of an mth order all-dielectric Fabry-Perot filter is given by
Macleod: Optical Coatings Page 85 Design through Manufacture
∆λH 4n0 nL ( nH − nL )
2q

= high − index spacers (7.7)


λ0 m π nH2 q +1 ( nH − nL + nL / m )

∆λH 4n0 nL2 q −1 ( nH − nL )


and = low − index spacers (7.8)
λ0 m π nH2 q ( nH − nL + nL / m )

High-index layers are assumed to be outermost, there being q high-index layers in each reflector
not counting the spacer layer. A typical all-dielectric Fabry-Perot filter is shown in Figure 7-2.

7.4 Multiple-Cavity Filters


The shape of the Fabry-Perot peak is triangular. It is possible to couple several Fabry-Perot
elements together to give a pass band of more nearly rectangular shape. Such filters are known
as multiple-cavity filters. For coupling two such Fabry-Perot elements together it is sufficient to
deposit the second over the first, with a single quarterwave coupling layer in between. A two-
cavity filter is sometimes known as a double-half-wave filter or DHW. Two cavities coupled in
this way almost always give an acceptable result. When, however, the number of cavities is
increased, pronounced ripple is frequently present in the pass band which can make the designs
unusable, and a more systematic design technique is required. Again this can be based on
symmetrical periods as described by Thelen[1].
For calculations that do not involve a change in the angle of incidence, a symmetrical system
of thin films can be replaced by a single equivalent layer with an equivalent admittance E and
equivalent phase thickness γ that behave in similar fashion to the y and δ of a real film[2-4].
Repeated similar symmetrical periods, say q of them, may then be represented by greater
thicknesses, qγ, of the single equivalent layer. If the original layers are wholly dielectric then the
equivalent layer exhibits regions of dielectric behavior with real E and γ separated by region of
pseudo-metallic behavior with imaginary values for E and γ. The pseudo-metallic regions
correspond to regions of potential high reflectance where addition of dielectric layers has little
effect. The dielectric regions correspond to potential pass bands where suitable dielectric
matching layers will assure high transmittance.
Instead of dealing with cavities, the filter is broken down into a series of symmetrical
assemblies of quarterwaves such as (AB)qA, where A and B represent quarterwaves. It can readily
be shown that such a symmetrical assembly exhibits a narrow region around g=1 where the
admittance and phase thickness are real. This is a potential pass band. Repeating the basic period
many times simply creates an assembly that has the attributes of a single real layer with the
equivalent admittance of the period and a phase thickness that is that of the basic period
multiplied by the number of periods. Completion of the design then consists of the construction
of matching assemblies between the periods and the substrate and the incident medium.
The equivalent admittance of a symmetrical assembly of design (AB)qA can be shown to be
E = yAq+1/yBq. The symmetrical period analysis also shows that the bandwidth of the region
where the equivalent admittance is real is

∆λ 4(1 − yL / yH )
= (7.9)
λο π ( yH / yL ) q
Macleod: Optical Coatings Page 86 Design through Manufacture

which is a sufficiently accurate measure of the halfwidth of the filter. Note that the expression is
identical for either high or low index layers outermost in the symmetrical period.
We can now attempt the design of a simple filter. Let us base it on a period

HLHLHLHLHLHLHLH

where H indicates a quarterwave of admittance 2.35, and L an admittance of 1.35. The filter is to
be deposited on glass, admittance 1.52, and a glass cover is to be cemented over it. We can
therefore assume that glass is both incident medium and substrate. The equivalent admittance of
this assembly is yH8/yL7. The matching problem is then to match yH8/yL7 to yglass, 1.52. We use
quarterwaves, limiting ourselves to H and L for simplicity in manufacture. It is fairly easy, using
the quarterwave rule, to show that a system, HLHLHLH, will transform the admittance of the
basic symmetrical period to yL, that is 1.35. This is a fairly good match for glass, and certainly it
is difficult to do much better with these materials. The complete design of the filter is then

Glass | HLHLHLH ( HLHLHLHLHLHLHLH ) q HLHLHLH | Glass

Substituting the appropriate values in (7.8), we have for ∆λ/λ0 , 0.011. This implies for a
reference wavelength of 1000nm, a width of 11nm, that is edges of the pass band at 994.5nm and
1005.5nm. The calculated transmittance of a multiple-cavity filter of this design, with q=3, is
illustrated in Figure 7-4. Note that the above figures do not quite represent the halfwidth but
rather a width lower down the passband edges. Nevertheless it is sufficiently close to the
halfwidth to make it unnecessary to attempt to derive a more precise formula which would
inevitably add great complication. Sideband blocking filters must be added in exactly the same
way as for the Fabry-Perot filter.

Multiple-cavity
100

80
Transmittance (%)

60

40

20

0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-4. A multiple-cavity filter. Design:
Glass | HLHLHLH ( HLHLHLHLHLHLHLH )3 HLHLHLH | Glass , with yH=2.35, yL=1.35,
yglass=1.52. λ0=1000nm.
Macleod: Optical Coatings Page 87 Design through Manufacture

7.5 Higher Performance in Multiple Cavity Filters


The curve of Figure 7-4 shows the square shape of the passband of a multiple-cavity filter but
also illustrates one of the problems inherent in this type of design, the "rabbit's ears", or the
rather prominent peaks at either side of the passband. This can become even worse with
increasing numbers of periods. Figure 7-5 with a greater number of symmetrical periods, but
otherwise similar to the filter of Figure 7-4, shows this clearly.

Multiple-cavity
100

80
Transmittance (%)

60

40

20

0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-5. A multiple-cavity filter similar to that of Figure 7-4 but with a greater number of symmetrical
periods. Design:
Glass | HLHLHLH ( HLHLHLHLHLHLHLH )5 HLHLHLH | Glass , with yH=2.35, yL=1.35,
yglass=1.52. λ0=1000nm. Note the very prominent peaks at the edges of the passband.
The reason for this problem feature is the dispersion of the equivalent admittance of the
symmetrical period. In the design approach, this is assumed to be a constant across the passband
but in reality it varies considerably, tending to either zero or infinity at the passband edges, see
Figure 7-6. It is, in fact, exactly the same problem as in edge filters where better ripple
suppression near the edge demands a matching system that exhibits similar dispersion. Shifted
periods are, however, difficult to arrange in the case of bandpass filters because of the need for
ripple suppression at both edges of the passband. However, inspired by the shifted periods
technique, we seek a similar solution, where part of the matching is due to a symmetrical system
that has a dispersion of the appropriate form so that its matching remains reasonably good even
when the equivalent admittance to be matched to the surrounding media is varying. Any of the
symmetrical periods we are dealing with will have an odd number of quarterwaves so that the
equivalent phase thickness at g=1 will be an odd number of π/2. This implies that the period
could, itself, be used as a simple matching assembly. Since the passband in this type of filter is
usually narrow, the matching condition will not vary too much over the passband width. In order
to make use of this possibility, we have to find at least pairs of symmetrical periods that will
permit one to be used as a matching assembly for the other. Attempting to find two, or more,
periods that have the correct relationship at g=1 for one to match the other to the substrate or
incident medium is difficult. If we could find two periods of different width but with the same
Macleod: Optical Coatings Page 88 Design through Manufacture

central admittance, then we could continue to use the straightforward matching illustrated in
(AB) which uses a series of quarterwave layers and is perfectly satisfactory at the center of the
passband. A solution lies with higher order periods.

(HL)^7 H - Equivalent Admittance


1000

800

600
E

400

200

0
994 996 998 1000 1002 1004 1006
Wavelength (nm)
Figure 7-6. The equivalent admittance of (HL)7 H over the potential passband. Note the rapid change near
the edges of the passband. This dispersion of equivalent admittance is very difficult to match to an essentially
dispersionless medium.

Equivalent admittances
400

300
E

200

100

0
980 990 1000 1010 1020
Wavelength (nm)
Figure 7-7. The equivalent admittances of symmetrical periods from narrower to broader in order
HHHLLLHLHLHLHLLLHHH, HHHLHLHLHLHLHHH, HLLLHLHLHLHLLLH and HLHLHLHLHLH with
H representing characteristic admittance 3.5 and L 1.35. The straight line represents the admittance that all
have at g=1.
The addition of further halfwave layers to the outside of a symmetrical period, does not
change its equivalent admittance at the passband center nor does it change the sense of curvature
of the variation of equivalent admittance. Figure 7-7 shows the admittances of HLHLHLHLHLH,
HHHLHLHLHLHLHHH, HLLLHLHLHLHLLLH and HHHLLLHLHLHLHLLLHHH. All have
Macleod: Optical Coatings Page 89 Design through Manufacture

the value yH6/yL5 at g=1 and all exhibit a gradually increasing admittance as the value of g moves
away from unity. The wider curves have values of admittance intermediate between the narrower
curves and the value that all possess at g=1. All represent an odd number of quarterwaves at g=1
but the broader curves remain closer to an odd number of quarterwaves than the narrower as g
varies. They could therefore be used to match the narrower ones to a notional medium of
constant admittance, yH6/yL5. Matching of the dispersionless notional medium to the incident and
emergent media is then a straightforward matter of a series of quarterwaves, as before.
A simple example uses just two of the periods from Figure 7-7, HLHLHLHLHLH and
HHHLHLHLHLHLHHH:

q
Glass | HLHLH HLHLHLHLHLH ( HHHLHLHLHLHLHHH ) HLHLHLHLHLH HLHLH | Glass

The characteristic curves of two such filters are shown in Figure 7-8 and Figure 7-9.

Multiple-cavity
100

80
Transmittance (%)

60

40

20

0
980 990 1000 1010 1020
Wavelength (nm)
Figure 7-8. A multiple-cavity filter similar to that of Figure 7-4 but using periods of increasing order to
improve the passband ripple. Design:
Glass | HLHLH HLHLHLHLHLH (HHHLHLHLHLHLHHH)2 HLHLHLHLHLH HLHLH | Glass
with yH=2.35, yL=1.35, yglass=1.52. λ0=700nm. Note the much flatter passband top compared with Figure 7-4
and Figure 7-5.
We need an expression for the width of such filters. This is determined principally by the
highest order periods. If we write the expression for the highest order period as:

mABABA… BAmA
Macleod: Optical Coatings Page 90 Design through Manufacture

where there are 2x+1 layers including the layers mA, then we can show that the bandwidth,
defined in the same way as before, is given by:

x
∆λ 4 ⎛ yL ⎞ ( yH − yL )
= ⎜ ⎟ (7.10)
λο mπ ⎝ yH ⎠ ( yH − yL + yL / m)

This expression reduces to (7.9) if m=1. Using the expression to calculate the bandwidth of
the filters of Figure 7-8 and Figure 7-9, we find 0.018, implying passband edges at 706.5nm and
693.6nm.

Multiple-cavity
100

80
Transmittance (%)

60

40

20

0
980 990 1000 1010 1020
Wavelength (nm)
Figure 7-9. A multiple-cavity filter similar to that of Figure 7-8. but with three central high order periods
rather than two. Design:
Glass | HLHLH HLHLHLHLHLH (HHHLHLHLHLHLHHH)3 HLHLHLHLHLH HLHLH | Glass
with yH=2.35, yL=1.35, yglass=1.52. λ0=700nm.
Designs arrived at in this way will be satisfactory for a wide range of applications where
ripple within the passband must be small. However there are applications where even the
performance in Figure 7-8 and Figure 7-9 is inadequate. There are requirements in dense
wavelength division multiplexing for peak transmittances in excess of 99%, for example. A
useful technique that is somewhat empirical uses additional matching layers. In the following
filter H and L indicate admittances of 2.35 and 1.35 and where the substrate is glass of
admittance 1.52 and the incident medium is air of admittance 1.00. These correspond to the
materials we have been using so far and we prefer not to change at this stage. Recently much use
has been made of dense silica and tantala in the manufacture of narrowband filters, particularly
for wavelength division multiplexing but the design techniques are similar. The filter we use as
an example is given by:

Air | L (HL)3 H (HL)7 H (HH(HL)7 HHH)2 (HL)7 H H(LH)3 | Glass


Macleod: Optical Coatings Page 91 Design through Manufacture

The performance is shown in Figure 7-10. The filter is matched to an incident medium of air
rather than the glass we have been using and so there is an extra L layer next to the incident
medium.

Basic filter transmittance


100

80
Transmittance (%)

60

40

20

0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-10. Transmittance of filter: Air | L (HL)3 H (HL)7 H (HH(HL)7 HHH)2 (HL)7 H H(LH)3 | Glass.
The loss is purely a reflection loss. No absorption is involved. Thus it should be possible by
correct matching to reduce the loss to zero and increase the transmittance to 100%. However, we
need to accomplish this in as simple a fashion as possible. We take as our target, therefore, to
increase the transmittance so that it is greater than 99% over the entire passband top. An
analytical approach is unlikely to be profitable and so we use some logic and then rely on
automatic methods.
The filter structure is thick and complicated and the matching must be capable of
accommodating considerable dispersion of the admittance of the assembly. This implies that a
very thin system of layers is unlikely to be of much value. We therefore assume from the start
that the matching layer will be fairly thick.
3-layer system
Index Optical thickness
(full waves)
Air Incident medium
1.3500 4.6377
2.3500 4.4624
1.3500 4.7197
Filter structure
Macleod: Optical Coatings Page 92 Design through Manufacture

5-layer system
Index Optical thickness
(full waves)
Air Incident medium
1.3500 3.0727
2.3500 2.9991
1.3500 3.0674
2.3500 2.9651
1.3500 3.1973
Filter structure

We try two different starting designs for the matching layer, a three-layer LHL and a five-
layer LHLHL arrangement to replace the single L matching layer of the original design (the layer
next to the air). However, we find single quarterwave thicknesses insufficient for a good match.
Some trial and error finds preferred starting designs of 18L18H18L and 12L12H12L12H12L
although the final result is not very sensitive to the exact starting design thicknesses. Some
gentle refinement with only the matching layers taking part then yields final matching systems as
shown in the tables and filter characteristics in Figure 7-11, Figure 7-13 and Figure 7-14. The
admittance locus of the three-layer matching system is shown in Figure 7-12

Extra matching
100

80
Transmittance (%)

60

40

20

0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-11. The two curves when the additional matching 3 and 5-layer systems are added. The 5-layer
system is superimposed over the 3-layer, which is almost invisible in the scale of the figure. An expanded
transmittance is shown in Figure 7-13.
Macleod: Optical Coatings Page 93 Design through Manufacture

Three-layer matching
2

Im(Admittance)
0

-1

-2
0 1 2 3 4 5
Re(Admittance)
Figure 7-12. The admittance locus of the three-layer matching system plotted at the center wavelength of the
filter.

Extra matching
100.0
Transmittance (%)

99.5

99.0

98.5

98.0
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-13. The two curves of Figure 7-11 plotted on an expanded scale to show the differences. The 3-layer
system (lighter curve) is slightly inferior to the 5-layer system.
Macleod: Optical Coatings Page 94 Design through Manufacture

5-layer extra matching


0
-5

log(Transmittance) (dB)
-10
-15
-20
-25
-30
-35
-40
990 995 1000 1005 1010
Wavelength (nm)
Figure 7-14. The performance of the filter with 5-layer matching from Figure 7-11 and Figure 7-13 shown on
a logarithmic scale.
It is very difficult to carry out this type of design in a completely systematic way. There are
other techniques but the quarterwave thicknesses of the basic filter design help considerably in
the thickness control of the deposition process. The final matching layers are much thicker than
quarterwaves and are the final layers of the structure and so their monitoring signals are also
quite favorable.

7.6 Tilted Performance of Narrowband Filters


7.6.1 Fabry-Perot Filters
For tilts up to around 20° in collimated light, the principal effect on a Fabry-Perot filter is a
shift in the pass-band position towards shorter wavelengths with peak transmittance and
halfwidth virtually unchanged. The shift can be calculated as being entirely due to the variation
in phase thickness of a spacer layer of the same order but with effective index n . The effective
index is given by
1/2
∗ ⎡ m − (m − 1)(nL / nH ) ⎤
n = nH ⎢ ⎥ (7.11)
⎣ (m − 1) − (m − 1)(nL / nH ) + nH / nL ⎦
for filters of order m with high index spacers, and
1/2
∗ ⎡ m − (m − 1)(nL / nH ) ⎤
n = nL ⎢ 2⎥
(7.12)
⎣ m − m(nL / nH ) + (nL / nH ) ⎦
for low index spacers. For first order filters, these reduce to

n∗ = (nH nL )1/2 (7.13)


Macleod: Optical Coatings Page 95 Design through Manufacture

1/2
∗⎡ nL ⎤
and n =⎢ 2⎥
(7.14)
⎣1 − (nL / nH ) + (nL / nH ) ⎦
for high and low index spacers respectively. Then,

λ peak = λ0 cosθ (7.15)

where θ = arcsin{(no sin θ o ) / n∗} (7.16)

and n0 and θ0 refer to the incident medium. For small angles,

∆λο θ 02
= (7.17)
λ 2 ( n∗ )
2

with λ peak = λ0 − ∆λ0

and θ0 in radians. With θ0 in degrees, this becomes

∆λ0 −4 θ 02
= 1.5 x10 (7.18)
λ (n ) ∗ 2

This gives the shift in peak wavelength with angle θ0. In practice a filter will be exposed
simultaneously to a range of angles of incidence. The simplest case is a cone of semiangle ψ.
The analysis is simpler in terms of wavenumber ν (= 1/ λ ) . The peak shifts to a new position

ν peak = 0.5[ν 0 + (ν 0 + ∆ν ′)] = ν 0 + 0.5∆ν ′ (7.19)

where ∆ν ′ is the shift corresponding to angle ψ (degrees)

∆ν ′ −4 ψ2
= 1.5 x10 (7.20)
ν0 (n ) ∗ 2

The new halfwidth is given by

Vψ2 = V02 + (∆ν ′) 2 (7.21)

and the peak transmittance by Θ


2
T ′ V0 ⎧ ∆ν ′ ⎫ 1 ⎧ ∆ν ′ ⎫
= arctan ⎨ ⎬ ≈ 1− ⎨ ⎬ (7.22)
T0 ∆ν ′ ⎩ V0 ⎭ 3 ⎩ V0 ⎭
Macleod: Optical Coatings Page 96 Design through Manufacture

where T0 and V0 refer to normal incidence.


This analysis can be further extended to a cone incident other than normally, provided that
some simplifying assumptions are made. Let the angle of incidence be Θ and the cone semiangle
be ψ. Then the range of angles encompassed by the filter is Θ ± ψ . If Θ < ψ then the result is
simply that for a filter illuminated normally by a cone of semiangle Θ + ψ . If Θ > ψ then we
have three frequencies, ν0 corresponding to normal incidence, ν1 the peak corresponding to
angle of incidence Θ −ψ and ν2 the peak corresponding to angle of incidence Θ + ψ . Then the
new filter peak is simply

0.5(ν 1 + ν 2 ) (7.23)

The halfwidth is V′ = [V02 + (ν 2 −ν 1 ) 2 ]1/2 (7.24)

and the peak transmittance


2
T' V0 ⎧ν − ν ⎫ 1 ⎧ν − ν ⎫
= arctan ⎨ 2 1 ⎬ = 1 − ⎨ 2 1 ⎬ (7.25)
T ν 2 −ν 1 ⎩ V0 ⎭ 3 ⎩ V0 ⎭

(ν 2 − ν 1 ) is proportional to Θ ⋅ψ and excellent agreement has been found between results


calculated from these assumptions and measurements made on real filters, for values of Θ ⋅ψ up
to 100 degree2.
A significant result is that if a filter is to be used at maximum efficiency in a cone of light at
normal incidence, then its peak wavelength in collimated light at normal incidence should be
slightly longer than the required wavelength, to allow for the shift in equation (7.18).

7.6.2 Multiple-Cavity Filters


Multiple-cavity filters behave in exactly the same way as Fabry-Perot filters in collimated
light. The expressions for effective index are identical. The effects of an incident cone of light
are similar, but much depends on the sharpness of the rectangular profile of the multiple-cavity
filter. Peak transmittance will, in general, fall off less rapidly until the total shift across the cone
of incident light is equivalent to the halfwidth of the filter. A slightly pessimistic result will,
therefore, be obtained with the Fabry-Perot expressions.

7.7 Chapter 7 References


1. Thelen, A, Design of multilayer interference filters, in Physics of Thin Films, G. Hass
and R.E. Thun, Editors. 1969, Academic Press: New York and London. p. 47-86.
2. Epstein, L I, The design of optical filters. Journal of the Optical Society of America,
1952. 42: p. 806-810.
Macleod: Optical Coatings Page 97 Design through Manufacture

3. Epstein, L I, Improvements in heat reflecting filter. Journal of the Optical Society of


America, 1955. 45: p. 360-362.
4. Epstein, L Ivan, Design of optical filters. Part 2. Applied Optics, 1979. 18(10): p. 1478-
1479.
Macleod: Optical Coatings Page 99 Design through Manufacture

8 COATINGS WITH METALLIC LAYERS


The simplest type of reflection-increasing coating is a single opaque metal layer, and the
commonest coating of this type is aluminum, which is used for the ultraviolet, visible and
infrared, sometimes unprotected but more frequently with a single dielectric overcoat to improve
the abrasion resistance. Metal layers all suffer from some absorption and can never, therefore,
achieve the very high reflectances possible with purely dielectric multilayers.

Design1: Reflectance
100
Reflectance (%)

95

90

85

80
200 400 600 800 1000 1200 1400 1600 1800 2000
Wavelength (nm)
Figure 8-1. The reflectance of an aluminum coating calculated from the measured optical constants. Compare
with the reflectance of boosted aluminum shown earlier.
The reflectance of a metal surface with admittance n-ik in an incident medium of admittance
y0 is given by:

( y 0 − n )2 + k 2
R= (8.1)
( y 0 + n )2 + k 2

For an incident medium of air, y0 will be unity. Then the term k 2 dominates the expression that
is then close to unity, that is 100% reflectance. In fact, the reflectance of aluminum, calculated
from the optical constants over a range 200nm to 2µm, is shown in Figure 8-1. The drop in
reflectance at 800nm is a consequence of the dip in the values of extinction coefficient at the
same wavelength, and is an unfortunate feature of an aluminum reflector. This leads to problems
in optical instruments that have many mirrors in series. Silver has better optical properties in this
region, as demonstrated by the reflectance values in Figure 8-2, but silver tarnishes readily when
exposed to a normal atmosphere and it also has very poor reflectance in the ultraviolet.
The environmental resistance of a metal can be improved by overcoating with a dielectric.
Aluminum is protected by a naturally produced thin transparent oxide film and is therefore
sometimes used with no other protection - astronomical reflectors are examples of components
with such a coating. Otherwise aluminum is frequently overcoated with a halfwave of silicon
dioxide. (The halfwave is explained later). Aluminum oxide has been found to be a useful
overcoat for silver because it sticks well to it. It is not a very effective moisture barrier by itself,
and can be used as a bonding layer between the silver and a thicker overcoat of silicon oxide,
which is a good barrier material. Unfortunately, the silicon oxide absorbs heavily beyond 8µm
and is, therefore, less suitable for longer wavelengths, especially as it shows a sharp dip in
Macleod: Optical Coatings Page 100 Design through Manufacture

reflectance for p-polarized light for oblique incidence. For the ultraviolet, aluminum is normally
used, frequently with an overcoat of quartz, for the near ultraviolet, or magnesium fluoride or
lithium fluoride for the extreme ultraviolet. These overcoats are partly to protect the coating
against mechanical damage, but more to prevent the growth of the natural aluminum oxide film,
which can cause a significant drop in ultraviolet reflectance.

Design1: Reflectance
100
Reflectance (%)

95

90

85

80
200 400 600 800 1000 1200 1400 1600 1800 2000
Wavelength (nm)
Figure 8-2. The reflectance of a silver coating. Note the steep dip in the near ultraviolet, which continues to
virtually zero reflectance.

Design1: Reflectance
100
Reflectance (%)

95

90

85

80
200 400 600 800 1000 1200 1400 1600 1800 2000
Wavelength (nm)
Figure 8-3. The reflectance of a gold-coated mirror.
The effect of an added thin dielectric layer on the reflectance of an underlying metal is always
to reduce the reflectance. The resultant reflectance is the result of multiple beam interference in
the dielectric layer but a good approximation is given by the two-beam interference illustrated in
Figure 8-4. The beams a and b are almost in phase and when the layer is of vanishing thickness
add to give the reflectance of the uncoated metal. As the dielectric layer increases in thickness
the phase difference between the beams increases and the interference becomes more and more
destructive. Only when the layer is close to a half wave does the reflectance return to that of the
uncoated metal. An admittance locus also predicts the fall in reflectance. Since the metal
admittance is in the fourth quadrant of the complex plane, any overcoating dielectric with
admittance greater than that of the incident medium (this will certainly be the case at normal
incidence) gives reducing reflectance with increasing thickness until the intersection with the
real axis.
Macleod: Optical Coatings Page 101 Design through Manufacture

a b

Dielectric Layer

Bulk Metal

Figure 8-4. The reflectance of an overcoated metal is given approximately by the interference of the beams
reflected from the outer dielectric surface and the inner dielectric/metal interface.
The reflectance reduction in front surface reflectors due to the overcoating materials can be
serious. Best optical results are obtained with the lowest possible index of overcoat. The
calculated results in the following table show this problem clearly. For aluminum the drop in
reflectance is potentially very large and so it is usual to deposit roughly a halfwave of the lowest
index material available, normally silicon dioxide. Silver is not nearly greatly affected.
Table 8-1
Aluminum Runcoated Rmin Dmin Rmax Dmax
(0.82-i5.99)
SiO2 91.63 83.64 0.2128 91.86 0.4628
n=1.45
CeO2 91.63 65.90 0.1925 92.44 0.4425
n=2.30
Note: Dmin and Dmax are measured in full waves.

In the case of metals that reflect poorly the two beams a and b can be closely matched in
magnitude and then the interference at the correct thickness of dielectric (close to a quarterwave)
can result in almost completely destructive interference and, therefore, very low reflectance.
Metals like titanium can be almost completely antireflected by an appropriate thickness of their
own oxides. This leads to the appearance of brilliantly colored fringes that are often used
artistically.
A high performance metal can be antireflected in a similar way but the outer surface of the
dielectric layer has insufficient reflectance on its own. The addition of a thin metal layer to the
outer dielectric surface gives a simple antireflection coating. Adjustment of both the outer metal
thickness and the dielectric layer thickness permits perfect antireflection to be attained. Because
it is used for matching the phases of the two beams the dielectric layer is sometimes known as a
phase-matching layer.
Macleod: Optical Coatings Page 102 Design through Manufacture

a b

Metal Reflector

Phase Matching

Bulk Metal

Figure 8-5. An antireflection coating for virtually any high performance metal consists of a phase matching
dielectric layer and outer metallic reflector. Variation of the thickness of the outer metal changes the
magnitude of the reflected beam and alteration of the phase matching layer adjusts the phase so that the
interference with the beam from the bulk metal surface can become completely destructive giving perfect
antireflection properties.
An antireflection coating for bulk aluminum illustrated in Figure 8-6 is constructed from a
silica phase matcher followed by a thin chromium reflector. This is the principle behind many
different types of coatings for optical data storage. For example a thermoplastic phase matching
layer can become so altered by the application of heat from a laser beam that the antireflection
properties are destroyed and the irradiated spot reflects strongly.

Figure 8-6. The performance of a two-layer dielectric/metal antireflection coating for aluminum.
The same techniques can be used to antireflect a thin metal film on both sides. The metal
thickness must be such that the reflectances at the outer surface and at the metal/phase matching
layer interface are equal. The lower the outer reflectance the thinner must be the metal layer.
Since the transmittance of this component will be important the outer reflectors should not be
absorbing. This implies the use of dielectric systems.
Macleod: Optical Coatings Page 103 Design through Manufacture

Reflector

Phase Matching

Metal

Phase Matching

Reflector

Figure 8-7. A thin metal film antireflected on both sides by a phase matching layer and reflector combination.
For a very thin metal the mismatch between high-admittance phase matching layers and the
surrounding media can give a sufficiently high matching reflectance leading to a three-layer
coating. Performance of such a coating is plotted in Figure 8-8.

Table 8-2
Incident Phase matcher Metal Phase matcher Emergent
medium medium
Glass TiO2 Ag TiO2 Glass
Massive 28nm 10nm 28nm Massive

Figure 8-8. The three-layer coating consisting of a thin layer of silver with a phase matching layer of titania
on either side and sandwiched in glass. (Wavelength units nm and reflectance and transmittance in %)
The performance in Figure 8-8 is rather broad and can be narrowed by using more silver,
which in turn demands a higher matching reflectance from the outer surface. An additional
quarterwave of magnesium fluoride on either side, making a five-layer coating, increases the
matching reflectance and permits the use of a thicker silver layer.
Macleod: Optical Coatings Page 104 Design through Manufacture

Figure 8-9. The performance of the five layer coating.


Increasing the outer reflectance still further, and therefore the silver thickness, by adding an
additional titania quarterwave narrows the transmittance zone and increases the reflectance. The
coating now has seven layers, three dielectric layers on either side of the metal layer.
With a reflecting system consisting of four quarterwaves of titania and magnesium fluoride on
either side, together with the phase matching layers, the coating is now a narrowband filter.

Glass HLHLH ′Ag H ′LHLH Glass (8.2)

Figure 8-10. The performance of the seven-layer coating.

Figure 8-11. The performance of the metal-dielectric narrowband filter.


Macleod: Optical Coatings Page 105 Design through Manufacture

Figure 8-12. Comparison between the reflectance of aluminum boosted by a 4-layer silica titania centered on
510nm and uncoated aluminum. Note the limited region over which enhanced performance is obtained.
It is possible to use a dielectric system to increase the reflectance of a metal rather than reduce
it. We can look on this as a special case of the above coatings where the phase matching layer is
adjusted to give constructive rather than destructive interference. When the phase matching layer
is of low index and the outer reflector is a quarterwave stack the correct adjustment of the phase
matching layer is close to a quarterwave. In practice it is usually made exactly a quarterwave.
Since this coating is now made up entirely of quarterwaves we can use the quarterwave rule to
calculate its reflectance. The substrate can be the metal layer. The quarterwave rule works just as
well for absorbing substrates as it does for dielectric ones, although the layers must be dielectric.
Thus the addition of four quarterwave layers of alternate high and low admittance

Air| HLHL| metal (8.3)

leads to a reflectance of

[ y0 − ( y H4 / y L4 )n]2 + ( y H4 / y L4 )k 2
R= (8.4)
[ y0 + ( y H4 / y L4 )n]2 + ( y H4 / y L4 )k 2

Since n is small compared with k, the effect of the great increase in the k-term is much greater
than that of the term containing n and the reflectance increases. If the order of the layers were
inverted then the effect would be a reduction of reflectance. This dielectric enhancement of the
reflectance of a metal layer is sometimes called boosting.

8.1 Induced Transmission Filters


Metal-dielectric filters suffer from absorption losses that, for a given halfwidth, limit the
achievable transmittance. So far we have not considered the question of just how much they limit
the transmittance, except to estimate, quite crudely, the peak transmittance of a metal-dielectric
Fabry-Perot filter. The question can be posed in an alternative way: Given a metal layer of
known thickness and optical constants, what is the maximum possible transmittance at
wavelength λ0, and how can it be attained? This leads to the concept of potential transmittance
that was devised in 1957 by Berning and Turner[1].
Potential transmittance is the ratio of the irradiance transmitted by the layer, or assembly of
layers, to the irradiance that actually enters. If the transmittance is T and the reflectance R, then
Macleod: Optical Coatings Page 106 Design through Manufacture

the fraction of the incident irradiance that actually enters the assembly is (1-R) and the potential
transmittance is:
T
ψ= (8.5)
(1 − R )

To make the actual transmittance equal the potential transmittance it is sufficient to make R
zero. If we are dealing with dielectric materials then T=(1-R) and the potential transmittance is
always unity. It can be shown that at any given wavelength, the potential transmittance of a layer
depends only on the optical constants and thickness of the layer and on the optical admittance of
the assembly at the rear interface, the exit admittance. It does not depend on anything on the
incident side of the layer. Further, the potential transmittance of an assembly of layers is the
product of the potential transmittances of the individual layers, each calculated for the
appropriate exit admittance. Since, once the parameters of a given layer have been defined, its
potential transmittance is a function only of the exit admittance, then we must be able to find a
value of the exit admittance which maximizes the potential transmittance. A metal-dielectric
filter that has been designed using this concept, is known as an Induced-Transmission Filter
because it will have the maximum possible transmittance of which the metal layers are capable.
The procedure for the design of such a filter is:
(1) Choose the metal thickness
(2) Find the optimum admittance and maximum potential transmittance, ψmax.
(3) Design a dielectric structure for the rear interface to give the optimum admittance.
(4) Add the metal layer.
(5) Design a dielectric assembly to reduce the reflectance of the multilayer to zero
and ensure that T = ψmax.
This, of course, gives a dielectric-metal-dielectric arrangement. If more metal layers are required
then two extra steps should be inserted between (4) and (5):
(4a) Design a dielectric assembly to change the admittance of the multilayer to the
optimum exit admittance for the second metal layer.
(4b) Add the second metal layer.
We now derive expressions for ψ, ψmax, and (X0+iZ0), the optimum exit admittance. These
expressions cannot be expected to be simple. We consider just one absorbing layer. The potential
transmittance is given by:

Re(Ye )
ψ= (8.6)
Re( Bi Ci *)

⎡ Bi ⎤ ⎡1⎤
where ⎢C ⎥ = [ M ] ⎢Y ⎥
⎣ i⎦ ⎣ e⎦
Macleod: Optical Coatings Page 107 Design through Manufacture

and [M] is the characteristic matrix of the metal layer.


Let Ye, the exit admittance, be X+iZ, and let the metal parameters be the usual n, k, d. Then

⎡ i sin δ ⎤
⎡ B ⎤ ⎢ cos δ ⎡ 1 ⎤
⎢ ⎥= y ⎥⎢
⎣C ⎦ ⎢iy sin δ ⎥ X + iZ ⎦⎥
⎣⎢ cos δ ⎦⎥ ⎣

2π (n − ik )d 2π nd 2π kd
where δ= = −i = α − iβ
λ λ λ

Let free space units be used, then y=n-ik, and

⎡⎧ i sin δ ⎫ ∗⎤
Re( Bi Ci∗ ) = Re ⎢ ⎨cos δ + ( )( X + iZ ) ⎬ {iy sin δ + ( X + iZ ) cos δ } ⎥ (8.7)
⎣⎩ λ ⎭ ⎦

There is nothing for it but to take each term in turn and find the real part - not difficult but very
tedious. We therefore skip to the answer:
Re(Ye ) X
ψ= =
Re( Bi Ci ) Re( Bi Ci∗ )

An expression for 1/ψ is slightly more easily obtained than ψ:

1 (n 2 − k 2 ) − 2nkZ / X
= (sin 2 α cosh 2 β + cos 2 α sinh 2 β )
ψ (n 2 + k 2 )
+(cos2 α cosh 2 β + sin 2 α sinh 2 β )
(8.8)
+(1/ X )(n sinh β cosh β + k cos α sin α )
X 2 + Z2
+ (n sinh β cosh β − k cos α sin α )
X (n 2 + k 2 )

This is a complicated expression, but it can be readily evaluated with a scientific calculator. So
far we have the potential transmittance, but we need also the optimum exit admittance. ψ is
always positive and is a well-behaved function, so for an extremum in ψ we have also an
extremum in 1/ψ, of the opposite sense. We can safely write that the optimum exit admittance is
given by:

⎛ ∂ ⎞⎛ 1 ⎞ ⎛ ∂ ⎞⎛ 1 ⎞
⎜ ⎟ ⎜ ⎟ = 0 and ⎜ ⎟⎜ ⎟ = 0
⎝ ∂X ⎠ ⎝ψ ⎠ ⎝ ∂Z ⎠ ⎝ ψ ⎠

Let us write

1 ⎡ p (n 2 − k 2 − 2nkZ / X ) r s( X 2 + Z 2 ) ⎤
=⎢ + q + + ⎥ (8.9)
ψ ⎣ (n 2 + k 2 ) X X (n 2 + k 2 ) ⎦
Macleod: Optical Coatings Page 108 Design through Manufacture

where p, q, r, s correspond to the appropriate expressions in (8.8). Then

p.2nkZ r s sZ 2
− 2+ 2 − =0 (8.10)
X 2 ( n2 + k 2 ) X ( n + k 2 ) X 2 ( n2 + k 2 )
p(−2nk ) 2 sZ
and + =0 (8.11)
X ( n + k ) X ( n2 + k 2 )
2 2

so that

nkp
Z= (8.12)
s

r (n 2 + k 2 ) n 2 k 2 p 2
and X2 = − (8.13)
s s2

We take, as usual, the positive root for X. Then, inserting the appropriate expressions,
1
⎡ 2 n 2 k 2 ( sin 2 α cosh 2 β + cos 2 α sinh 2 β ) ⎤
2 2
( n + k 2
)( n sinh β cosh β + k sin α cos α )
X opt =⎢ − ⎥
⎢ 2(n sinh β cosh β − k sin α cos α ) 2 ( n sinh β cosh β − k sin α cos α ) ⎥
2

⎣ ⎦

(8.14)

nk (sin 2 α cosh 2 β + cos 2 α sinh 2 β )


and Z opt = (8.15)
(n sinh β cosh β − k sin α cos α )

Note that for β large, Xopt→n, and Zopt→k, that is Yopt→(n-ik)*. ψmax can then be found by
substituting (8.14) and (8.15) in (8.8). There is little gain and too much labor in finding a
separate analytical expression for ψmax.
To devise a matching stack we must find an assembly of dielectric layers which will give an
input admittance of (Xopt+iZopt) when deposited on the substrate, ysub. It is easier to turn this
problem round. If we consider the admittance diagram, we see that this is the identical assembly
which, when deposited in reverse order on a surface of admittance Yexit*=(Xopt-iZopt) would give
an input admittance of ysub. Then the problem becomes that of adding a dielectric layer to
(Xopt-iZopt) to make the admittance real, followed by a series of quarterwave layers to transform
that real admittance into ysub. By considering a single dielectric layer of index nf on (Xopt-iZopt),
we find that the thickness of the matching layer is given by
Macleod: Optical Coatings Page 109 Design through Manufacture

nf d f 1 ⎡ 2Z opt n f ⎤
= arctan ⎢ 2 ⎥ (8.16)
λ 4π 2 2
⎣⎢ (n f − X opt − Z opt ) ⎦⎥

where the arc tangent is to be taken in the first or second quadrant, and the real admittance which
the addition of this layer gives, is

2 X opt n 2f
µ= 1
(8.17)
(X 2
opt +Z 2
opt + n ) + {( X
2
f
2
opt +Z 2
opt
2 2 2
+ n ) − 4X n
f opt
2
f } 2

nf can be of high or low value but µ will always be of lower admittance than the substrate
(except in very unlikely cases) because it is the first intersection of the locus of nf with the real
axis which gives (8.16). Then, since the substrate will always have an admittance greater than
unity, the stack of matching quarterwave layers should start with a low index. Alternate high and
low indices follow, the precise number being found by trial and error and the combination, which
gives the closest match being selected. Whether or not the match is satisfactory can be estimated
by calculating the reflectance which any mismatch represents. This will be the loss in
transmittance that the final filter design will suffer.
The matching of the front surface of the metal layer is the final part of the design process.
This is made easy by the fact that if the exit admittance of the metal is (Xopt+iZopt) then the
admittance at the front surface will be (Xopt-iZopt). We can see that this must be so, by a logical
argument. Assume that the filter is correctly designed so that the transmittance is ψmax.
Transmittance in the opposite direction must be equal and therefore also ψmax. In this case what
was the front surface of the metal is now the rear, and, since we have ψmax, the exit admittance
must be Yexit=(Xopt+iZopt). The rest follows.
Now we have already constructed a matching assembly to match (Xopt-iZopt) to ysub. If the
filter is to have a cover slip cemented over it, as will usually be the case, then we can assume that
the cement will have the same index as the substrate and the matching assembly on the front is
exactly the same as that at the rear. If there is no cover slip, then we add the first dielectric layer
to make the admittance real and then add quarterwave layers as before until we reach a good
enough match with the incident medium. Again, the loss suffered by the final filter is just the
reflectance mismatch.
An example may help to make it clear.
A filter with peak wavelength 550nm is to use dielectric materials of indices 2.35 and 1.35,
and a silver layer with optical constants 0.055-i3.32 and thickness 70nm. The substrate has index
1.52 and the filter is to be protected by a cemented cover so that the index of the incident
medium is also 1.52.

2π nd
α= = 0.04398 (8.18)
λ
Macleod: Optical Coatings Page 110 Design through Manufacture
2π kd
β= = 2.6549 (8.19)
λ

and from (8.14) and (8.15) we find the optimum admittance:

( X opt + iZ opt ) = 0.4572 + i3.4693

Substituting in (8.8) gives

ψ max = 80.5%

We can choose a low or high index initial matching layer. Let us arbitrarily choose low index.
Then (8.16) gives a thickness for the layer of 0.19174 full waves. (8.17) gives a value for µ of
0.05934. All the steps up to this point can be carried out straightforwardly by a programmable
calculator or computer. The next step is to match the value of µ to the substrate admittance of
1.52. The best arrangement is found to be HLHLHL/µ (note that a low index layer is next to µ )
which has an admittance of

nH6 µ
= 1.6511
nL6

The reflectance loss due to the mismatch with 1.52 is 0.2%. The structure so far is:

/ Ag / L′LHLHLH / Substrate

(note that the a low index quarterwave is next to L' or µ ). Since the incident medium is of
identical index to the substrate, the complete design is:

Glass / HLHLHLL′ / Ag / L′LHLHLH / Glass

or, combining the adjacent low index layers,

Glass / HLHLHL′′ / Ag / L′′HLHLH / Glass

where Ag = 70nm geometrical thickness of material of


optical constants 0.055-i3.32
L = quarterwave of 1.35
H = quarterwave of 2.35
L" = (0.19174+0.25) = 0.44174 full waves
λ0 = 550nm
If the performance of this design is computed, it will be found to be a band pass filter centered
indeed on 550nm. The design appears very similar to a double-cavity filter with a metallic
central reflector, Figure 8-13.
Macleod: Optical Coatings Page 111 Design through Manufacture

We have said nothing about halfwidth of the filter. This is difficult to predict. A rough idea
can be obtained from the halfwidth of the multiple-cavity all-dielectric filter that uses the same
dielectric stacks as the induced transmission filter. But the best way of estimating the halfwidth
is to compute the characteristic.
The rejection of a single metal layer filter is not usually very high outside the high reflectance
zone of the quarterwave stacks. There is also a small residual leak, which can be as high as 1%
transmittance, at the long wave side of the pass band where the quarterwave stacks are failing
and the metal rejection is taking over. To improve the rejection to a very satisfactory level, at
least two metal layers should be used. The design of a two metal layer filter follows simply from
the design we have already established for the single metal layer filter. The admittance must be
returned to the optimum exit value after the first metal layer and this can be accomplished by
inserting twice the thickness of the matching layer. One thickness will return the admittance to
the real axis and the second thickness, therefore, to the correct exit admittance for the second
metal layer. The second metal layer then follows along with the remainder of the structure. The
design is:

Glass / HLHLHLL′ / Ag / L′L′ / Ag / L′LHLHLH / Glass

and the performance is shown in Figure 8-14.

Transmittance (%)
90
80
70
60
50
40
30
20
10
0
500 510 520 530 540 550 560 570 580 590 600
Wavelength (nm)

Figure 8-13. The characteristic curve of the induced transmission filter designed in the text.
Macleod: Optical Coatings Page 112 Design through Manufacture

Transmittance (%)
70

60

50

40

30

20

10

0
500 510 520 530 540 550 560 570 580 590 600
Wavelength (nm)

Figure 8-14. The performance of the two metal layer induced transmission filter:
Glass/HLHLHL L' /Ag/ L'L' /Ag/ L' LHLHLH/Glass.
Note the appearance of a severe shoulder on the pass band.
The rejection of this filter is excellent but there is a shoulder on the pass band and this is a
common feature of such a design. The shoulder can be removed by adding a decoupling
halfwave layer to the already existing L'L' combination between the two metal layers. This yields
the design:

Glass / HLHLHLL′ / Ag / L′LLL′ / Ag / L′LHLHLH / Glass

and the performance of this design is shown in Figure 8-15. Now the pass band shape is much
improved but close examination of the rejection at longer wavelengths will show a very small
transmittance increase, still far below the peak in the single metal filter, but nevertheless above
the transmittance in the remainder of the rejection region. Which design is better will depend on
the application. For the suppression of long wave sidebands of narrowband all-dielectric filters,
the misshapen peak of Figure 8-14 is to be preferred because of its excellent rejection. For
applications where the induced transmission filter pass band will define the transmitted region,
the version in Figure 8-15 is better. More metal layers can, of course, be added to improve
matters still further. Each time a metal layer is added, the peak transmittance will fall by a
further factor ψ. With several metal layers it will probably be necessary to return to the starting
calculation to reduce the metal layer thickness and increase the peak transmittance. It can readily
be shown that several thinner metal layers are better than a single thick metal layer in achieving
a certain level of rejection with smaller drop in peak transmittance.
Macleod: Optical Coatings Page 113 Design through Manufacture

Transmittance (%)
70

60

50

40

30

20

10

0
500 510 520 530 540 550 560 570 580 590 600
Wavelength (nm)

Figure 8-15. The performance of the two metal layer induced transmission filter:
Glass/HLHLHL L' /Ag/ L' LL L' /Ag/ L' LHLHLH/Glass.
The shoulder of Figure 8-14 has disappeared.

8.2 Chapter 8 References


[1] P. H. Berning and A. F. Turner, “Induced transmission in absorbing films applied to band
pass filter design,” Journal of the Optical Society of America, vol. 47, pp. 230-239, 1957.
Macleod: Optical Coatings Page 115 Design through Manufacture

9 OBLIQUE INCIDENCE
Coatings at angles of incidence other than normal present special problems. First there is the
shift in properties with angle of incidence that follows roughly a cosine law. At higher angles of
incidence the cosine curve becomes very steep implying a much more rapid variation in
properties with angle of incidence than there is near normal incidence. Then there are the two
principal planes of polarization that must be considered. Performance is usually rather different
for these two planes. In many cases the design aim will be to minimize the difference in
performance for these two planes but in others the aim will be to enhance the difference. In those
fewer cases where the specification is applied to only one plane of polarization the design is a
simple extension of the normal incidence techniques. We begin by discussing the admittance
diagram and its use at oblique incidence
The admittance diagram is a useful visualization tool. It becomes more complicated to use at
oblique incidence and the complications can be reduced somewhat by adopting the scheme of
modified admittances described briefly earlier but repeated here for convenience.
The modified admittances are the usual oblique incidence admittances that have been
normalized by the dividing of the s-polarized admittances, and the multiplying of the p-polarized
admittances, by cosθ0. This has the effect of preserving, for both s- and p-polarization, the
admittance of the incident medium at its normal incidence value, regardless of the angle of
incidence, and means that the isoreflectance and isophase contours of the admittance diagram
retain their normal incidence values whatever the angle of incidence or plane of polarization.
The expressions are:

no sin θ 0 = n sin θ (Snell's Law) (9.1)

cosθ
ηs = y ⋅ (9.2)
cosθ 0

cos θ 0 y 2
and ηp = y ⋅ = (9.3)
cos θ ηs

The phase thickness δ remains as


2πd
δ =( )(n − ik ) cosθ
λ (9.4)

For metal layers, the solution of Snell's Law becomes complex and although (9.1)-(9.3) give
the correct answer, it avoids possible ambiguities to put the expressions in the form:

δ = (2πd / λ )(n 2 − k 2 − n02 sin 2 θ 0 − 2ink )1/ 2 (9.5)


Macleod: Optical Coatings Page 116 Design through Manufacture

the correct root being in the fourth quadrant,

η s = (n 2 − k 2 − n02 sin 2 θ 0 − 2ink )1/ 2 (9.6)


again in the fourth quadrant, and

η p = (n − ik ) 2 / η s
(9.7)

The values of reflectance, transmittance, absorptance and phase changes on either


transmission or reflection are completely unchanged by the adoption of these values for the
admittances.

9.1 Metal Layers at Oblique Incidence


It is useful to illustrate some of the properties of the oblique incidence admittance diagram by
considering the properties of metals at oblique incidence. This will be helpful in the design of
coatings involving metal layers.

9.1.1 Below Critical Angle


Let us first take air of index unity as the incident medium. We recall that all transparent thin-
film materials have refractive index greater than unity, thus the p-admittances of all transparent
materials fall while the s–admittances increase.
In metal layers that are characterized by low n and high k the variation of modified admittance
with angle of incidence is almost entirely due to the cosϑ0 term. Thus an approximation good
enough for our purpose is to write

η s = (n − ik ) / cosθ 0 (9.8)

η p = (n − ik ) cosθ 0
and (9.9)

while δ remains virtually unchanged from its normal incidence value.


As a first example let us consider a single thick metal layer. The trajectory of the admittance
of a metal with increasing angle of incidence in air is shown in Figure -9-1. It is simply a line
joining the start point to the origin. For s-polarization the admittance moves from start point A
towards B so that Rs rises with ϑ0. At the same time, ϕs the phase shift on reflection, starts in the
second quadrant, a little short of 180° and rises towards 180°. p-behavior is more involved. First
of all we can show that the line is tangent to an isoreflectance circle at the point C. This therefore
must mark a minimum of reflectance. Furthermore this point is on the boundary between the first
and second quadrants of ϕp which is therefore 90°. Thus, for increasing ϑ0 Rp first falls while ϕp
decreases. Rp reaches a minimum at C where ϕp is equal to 90°. We can estimate that
cos ϑ0 ≈ ( y0 / k ) at this point. Beyond the point C, Rp rises while ϕp continues to fall. Point C is
sometimes called the pseudo-Brewster angle. Note that ϕ p − ϕ s < 90° at this point. At a slightly
Macleod: Optical Coatings Page 117 Design through Manufacture

greater angle, as ϕp continues to fall and ϕs to rise, the difference does become 90° at an angle
known as the principal angle. The behavior of silver with optical constants (0.075-i3.41) in air as
the incident medium is illustrated in Figure -9-1. The pseudo-Brewster angle, C, is given
approximately by arc cos(1/3.41), i.e. 72.9°.

Design1: Reflectance
100

99

Reflectance (%)
98

97

96
0 15 30 45 60 75 90
Incident Angle

Figure -9-1. The left hand side is a sketch of the behavior of a metal surface as the angle of incidence is
varied from zero to 90°. The oblique line through the origin and the point (n-ik) marks the locus of the metal
admittance. The right hand side shows the calculated behavior of silver at around 550nm with optical
Macleod: Optical Coatings Page 118 Design through Manufacture

constants (0.055-i3.32). The left hand diagram shows s-reflectance as the upper curve and p-reflectance as
the lower. See the text for a fuller description.

9.1.2 Beyond Critical Angle


We now consider an incident medium of higher index where angles of incidence may exceed
the critical angle. Metal layers are still scarcely affected except for the cosθ0 dependence of the
modified admittance and dielectric layers of index higher than that of the incident medium have
behavior similar to that with low-index incident medium. The major differences occur with
dielectric materials of index lower than that of the incident medium. Equations (9.4), (9.6) and
(9.7) are the relevant ones with k=0 and n0sinϑ0>n. The phase thickness, δ, 2πnd/λ at normal
incidence, becomes at oblique incidence, from (9.5),

2π (n 2 − n02 sin 2 θ 0 )1/ 2 d / λ (9.10)

that is −i 2π (n02 sin 2 θ 0 − n 2 )1/ 2 d / λ (9.11)

where, again, the fourth rather than second-quadrant solution has been chosen. The modified
admittances are then

−i (n02 sin 2 θ 0 − n 2 )1/ 2


ηs =
cosθ 0 (9.12)

i.e. in the fourth quadrant)

η p = n2 / η s
and (9.13)

Since ηs is negative imaginary, ηp must be positive imaginary. The behavior of the modified
admittance is shown diagrammatically in Figure 9-2.

Figure 9-2: The variation of the s- and p-admittances (modified) as the angle of incidence varies in a medium
of higher admittance. The s-admittance falls to zero at the critical angle then becomes negative imaginary and
Macleod: Optical Coatings Page 119 Design through Manufacture

falls down the imaginary axis towards negative infinity as the angle increases. The p-admittance moves
towards infinity on the real axis then, with angle of incidence above critical, falls down the imaginary axis
from positive infinity to zero.
For a thin film of material used beyond the critical angle, then, the s-polarized behavior is
indistinguishable from that of an ideal metal. We have a set of circles centered on the real axis,
described clockwise and ending on the point ηs which is on the negative imaginary axis. For p-
polarized light, the behavior is, in one important respect, different. Here the combination of
negative imaginary phase thickness and positive imaginary admittance inverts the way in which
the circles are described, so that although they are still centered on the origin, they are
counterclockwise and terminate at ηp on the positive imaginary axis. This behavior plays a
significant part in what follows.
In subsequent sections when dealing with phenomena beyond the critical angle we will
usually assume a beam of light incident on the hypotenuse of a glass prism beyond the critical
angle and take the index of the incident medium as 1.52. This implies that the circle separating
the first and second quadrants and the third and fourth quadrants, which has center the origin,
will have radius 1.52. Frequently, the exit medium, taking the place of the substrate, will be air
of refractive index unity. Mostly we will consider p-polarized light because that has the more
interesting properties. The modified admittance for p-polarization will be positive imaginary
and, as θ0 increases, will falls down the imaginary axis towards the origin. The reflectance will
be unity and reference to Figure 9-2 shows that the phase shift will vary from 180°, just beyond
the critical angle, through the third and fourth quadrants towards 0° at grazing incidence. The s-
polarized reflectance will be likewise unity, but the admittance negative imaginary, falling from
zero to infinity along the imaginary axis so that the s-polarized phase shift will increase with θ0
from zero, through the first and second quadrants towards 180°.

9.2 Coatings for One Plane of Polarization


The design of coatings for one single plane of polarization is relatively straightforward. The
admittances of all available materials are changed to those for the particular angle of incidence
and appropriate polarization. The design then proceeds as for normal incidence. For p-
polarization the range of available admittances shrinks and so it becomes more difficult to have
strong reflectance, narrow pass bands and so on while for s-polarization the difference increases
and so it becomes more difficult to have antireflection coatings with high performance over an
extended region.

9.3 Polarizers
There is a number of different techniques which have been or are used for the design of
Polaris's, but two are much more common than any of the others, and will be described here.
Polaris's designed using either of these techniques can also be used as polarizing beam splitters.

9.3.1 Brewster-Angle Polarizers


In p-polarized light, the reflectance of a single surface falls to zero when the angle of
incidence is equal to the Brewster angle, while the reflectance in s-polarized light simply
Macleod: Optical Coatings Page 120 Design through Manufacture

increases with angle of incidence. A quarterwave stack, therefore, tilted so that the propagation
of light within the layers is at the Brewster angle, will reflect strongly the light which is s-
polarized, but will transmit the p-polarized component. For zinc sulfide (n=2.35) and cryolite
(n=1.35), for example, the Brewster angle is given by

θ L = arctan(nH / n L ) = 6012
. ° (9.14)

measured in the low-index material. Unfortunately, when we attempt to refer this angle of
incidence to air as the incident medium, we find that it is beyond the critical angle. Indeed, given
the usual range of film indices, the angle is always beyond critical for any but combinations of
such similar low indices as to be quite impractical. When, however, the angle of incidence is
referred to glass, (n=1.52), as the incident medium, even with zinc sulfide and cryolite as the
layer materials, a useful angle of incidence, 50.5°, is obtained. A simple arrangement of
cemented prisms, with a quarterwave stack, tuned to the appropriate angle of incidence,
deposited on the hypotenuse of one of them, acts as a polarizing beam splitter with useful
performance over a wide wavelength range[1]. A more suitable angle of incidence would be 45°
and this can be obtained either by using glass of higher refractive index or by reducing the index
of the higher-index layer in the coating.

Figure 9-3. Construction of a polarizing beam splitter.


Macleod: Optical Coatings Page 121 Design through Manufacture

Figure 9-4. A 29-layer plate polarizes. The design is


Air | (0.505H 1.146L 0.505H)3 (0.538H 1.22L 0.538H)8 (0.505H 1.146L 0.505H)3 | glass, with
yH=2.3, yL=1.45, yglass=1.52 and λ0=900nm. The angle of incidence in air is ϑo=56.5° chosen
to be the Brewster angle for the rear surface of the substrate.

9.3.2 Plate Polarizer


The width of the high-reflectance zone of a quarterwave stack is a function of the ratio of the
admittances of the two coating materials. When tilted, the ratio of the indices for p-polarized
light is reduced from its value at normal incidence while the ratio for s-polarized light increases.
Thus the width of the high-reflectance zone is less in p-polarized light than in s-polarized. Over a
limited region, therefore, just outside the p-polarized zone, the quarterwave stack will reflect s-
polarized light while transmitting p-polarized. The principal problem to be overcome is the
ripple in the transmission zone which will cause a proportion of the p-polarized light to be
reflected, reducing the polarization efficiency. This ripple can be reduced by using exactly those
techniques for ripple reduction in edge filters[2]. Usually the long-wave edge of the high-
reflectance zone will be used since this gives a slightly wider region over which polarization
splitting exists and since long-wave-pass edge filters are easier to manufacture than short-wave-
pass. Figure 9-4 shows the calculated performance of a typical design.

9.4 Non-Polarizing Coatings


This is a much more difficult task than the design of polarizing components and there is no
completely effective method. Here we collect a few different techniques together.

9.4.1 Reflecting or Beam-Splitting Coatings


The major problem is the catastrophic drop in the reflectance for p-polarization that occurs at
high angles of incidence.
The reflectance of a quarterwave stack at oblique incidence is given by
Macleod: Optical Coatings Page 122 Design through Manufacture

R=
LMη 0 −Η OP 2

Nη 0 +Η Q
η12η 23η 52 …η sub
Η=
η 22η 24η 26 … (9.15)

where the various η's are the appropriate s- or p-polarized versions and where ηsub will be in the
numerator as shown if the number of layers is even, or in the denominator if the number is odd.
If we are using the modified admittances then the value of η0 for each plane of polarization is the
same, and for the reflectances for p- and s-polarizations to be equal it is sufficient for

∆21∆23 ∆25… ∆ sub


∆Η = =1
∆22 ∆24 ∆26… (9.16)

where ∆ 1 = η 1 p / η 1s and so on. The objective in the design is then to find a combination of
materials such that condition (9.16) is satisfied at the same time as the value of Η in (9.15) gives
the correct reflectance. This will take usually three materials although Mahlein[3] has shown that
a system of even number of quarterwaves with a low index layer outermost will give equal s- and
p-reflectances at one particular angle of incidence and this simplification can sometimes be
adopted if the materials and angle of incidence correspond to what is possible and acceptable.
We can take as an example a reflector for 45° angle of incidence. In order to simplify slightly the
design we assume that any material with characteristic admittance from 1.35 to 2.45 at intervals
of 0.05 is available, that the incident medium is air of admittance unity and that the substrate is
glass of admittance 1.5. The modified admittances together with their ratios are listed in Table 9-
1.
The value of ∆ associated with the substrate is 0.642857 from the table. We can take a value
of 2.4 as the high admittance value with a ∆ of 0.547528 and 1.60 as the low admittance with ∆
of 0.621359. Then the combination LHLHL | Glass has a value of ∆Η of
0.6213596/(0.5475284x0.642857), i.e. 0.99613. This is sufficiently near unity for our purposes.
The layers must be quarterwaves at 45° and so the design finally becomes

Air |(11148
. L10465
. H ) 2 11148
. L| Glass

with L a quarterwave of admittance 1.60 and H a quarterwave of admittance 2.4. The


performance of this design is shown in Figure 9-5 where the reflectance can be seen to be quite
low. This is a basic problem with this type of two-material design. Higher reflectance demands
more pairs of high and low admittance but the ∆ for the substrate is the main compensation
factor in the expression for ∆Η and so only modest reflectances can be achieved.

Table 9-1. Modified admittances for 45° incidence in air


y 1/cosθ ηp ηs ∆=ηp/ηs
Macleod: Optical Coatings Page 123 Design through Manufacture

y 1/cosθ ηp ηs ∆=ηp/ηs
1.35 1.17391 1.12061 1.62635 0.689036
1.40 1.15865 1.14700 1.70880 0.671233
1.45 1.14543 1.17442 1.79025 0.656006
1.50 1.13389 1.20268 1.87083 0.642857
1.55 1.12375 1.23165 1.95064 0.631406
1.60 1.11477 1.26122 2.02978 0.621359
1.65 1.10678 1.29131 2.10832 0.612486
1.70 1.09964 1.32186 2.18632 0.604602
1.75 1.09322 1.35279 2.26385 0.597561
1.80 1.08742 1.38406 2.34094 0.591241
1.85 1.08217 1.41563 2.41764 0.585543
1.90 1.07739 1.44748 2.49399 0.580386
1.95 1.07303 1.47956 2.57002 0.575700
2.00 1.06904 1.51186 2.64575 0.571429
2.05 1.06538 1.54435 2.72121 0.567522
2.10 1.06202 1.57701 2.79643 0.563939
2.15 1.05891 1.60984 2.87141 0.560643
2.20 1.05603 1.64280 2.94618 0.557604
2.25 1.05337 1.67590 3.02076 0.554794
2.30 1.05090 1.70912 3.09516 0.552192
2.35 1.04860 1.74245 3.16938 0.549776
2.40 1.04645 1.77588 3.24345 0.547528
2.45 1.04445 1.80941 3.31738 0.545434
Macleod: Optical Coatings Page 124 Design through Manufacture

Figure 9-5. The nonpolarizing beam splitter of design


Air | (1.1148L 1.0465H)2 1.1148L | Glass. λ0=510nm and ϑ0=45°.
For higher reflectances at this angle of incidence it is possible to have greater flexibility if
three materials rather than two are introduced[4]. These materials, H, L and M with the
admittance of M intermediate between L and H, can be arranged in a scheme (HMLM)q such that
∆H∆L/∆M2 is unity and q can be chosen as desired. In fact a slight adjustment must be made since
this neglects ∆sub. Choosing 1.45, 2.4 and 1.65 as the admittances of L, H and M respectively, we
then have ∆H∆L/∆M2 of 0.9575. This implies that the overall ∆Η of Air|(MHML)5|Glass, where
the quarterwaves are at 45°, is 0.993. The reflectance is (calculating for s-polarization) 59%. The
calculated performance is shown in Figure 9-6. Clearly an overcorrection could be made to lift
Rp a little higher than Rs and then the two curves would cross on either side of the central
wavelength. This might be a preferable characteristic.
Macleod: Optical Coatings Page 125 Design through Manufacture

Figure 9-6. the performance of the design Air | (1.10678M 1.04645H 1.10678M 1.14543L)5 | Glass where
the admittances of M, H and L are, respectively, 1.65, 2.4 and 1.45, of glass 1.50 and of air 1.00. λ0=510nm
and ϑ0=45°.

9.4.2 Edge Filters


The same approach can be used to design edge filters[5]. We can take the ideas behind the
previous design and turn it into a symmetrical assembly by writing (.5HMLM.5H) or
(.5LMHM.5L). Equivalent admittances can be calculated in the same way as for normal
incidence simply by using the admittances in the table for either s- or p-polarization.
s-polarization will usually be the poorer in performance and so it will normally be optimized.
Figure 9-7 shows the performance of a filter designed in this way for a very high angle of
incidence of 45° in glass. The design is

Glass|1877
. L(0.826 L13372
. M 1138
. H1372
. M 0.826 L) 20 1877
. L| Glass

where L, M and H represent materials of admittances 1.35, 1.57 and 2.25 respectively. Glass has
admittance 1.52 and the value of λ0 is 510nm.
Macleod: Optical Coatings Page 126 Design through Manufacture

Figure 9-7. The calculated performance of a polarization-free edge filter at an angle of incidence of 45° in
glass. The design of the filter is given in the text. s-polarization is indicated by the dark line and
p-polarization by the grey.
A further technique that can be used for polarization insensitive edge filters involves
narrowband filters. The peak position of a narrow band filter is fixed by the thickness of the
spacer layer and also the phase shift on reflection from the surrounding reflectors. The phase
shift varies with g and since the admittances differ for p- and s-polarization the dispersion of
phase shift with g will differ for the two planes of polarization. This implies that since the phase
thickness of the spacer layer will be the same for both polarizations, the phase shift differences
will displace the pass bands different amounts. In particular it should be possible to bring into
coincidence, at any given angle of incidence, the longwave or shortwave edges of the pass band
of a bandpass filter. If the bandpass filter is fairly broad, then it can serve as an edge filter.
To design such a bandpass/edge filter we begin with the method of symmetrical periods. We take
a basic structure:

fBABAB... BAfB

where A and B are quarterwave layers and f is a correction factor. We take the number of layers
in the central stack of quarterwaves as 2x+1 so that the total number of layers in the period is
2x+3. Let us represent the central quarterwave stack by matrix M then the overall matrix is given
by

LM cosα OPL M
i sin α j
iM12 OPLM cosα OP
i sin α j

MNiη sin α η PM
11

B cos α Q N
B
iM 21 M 22 QMNiη sin α
B
η P
cos α Q
B

(9.17)

If we denote the product matrix by N, where


Macleod: Optical Coatings Page 127 Design through Manufacture

N=
LM N11 iN12 OP
NiN 21 N 22 Q (9.18)

then the edges of the pass band are given by

N11 = N 22 = ±1

and

N11 = N 22 = M11 cos 2α − 0.5( M12η B + M12 / η B ) sin 2α ± 1 (9.19)

where

α = (π / 2)( λ R / λ ) = (π / 2)( λ R / λ 0 ) g = (π / 2) fg

If we write

ε = (π / 2)(1 − g )

we can show that

b g x
M11 = M 22 = ( −1) x ( −ε ) η A / η B +…+ η B / η A b g x

i ( −1) x
iM12 =
(η A / η B ) x η A
iM 21 = i ( −1) x (η A / η B ) x η A
(9.20)

Then, to a reasonable approximation (where we neglect terms such as (η L / η H ) )


x +1

( −1) x ( −ε )(η H / η L ) x
M11 =
(1 − η L / η H )
= ( −1) x ( −ε ) P (9.21)

and
0.5(M 12 η B + M 12 / η B ) = 0.5( − 1)x ((ηA / η B )x + 1 + (η B / ηA )x + 1
= 0.5( − 1)x (ηH / ηL )
x +1

= ( − 1)x Q (9.22)

where
Macleod: Optical Coatings Page 128 Design through Manufacture

(η H / η L ) x
P=
(1 − η L / η H )

Q = 0.5(η H / η L ) x +1 (9.23)

Then for s- and p-polarization we have the equations for the edges
±1 = εPp cos 2α + Qp sin 2α
±1 = εPs cos 2α + Qs sin 2α (9.24)

which give for α and ε:

Ps − Pp
sin 2α = ±
( Ps Qp − Pp Qs )
±1 − Qp sin 2α
ε=
( Pp cos 2α )
(9.25)

Then

f = α / (πg / 2) = α / (π / 2 − ε ) (9.26)

There will be two solutions the larger usually corresponding to a shortwave pass filter and the
smaller to a longwave pass.
A numerical example will perhaps help to make things clearer. Let us consider the design of a
longwave pass edge filter at 45° incidence in air. The substrate can be assumed to be glass of
index 1.52. We base the design on a period:

fLHLHLHLHLfL

where x is 3 and the admittances of H and L are 2.35 and 1.35 respectively. Modified
admittances, P and Q, are given in Table 9-2.

Table 9-2
s-polarization p-polarization
ηH 3.1694 1.7425
ηL 1.6264 1.1206
η0 1.000 1.000
Macleod: Optical Coatings Page 129 Design through Manufacture

ηsub 1.9028 1.2142


P 15.201 10.535
Q 7.211 2.923

From these values sin α = ±01480


. .
Since the outer tuning layers are quarterwaves in their unperturbed state, we look for solutions
near 2α=π. Then

2α = π ± 01485
. = 2.9931 or 3.2901

Then, in both cases, cos 2α = −0.9890 and so

(1 + 2.923 × 0148
. )
ε=± = ± ( −01375
. )
( −10.535 × 0.9890)

whence
(2.9931 / 2)
f = = 1148
.
(π / 2) − 01375
.

with

g = 1 − 2 × 01375
. / π = 0.9125

and
(2.9931 / 2)
f = = 0.876
(π / 2) + 01375
.

with

g = 1 + 2 × 01375
. / π = 1088
.

We take the second of these which corresponds to a longwave pass filter. The basic period is
now

0.876 LHLHLHLH 0.876 L

We calculate the equivalent admittances and phase thicknesses for the arrangement as shown in
Table 9-3.
Macleod: Optical Coatings Page 130 Design through Manufacture

Table 9-3
s-polarization p-polarization
g E(modified) γ/π E(modified) γ/π
1.05 0.0949 4.2955 0.2018 4.4372
1.06 0.1190 4.4454 0.1993 4.5884
1.07 0.1202 4.5786 0.1861 4.6652
1.08 0.0982 4.7211 0.1588 4.7486
1.09 Imaginary values 0.1049 4.8530
1.10 Imaginary values Imaginary values

We arrange matching at g=1.08. Adding a HLHL combination to the period with the L layer next
to it yields admittances of 0.9625 for p-polarization and 1.416 for s-polarization. Since we have
to match 1.00 for air (modified admittance) for both s- and p-polarization, and, for substrate,
1.214 for p-polarization and 1.903 for s-polarization, this matching is probably satisfactory. The
matching is at g=1.08 and therefore the thicknesses of the matching assembly must be corrected
by the factor 1/1.08. The final design, with all layers tuned for 45° by multiplying by the
appropriate 1/cosθ, is then:
q
Air |(0.971H1087
. L) 2 (1028
. L(1049
. . L) 31049
H1174 . H1028
. L)
(1087
. L0.971H ) 2 | Glass

The performance with q=9 is shown in Figure 9-8. The pass region is, unfortunately, not broad,
particularly for s-polarization.
Macleod: Optical Coatings Page 131 Design through Manufacture

Figure 9-8. The performance of the edge filter with a reference wavelength of 655nm.
A more general technique for the design of polarization insensitive edge filters is also due to
Thelen. It concerns the use of a symmetrical period of a more general type than the above:

aAbBcC… mMnNnNmM … cCbBaA

This is made up of two half systems and the central layer is therefore split in the middle as
shown. Let the right hand half period be written as

aAbBcC… mMnN = XmMnN = W

where W represents the layers to the left of mM. Let Z be the matrix representing the total
symmetrical period, then

Z=
LM Z iZ OP = LM W iW OPLM W iW OP
11 12 11 12 22 12

NiZ Z Q NiW W QNiW W Q


21 22 21 22 21 11

=M
LW W − W W 2iW W OP
11 22 12 21 11 12

N 2iW W W W − W W Q
22 21 11 22 12 21 (9.27)

where W is given by

LM cosδ i
sin δ m OPLM cosδ i
sin δ n OP
W=X
MNiη sin δ PQMNiη sin δ PQ
m n
ηm ηn
m m cos δ m n n cos δ n

so that
Macleod: Optical Coatings Page 132 Design through Manufacture

RS
W11 = X 11 cos δ m cos δ n −
η n sin δ m sin δ n UV l
− X 12 η m sin δ m cos δ n + η m cosδ m sin δ n q
T ηm W
(9.28)

and similar expressions for the other elements. At the edge of the stopband either Z12 or Z21 is
zero and this implies, from (9.27) that any one of the matrix elements W, is zero. If, for example,
it is W11 that is zero, then, from (9.28) we have

X 11 / X 12 − η n tan δ n
tan δ m =
η m + X 11η n tan δ n / FH X 12η m IK

For the edges of the stop band for s- and p-polarizations to be coincident, we must therefore
have:

LM X 11 / X 12 − η n tan δ n OP = LM X 11 / X 12 − η n tan δ n OP
Nη F I
m + X 11η n tan δ n / H X 12η m K Q Nη
s
F I
m + X 11η n tan δ n / H X 12η m K Q p
(9.29)

This is a quadratic equation in tanδn. Three other similar equations can be obtained from the
other elements of Z. Each has two solutions and there are eight solutions in tanδn. Some
solutions are better than others, some are degenerate and some are for longwave pass while
others are for shortwave pass. Trial and error will decide which of the eight solutions is best. The
aAbBcC . . section can be chosen at will and this opens up further possibilities. The best
solutions will depend on the criterion chosen for the definition of best but it may be the width of
the stop band, the width of the pass band or a combination of both.
The simplest arrangement will consist of a five layer period,

aAmMnNnNmMaA

where the two nN layers are combined into one central one. The filter design will then consist of
this period, repeated a sufficient number of times to achieve the desired rejection, together with
matching assemblies for the incident medium and substrate. The matching assemblies can be for
either s-polarization, p-polarization or a compromise between both.
Three examples of such designs follow. In all cases the layer X is of admittance 1.38 and Y of
admittance 2.45. the incident medium is air of admittance 1.00 and the substrate glass of
admittance 1.52. The designs are given in the captions to the figures. It can be seen fairly easily
that the performance of polarization-insensitive shortwave pass filters is rather better than that
for longwave pass. This is quite normal although it may simply be a case that the better
longwave pass designs are still to be found.
Macleod: Optical Coatings Page 133 Design through Manufacture

Figure 9-9. A longwave pass filter of design:


Air | (0.75X 1.497Y 2.038X 1.497Y 0.75X)17 | Glass
Reference wavelength is 500nm. Angle of incidence in air is 45°. X has admittance 1.38 and Y, 2.45.

Figure 9-10. A longwave pass filter of design:


Air | (0.6X 1.2Y 1.615X 1.732Y 1.615X 1.2Y 0.6X)13 | Glass
Reference wavelength is 500nm. Angle of incidence in air is 45°. X has admittance 1.38 and Y, 2.45.
Macleod: Optical Coatings Page 134 Design through Manufacture

Figure 9-11. A shortwave pass filter of design:


Air | (0.45X 0.547Y 0.322X 0.547Y 0.45X)17 | Glass
Reference wavelength is 700nm. Angle of incidence in air is 45°. X has admittance 1.38 and Y, 2.45.

9.5 Surface Plasmon Resonances


Let a thin film be added to the hypotenuse of a prism illuminated internally beyond critical
angle. The glass prism is the incident medium and the surrounding air is the emergent medium.
Thus the starting admittance for the film is on the imaginary axis, on the negative part for s-
polarization and the positive part for p-polarization. Provided the added thin film has no losses,
then the admittance of film-substrate combination must remain on the imaginary axis. If the film
admittance is imaginary, the combination admittance will simply move towards the film
admittance. If, however, the film admittance is real, the admittance of the combination will move
along the imaginary axis in a positive direction, returning to the starting point every half wave.
Given then that the starting point is on the axis, then the only way in which the admittance can
be made to leave it is by an absorbing layer.
Let us consider a particularly straightforward case where the hypotenuse of the prism is
coated with a single high-performance metal layer. The substrate is still effectively air beyond
the critical angle and therefore is of imaginary admittance, positive for p-polarization and
negative for s-polarization. The form of the metal locus was given earlier and we know that for a
range of values of starting admittance on the imaginary axis, the metal loci loop around, away
from the axis, to cut the real axis. Although the earlier figures show the behavior of metal layers
for an incident medium of unity at normal incidence, the tilted behavior for an incident
admittance of 1.52 is quite similar for both p- and s-polarizations. There is a possibility, then, of
leaving the positive section of the imaginary axis and approaching the real axis in the vicinity of
the admittance of the incident medium for p-polarization provided that the admittances of air and
metal are close. Figure 9-12 shows the illuminating arrangement and the loci. For a very narrow
range of starting values, the metal locus cuts the real axis in the vicinity of the incident
admittance, and, if the metal thickness is such that the locus terminates there, then the reflectance
of the combination will be low. For one particular angle of incidence and metal thickness the
reflectance will be zero. The condition is very sensitive to angle of incidence. Since the
Macleod: Optical Coatings Page 135 Design through Manufacture

admittance of the metal varies much more slowly than the air substrate, the zero reflectance
condition will no longer hold, even for quite small tilts. This very narrow drop in reflectance to a
very low value, with all the hallmarks of a sharp resonance, can be interpreted as the generation
of a surface plasma wave on the metal film. The calculated resonance associated with the
condition is shown in Figure 9-13.

Figure 9-12. Loci of a single silver film on the base of a prism for p-polarization and angles of incidence
equal to and 0.05°on either side of 42.98°. The wavelength is 632.8nm and the optical constants of glass and
silver, 1.5151 and 0.0666-i4.045 respectively.

Figure 9-13. The calculated resonance associated with the single silver film on the base of a prism of Figure
9-12.
This coupling arrangement, devised by Kretschmann and Raether[6], cannot operate for s-
polarized light without modification. The admittance of the substrate for s-polarization is now on
the negative part of the real axis and, therefore, any metal which is deposited will simply move
the admittance of the combination towards the admittance of the bulk metal.
Macleod: Optical Coatings Page 136 Design through Manufacture

An alternative coupling arrangement, devised by Otto[7], involves the excitation of surface


waves through an evanescent wave in an FTR layer (frustrated total reflectance). We recall that
the admittance locus for p-polarization of a layer used beyond the critical angle, is a circle which
is described in an anticlockwise direction. This means that such a layer can be used to couple
into a massive metal. Here the metal acts as the substrate, with a starting admittance in the fourth
quadrant of the complex plane. For p-polarized light, the dielectric FTR layer has a circular locus
which cuts the real axis. Clearly, then, for the correct angle of incidence and dielectric layer
thickness, the reflectance can be made zero.

Figure 9-14. A sketch of the admittance diagram for the Otto form of coupling into a surface plasmon. Here
the coupling layer is air beyond the critical angle. The wave is evanescent and because we are dealing with p-
polarization the circular locus is described counter clockwise.
Macleod: Optical Coatings Page 137 Design through Manufacture

Figure 9-15. The appearance of the resonance with Otto coupling.

Figure 9-16. The calculated appearance of the resonance when the silver film has been overcoated with
0.5nm of material of refractive index 2.0. The measuring wavelength is 632.8nm.
Now let us return to the first case of coupling and let us examine what happens when a thin
layer is deposited over the metal next to the surrounding air. The starting admittance is, as
before, on the imaginary axis, but now the dielectric layer modifies that position, so that the
starting point for the metal locus is changed. Because the metal loci at the imaginary axis are
clustered closely together, almost intersecting, a small change in starting point produces an
enormous change in the locus, and hence in the point at which it cuts the real axis, leading to a
substantial change in reflectance, Figure 9-16. Provided that the film is very thin, then an
additional tilt of the system will be sufficient to pull the intersection of the metal locus with the
Macleod: Optical Coatings Page 138 Design through Manufacture

real axis, back to the incident admittance, and so the effect can be interpreted as a shift in the
resonance rather than a damping. This very sensitive effect has been used in the study of
contaminant films adsorbed on metal surfaces[8, 9]. Film thicknesses of a few Angstroms can be
detected.
Surface plasma oscillations and their applications are extensively reviewed by Raether[10].
Abelès[11] includes an account of the optical features of such effects in his review of the optical
properties of very thin films.

Figure 9-17. The long-range surface plasmon. The materials on either side of the metal layer are beyond the
critical angle. Thus the metal starts on the positive imaginary axis and loops round near the negative
imaginary axis where the evanescent wave in the coupling layer sweeps down to the real axis and the
admittance of the incident medium. The calculated admittance diagram on the right has the same lettering as
the sketch. The values used in the calculation are outermost medium and coupling layer indices of 1.45.
coupling layer geometrical thickness of 1.525λ. Metal layer silver with (n-ik) = (0.075-i3.41) and d/λ = 0.03.
The angle of incidence of the resonance is 54.591° in the incident medium of index 1.8. The light is
p-polarized.
The long-range surface plasmon is a phenomenon that has been of some interest[12, 13]. Once
again the light must be p-polarized for the effect to be possible. A thinner metal layer than in the
classical case, and it is surrounded by dielectric media that are beyond the critical angle so that
they support evanescent waves. A resonance that is very narrow is induced in the reflectance of
the assembly. The usual way of explaining the effect involves two surface waves, one on each
side of the metal layer. Here we consider an alternative explanation involving the admittance
diagram.
Macleod: Optical Coatings Page 139 Design through Manufacture

The starting point is the positive imaginary admittance of the final medium. The metal layer
starts at this point and loops round to a termination point near the negative imaginary axis. There
the evanescent wave in the coupling layer sweeps the locus down to the real axis ending at the
admittance of the incident medium and zero reflectance. A calculated resonance uses a prism of
index 1.8 with outermost medium and coupling layer having indices of 1.45. The coupling layer
geometrical thickness is 1.525λ. The metal layer is assumed to be silver with
(n-ik) = (0.075-i3.41) and d/λ = 0.03. The angle of incidence of the resonance is 54.591° in the
incident medium.

Figure 9-18. The resonance associated with the long-range surface plasmon. This resonance was designed to
be exactly at 54° in a prism of index 1.8. The silver has optical constants 0.0666-i4.045 and the dielectric
layers on either side have index 1.45. Thicknesses are of silver 19.32nm, layer next to the prism, 1347.71nm
and next to air, 343.97. Note how much narrower the resonance is compared with Figure 9-13.

9.6 Coupled plasmon-waveguide resonator


The changes that can be measured in the surface plasmon resonance when small amounts of
material is added can be used in the characterization of thin biological films and especially in
biochemical investigations of membrane protein processes[14, 15]. These measurements are
made in an aqueous medium and so are at rather higher angles of incidence than measurements
made with air as the emergent medium. Figure 9-19 shows a typical resonance in silver with the
displacement that occurs when a thin layer of dielectric material is added.
Silver gives narrowest resonances, and therefore greatest detection sensitivity in the visible and
near infrared, but it is not a rugged material and can deteriorate even in a normal laboratory
environment. An overcoat is indicated but it must not cause any deterioration in the detection
capabilities. Even more satisfactory would be to enhance the detection. We look for structures,
preferably simple, that could be used as overcoats but also to narrow the resonance.
There are many possibilities but a particularly effective coating is a thick layer of material of
refractive index slightly higher than that of water. The action is most easily understood in terms
of our optical admittance diagram. We use the modified tilted admittances as usual.
Macleod: Optical Coatings Page 140 Design through Manufacture

OSA#01: Reflectance
100

80

Reflectance (%) 60

40

20

0
65 66 67 68 69 70 71 72 73 74 75
Incident Angle (deg)
Figure 9-19. A typical surface plasmon resonance associated with a 50nm silver film at 632.8nm. The prism
is made of glass and the outer medium is water. The grey curve shows the effect of adding 1nm of high-index
material to the outer surface of the silver.

We consider first the classical resonance involving only the thin metal film. We are beyond
the critical angle and so the emergent medium is represented by a positive imaginary admittance
for p-polarized light that reduces rapidly in magnitude with increasing angle of incidence. Metals
are characterized by loci that for both p- and s-polarization start on the positive limb of the
imaginary axis and sweep down across the real axis into the fourth quadrant. If the starting point
on the imaginary axis is correctly chosen, then the metal locus will cross the real axis at the point
corresponding to the admittance of the incident medium and, if it terminates there, the
reflectance will be zero. The conditions are extremely sensitive to angle of incidence yielding the
narrow p-polarized resonance. Since the emergent medium admittance for s-polarized light is
negative imaginary, the subsequent metal locus is confined to the fourth quadrant and completely
unable to cross the real axis implying no resonance effect.
A dielectric layer, silica in this case, inserted between the emergent medium and the outer
metal layer, has a locus that starts on the imaginary axis and must remain on it. Since this layer is
below critical angle, the locus actually consists of a circle with infinite diameter implying that
the admittance climbs up the positive direction of the axis. The circle is completed for every half
wave of optical thickness at the appropriate angle of incidence and so a half wave thickness at
exactly the resonant angle will leave the condition for the p-polarized resonance unchanged. A
change in thickness alters the angle of the resonance and a variation in overcoat thickness can be
accommodated in this way. The cosϑ dependence of the phase thickness of the dielectric layer,

δ = 2π nd cos ϑ λ (9.30)

is rapid, and greatly narrows the response peak. Since the overcoat is of low index, the high
angle of incidence splits the p- and s-admittances considerably so that the s-admittance
(modified) is very much greater than the p-admittance. This means that in the region around the
origin the s-admittance locus of the dielectric that starts below the real axis tends to catch up
Macleod: Optical Coatings Page 141 Design through Manufacture

with the p-admittance. It becomes possible not only to have an s-polarization resonance, but to
have the resonance at an angle of incidence close to that for p-polarization. This is shown in
Figure 9-20. This enables both s- and p-polarization measurements with increased
sensitivity[16]. Such coatings have been successfully manufactured and used.

OSA#02: Reflectance
100

80
Reflectance (%)

60

40

20

0
65 66 67 68 69 70 71 72 73 74 75
Incident Angle (deg)
Figure 9-20. The original 50nm silver film is now covered by 800nm of SiO2. The new resonances for both
p-polarization (leftmost resonance) and s-polarization (rightmost resonance) are compared with the original
classical surface plasmon resonance (thin line).

9.7 Ellipsometry
The measurement of the absolute properties of an optical surface with or without a coating is
often desirable. Values of R and ϕ for an opaque surface, for example, define completely and
unambiguously the optical constants of the surface. Absolute reflectance is a difficult
measurement and it is more usual to measure the way in which the unknown surface compares
with a known reference - which introduces further difficulties. Phase is even more involved,
requiring an interferometric operation as well as a known standard. Phase measurements are,
therefore, quite rare and routine measurements of reflectance are almost always comparative. A
major problem is the calibration and maintenance of suitable standards. There is a way, however,
a way of avoiding such difficulties. At normal incidence there is only one value of reflectance
and one of phase but at oblique incidence there are two, one pair for s-polarization and the other
for p-. In principle, therefore, it should be possible to use as one as a reference for the other and
this leads to the method known as ellipsometry. Ellipsometry has many advantages, especially
the ability to use a single illuminated spot for both measurements. Two quantities are involved,
the ratio of p and s reflectances |ρp/ρs| and denoted by tanψ where ψ is an angle in the first
quadrant, and (ϕp-ϕs) the relative phase shift denoted by ∆. The effect of ψ and ∆ is to change
the state of polarization of the reflected light with respect to the polarization of the incident, and
so they are directly and simply related to the ellipticity and orientation of the polarization of the
beam. ψ and ∆ are therefore known as the ellipsometric parameters and their study is known as
ellipsometry.
Macleod: Optical Coatings Page 142 Design through Manufacture

Ellipsometry possesses several advantages and disadvantages over other measurement


techniques. Advantages are the absence of any reference samples that must be maintained and,
although high accuracy is required, the measurement is simple involving straightforward
manipulations of polarized light. Disadvantages are that the measurement is at oblique incident,
quite far from more normal measurements of performance, making it difficult to exercise instinct
in judging the results. A limitation is that there are two parameters only, rather less than the
number that must often be established for a complete description of the system. This limitation is
frequently addressed by additionally varying the angle of incidence, or the wavelength, or both.
The combination is known as Variable Angle Spectroscopic Ellipsometry, frequently
abbreviated to VASE.
A full description of ellipsometry and its techniques is beyond the scope of this course but
some observations are appropriate.
First of all the ellipsometric convention for phase shift is different from that normally used in
optical coatings. The p-polarization reference direction in the reflected beam is reversed
implying a difference of 180° in the values for p-polarized phase shift on reflection. The reason
for this difference is the desirability in ellipsometry of arranging that the reference directions for
s- and p-polarizations should coincide with the reference directions used in defining the elliptical
polarization state. It would be very difficult if these were changed in the reflected beam.
Only two parameters, refractive index and extinction coefficient are sufficient to define a
simple surface. Since there are two ellipsometric parameters ψ and ∆, then it should be possible
to make a determination of the surface parameters from a single ellipsometric measurement. This
is indeed the case and there is a direct analytical connection between the two sets of parameters.
Unfortunately, this is not the case with a thin film on a substrate. Even with the simplest film on
a substrate that is already characterized, there are at least three parameters, n, k and d, necessary
to define the film. The properties of films that are absorbing may depart only slightly from a
surface of bulk material. In such cases it is often assumed that the extraction techniques used for
a simple surface are applicable. The parameters, n and k, that are extracted in this way are
usually referred to as the pseudo-optical constants. They exhibit, usually, the gross features of
the real optical constants, although they may not be suitable for thin film calculations and
predictions.
In spectroscopic ellipsometry, the wavelength is varied. Since the film physical thickness is
not sensitive to wavelength, this introduces an element of redundancy. It is then sufficient to
introduce a small amount of additional information. This frequently takes the form of a
prescribed spectral variation of optical constants. Other film parameters may then also be
included, and the redundancy in the measurement can be so great that even simple multilayers
may be evaluated.
We illustrate the extraction process by considering the simple case of a single wavelength,
single angle measurement of a surface characterized by refractive index n and extinction
coefficient k.
Let the incident medium be of index unity and let ε = tan ψ expi∆ . Then
Macleod: Optical Coatings Page 143 Design through Manufacture
ρ
ε= p =
d
η op − η p η 0 s + η s

ib g
ρs d
η op + η p η os − η s ib g (9.31)

where the symbols may be taken as the modified admittances and the sign convention of ∆ may
be considered corrected to the usual thin film convention by adding or subtracting 180°. Then,

ε=
c1 − y h − dη2
p i
− ηs
c1 − y h + dη2
p −η i
s
(9.32)

where we have replaced the incident medium admittance by unity. Now

y 2 − sin 2 ϑ 0
ηs =
cosϑ 0 (9.33)

y 2 cosϑ 0
ηp =
y 2 − sin 2 ϑ 0
and (9.34)

so that after some manipulation we can write

F η I η F 1 − c y − sin ϑ h I
2 2

1 − ε dη − η i
p
η G1 − J
Hs η K
GH y cos ϑ JK
p
s

p
p 2 2
0
0

γ = =
1+ ε c1 − y h = c1 − y h =
2
c1 − y h
2 2
(9.35)

η p sin 2 ϑ 0 sin 2 ϑ 0
γ = 2 2 =
y cos ϑ 0 cosϑ 0 y 2 − sin 2 ϑ 0
i.e. (9.36)

sin 4 ϑ 0
2
y = 2 + sin 2 ϑ 0
This gives γ cos ϑ 0
2
(9.37)

There will be two solutions and the fourth quadrant solution will be the correct one.

9.8 Chapter 9 References


1. MacNeil, S M. Beam splitter. 1946. USA. Patent 2,403,731.
2. Songer, L, The design and fabrication of a thin film polarizer. Optical Spectra, 1978.
12(10): p. 45-50.
3. Mahlein, H F, Wavelength-selective beam splitters with minimum polarizing effects for
wavelength-division multiplexing in optical communication systems. Optica Acta, 1981.
Macleod: Optical Coatings Page 144 Design through Manufacture

28: p. 29-41.

4. Thelen, A, Nonpolarizing interference films inside a glass cube. Applied Optics, 1976.
15: p. 2983-2985.
5. Thelen, Alfred, Design of Optical Interference Coatings. First ed. McGraw-Hill Optical
and Electro-Optical Engineering Series, ed. R.E. Fischer and W.J. Smith. 1988, New
York: McGraw-Hill Book Company.
6. Kretschmann, E and H Raether, Radiative decay of non-radiative surface plasmons
excited by light. Zeitschrift für Naturforschung, 1968. 23A: p. 2135-2136.
7. Otto, A, Excitation of non-radiative surface plasma waves in silver by the method of
frustrated total reflection. Zeitschrift für Physik, 1968. 216: p. 398-410.
8. Abelès, F and T Lopez-Rios, Ellipsometry with surface plasmons for the investigation of
superficial modifications of solid plasmas, in Polaritons. Proceedings of the First
Taormina Research Conference on the Structure of Matter, 1972, Taormina, Italy, E.
Burstein and F.D. Martini, Editors. 1974, Pergamon Press: New York. p. 241 - 246.
9. Song, D Y, F S Zhang, H A Macleod, and M R Jacobson, Study of surface contamination
by surface plasmons. Proceedings of the Society of Photo-Optical Instrumentation
Engineers, 1986. 678: p. 211-218.
10. Raether, H, Surface plasma oscillations and their applications, in Physics of Thin Films,
G. Hass and M.H. Francombe, Editors. 1977, Academic Press: New York, San Francisco
and London. p. 145-261.
11. Abelès, F, Optical properties of very thin films. Thin Solid Films, 1976. 34: p. 291-302.
12. Craig, A E, G A Olson, and D Sarid, Experimental observation of the long-range
surface-plasmon polariton. Optics Letters, 1983. 8: p. 380-382.
13. Quail, J C, J G Rako, and H J Simon, Long-range surface-plasmon modes in silver and
aluminum films. Optics Letters, 1983. 8: p. 377-379.
14. Salamon, Zdzislaw, H Angus Macleod, and Gordon Tollin, Surface plasmon resonance
spectroscopy as a tool for investigating the biochemical and biophysical properties of
membrane protein systems. I: Theoretical principles. Biochimica et Biophysica Acta,
1997. 1331: p. 117-129.
15. Salamon, Zdzislaw, H Angus Macleod, and Gordon Tollin, Surface plasmon resonance
spectroscopy as a tool for investigating the biochemical and biophysical properties of
membrane protein systems. II: Applications to biological systems. Biochimica et
Biophysica Acta, 1997. 1331: p. 131-152.
16. Salamon, Zdzislaw, H Angus Macleod, and Gordon Tollin, Coupled plasmon-waveguide
resonators: A new spectroscopic tool for probing proteolipid film structure and
properties. Biophysical Journal, 1997. 73: p. 2791-2797.
Macleod: Optical Coatings Page 145 Design through Manufacture

10 INHOMOGENEOUS LAYERS AND RUGATE


FILTERS
The term rugate is gradually becoming common in optical thin films. The definition of the
term is not clear because different workers use it in different ways. To understand what is meant
by it we need to consider an inhomogeneous layer. Imagine an incident medium of admittance y0
and an emergent medium of higher admittance ysub. Let there be a gradual transition in
admittance from that of the emergent medium to that of the incident. Further, let this layer be
sliced into very thin sections. Now we apply the vector method to this system. There will be
many vectors, all of equal length and negative (ysub is higher than y0) and very small. The
difference in angle between the first and the last will depend on the total phase thickness of the
coating and will be twice that value. Now let the total phase thickness of the system be
equivalent to one halfwave at g=1. As g grows from zero towards unity, the amplitude reflection
coefficient will gradually fall from the value corresponding to the uncoated substrate (when all
the vectors are aligned) to exactly zero at g=1 (where they form a complete circle). As g grows
beyond unity the reflectance will rise until the circle is described one and one half times at a
value of g of 1.5. Here the reflectance peak will be given by the square of the diameter of the
circle. Since the total length of the locus is equivalent to the original substrate amplitude
reflection coefficient (We can assume glass with an admittance of 1.52 in air of 1.00 giving a
reflectance of 4.25% and an amplitude reflection coefficient of (0.0425)1/2). One and one half
circumferences of a circle is 1.5πD where D is the diameter and so the reflectance will roughly
be R0/(1.5π)2 that is roughly R0/22 or, for glass in air, about 0.2%. Increasing g further makes the
circular locus alternately close and open with reducing peak reflectances. An accurate
calculation is shown in Figure 10-1. Different forms of transition have slightly different profiles
but we can roughly say that the halfwave inhomogeneous layer is an excellent antireflection
coating for all values of g for which the phase thickness exceeds a halfwave.
Since the profile is not very dependent on the actual form of the variation as long as it is
monotonic (that is it does not have any extrema or points of inflexion) the inhomogeneous layer
is also insensitive to variations in angle of incidence. It becomes effectively thinner, of course,
and this eventually limits the angle that can be attained before the performance begins to
degrade.
Unfortunately, the normal antireflection problem of matching an optical medium to air,
requires a coating that has an outermost admittance of unity. Normal thin film materials have
much higher admittances and a conventional approach is not possible. Microstructural
approaches have been demonstrated. It is possible to construct a coating that has increasing void
volume towards the outer surface and in this way approximate the required admittance
distribution. This variation of void volume can be achieved in different ways. The looseness of
packing of a sol-gel coating is one example[1]. Arrays of pyramids, usually constructed by a
process of masking and etching, sometimes called moth's eye coatings[2], represent another. The
performance of such coatings can be impressive but there are two problems. The first is that the
outer parts of the coating are exceedingly weak and easily damaged and/or contaminated. The
second is that at shorter wavelengths the structural features increasingly scatter light. For the
Macleod: Optical Coatings Page 146 Design through Manufacture

moth eye coating this implies rather narrow and long pyramids if the region of high performance
is to be reasonably broad.

Figure 10-1. The reflectance of an inhomogeneous linear transition from the emergent medium of admittance
1.52 to the incident medium of admittance 1.00. The thickness of the layer is a full wave at g=1 so that g=0.5
corresponds to a halfwave.
The inhomogeneous layer can, however, be used to advantage to solve internal matching
problems in multilayers where the lower admittances are attainable[3]. Rugate filters[4] probably
represent the most extreme example of this. Reflectors that consist of discrete layers exhibit
higher-order high reflectance zones at shorter wavelengths and these can present problems in
short wave pass filters where wide regions of thigh transmittance are required. The wider the
long wave rejection zone must be, the more severe the problem. To eliminate higher order high
reflectance, antireflection coatings are needed at the interfaces between the various layers. These
antireflection coatings must be ineffective at the fundamental wavelength but must operate at all
wavelengths shorter than fundamental. Clearly, an inhomogeneous layer with thickness
sufficiently small to prevent operation at the fundamental, that is rather less than a halfwave, but
with thickness sufficient to suppress reflectance at shorter wavelengths, is a good candidate. It is
used in the classical rugate that has a smooth sinusoidal (strictly logarithmic sinusoid) variation
of admittance that suppresses all orders greater than the first.
The concept of matching using inhomogeneous layers can be considerably extended. The
quality of matching depends on the rate at which the admittance changes as a function of the
optical thickness. The more rapid the admittance change, the poorer the matching and the
greater, therefore, the light reflected by that part of the coating. If the gradient in question is
within a coating then the reflected light, when it reaches the front of the system will have a phase
associated with it that includes the passage from front surface into the coating and back out
again. At the front surface, therefore, the performance can be considered to be the combination
Macleod: Optical Coatings Page 147 Design through Manufacture

of all such beams and will appear as the Fourier transform of the derivative of the admittance as
a function of the optical depth. In fact the correct expression is[5]:

z

dn 1
dx
b g bg bg
⋅ ⋅ exp ikx ⋅ dx = Q k ⋅ exp iϕ k
2n
(10.1)
−∞

where Q is a function of performance, k = 2π/λ and x is twice the optical path. ϕ(k) is a phase
factor that must be an odd function to ensure that n(k) is real. Multiple beam effects are
neglected and judicious choice of Q can reduce the errors that arise from this approximation.
Functions that have been proposed and used for Q include:

Q= R
R
Q=
T
1 1FG IJ (10.2)
Q=
2 T H
−T
K
1
Q= − T
T

The great advantage of this approach is that it permits an analytical connection in either
direction of a function of design with a function of performance. If we know the performance we
can find a design and vice versa. Disadvantages are that the technique is approximate and
considerable skill and experience are required in the choice of the appropriate Q function and
phase factor ϕ. Although the resulting design is a continuously varying admittance profile, it can
be converted into a discrete-layer design, the thicknesses being chosen thin enough not to affect
performance at the shortest wavelength of interest. The "Flip-flop" technique has also been
proposed as a useful way of achieving such designs.
Finally we note that the term "rugate" is often used for any layer system in which there is a
deliberate attempt to induce an inhomogeneity although the original word was adopted from
biological systems to describe a film system in which there was a sinusoidal variation of
admittance so that only the fundamental first order reflection peak was present, the other being
suppressed.

10.1 Chapter 10 References


[1] B. E. Yoldas and D. P. Partlow, “Wide spectrum antireflective coating for fused silica
and other glasses,” Applied Optics, vol. 23, pp. 1418-1424, 1984.
Macleod: Optical Coatings Page 148 Design through Manufacture

[2] P. B. Clapham and M. C. Hutley, “The moth's eye coating,” Nature, vol. 244, pp. 281,
1973.

[3] R. Jacobsson, “Matching a multilayer stack to a high-refractive-index substrate by means


of an inhomogeneous layer,” Journal of the Optical Society of America, vol. 54, pp. 422-
423, 1964.
[4] B. G. Bovard, “Derivation of a matrix describing a rugate dielectric thin film,” Applied
Optics, vol. 27, pp. 1998-2005, 1988.
[5] J. A. Dobrowolski and D. Lowe, “Optical thin film synthesis program based on the use of
Fourier transforms,” Applied Optics, vol. 17, pp. 3039-3050, 1978.
Macleod: Optical Coatings Page 149 Design through Manufacture

11 SHORT-PULSE PHENOMENA IN COATINGS


Traditionally, coating designers have been able to rely on the steady-state nature of the effects
they seek to produce. There are now laser systems, known as ultrafast, capable of generating
pulses of light that are short enough for transient response to become significant. Optical
communication uses trains of very short pulses that are manipulated by beam splitters in the form
of multiple-cavity narrow band filters. These pulses are rather longer than those in ultrafast
systems but on the other hand the effect of the multiple-cavity filters is correspondingly greater.
The theory of the effect in either case is similar.
There is no existing large body of knowledge in optical coatings that deals with short pulse
phenomena and coating designers are largely unfamiliar even with the terminology, which
includes such parameters as Group Delay and Group Delay Dispersion, and, in communication,
Chromatic Dispersion. These notes attempt to explain the effects in terms that are familiar to
optical coating practitioners.
A normal high reflector consisting of a quarterwave stack might be some twenty-five
quarterwaves in thickness. At a wavelength of 1µm this implies a trip length for light traveling
from the front to the rear of the coating and back again of 12.5µm or a trip time of around 42fs.
Pulses that are around 50fs in length are now becoming common in the ultrafast field and pulses
can be even considerably shorter. It is clear that transient response of coatings must now be
considered important in such applications, but the effects, in fact, can be significant even with
pulses some two or three orders of magnitude longer. The idea that coating properties should
have an influence on short pulses and that they might be engineered to have prescribed effects is
not new. It is, however, only recently that the field has expanded and the technology advanced to
the stage where the application is becoming of major importance. In the communication field the
components in question are multiple-cavity narrowband filters with extremely narrow widths.
The pulses are reflected backwards and forwards many times before they emerge from these
components and so the transient effects can be measured in terms of picoseconds rather than
femtoseconds.

11.1 Short pulses


A short pulse can be thought of as an envelope over a carrier. The carrier contains the phase
information associated with the pulse and it travels at what is known as the phase velocity. The
energy is obviously associated with the envelope that travels at what is known as the group
velocity. The group velocity and the phase velocity are generally different, the phase velocity
being the greater. Thus the carrier appears to run through the pulse envelope. A short pulse with
Gaussian envelope is shown in Figure 11-1.
The pulse may also be visualized in a different way, as a collection of monochromatic
component waves with a continuous distribution of frequencies over a given band. The coherent
combination of these monochromatic waves yields the envelope and carrier of the alternative
model. Both of the models are entirely equivalent and if we wish, we can pass from one to the
other by way of a Fourier transform.
Macleod: Optical Coatings Page 150 Design through Manufacture

Pulse envelopes frequently have a Gaussian shape[1, 2]. For simplicity we can look at the
temporal variation at the origin of our coordinates, t = 0, z = 0, and then
t2

F ( t ) = Ae 2µ2
(11.1)

where µ has the dimensions of time. The Fourier transform gives the frequency distribution and
it is also a Gaussian function,
2
µ 2 (ω −ω0 )

G (ω ) = Be 2
(11.2)

and it gives the form of the amplitude of the component wave of frequency ω. If the time
between the half-maximum points [F and G having values of 1/√2] is τ and the full width of the
pulse (angular) frequency distribution also at half-maximum is 2∆ω [we will be using ∆ω later as
the frequency width from the peak to the half-peak point] then

τ ⋅ 2∆ω = 4 ⋅ log e 2 (11.3)

Note that both these quantities are functions of µ. For example,

τ = 2 log e 2 ⋅ µ (11.4)

Gaussian light pulse


1.2
0.8
Disturbance

0.4
0.0
-0.4
-0.8
-1.2
-60 -40 -20 0 20 40 60
Time or distance
Figure 11-1. A short Gaussian-shaped pulse consisting of an envelope over a carrier of constant frequency.
The carrier phase may move faster than the pulse when it will appear to run through the envelope as it travels.
The center of the pulse is the point where all of the component waves are exactly of identical
phase. If all the component waves travel at the speed of light in vacuo then the phase coincidence
will also travel at that speed and the center of the pulse will move with it. Similarly if all waves
slow down equally then the pulse will slow down to the same extent but will otherwise be
unchanged.
Macleod: Optical Coatings Page 151 Design through Manufacture

11.2 Group Velocity


The relative phase of the carrier within the pulse is set by the value of phase where all the
component waves coincide. If the phase of the waves is zero then the carrier will have a peak
exactly at the peak of the pulse. We can find the position of the pulse peak at any time by a
simple procedure.
The pulse can be considered to be made up of monochromatic component waves. As these
propagate the phase relationships between them will change, but if the pulse shape is unaltered
as it propagates then at any particular time there must be a distance along the path where the
phase is identical for all the component waves, and this must correspond to the pulse center. We
use the normal thin film convention of (ωτ-κz) in the phase factor where κ = 2π n λ with λ the
free space wavelength. We write the component wave phase at distance z and time t as
ϕ = ϕ 0 + ∆ϕ . Then for coincidence of all component phases, ∆ϕ must be zero.
This condition is

(ω 0 + ∆ω ) t − (κ 0 + ∆κ ) z = ϕ 0 + ∆ϕ
ω 0t − κ 0 z = ϕ 0
∆ϕ = 0 = ∆ω t − ∆κ z (11.5)
∆ω
z= t = vg t
∆κ

The quantity ∆ω/∆κ is known as the group velocity, vg , and clearly it must remain constant if
the position z is to be the same for all the component waves and the shape of the pulse
unchanged.

Widt h of
Pulse dir ect ion Component waves
ω fr equency
dist r ibut ion along z-dir ect ion

Lines of const ant ϕ


Pulse posit ion
z
Figure 11-2. Sketch showing the component waves of the pulse as horizontal lines along the direction of
propagation and with their relative phases marked as contour lines across them. The pulse peak coincides
with the position where the phase of all the components is exactly equal.
An alternative visualization involves a simple diagram. We plot the z-direction horizontally
and ω vertically. We sketch the bundle of component waves making up the pulse, as a set of
lines through the appropriate values of ω and parallel to the z-axis. We mark contours of constant
Macleod: Optical Coatings Page 152 Design through Manufacture

ϕ on the lines. Provided there is one contour that runs normally across the lines then the pulse
peak will be positioned there and the pulse shape will be unchanged.
In a nondispersive medium the phase at the peak will be zero because all the component
waves will be traveling at identical velocity even though it may be less than the velocity in free
space. In a dispersive medium, the component waves travel at different velocities according to
the particular value of refractive index. Provided the variation in velocities still permits a phase
coincidence somewhere, then the pulse will appear there and will be unchanged in shape
although the phase of the carrier wave will be altered. It is clear from equation (11.5) that the
critical condition is for the group velocity to remain constant across the frequency spectrum of
the pulse.
In a dispersive medium, the refractive index changes with frequency. We can calculate the
group velocity in terms of this change.

2π n (ω ) ω n (ω )
κ= =
λ c

dk n(ω ) ω dn (ω )
= + ⋅
dω c c dω

c
vg = (11.6)
dn (ω )
n (ω ) + ω ⋅

In a medium with normal dispersion, this is not constant.

11.3 Group Velocity Dispersion


There is no guarantee that the group velocity be constant with changing frequency. If the
second derivative of ω with respect to κ is non zero then there can be no phase coincidence and
the pulse will be perturbed. If we limit ourselves to the second derivative then we can write the
expression for the phase of an arbitrary component wave as:

⎛ dκ 1 2 d κ
2 ⎞
(ω 0 + ∆ω ) ⋅ t − ⎜⎜ κ 0 + ∆ω ⋅ + ( ∆ω ) ⋅ ⎟⎟ ⋅ z = ϕ + ∆ϕ (11.7)
⎝ dω 0 2 dω 2 0⎠

and we can immediately identify a problem. The third term in the coefficient of z is even in ∆ω
and so cannot be compensated by the other terms. A simple way of handling this problem is to
pair the component corresponding to +∆ω with that corresponding to -∆ω. In each case we
ensure that the value of ∆ϕ is zero. This implies introducing a different time t for each member
of the pair although we retain the same value of z, and, therefore, gives two equations for each
∆ω instead of the usual one.
Macleod: Optical Coatings Page 153 Design through Manufacture

dκ 1 2 d κ
2
∆ω ⋅ t1 − ∆ω ⋅ ⋅ z − ( ∆ω ) ⋅ ⋅z =0 (11.8)
dω 0 2 dω 2 0

dκ 1 2 d κ
2
−∆ω ⋅ t2 + ∆ω ⋅ ⋅ z − ( ∆ω ) ⋅ ⋅z =0 (11.9)
dω 0 2 dω 2 0

d 2κ
Then ∆t = ( t1 − t2 ) = ∆ω ⋅ ⋅z (11.10)
dω 2

The component pairs are, therefore, displaced from each other in time and, of course, the pulse
must be broadened in consequence. We could treat the ∆ω in expression (11.10)as half the width
of the frequency distribution of the basic initial pulse, so that the range of frequencies would
then be from ω0-∆ω to ω0+∆ω. The pulse would then be smeared out by the ∆t in (11.10).
Actually, more rigorous calculation shows that we should better use a value of 2∆t instead of just
∆t
Then the new pulse width is given by

τ 2 = τ 02 + ( 2∆t )
2

If we write

d 2κ
τ g2 = ⋅z
dω 2 0

then,

τ 2 = τ 02 + ( 2∆ωτ g2 )
2

and since ∆ω = (√loge2)/µ2 and τ0 = 4(√loge2)µ2,


12
⎛ τ4 ⎞
τ = τ 0 ⎜⎜1 + g4 ⎟⎟ (11.11)
⎝ τ0 ⎠

Note that (τg/µ)4 rapidly becomes negligible when τg becomes less than µ but rises sharply when
it is greater. This means that as long as long as τg is comfortably less than µ its effect can be
neglected.
Figure 11-3 shows a sketch of the spectral components of the pulse in the presence of
appreciable group velocity dispersion. At the center of the pulse there is no exact phase
coincidence. All that can be done is to pick the best position where the phase varies least in total
across the pulse. We can think of the pulse as split into two parts each of which represents a
pulse on its own, with its own group velocity. Because these group velocities are different, there
Macleod: Optical Coatings Page 154 Design through Manufacture

is a difference in the arrival times of the two sub-pulses representing a broadening of the
resultant.

Widt h of
Pulse dir ect ion Component waves
ω fr equency
dist r ibut ion along z-dir ect ion

Peaks of
component Lines of const ant ϕ
pulses Pulse posit ion z
Figure 11-3. The pulse frequency distribution is now split into two parts, each of which represents a
component pulse with its own center position. Since the group velocity is different for the two component
pulses they separate such that one lags behind the other and the combined pulse is broadened.
A more rigorous calculation of the effect of the group velocity dispersion on the pulse is a
little involved [2]. We choose the position z where we will measure the pulse. We will then treat
z as a constant in the analysis. Then, at z, each individual component of the pulse can be written
in the form of an amplitude that depends on the frequency together with the phase factor. This is:

⎡ ( )2 µ 2 ⎤ exp ⎡i ⎧ ⎛ dκ ⎞ ( ∆ω ) d κ ⎫ ⎤ (11.12)
2 2
B exp ⎢ − ∆ω ⎥ ⎢ ⎨( ω 0 t − κ 0 z ) + ∆ω ⎜ t − ⋅ z ⎟ − ⋅ ⋅ z ⎬⎥
⎣ 2 ⎦ ⎣ ⎩ ⎝ dω ⎠ 2 dω 2 ⎭ ⎦

The amplitude of the component is taken straight from (11.2). To find the temporal profile of the
pulse we need to combine all the spectral components and this implies integrating (11.12) over
all frequencies. To make it easier to see what we are doing, we will define a reference time, t0, as
the time taken by the pulse to reach z assuming that it is traveling at its group velocity. Then any
arbitrary time, t, will be expressed as t0+∆t. This implies


t− ⋅ z = t − t0 = ∆t (11.13)

Then

B exp ⎡⎣i (ω 0t − κ 0 z ) ⎤⎦
+∞
⎧ ⎡ ⎤ ⎡ ⎧ ( )2 d 2κ ⎫⎤ ⎫d ( ) (11.14)
× ∫ ⎨exp ⎢ − ( ∆ω ) µ ⎥ exp ⎢i ⎨ ∆ω ⋅ ∆t − ∆ω ⋅
2 2
⋅ z ⎬⎥ ⎬ ∆ω
−∞ ⎩ ⎣ 2 ⎦ ⎣ ⎩ 2 dω 2 ⎭ ⎦ ⎭

which, with a small rearrangement, becomes


+∞
⎧ ⎡ ⎧µ2 1 d 2κ ⎫ ⎤⎫
B exp ⎡⎣i (ω 0t − κ 0 z ) ⎤⎦ ∫−∞ ⎨⎩ ⎢⎣ ⎨⎩ 2 + i 2 ⋅ dω 2 ⋅ z ⎬⎭ ( ∆ω ) + i∆t ⋅ ∆ω ⎥⎦ ⎬⎭d ( ∆ω ) (11.15)
exp − 2
Macleod: Optical Coatings Page 155 Design through Manufacture

This integral is a standard form[2]:


+∞
π ⎛ b2 ⎞
∫ exp {− ( ax 2 + bx )} dx =
−∞
a
exp ⎜ ⎟
⎝ 4a ⎠
(11.16)

Putting (11.15) in the form of (11.16), we obtain

⎡ − ( ∆t )2 ⎤
2π ⎢ ⎥
⋅ B ⋅ exp ⎛ 2
⎢ d 2κ ⎞ ⎥ ⋅ exp ⎡⎣i (ω 0t − κ 0 z ) ⎤⎦ (11.17)
dκ2
2 ⎜µ +i ⋅z⎟
µ2 + i 2 ⋅ z ⎢ ⎝
⎣ dω 2 ⎠ ⎦⎥

the envelope of which simplifies to (11.1) if d2κ/dω2 is zero. The irradiance is given by the
product of (11.17) with its complex conjugate. This is

⎡ ⎧ 1 1 ⎫⎤
2π 2⎢ ( ∆t )2 ⎪ + ⎪⎥
⋅ B ⋅ exp − ⋅ ⎨⎛ 2 dκ ⎞ ⎛ 2
2
d κ ⎞ ⎬ ⎥ (11.18)
2
⎢ 2 ⎪⎩ ⎜⎝ µ + i dω 2 ⋅ z ⎟⎠ ⎜⎝ µ − i dω 2 ⋅ z ⎟⎠ ⎭⎪ ⎥
2
⎛ d 2
κ ⎞ ⎢
⎣ ⎦
µ +⎜ 2 ⋅z⎟
4

⎝ dω ⎠

that is,

⎡ ⎧ 2µ 2 ⎫⎤
2π ⎢ ( ∆t )2 ⎪ ⎪⎥
2
⋅ B ⋅ exp ⎢ − ⋅ ⎨ ⎛ 4 ⎡ d 2κ ⎤ ⎞ ⎬⎥
2
(11.19)
2 ⎪ ⎜ µ + ⎢ 2 ⋅ z ⎥ ⎟ ⎪⎥
2
⎛ d 2
κ ⎞ ⎢ ⎣ dω ⎦ ⎠ ⎭⎦
µ4 + ⎜ 2 ⋅ z ⎟ ⎣ ⎩⎝
⎝ dω ⎠

Since this is now an expression in terms of irradiance rather than amplitude the width of the
pulse will be measured between the half-peak points rather than the 1/√2 points we used earlier.
The full width of the pulse, τ at the half-peak points is given by 2∆t where

⎛ 4 ⎡ d 2κ ⎤ 2 ⎞
⎜ µ + ⎢ 2 ⋅ z⎥ ⎟
⎣ dω ⎦ ⎠
( ∆t ) = ( log e 2 ) ⋅ ⎝
2
(11.20)
2µ 2

1
⎛ ⎡ d 2κ ⎤ 2 ⎞ 2
τ = 2∆t = 2 ( log e 2 ) ⋅ µ ⋅ ⎜⎜ ⎢⎣ dω 2 ⋅ z ⎥⎦ ⎟⎟
⎜1 + µ4 ⎟
⎝ ⎠ (11.21)
1
⎛ τ g4 ⎞ 2
= τ 0 ⎜1 + 4 ⎟
⎝ µ ⎠
Macleod: Optical Coatings Page 156 Design through Manufacture

This is exactly the expression quoted earlier. Note that the imaginary part of the argument of the
first exponential in (11.17) effectively adds a phase to the carrier, ω0, that depends on (∆t)2. This
is the “chirping” that is referred to in the next section.
All of these effects are linear and so they can be undone by a similar but opposite effect.
Further, the order in which the effects occur is unimportant. A dispersive broadening may be
canceled by an opposite dispersion.

11.4 Chirping
A pulse, consisting of an envelope over a carrier, may be subjected to a modification, by
passing through a crystal modulator for example, in which the phase of the carrier is gradually
varied throughout the length of the pulse. If this variation is a linear function of time then the
effect is just as though the frequency of the carrier had been changed. There is little other effect.
However if the phase is changed as a quadratic function of time then it is as though the
frequency of the carrier were shifted gradually throughout the length of the pulse[2]. The pulse
with sliding frequency is said to be chirped.

cos (ω t + at 2 ) = cos ⎡⎣(ω + at ) t ⎤⎦ (11.22)

has frequency (ω+at). This chirped pulse appears indistinguishable from a short pulse that has
been dispersion broadened, except that the apparent dispersion can be opposite in sign to normal
dispersion. The pulse can then be subjected to the action of a dispersive medium where there is
significant group velocity dispersion. Provided this dispersion is of the correct magnitude and
sense then it will undo the artificially induced effect in the pulse leaving it considerably
narrowed. Various components have been used for this purpose but the flexibility of optical
coatings makes them particularly attractive in this application [3] [5].

11.5 Optical Coatings - Phase Change


Optical coatings affect both the amplitude and the phase of incident light. They can therefore,
in principle, make the kinds of adjustments to incident light that we have been considering. They
have an advantage over dispersive systems in that the correction is made immediately. We first
must consider the nature of the effect that thin film coatings have on the pulse.
Amplitude reduction over part of the range of frequencies leads to pulse broadening because
the narrower the frequency spectrum the broader is the pulse. We therefore limit ourselves to
consideration of those systems that have flat performance in terms of either transmittance or
reflectance and that make adjustments to the phase.
The sign convention is important. We use the normal thin film convention[6]. The coordinate
system has its origin at the surface where the reflection is said to be taking place and the phase
shift is measured at that surface. The electric field retains its incident positive direction. An
incident wave E cos (ω t − κ z + ϕ inc ) suffers a phase change ϕ ref at the surface z = 0 . The
electric field at that surface for the reflected beam therefore becomes E cos (ω t + ϕ inc + ϕ ref ) .
This then forms a reflected beam that has expression E cos (ω t + κ z + ϕ inc + ϕ ref )
Macleod: Optical Coatings Page 157 Design through Manufacture

The returned beam is now propagating along the negative direction of the z -axis. We can
rename the negative z -axis as the positive direction of the x -axis to give

E cos (ω t − κ x + ϕ inc + ϕ ref ) (11.23)

11.6 Group Delay and Group Delay Dispersion


Now let us examine the effects of the various phase angles on the pulse and its components.
We take equations (11.6) and we rewrite the left-hand side to include a change of phase on
reflection. We take the value of x as zero and so we are evaluating the functions at the surface of
the coating. We assume that the value of x, representing distance already traveled, can be given
by a distance L and so we replace it by that.

dκ 1 2 d κ
2
dϕ 1 2 d ϕ
2
ω t − κ L + ∆ω ⋅ t − ∆ω ⋅ L − ( ∆ω ) ⋅ L + ϕ 0 + ∆ω ⋅ + ( ∆ω ) ⋅ (11.24)
dω 0 2 dω 2 0 dω 0 2 dω 2 0

⎛ dκ dϕ ⎞ 1 2⎛ d κ d 2ϕ ⎞
2
= (ω t − κ L ) + ∆ω ⋅ ⎜ t − L+ ⎟ − ( ∆ω ) ⎜ L − ⎟⎟
⎜ dω 2
⎝ dω 0 dω 0 ⎠ 2 ⎝ 0
dω 2 0⎠

This expression is usually rewritten so that the first derivative of phase has a negative sign.

⎡ dκ ⎛ dϕ ⎞ ⎤ 1 2⎛ d κ d 2ϕ ⎞
2
(ω t − κ L ) + ∆ω ⋅ ⎢t − L −⎜− ⎟⎥ − ( ∆ω ) ⎜
⎜ dω 2
L − ⎟⎟ (11.25)
⎣ dω 0 ⎝ dω 0 ⎠ ⎦ 2 ⎝ 0
dω 2 0⎠


Now − has units of time and we can identify it as equivalent in its effect to the group delay
dω 0
due to dispersion and it is therefore known as the Group Delay, sometimes abbreviated to GD.
d 2ϕ
The units are time. The next term shows the equivalence in effect of − and the Group
dω 2 0
Velocity Dispersion. Since the negative first derivative is known as Group Delay this second
derivative is known as Group Delay Dispersion, abbreviated to GDD, and has units of (time)2.
Although we have said little about it here, the third derivative is sometimes called the Third
Order Dispersion, with units of (time)3, and abbreviated to TOD. Third Order Dispersion is
usually small but if it is significant, it can adversely affect the shape of the pulse.

11.7 Chromatic Dispersion


Chromatic dispersion in an optical medium refers to the change in group velocity with respect
to wavelength. Like Group Velocity Dispersion, the effect is to broaden a pulse.
It is tempting to write the phase expression (11.5) in terms of wavelength rather than
frequency and then to expand as a Taylor series in ∆λ rather than ∆ω. This, however, leads to
Macleod: Optical Coatings Page 158 Design through Manufacture

quite clumsy and involved expressions since a zero value of d2κ/dλ2 does not imply zero Group
Velocity Dispersion and zero pulse broadening. We therefore start with equation (11.10) and
translate it into terms of wavelength. We will do this in several steps.

⎛ 2π c ⎞ 2π c
∆ω = ∆ ⎜ ⎟ = − 2 ⋅ ∆λ (11.26)
⎝ λ ⎠ λ

d 2κ d ⎡ d ⎛ 2π n ⎞ d λ ⎤ d λ
= ⎜ ⎟⋅ ⋅
dω 2
d λ ⎢⎣ d λ ⎝ λ ⎠ dω ⎥⎦ dω
d ⎡ ⎛ 2π dn 2π n ⎞ ⎛ λ 2 ⎞ ⎤ ⎛ λ 2 ⎞
= ⎢⎜ ⋅ − ⎟⋅ − ⎥⋅ −
d λ ⎣ ⎝ λ d λ λ 2 ⎠ ⎜⎝ 2π c ⎟⎠ ⎦ ⎜⎝ 2π c ⎟⎠
(11.27)
d ⎡ ⎛ λ dn n ⎞ ⎤ ⎛ λ 2 ⎞
= ⎜ ⋅ − ⎟ ⋅
d λ ⎢⎣ ⎝ c d λ c ⎠ ⎥⎦ ⎜⎝ 2π c ⎟⎠
λ 3 d 2n
= ⋅
2π c 2 d λ 2

Then, substituting in (11.10), we obtain

d 2κ
∆t = ∆ω ⋅ ⋅z
dω 2
⎡ λ d 2n ⎤
∆t = − ⎢ ⋅ 2 ⎥ ⋅ ∆λ ⋅ z (11.28)
⎣ c dλ ⎦
= − D ⋅ ∆λ ⋅ z

and D is known as the Chromatic Dispersion Parameter. In a fiber or waveguide, group velocity
is often given in terms of the Group Effective Index, ng. This is defined in much the same way as
the normal refractive index.

ng = c vg (11.29)

where vg is the group, or pulse propagation, velocity. Then D is given in terms of ng as

1 dn
D=− ⋅ g (11.30)
c dλ

that can also be written


Macleod: Optical Coatings Page 159 Design through Manufacture
d ⎛1⎞
D=− ⎜ ⎟ (11.31)
d λ ⎜⎝ vg ⎟⎠

D has dimensions of time/(distance×wavelength) usually units of ps km-1 nm-1.


When the effects of a coating are included then equation (11.25) is the relevant one. The
definition for ∆t involving Group Velocity Dispersion, (11.10), is replaced now by (11.28), but
we also need to consider a suitable replacement for the Group Delay Dispersion, -d2ϕ/dω2.

⎛ d 2κ d 2ϕ ⎞
∆t = ⎜ ⋅ z − ⎟ ⋅ ∆ω
⎝ dω dω 2 ⎠
2

⎛ d 2κ d 2ϕ ⎞
=⎜ ⋅z− ⎟ ⋅ ∆ω
⎝ dω dω 2 ⎠
2

(11.32)
⎛ d 2ϕ dω ⎞
= −⎜ D⋅ z + ⋅ ⎟ ⋅ ∆λ
⎝ dω 2 d λ ⎠
⎡ ⎛ d 2ϕ 2π c ⎞ ⎤
= − ⎢D ⋅ z + ⎜ − ⋅ 2 ⎟ ⎥ ⋅ ∆λ
⎝ dω λ ⎠ ⎦
2

The second term inside the brackets is just

d 2ϕ 2π c 2π c
− ⋅ = 2 ⋅ GDD (11.33)
dω 2 λ 2 λ

Its sign is such that it simply adds to the chromatic dispersion parameter times the propagation
distance. It is known as the Chromatic Dispersion Coefficient, abbreviated to CDC with units of
time/length usually ps/nm.

2π c d ⎡ dϕ d λ ⎤ d λ
CDC = − ⋅ ⋅ ⋅
λ 2 d λ ⎢⎣ d λ dω ⎥⎦ dω
(11.34)
λ ⎡ dϕ d 2ϕ ⎤
=− 2 ⋅ + λ ⋅
2π c ⎢⎣ d λ d λ 2 ⎥⎦

If ∆λ is such that the range of wavelengths in the pulse spectrum is from λ-∆λ to λ+∆λ then
the time ∆λ×CDC will be equivalent to the time ∆ω×GDD, of course with ∆λ and ∆ω in the
correct units.
Chromatic Dispersion Coefficient will often be associated with multiple-cavity filters in
Wavelength Division Multiplexing. The default time units are femtoseconds (fs), consistent with
ultrafast phenomena. For narrow-band filters picoseconds (ps) is a more convenient and usual
unit.
Macleod: Optical Coatings Page 160 Design through Manufacture

Now let us look at the relationship between CDC and GDD using, as an example, the three-
cavity narrow-band filter:

Air | (HL)8 HH (LH)8 L(HL)8 H 6L H(LH)8 L(HL)8 HH (LH)8 L | Glass

where H indicates a quarterwave of Ta2O5 and L of SiO2.


Units are important in conversion factors. 2πc/λ2 has dimensions of 1/(length×time), but the
units must be consistent with those of GDD and CDC, ps2 and ps/nm, respectively. At 1550nm
the conversion factor is then

2π c 2π × 2.998 × 108 × 109 ÷ 1012 nm / ps


=
λ2 15502 nm 2 (11.35)
= 0.7841 ( nm ⋅ ps )
−1

The factor varies with wavelength, but it is confirmed when the plots of Figure 11-4 are
compared.
Note that at a pulse rate of 10 Gb s-1 the pulse length for a mark/space ratio of 1.0 is 50ps. For
the onset of serious pulse spreading the Group Delay Dispersion should be around 2500 ps2. The
filter used in the calculations of Figure 11-4 would have negligible effect on such pulses.
However, at 50 Gb s-1 the effect is becoming much more important.

Transmitta nce Chromatic Dispersion (ps/nm) Transmitta nce Group De la y Dispersion (ps^2)
100 150
100
50
50
0 0
-50
-50
-100
-100 -150
1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
Wa velength (nm) Wave le ngth (nm)

Figure 11-4. The Chromatic Dispersion Coefficient (upper plot) compared with the Group Delay Dispersion
(lower plot) for the three-cavity narrow-band filter.
To calculate τg2 from the Chromatic Dispersion Coefficient, it should be multiplied by
2
λ /(2πc) in the appropriate units. This is a somewhat tedious calculation. For calculations
involving τg2 it is better to use the Group Delay Dispersion. However, for an estimate of the
pulse spreading where ∆λ is known then the Chromatic Dispersion Coefficient is more useful.
This is often the case when optical fiber communication is involved.
Macleod: Optical Coatings Page 161 Design through Manufacture

11.8 References
[1] Saleh, B.E.A. and M.C. Teich, Fundamentals of Photonics, First ed., Wiley Series in
Pure and Applied Optics, ed. J.W. Goodman, John Wiley and Sons Inc, New York,
1991.
[2] Yariv, A. and P. Yeh, Optical Waves in Crystals, First ed., Pure and Applied Optics, ed.
J.W. Goodman, John Wiley & Sons, New York, 1984.
[3] Ferencz, K. and R. Szipocs, “Recent developments of laser optical coatings in Hungary,”
Optical Engineering, 32(10), 2525-2538 (1993).
[4] Stingl, A., et al., “Generation of 11-fs pulses from a Ti:sapphire laser without the use of
prisms,” Optics Letters, 19(3), 204-206 (1994).
[5] Szipöcs, R., et al., “Chirped multilayer coatings for broadband dispersion control in
femtosecond lasers,” Optics Letters, 19(3), 201-203 (1994).
[6] Macleod, H.A., Thin-film optical filters, Second ed., ed. Adam Hilger, Bristol, 1986.
[7] Schott Optical Glass, Schott Glass Technologies Inc, Duryea, Pennsylvania, 1992.
[8] Gires, F. and P. Tournois, “Interféromètre utilisable pour la compression d’impulsions
lumineuses modulées en fréquence,” C R de l’Acad Sci, Paris, 258, 6112-6115 (1964).
[9] Kuhl, J. and J. Heppner, “Compression of femtosecond optical pulses with dielectric
multilayer interferometers,” IEEE Transactions on Quantum Electronics, QE-22(1), 182-
185 (1986).
Macleod: Optical Coatings Page 163 Design through Manufacture

12 PERFORMANCE ENVELOPES
An envelope is a useful concept in many aspects of optical coating design and performance
analysis. Envelopes represent performance limits and they can often be calculated very rapidly.
Their original usefulness was very much in economy of calculation and they have become less
used as computing has advanced. A major current application is in the various envelope methods
of n, k and d extraction but they can be valuable in areas well beyond n and k particularly in
connection with multilayers. In this section we look at the Performance Envelopes of a chosen
individual layer in a design.
Let us consider a multilayer optical coating which contains at least one dielectric, or lossless,
layer. If we choose any dielectric layer in this assembly we find that as we vary its thickness the
transmittance of the assembly varies between two definite limits. These two limits are the
performance envelopes. If the multilayer is entirely dielectric then the complement of the
performance envelopes gives the reflectance information. If the multilayer is absorbing then
there are no reflectance envelopes.
Performance envelopes are much more informative and useful than the concept of error
sensitivity. Error sensitivity is an attempt at estimating the degree of caution required in the
deposition of a particular layer. It is normally calculated as the derivative of transmittance (or
reflectance) with respect to the thickness of a particular layer and it is usually shown as a
function of wavelength. Unfortunately it is useful only for vanishingly small changes in layer
thickness. Larger changes include nonlinearities that demand the inclusion of higher-order terms.
Errors committed in practice usually exceed by orders of magnitude the limits of linearity in the
relationship between response and layer thickness. Since the performance envelopes mark the
limits of performance whatever the variation in thickness of the chosen dielectric layer they
represent the limits of performance for any possible thickness error, no matter how large, in the
given layer. We shall see that with the addition to the envelopes of the Round-Trip Phase
Change in the layer the effect on the performance of an error of virtually any magnitude
whatsoever can quickly be visualized.
Fabry and Perot described their famous interferometer in 1899[1]. The theory of multiple-
beam fringes was not new even then, dating back at least to Poisson[2] 70 years earlier. It was
Poisson who first pointed out that multiple-beam interference in a single thin film was
responsible for the high transmittance and vanishing reflectance in a halfwave layer[2]. Since
then, the Poisson ideas have been used and elaborated by very many workers. Smith [3] adapted
the technique for multiple-cavity filter design. Musset and Thelen [4] used it for antireflection
coatings. We use it to derive the performance limits, the envelopes.
Macleod: Optical Coatings Page 164 Design through Manufacture

Ta Rb
ϕb

ϕa T
Incident
light Ra Tb

System a System b
Figure 12-1. Coupling layer between two structures. The symbols R and T indicate reflectance and
transmittance while ϕ is the phase shift on reflection at the appropriate surface.
We imagine two structures, System a and System b, each exhibiting reflectance, transmittance
and absorptance placed on either side of a coupling layer, Figure 12-1. The coupling layer
supports an infinite number of beams that are reflected back and forth between the two systems.
The transmittance of the overall assembly is given [5] by expression (12.1) where the symbols
are defined in Figure 12-1.
TaTb 1
T= × (12.1)
⎡1 − ( Ra Rb ) ⎤
12 2
⎡ ⎤
⎢1 + 4 ( Ra Rb )
12
⎣ ⎦ 2 ⎛ ϕ a + ϕb ⎞⎥
⎢ ⋅ sin ⎜ − δ ⎟⎥
12 2 2
⎡ ⎤ ⎝ ⎠
⎢⎣ ⎣1 − ( Ra Rb ) ⎦ ⎥⎦

The reflectance of the assembly can be represented by a somewhat more complex expression
but there is an additional dependence on the outer reflectance of System a and on the phase
relationship between the primary beam reflected at that surface and the resultant beam
reemerging through it that makes it difficult to draw general conclusions and, in fact, usually
prevents the expression of performance envelopes at all. In the case of dielectric systems,
reflectance is simply the complement of equation (12.1).
From (12.1) we can show that the transmittance of the combined system with a lossless
coupling layer must lie between two envelopes given by
TaTb
Tmax/min = 2
(12.2)
⎡1 ∓ ( Ra Rb )1 2 ⎤
⎣ ⎦

with the negative sign in the denominator corresponding to the maximum envelope. The
transmittance must always lie between these limits and it is the round-trip phase change that
determines the exact value. If the phase change is an even integral multiple of π then the
transmittance must coincide with the maximum and, if an odd multiple, with the minimum.
To draw the envelopes and the round-trip phase change we introduce a cut of zero thickness in
the chosen layer. This cut is assumed to be filled with material identical to the given layer but it
is of zero thickness. An increase in thickness of the given layer implies a subtraction of a
Macleod: Optical Coatings Page 165 Design through Manufacture

corresponding quantity from the round-trip phase change so that the curve slips downwards. A
decrease in thickness will slide the curve upwards. The crossing points of the curve with the zero
phase axis mark those points where the actual performance curve will touch the maximum
envelope. The slope of the round-trip phase curve at the axis crossing indicates the sensitivity of
that particular feature to layer thickness. When the slope is very steep the change in feature
position with layer thickness change will be small.
Let us take first the example of a 43-layer edge filter with starting design
Air | (0.5HL0.5H)21 | 1.52 with L = 1.45, H = 2.45 where the outermost five layers on either side
have been refined to reduce ripple in the long-wave pass region. Let us split this filter into two
parts at the inner edge of layer 21 counting from the air incident medium. This is a high-index
layer virtually at the center of the stack. We will also make the cut in the filter of zero thickness,
initially, so that δ in (12.1) is zero. The maximum and minimum envelopes of performance
relative to layer 21 are shown in Figure 12-2 and the round-trip phase change in the zero-
thickness cut is plotted in Figure 12-3.
Only when the round-trip phase change is in the immediate vicinity of 0 or 2mπ, where m is
an integer, does the coating performance match the maximum envelope and when the gap
between the maximum and minimum envelopes is large, as in the center of the rejection zone,
then the performance will tend to be close to the minimum except when the 0 or 2mπ condition
is very closely satisfied. Figure 12-3 shows a round-trip phase change that is far away from zero
throughout the high-reflectance zone and, accordingly, the minimum envelope marks the
performance in the rejection region.

Figure 12-2. The performance envelopes for layer 21 in the 43-layer edge filter. The theoretical and unaltered
performance of the filter is also shown.
Macleod: Optical Coatings Page 166 Design through Manufacture

Figure 12-3. The round-trip phase change in the cut inside layer 21.
Now let us reduce the thickness of layer 21 where the envelopes are calculated. This will
introduce a non-zero δ into the round-trip phase change. Since δ is subtracted in the expression
for phase an increased thickness will move the phase curve downwards while a reduced
thickness will move it upwards. Reducing layer 21 thickness, therefore, will introduce a zero into
the right-hand side of the rejection region. We can see this clearly in Figure 12-4 where a narrow
leakage peak has appeared at the right-hand side of the rejection zone. The peak of the leak lies
exactly on the upper envelope.
The envelopes can also be plotted for any layer in the system and also in terms of optical
density. Figure 12-5 shows the envelopes for layer 11. The maximum envelope drops down more
in the center than that in Figure 12-2 but the minimum envelope is similar and the round-trip
phase change across the rejection zone (not shown) is indistinguishable from that of Figure 12-3.
Figure 12-5 also shows that increasing the thickness of layer 11 pulls a narrow leak across from
the left-hand side of the high reflectance region, as we would expect.
Macleod: Optical Coatings Page 167 Design through Manufacture

Figure 12-4. The layer 21 envelopes together with the performance when the optical thickness of the layer is
changed from 0.25 to 0.15.

Figure 12-5. The envelopes in terms of optical density corresponding to layer 11 from the air incident
medium. The black curve shows the performance with layer 11 set to an optical thickness of 0.4 rather than
0.25. The leakage peak on the left-hand side of the rejection region marks the round-trip phase change of
zero that has moved to the right as a consequence of the thickness change.
The remaining figures show some results for a three-cavity narrowband filter consisting of 91
layers. Here the rate of change of round-trip phase varies enormously with the position of the
particular layer, the cavity layers having almost no slope across the transmission region at all.
Small shifts in thickness of the cavity layers are therefore very damaging to the performance.
Shifts in the thickness of the coupling layers mainly distort the passband shape.
Macleod: Optical Coatings Page 168 Design through Manufacture

Figure 12-6. Performance envelopes for a coupling layer in a three-cavity narrowband filter. Note the very
high transmittance at the central wavelength for the minimum envelope. The black curve shows the distorted
passband shape when the coupling layer optical thickness is changed from 0.25 to 0.35.

Figure 12-7. The envelopes for a 101-layer three-cavity narrow band filter corresponding to a layer near the
center of the outermost reflecting stack (layer 9). The minimum envelope is quite low. The ideal performance
of the filter is shown.
Macleod: Optical Coatings Page 169 Design through Manufacture

Figure 12-8. A shift in thickness of layer 9 from 0.25 to 0.26 shows a considerable drop in transmittance and
a shape distortion. Note that the increase in thickness will depress the layer 9 phase characteristic in Figure
12-9 making the zero shift to the right corresponding to the result in this figure.

Figure 12-9. The round-trip phase change for various layers in the 91-layer three-cavity filter. The cavity
layer is very close to zero over the whole range so a small change in δ will have a severe effect. It will be
slightly less severe for layer 9. For the coupling layer three zeros remain over a quite large range of δ but a
shift in the outer two will cause a distorted pass-band. Larger changes δ will give only one zero but the filter
will still show a central peak because of the shape of the coupling-layer minimum-envelope shape.

12.1 References
1. Fabry, C and A Perot, Theory et applications d'une nouvelle méthode de spectroscopie
interférentielle. Ann Chim Phys, Paris, 7th series, 1899. 16: p. 115-144.
Macleod: Optical Coatings Page 170 Design through Manufacture

2. Knittl, Z, Fresnel historique et actuel. Optica Acta, 1978. 25: p. 167-173.

3. Smith, S D, Design of multilayer filters by considering two effective interfaces. Journal of


the Optical Society of America, 1958. 48: p. 43-50.
4. Musset, A and Thelen, Multilayer antireflection coatings, in Progress in Optics, E. Wolf,
Editor. 1966, North Holland: Amsterdam. p. 201-237.
5. Macleod, H A, Thin-Film Optical Filters. Second ed. 1986, Bristol: Adam Hilger.
Macleod: Optical Coatings Page 171 Design through Manufacture

13 COATING MANUFACTURE
There is a considerable number of processes that can be and are used for the deposition of
optical coatings. The commonest processes take place under vacuum and can be classified as
physical vapor deposition. In these processes, the thin film condenses directly in the solid phase
from the vapor. The word “physical” as distinct from “chemical” is intended to indicate the
absence of any chemical reactions in the formation of the film. This is an oversimplification.
Chemical reactions are, in fact, involved but the term chemical vapor deposition (CVD) is
reserved for a family of techniques where the growing film differs substantially in composition
and properties from the components of the vapor phase. The physical vapor deposition processes
can be classified in various ways but the most useful classifications for our purposes are based
on the methods used for producing the vapor and on the energy that is involved in the deposition
and growth of the films.

13.1 Thermal Evaporation


Vacuum, or thermal, evaporation has for years been the principal physical vapor deposition
process and because of its simplicity, its flexibility and its relatively low cost, and because of the
enormous number of existing deposition systems it is likely so to continue for some considerable
time. It is, however, clear that it possesses major shortcomings, especially in respect of the
microstructure of the films, and, particularly for high-performance specialized coatings,
alternative processes, such as sputtering, are being adopted. In thermal evaporation, the material
to be deposited, the evaporant, is simply heated to a temperature at which it vaporizes. The vapor
then condenses as a solid film on the substrates, which are maintained at temperatures below the
melting point of the evaporant. Molecules travel virtually in straight lines between source and
substrate and the laws governing the thickness of deposit are similar to the laws that govern
illumination.
The details of a typical thermal evaporation plant are sketched in Figure 13-1. Although there
could be considerable freedom in placing sources and substrates, it is usual to mount sources low
in the chamber and evaporate upwards onto the substrates. Gravity holds the charge in place in
the source and the substrates are protected to some extent from any rubbish that may fall to the
bottom. The material to be evaporated is held in a crucible and heated, either indirectly by
radiation from a hot filament, directly by passing a current through the crucible, known in that
case as a boat and made of refractory metal, or by electron bombardment, the preferred technique
nowadays. Here the crucible is copper and water-cooled. The charge in the crucible is
bombarded with a beam of electrons, usually around 6 to 10 kV with beam currents of up to
several amps. The material in contact with the cooled crucible remains unmelted and effectively
forms a container that has no adverse reaction with the molten material.
Metal oxide films were long known to be very tough and durable but difficult to produce
partly because of the very high temperatures required to evaporate them, and partly because they
tend to lose oxygen so that the films are absorbing. The electron beam sources made it possible
to evaporate even very high melting point materials with relative ease, and it has been found
possible to make up the oxygen deficiency by a process known as reactive evaporation. In this
technique, a slight residual atmosphere of oxygen is maintained within the plant by bleeding
Macleod: Optical Coatings Page 172 Design through Manufacture

oxygen through a needle valve. The extra oxygen molecules that bombard the substrate replace
much of the oxygen lost by the material during evaporation, and the resulting films exhibit only
slight residual absorption. Heitmann[1] showed that the absorption could be still further reduced
if the residual oxygen were ionized, or activated, by passing it into the chamber through an
electrical discharge. This variant of the process is known as activated reactive evaporation and
has been developed to a high degree of perfection by Ebert[2].

Figure 13-1. A typical coating plant for thermal evaporation.


The emission from a source is a function of the direction. A reasonable approximation in
many cases is a simple cosine law, although some sources can be more directional, particularly
the electron-beam source. Thus the thickness of film built up on a flat stationary substrate will
vary usually to a degree which cannot be tolerated in any but the simplest coatings. Increasing
the dimensions of a coating plant to the point where the variation in the thickness of the deposit
can be neglected is not practical and so a system of rotating jigs has been devised to improve
uniformity. For many purposes a simple jig rotating about the center of the plant above an offset
source is sufficient. For applications where uniformity is of major importance an arrangement
known as planetary, Figure 13-2, in which small substrate carriers rotate rapidly about their own
axes while rotating more slowly about the center of the plant, is more common.
For increased durability, substrates are frequently heated, almost invariably by radiation either
from behind the substrates, Figure 13-1, or, more often today, from in front, Figure 13-2.
Temperature measurement is still a difficult problem in optical coating. Frequently a
thermocouple probe is simply placed somewhere in the plant near the substrates. A more modern
method that is gaining in popularity uses a remote sensing infrared thermometer. It is clear that
substrate temperatures may vary considerably over the jig from one substrate to another and
change significantly during the deposition of even one single layer. Variations of 30°C from one
substrate to another are not uncommon and Hacman[3] quotes a 15°C rise in the temperature of a
glass substrate, held nominally at 300°C, due solely to radiation from the vapor source during the
Macleod: Optical Coatings Page 173 Design through Manufacture

deposition of one single quarterwave of MgF2. Substrate temperature is a particularly important


parameter and there is a definite need for more precise measurement and control techniques.

Input
fiber

Monitor
Planetary chip To pumping
substrate carrier system
Radiant Radiant
heater heater

E-beam Boat
source source

Output
fiber

Figure 13-2. Arrangement of thermal evaporation including planetary work holder. In this configuration
front-surface heaters are shown.

13.2 Critical Areas


Coatings consist of a succession of dielectric or metal layers on a substrate. Both refractive
index and thickness of each layer are of importance in controlling the amplitudes and phases of
the various beams involved in the interference that determines the overall characteristic, and
therefore must be achieved with precision in manufacture. The greater the number of different
materials in a design, the more complex its manufacture and coating designs attempt to limit the
total number of materials as far as possible, ideally to two but sometimes three or even more
different materials may be necessary.
Optical thickness is traditionally the parameter that is controlled during deposition. Refractive
index is difficult to measure and therefore to control and so the assumption is made that it will be
characteristic of the particular material being deposited. To some extent this is true but it is only
an approximation. The index achieved with a particular film composition can still vary within
wide limits depending on many deposition parameters. This will be considered later. Optical
thickness is controlled simply by stopping deposition when the correct value is reached. No
satisfactory way of removing already-deposited material has yet been devised. Two principal
techniques exist for the measurement of thickness, optical methods and the measurement of mass
by the quartz crystal microbalance.
Macleod: Optical Coatings Page 174 Design through Manufacture

The commoner technique is optical monitoring, and this is suitable for all optical coatings. In
optical monitoring the reflectance or transmittance, or sometimes both, of the film are measured.
The film may be on a separate test glass, known as indirect monitoring, or may be on an actual
component forming part of the batch to be coated, known as direct monitoring. One single
wavelength may be used or a range of wavelengths. Sometimes the layers are simply observed
visually and the color of light reflected is used as an indication of layer thickness. The
reflectance of a single layer varies in a sinusoidal fashion with thickness, passing through an
extremum every quarter wave and an especially simple method of monitoring the thickness is
simply to terminate deposition at a turning value when the layer will be an integral number of
quarterwaves thick. More demanding designs may require the use of many wavelengths
simultaneously with absolute transmittance and/or reflectance measurements
The quartz crystal microbalance measures the natural frequency of a thin slice of quartz (AT
cut) vibrating in a shear mode. As a film is deposited on the surface of the quartz, the resonant
frequency is reduced and the reduction is a measure of the mass of the film. Provided all the film
parameters remain constant then the mass may be taken as an indication of optical thickness. The
quartz crystal microbalance has been used with great success in low to medium accuracy
applications, particularly where repeat runs of the same component are being performed. It is
especially useful when layer thicknesses are small fractions of a quarterwave. A further
advantage of the quartz crystal is that the signal is suitable for use by automatic control systems
with very little alteration. The optical signal oscillates between the limits associated with quarter
and halfwaves and these limits vary with the admittance of the layers already deposited on the
monitoring substrate, making the task of interpretation still more difficult. The signal from the
quartz crystal monitor simply increases with total coating thickness. Antireflection coatings and
edge filters are commonly produced using quartz crystal monitoring but narrowband filters
require the greater wavelength precision of optical techniques.
The production of thin-film coatings starts with the substrate. High quality coatings cannot be
produced on a poor substrate. Substrate defects invariably lead to coating defects, usually
enhanced. Cleanliness is of major importance. The forces that hold thin films onto a substrate are
all short-range forces between one atom and the next. Thus even a layer of contamination which
is only one molecule thick will completely change the character of the adhesion of the film.
Cleaning procedures are critical. The nature of many optical coatings is such that no
subsidiary coatings can be added for protection. Thus, optical coatings to be satisfactory must
not only satisfy optical performance criteria but must also be capable of withstanding
environmental disturbances typical for the particular application. Considerable effort has been
devoted to improving resistance of optical coatings to humidity, abrasion and chemical attack.
Since a typical quarterwave layer for the visible region will be around 100nm (approximately
four microinches), exceedingly thin compared with coatings designed specifically for surface
protection and wear resistance, it is clear that we should not expect too much in environmental
resistance. Normal environmental conditions are variable and predictable only in statistical
terms. Therefore environmental resistance of coatings is usually measured by subjecting them to
certain specified test environments, typical specifications including such treatments being MIL-
C-675, MIL-C-14806, MIL-C-4888497 and MIL-M-13508. A new international (ISO)
specification is about to be issued for optical coatings that is expected to be more suitable,
particularly for commercial coatings, that these military specifications that were not necessarily
Macleod: Optical Coatings Page 175 Design through Manufacture

directed specifically towards coatings or, in some cases, were originally intended only to cover
specific coating types.
In the production process, there are many parameters that are important in determining the
properties of the coating. Only a few of these parameters can be controlled in closed loop. The
remainder must be controlled simply by maintaining all process conditions as constant as
possible. Each coating run must be carried out in exactly the same way, at the same rate and
under the same conditions as the previous run of the batch. Test runs are important and are
always performed when any changes, even apparently minor ones, have been made either to the
plant or to the For a number of reasons, the real performance of thin-film optical devices will
always fall short of what is theoretically possible. The layers will never be deposited with
precisely the correct values of thickness and refractive index. This is a question of manufacturing
accuracy and allowable tolerances. Then there are defects in the layers themselves - that is, they
depart from the ideal parallel-sided homogeneous slab, usually assumed in design calculations.
The field of manufacturing tolerances is a difficult and involved one, in which many factors
play a part, and, although much work has been published on this subject, there is a great deal to
be done, and manufacturers still depend, to a large extent, on experience and subjective
judgment. Purely analytical methods of assessing tolerances have never been entirely successful,
because the complex interaction between errors in one layer with those in another, forces the
adoption of low-order approximations, valid only for very small errors, which are usually an
order of magnitude below those occurring in practice. The most fruitful method of assessing
tolerances and the effects of errors, is that of computer simulation. If a model of the appropriate
monitoring process is set up on a digital computer, it can be used in a Monte Carlo simulation of
the production of a series of batches of coatings with appropriate random errors included.
Examination of the series of characteristics produced for the coatings permits a rapid assessment
of the yield of the process in the presence of errors of a particular magnitude. A series of such
trials with varying error magnitude can be used to determine the permissible error levels for the
coating design in question. Studies such as this have shown that the tolerances in production are
not independent of the monitoring procedure that is to be used. Indeed, it is more meaningful to
talk in terms of the errors in the monitoring signal itself, rather than the errors in layer optical
thickness that are produced. In particular, there is a vast difference in optical monitoring between
techniques in which all layers are monitored on one single substrate and those in which a series
of test glasses are used. In single-wavelength monitoring, the former technique, known as direct
monitoring, has been shown to include an interaction between earlier errors and later ones which
can be advantageous, leading to great precision in peak wavelength centering in the case of
narrow band filters for example, and disadvantageous, in the case of coatings where high
performance over an extended spectral region is required, such as antireflection coatings.
It should not be forgotten that uniformity of layer thickness can cause problems quite apart
from those of random monitoring errors. As has already been said, rotation of the substrates
during deposition is almost always required. There can be severe problems when deeply curved
substrates are concerned, and these usually imply the design and construction of special jigs. The
extreme case of uniformity is probably narrow-band filters where uniformity errors should not
cause shifts over the surface of the filter greater than 0.3xhalfwidth. This is straightforward for
filters up to 50mm diameter with halfwidths of not less than 2nm in the visible region, but for
filters which are larger in diameter or narrower in bandwidth there can be problems and great
attention must be paid to source positioning and smoothness of rotation.
Macleod: Optical Coatings Page 176 Design through Manufacture

Layer structure is the most significant factor in determining the properties of thin films and
the way in which they differ from the same material in bulk form. We shall say more about this
microstructure and the way in which it determines the behavior of thin films. It is a columnar one
and the details are sensitive functions of deposition conditions and especially of substrate
temperature
Substrate temperature is a difficult parameter to measure and to control so that consistency in
technique, heating for the same period each batch, identical rates of deposition, pumping for the
same period before commencing deposition and so on, is of major importance in assuring a
stable and reproducible process. Changing the substrate dimensions, especially substrate
thickness, from on run to the next can cause appreciable changes in film properties. Such
changes are even more marked in the case of reactive processes where the residual gas pressure
is raised, and where a reaction between evaporant and residual atmosphere takes place at the
growing surface of the film. Thus it should not be surprising that a very high proportion of test
runs are required in any manufacturing sequence.
The variation in refractive index is not the only feature of film behavior associated with the
columnar structure. Moisture adsorption is a principal source of film instability and it is the
columnar structure that facilitates the entry of moisture into the multilayers. The
moisture-induced drift of the filters towards longer wavelengths that occurs on exposure to the
atmosphere is a well-known problem, frequently referred to as aging or settling.
Changes in temperature cause changes in the spectral characteristics of coatings, and narrow-
band filters are probably those most affected. For small changes in temperature, the principal
effect is a shift towards longer wavelengths with increasing temperature. For the materials
commonly used in the visible region of the spectrum, the shift is of the order of 0.003% per °C,
while for infrared filters it can be greater, and a useful figure is 0.005% per °C, although it can
be as high as 0.0125% per °C. It must be emphasized that these figures depend strongly on the
particular materials used. Filters of lead telluride and zinc sulfide can actually have negative
coefficients greater than 0.01% per °C and, using these materials, it is even possible to design a
filter which has zero temperature coefficient. With greater positive changes of, say, 60°C or
more, it is usual for the moisture in the filter to desorb partially, causing an abrupt shift towards
shorter wavelengths. We shall return to this later. Exposure to higher temperatures still, over
100°C, can cause permanent changes often related to structural alterations, sometimes rendering
layers less ready to adsorb moisture but sometimes leading to catastrophic failures.
One of the more amazing features of coatings produced by the energetic processes is their
exceptionally low coefficients of spectral shift with temperature. In many reported cases these
shifts are virtually zero over ranges as wide as 100ºC.
Coatings that are subjected to very low temperatures usually shift towards shorter
wavelengths, consistent with their behavior at elevated temperatures. Filters are not usually
affected mechanically except for laminated components that run the risk of breaking due to
differential contraction and/or expansion.
There are losses associated with all layers, which can be divided into scattering and
absorption. In absorption, the energy, which is lost from the primary beam, is dissipated within
the coating and usually appears as heat. In scattering, the flux lost is deflected and reemerges
from the coating in a different direction. Absorption is a material property which may be
Macleod: Optical Coatings Page 177 Design through Manufacture

intrinsic or due to impurities. A deficiency of oxygen, for example, can cause absorption in most
of the refractory oxide materials. Scattering is usually due to defects in the coating, which can be
classified into volume or surface defects. Surface defects are simply a departure from the smooth
flat surfaces of the ideal film. Such departures can be due to roughness of the substrate surface
which tends to be reproduced at each interface in a multilayer, or to the columnar structure of the
layers which results in a nodular appearance of the film boundaries. Volume defects are local
variations of optical constants and are usually dust particles, pinholes or fissures in the coating.
Losses in thin films are of particular importance in the laser field where they determine the
limiting performance of multilayers. For low powers, the scattering losses are of greater
importance. For high powers, absorption losses appear to be of greatest importance in continuous
exposure to high flux, while defects, which can act to increase peak electric fields, are more
important in short-pulse applications.
A major problem in the production of high quality laser coatings is dust that emanates from
the sources and from the powdery deposit that forms on the cold walls of the chamber. If this
dust can be eliminated, only possible if the strictest attention is paid to detail and the most
involved precautions are taken, then the remaining source of scattering loss is the roughness of
the interfaces between the layers and between multilayer and substrate. If great care is exercised,
then, in the visible and near infrared regions, the total losses, that is, absorption and scattering,
can be reduced below 0.01%.
Once the losses become exceedingly small other factors such as grain boundaries and surface
and interfacial roughness become much more important. Coatings with ultimate performance in
terms of scattering are invariably amorphous. Ion-beam sputtered reflectors for the visible and
near infrared have been constructed[4] with total losses (absorption and scattering) as low as
0.0001%. The process used was ion-beam sputtering and the materials silica and tantala.
Laser damage is still very much a research topic with no universally accepted detailed models
of damage mechanisms. It seems clear that thermal effects associated with absorption, either
local or general, are the principal source of damage in continuous-wave applications. Small
defects appear less important. In pulsed-laser applications, damage incidence seems to be related
much more to the peak electric field to which the coating is subjected and the importance of
small, local defects in the coatings, appears clear In those spectral regions where water absorbs
strongly considerable importance is now being attached to the presence of liquid water within the
films. In other parts of the spectrum its role is less clear, but it may well play a part. The nodules
that are a common feature in many coatings are thought to play an important part and there are
some recent reports that are particularly relevant[5, 6].

13.3 Dielectric Thin Film Coating Materials


The list which follows is far from being complete but gives the more important properties of
some commonly used materials. Often the properties of a particular material appear to vary from
plant to plant and even from operator to operator. This is a consequence of the difficulties
inherent in assuring the exceedingly tight control that is necessary for stable, reproducible
results. Usually the best that can be done is to make certain that the results in any one particular
plant are consistent and to use calibration factors when transferring processes from one plant to
another.
Macleod: Optical Coatings Page 178 Design through Manufacture

The material probably used more than any others in thin-film work is magnesium fluoride.
This has an index of approximately 1.38 in the visible and is used extensively in lens blooming.
In the simplest case this is generally a single layer. Best results are obtained when the substrate
is heated during deposition, often to 300°C or so. Probably the easiest materials of all to handle
are zinc sulfide and cryolite. They have a good refractive index contrast in the visible, the index
of zinc sulfide being around 2.35 and that of cryolite around 1.35. Although these materials are
not particularly robust, they are so easy to handle and they have such high optical quality that
they are very much used, especially for narrow band filters for the visible and near infrared
where the layers can be protected by a cemented cover. For this purpose the substrates are not
usually heated. Cryolite has been found recently to be a useful ultraviolet material in the 200-
400nm region, usually with aluminum fluoride in high reflecting multilayers. Zinc sulfide is also
a staple material in the infrared especially in the far infrared from around 6µm to 25µm where it
is used as a low index rather than high index material The high index material is usually a
semiconductor like germanium or lead telluride. At shorter wavelengths silicon monoxide is
perhaps slightly more popular because of its slightly lower index and more rugged mechanical
properties. In the infrared the substrates should be heated for the best performance. 150°C is a
useful temperature and glow discharge cleaning just prior to evaporation is recommended[7].
A trick, which has sometimes been used with zinc sulfide to improve its durability, is
bombardment of the growing film with electrons. This can be achieved by positioning a
negatively biased hot filament, somewhere near the substrate carrier, in such a way that the
filament is shielded from the arriving evaporant, but is in line-of-sight of the substrates. This
process is still not entirely understood, but some work[8] suggests that an important factor is the
modification of the crystal structure of the zinc sulfide layers by electron bombardment.
Resistively heated boats produce a mixture of the cubic zinc blende and the hexagonal wurtzite
structure, while electron-beam sources produce purely the zinc blende modification. The
hexagonal form is a high temperature modification which, it is suspected, will tend to transform
into the lower temperature cubic modification, particularly when water vapor is present, a
transformation accompanied by a weakening of adhesion, and even delamination. Deliberate
electron bombardment of growing zinc sulfide films from boat sources, results in films with
entirely cubic structure and with the improved stability expected from that structure.
For more durable films in the visible region, use can be made of cerium dioxide[9] as the high
index layer and magnesium fluoride as the low index layer although they are not as popular as
certain other combinations to be described. Cerium dioxide unfortunately is not an easy material
to control as it tends to inhomogeneity and variable refractive index. Magnesium fluoride too
suffers from high levels of stress. High performance can be achieved but most workers prefer to
use titanium dioxide and silicon dioxide.
Titanium dioxide is an extremely robust material but with a melting point of 1925°C which
makes it difficult to evaporate directly unless an electron-beam source is used. A very successful
early method was the evaporation of titanium metal with subsequent oxidation by heating in air
and indices of 2.65 were achieved in this way. The direct evaporation of the oxide usually results
in rather lower indices, 2.25 to 2.4. Like many refractory oxides, titanium dioxide loses oxygen
on heating to melting temperature and the film is oxygen deficient and therefore absorbing. To
replace this lost oxygen it is usual to add a residual atmosphere of oxygen to the coating
chamber, a process known as reactive evaporation.
Macleod: Optical Coatings Page 179 Design through Manufacture

The most complete account of the properties of titanium dioxide, and the way in which they
depend on deposition conditions, is that of Pulker, Paesold and Ritter[10]. The behavior is
exceedingly complicated and the results depend on starting material, oxygen pressure, rate of
deposition, and substrate temperature. The evaporation of Ti3O5 as starting material, gave more
consistent results than were obtained with other possible starting materials, probably because the
composition of the material in the source did not vary during the course of successive
evaporations. With other forms of titanium oxide, the composition varied as the material was
used up, tending in each case towards Ti3O5. Apfel[11] has pointed out the slight conflict
between high optical properties and durability. Optical absorption falls as the substrate
temperature is reduced and the residual gas pressure is raised. At the same time, the durability of
the layers is adversely affected, and a compromise, which depends on the actual application, is
usually necessary. Useful values are substrate temperatures between 200°C and 300°C with gas
pressures around 10-4torr (1.3x 10-4mb). An exactly similar reactive method can be used for
silicon oxide. Silicon monoxide is a convenient starting material which, in its own right, is a
useful material for the infrared. The silicon monoxide can be evaporated readily from a tantalum
boat or, as the material sublimes rather than melts, a howitzer source. Provided there is sufficient
oxygen present, the silicon monoxide will oxidize to the form Si2O3 that has a refractive index of
1.52 to 1.55 and exhibits excellent transmission from just on the longwave side of 300nm out to
8µm[12] An interesting effect involving the ultraviolet irradiation of films of Si2O3 has been
reported[13, 14] With ultraviolet intensity corresponding to a 435W quartz-envelope Hanovia
lamp at a distance of 20cm, the refractive index of the film, after around 5 hours' exposure, drops
to 1.48 (at 540nm). This change in refractive index appears to be due to an alteration in the
structure of the film, rather than in the composition that remains Si2O3. At the same time as the
reduction in refractive index, an improvement in the ultraviolet transmission is observed, the
films becoming transparent to beyond 200nm. Longer exposure to ultraviolet, around 150 hours,
does eventually alter the composition of the films to SiO2. These changes appear to be
permanent. Si2O3 is a particularly useful material for protecting aluminum mirrors, and this
method of improvement by ultraviolet irradiation, opens the way to greatly improved mirrors for
the quartz ultraviolet[13].
Heitmann[1] has made considerable improvements to the reactive process by ionizing the
oxygen in a small discharge tube through which the gas is admitted to the coating chamber. The
degree of ionization is not high, but the reactivity of the oxygen is improved enormously, and the
titanium oxide and silicon oxide films produced in this way have appreciably less absorption
than those deposited by the conventional reactive process. The silicon oxide films show infrared
absorption bands characteristic of the SiO2 form rather than the more usual Si2O3. The technique
has been further improved by Ebert[2, 15] and his colleagues who have developed a more
efficient hollow-cathode ion source, and extended the method to materials such as beryllium
oxide, with useful transmittance in the ultraviolet.
Other materials which have been found useful in thin films, are the oxides and fluorides of a
number of the rare earths. Ceric oxide, although possibly strictly not a rare earth, is examined
along with other materials by Hass, Ramsay and Thun[16].. Cerium fluoride forms very stable
films of index 1.63 at 550nm when evaporated from a tungsten boat. Similarly, the oxides of
lanthanum, praseodymium and yttrium, and their fluorides, form excellent layers when
evaporated from tungsten boats. The properties of the rare earth oxides has been the subject of a
Macleod: Optical Coatings Page 180 Design through Manufacture

study[17] that has shown that better transparency, especially in the ultraviolet, is obtained when
electron beam evaporation is used. Then there is a number of other hard oxide materials which
were extremely difficult to evaporate until the advent of the high-power electron-beam gun, and
so were used only relatively infrequently, if at all. Zirconium dioxide[18] is a very tough, hard
material which has good transparency from around 350nm to some 7µm. It tends to give
inhomogeneous layers, the degree of inhomogeneity depending principally on the substrate
temperature. Hafnium oxide[19, 20] has good transparency to around 235nm, and an index
around 2.0 at 300nm, so that it is a good high-index material for that region. Both yttrium and
hafnium oxide have been found to be good protecting layers for aluminum in the 8 to 12µm
region[21, 22], which avoid the drop in reflectance at high angles of incidence associated with
SiO2 and with Al2O3 [23]
A wide range of low index materials is used in the infrared. Zinc sulfide[23, 24] in
comparison with the high-index semiconductors has a relatively low index. If an electron-beam
source is not available, then zinc sulfide should be deposited from a tantalum boat, or, better still,
a howitzer, on substrates freshly cleaned by a glow discharge and held at temperatures of around
150°C, if the maximum durability is to be obtained. Zinc sulfide films so treated will withstand
boiling for several hours in 5% salt solution, cleaning with cotton wool, and exposure to moist
air, without damage. Silicon monoxide is another possibility[7, 24, 25]. It can also be deposited
from a tantalum boat or a howitzer. The deposition rate should be fast and the pressure low, of
the order of 10-5torr (1.3x 10-5mb) or less if possible. The refractive index is around 1.85 at 1µm
and falls to 1.6 at 7µm. A strong absorption band prevents use of the material beyond 8µm.
Thorium fluoride, unfortunately radioactive, is much used, and there are many other materials,
such as fluorides of lead, lanthanum, barium, cerium, for example, and oxides such as titanium,
yttrium, hafnium and cerium.
In the infrared many semiconductors make good film materials. Germanium is probably the
most common because it is relatively easy to evaporate and it has a useful index around 4.0 and
wide region of transparency beyond around 1.8µm. Lead telluride has an even higher index
around 5.5 and is transparent from 3.4µm out to beyond 20µm. It is difficult to evaporate lead
telluride from an electron-beam source and the best method is a tantalum boat with care taken to
avoid overheating the material. Tight temperature control of the substrates is necessary if good
performance at wavelengths beyond around 10µm is required. Ritchie[26] showed that the 15µm
performance was optimized with a substrate temperature of 250°C and that exceedingly tight
control of substrate temperature was necessary in order to achieve that optimum performance.
Unfortunately the zinc sulfide that normally accompanies the lead telluride as low index material
cannot stand temperatures of 250°C and so for less critical filters a compromise of 150°C is
frequently used although very narrow band filters must simply have the zinc sulfide and lead
telluride controlled at different temperatures.
Lead telluride can in some circumstances behave in a curious way immediately after
deposition[27]. The optical thickness of the material is observed to grow during a period of
around 15 minutes while the layer is still under vacuum. Typical gains in optical thickness of a
halfwave layer are of the order of 0.007 full waves although in any particular case it varies
considerably and can often be zero. The reasons for this behavior are not clear but the layers
once they have ceased growing do not exhibit any further instability. It is simply a matter of
allowing for this behavior in the monitoring process.
Macleod: Optical Coatings Page 181 Design through Manufacture

Mixtures of materials are now receiving attention both in deliberately inhomogeneous films
and in homogeneous films where an intermediate index between the two components of the
mixture is required to improve the evaporation properties of an otherwise difficult material.
Jacobsson and Martensson[28] used mixtures of cerium oxide and magnesium fluoride, of
zinc sulfide and cryolite and of germanium and magnesium fluoride with the relative
concentrations of the two components varying smoothly throughout the films to produce
inhomogeneous films with a refractive index variation of a prescribed law. To produce the
mixture, two separate sources, one for each material were used; they were evaporated
simultaneously but with independent rate controls. Apparently no difficulty in obtaining
reasonable films was experienced, the mixing taking place without causing absorption to appear.
Fujiwara[29, 30] was interested in the production of homogeneous films for antireflection
coatings. The three layer quarter half quarter coating for glass requires a film of intermediate
index which is rather difficult to obtain with a simple material and the solution adopted was a
mixture of two materials, one having a refractive index lower than that required and the other
higher. The two combinations which were tried successfully were cerium oxide and cerium
fluoride and zinc sulfide and cerium fluoride. they were simply mixed together in powder form
in a certain known proportion by weight and then evaporated from a single source. The mixture
evaporated giving an index which was sufficiently reproducible for antireflection coating
purposes. The range of indices obtainable with the cerium oxide cerium fluoride mixture was
1.60 to 2.13 and with the cerium fluoride and zinc sulfide mixture was 1.58 to 2.40. One
interesting feature of the second mixture was that, although zinc sulfide on its own is not
particularly robust, in the form of a mixture with more than 20% by weight of cerium fluoride,
the robustness was greatly increased, the films withstanding boiling in distilled water for 15
minutes without any deterioration. Curves are given for refractive index against mixing ratio in
the papers.
Mixtures of zinc sulfide and magnesium fluoride have also been studied by Yadava, Sharma
and Chopra[31]. The refractive index of the mixture varies between the indices of magnesium
fluoride and zinc sulfide, depending on the mixing ratio, and the absorption edge varies from that
of zinc sulfide to that of magnesium fluoride in a nonlinear fashion. The same authors have
studied the use of assemblies of large numbers of alternate very thin discrete layers of the
components instead of mixtures. For a wide range of material combinations, ZnS-MgF2, ZnS-
MgF2-SiO, Ge-ZnS, ZnS-Na3AlF6 for example, the results were similar to those expected from
the evaporation of mixtures of the same materials.
Feldman and his colleagues at the National Bureau of Standards[32, 33] have investigated
mixtures of zirconia and silica with the idea of forcing the formation of a glass. He found that
small amounts of silica added to the zirconia actually cause the refractive index of the resulting
mixture to increase. A maximum index is obtained with a mixture of some 20% silica and the
characteristics of the film are completely amorphous. This increase in refractive index is clearly
due to an increase in the packing density of the film. Zirconia has quite a low packing density on
its own and even though the effect of the silica admixture is to reduce the bulk refractive index
nevertheless the effect of the increased packing density is sufficiently strong to lead to an overall
increase of refractive index.
Macleod: Optical Coatings Page 182 Design through Manufacture

Quartz is a particularly difficult material to evaporate because of its high melting point and
also because of its transparency to infrared which makes it difficult to heat. It was found by the
Libby-Owens-Ford Glass company that quartz could be thermally evaporated readily provided
some pretreatment was carried out. This consisted of combining the quartz with a metallic oxide,
a vast number of different oxides being suitable. The oxide can be mixed intimately with the
quartz, coated on the outside or even mixed very crudely especially when the oxide has a rather
lower melting point than the quartz. Oxides mentioned in the original patent include aluminum,
titanium, iron, manganese, cobalt, copper, cerium and zinc. Working along similar lines it has
been discovered by workers at Balzers AG that ceric oxide mixed with other oxides improves the
oxidation and increases the transparency and ease of evaporation. Materials such as titanium
dioxide are difficult to evaporate without absorption and the most successful method is reactive
evaporation in oxygen although the process is somewhat time consuming because the
evaporation must proceed relatively slowly. With the addition of a small amount of cerium oxide
- the mixture can vary from 1:1 to 8:1 titanium oxide to cerium dioxide, - hard films free from
absorption even when evaporated quickly at pressures of 10-5Torr are readily obtained.
Apparently this effect is not limited to titanium dioxide and a vast range of different materials
which have been successfully tried is given. Other rare earth oxides can also take the place of the
cerium dioxide.
Stetter and his colleagues[18] have pointed out the advantage of oxygen depleted materials as
source material for electron-beam evaporation, in that composition changes little if at all during
evaporation, which leads to more consistent film properties. The extra oxygen is supplied, in the
usual way, from the residual atmosphere in the plant. The depleted materials also have higher
thermal and electrical conductivity. A mixture of ZrO2 and ZrTiO4, sintered at high temperature
under high vacuum and oxygen depleted, was developed. This material, designated "Substance
No 1," when evaporated from an electron-beam system in a residual oxygen pressure of 1-2x10-
4torr (1.3-2.5x10-4mb) with substrate temperature 270°C, and condensation rate of the order of
10nm/min, gives homogeneous layers of refractive index 2.15 (at 500nm). Such a value of index
is ideal for the quarter-half-quarter antireflection coating for the visible region. Butterfield[34]
has produced films of a mixture of germanium and selenium. For composition varying from 35
to 50 atomic percent of germanium, glassy films with refractive index in the range 2.4 to 3.1,
with good transparency from 1.5 to 15µm, could be produced. The starting material was an alloy
of germanium and selenium in the correct portions, produced by melting the pure substances in
an evacuated quartz tube. The evaporation source was a graphite boat. It is likely that much more
work will be carried out on mixtures, because of the apparent ease with which the deposition can
be performed to give a side range of refractive indices, many of which are not available by other
means. The theory of the optical properties of mixtures is covered in a useful review by
Jacobsson[35] who also gives further information on mixtures, and on inhomogeneous layers.

13.4 The Energetic Processes


The energetic processes, as the name suggests, are ones that involve energies rather greater
than thermal. The idea is that the columnar structure with its accompanying voids must be
disrupted and this requires extra energy. and this does work well. Some of the processes are old
ones that have always involved extra energy and are now recognized as having certain
Macleod: Optical Coatings Page 183 Design through Manufacture

advantages because of it. Others are either conventional thermal processes with additional
energy or are completely new.
Sputtering is an old process that predates thermal evaporation. Momentum transfer from
incident energetic ions is used to eject atoms and molecules from a target into the vapor phase.
The kinetic energy and momentum of the ejected particles are high and so the growing film is
subjected to a much greater impulse each time a fresh particle arrives, which disrupts the void
and columnar structure. In the conventional form of sputtering, the target is metallic so that it
conducts and the bombarding ions are derived from a DC discharge in the vicinity of the target.
This discharge may be confined by crossed electric and magnetic fields when it is known as
magnetron sputtering and this is the most common way in which the process is applied in optical
coating. DC planar magnetron targets are most common, Figure 13-3 shows a schematic form of
such a target. The great advantage of magnetron sputtering is the much longer path length of the
electrons so that the discharge can be maintained at a considerably lower pressure (0.3Pa or
0.3×10-2mb for example) than is required compared with conventional sputtering in the absence
of the magnetic field.

Figure 13-3. Schematic representation of a planar magnetron source. The target or cathode is connected to the
negative supply. The structure of the coating machine including the grounded shield is the positive side of the
supply. Electrons leaving the cathode surface move outwards but are turned into a cycloidal path by the field
of the magnets.
There are, however, some disadvantages. The arrangement of magnets concentrates the
discharge in the region between the pole pieces and the erosion of the target is greatest there
while other areas of the target show negligible erosion. With long rectangular targets, the
appearance of the eroded region is not unlike the shape of a race track, a term often used to
describe it. Target utilization is therefore not good and so used targets are usually recovered
rather than scrapped. Since the targets in DC magnetron sputtering are metallic, a process of
reactive sputtering must be used to produce oxides or nitrides and the sputtering gas therefore, is
usually a mixture of a noble gas such as argon and oxygen or nitrogen. This reactive gas reacts
also with the target to produce a skin of oxide or nitride and the skin tends to build up in the less
eroded regions. The dielectric skin acts as a capacitor, which, because it is thin, has a relatively
high capacitance. The charged particles collect there and charge it up so that considerable energy
is stored. The capacitor is not a particularly good one and it will often break down in a sudden
very violent discharge. The resultant arcing tends to produce molten droplets of material that are
often embedded in the film. In the worst case the discharge can actually damage the target so
Macleod: Optical Coatings Page 184 Design through Manufacture

badly as to render it unusable. The insulating skin also modifies the electrical properties of the
sputtering system so that hysteresis appears making control difficult. These effects are
particularly severe with silicon targets, and silicon oxide is the sole low index material really
suitable for sputtering. The problem is often called target poisoning.
There are several current solutions to the target poisoning problem.
The target surface may be moved with respect to the magnets so that the region of high
erosion moves over the surface and cleans it up. In the usual embodiment the target is made in
cylindrical form and rotated about a longitudinal axis around the magnets and inside the
grounded shield[36].
Another more recent form of solution[37] involves twin magnetron targets that are connected
to opposite poles of a mid-frequency power supply. The targets are now alternately the anode
and cathode of the system. This discharges the effective capacitors before they can cause
damage. Since the capacitance is high the time constant is long allowing relatively low
frequencies to be used. The arrangement also solves the problem of the disappearing anode. In
normal single-target sputtering the chamber structure is the anode of the supply. The build up of
insulating film over this structure gradually makes the anode less and less effective with all kinds
of implications for both control and deposition. The twin magnetron solution avoids this problem
because the alternate source is the anode. The frequency is usually of the order of 40kHz, high
enough to avoid the charging problems but low enough so that the targets are effectively
operating in the DC regime. Usually the twin magnetrons are planar but the process has also
been used with rotating magnetrons.

Figure 13-4. The twin magnetron arrangement in which two magnetron targets are connected to a mid-
frequency power supply so that each is alternately anode and cathode. The arrangement avoids the charging
problems of reactive DC sputtering without the complications of RF sputtering.
An alternative process uses a configuration that has much in common with the traditional
thermal evaporation arrangement. It avoids the interaction with the reactive gas by placing the
magnetron source inside a shroud where it can be operated in argon. The material escapes
through a large aperture above the source in the center of the shroud. Outside the shroud in the
Macleod: Optical Coatings Page 185 Design through Manufacture

main chamber the material coats the substrates while the growing film is bombarded with a beam
of oxygen or nitrogen ions in the manner of ion-assisted deposition, described shortly. A
disadvantage is that enormous quantities of gas are entering the deposition chamber and to
remove the gas very fast and high capacity pumps are necessary. The films that grow are
amorphous of very high packing density. Not a great deal is known about this process because it
was reserved for the inventors’ company. It is the subject of an issued patent[38] and is known as
Microplasma.
Radio-frequency sputtering is a process that avoids the problems of an insulating target. It is
much used in other areas of thin film deposition but has not been popular in optical coatings
mainly because of all the additional problems of radio-frequency systems such as screening and
matching.
The most advanced form of sputtering uses a separate chamber to generate the ions that are
then extracted and directed towards the target. This is known as ion-beam sputtering. It is
capable of a very high degree of film purity and the lowest published losses in optical coatings
have been achieved with this process[4]. Not all materials are suitable for sputtering. In
particular the fluorides present considerable difficulties because of preferential sputtering of
fluorine atoms. The film is then fluorine deficient and optically absorbing. The fluorine
vacancies can be filled with oxygen - there is usually plenty of oxygen around - which removes
the absorption, at least at longer wavelengths, but the film becomes an oxyfluoride with altered
(usually raised) index of refraction and frequently degraded environmental resistance. Since the
ion beam is usually neutralized by adding electrons, charging problems with insulating targets
can be avoided and the process is as useful for insulating materials as for conductors. Ion-beam
sputtering is slow compared with most other processes and it is not able to cope with deposition
over large areas. It has not been generally adopted and its use is largely limited to special
coatings where low loss is the important criterion.

Figure 13-5. Ion beam sputtering schematic. The ion-generating discharge is within the ion gun and therefore
removed from the deposition chamber. This gives much higher quality films.
Macleod: Optical Coatings Page 186 Design through Manufacture

In low-voltage ion plating, a high-current beam of low voltage electrons is directed into the
region above the hearth in an electron beam source. This results in a very high degree of
ionization of evaporant material. There is a complete circuit from ion gun to electron beam
source and back and it is completely isolated from the rest of the structure. The substrate carrier
is also electrically isolated. There are many electrons and they are very mobile and so the
isolated substrates acquire a charge that is negative with respect to the electron beam source.
This attracts the positive ions from the source so that they arrive at the film surface with
additional momentum that is transferred to the film and compacts it.

Figure 13-6. The low-voltage ion plating process. The negative bias on the electrically isolated
substrates is acquired from the free electrons in the chamber.
Laser evaporation is another process that has a much higher energy associated with arriving
atoms and molecules at the film surface. Here the vapor is produced thermally by directing the
beam from a pulsed laser onto the target. This produces a very high local temperature and the
vapor molecules have much higher energy than in conventional thermal evaporation. The laser is
external to the chamber and so there must be a window for entry of the beam. This presents a
problem. It is necessary for there to be a line of sight from laser to target. This means that the
final optical element in the laser beam train will be in line of sight of the target and therefore will
be coated along with the substrates. It is undesirable for this element to be the window and so a
turning mirror is usually provided. It becomes gradually coated during the operation with usually
reducing efficiency. Various techniques for limiting the effects of the deposition are practiced.
The mirror may be gradually rotated from behind a mask during the process so that a fresh
surface is continually presented to the beam. It is also possible to place the mirror such that the
deposition is minimized, although not suppressed. The process tends to be used for coatings that
are essentially inhomogeneous. In such cases the design can often be represented as a succession
of very thin homogeneous layers of either of two materials. Each laser pulse can be arranged to
deposit a further basic unit of the appropriate material. This can ease considerably the normally
severe control and monitoring problems that accompany inhomogeneous layer deposition. The
Macleod: Optical Coatings Page 187 Design through Manufacture

term Laser Pulsed Deposition (LPD) is becoming accepted for a similar process used in
semiconductor device manufacture and so may also be applied to the optical coating method.

Substrate
Plant
wall
Turning
mirror

Window

Laser Baffle
beam

Target
Figure 13-7. The arrangement for laser evaporation.
Ion-assisted deposition is an energetic process that has the great advantage that it is easy to
implement in conventional equipment. It consists of thermal evaporation to which has been
added bombardment of the growing film with a beam of energetic ions. All that is required to put
it into operation in a conventional plant, therefore, is the addition of an ion gun. The commonest
types of ion sources for this purpose are broad-beam often with extraction grids. Much of the
published work and reported successes have been with the Kaufman or gridded type of ion gun.
In that, the source of electrons is a hot filament and the extraction system consists of two closely
aligned grids, the inner floating and acquiring the potential of the discharge so that it confines it
within the gun, and the second applying a field to draw the positive ions out of the discharge
chamber through the apertures in the inner grid. The beam of ions is neutralized outside the
discharge chamber by adding electrons, usually from a hot filament immersed in the beam to
avoid space charge limitation. The grids are fragile and easily misaligned or damaged and so
some effort has been put into the development of sources that do not require extraction grids and
they are being used in increasing numbers in production. Fulton[39]. has written a particularly
useful account of work with gridless sources that includes much about ion assisted deposition in
general.
There are other energetic processes but the principal ones currently appear to be sputtering,
ion plating and ion assisted deposition. Other processes that are currently in use include chemical
vapor deposition, a variety of plasma assisted deposition processes and some liquid processes
mostly based on sol gel technology.
It seems clear that the major benefit of the energetic processes is an increase in film packing
density. The improvements are achieved at comparatively low substrate temperatures which
helps with the difficult coating of plastic substrates.
It has been theoretically demonstrated by advanced computer modeling[40, 41] that the major
effects are due to the additional momentum of the molecules, either supplied by collisions with
the incoming energetic ions, or derived from the additional kinetic energy of the evaporant.
Experimental evidence exists[42] that shows correlation of the effects with momentum rather
Macleod: Optical Coatings Page 188 Design through Manufacture

than energy of the bombarding ions. Major benefits of these processes are the increased packing
density of the films, making them more bulk-like and hence increasing their ruggedness, the
improved adhesion resulting from a mixing of materials at the interfaces between layers, and a
reduction of the sometimes quite high tensile stress in the layers. The increase in packing density
reduces also the moisture sensitivity and can actually eliminate it altogether[43]. The increased
packing density also improves the stability of the films in other ways. Magnesium fluoride films
resist high temperature oxidation better, for example[44]. The hardness and corrosion resistance
of metal films, especially with dielectric overcoats[45], is improved by ion-assisted deposition
but the optical properties tend to be adversely affected, possibly by the implantation of a small
fraction of the bombarding ions[46]. The increased reactivity of the bombarding ions permits the
deposition of compounds, such as nitrides[47], that are difficult or impossible by normal vacuum
evaporation.

Substrate

Ion
beam
Evaporant

Ion Electron
gun beam
source

Figure 13-8. The addition of ion bombardment of the growing film transforms conventional thermal
evaporation into ion-assisted deposition.
The Ionized Plasma Assisted Deposition process includes features of both ion-assisted
deposition and low-voltage ion plating. It makes use of what is known as an Advanced Plasma
Source[48-50]. The source which is insulated from the chamber and floats in potential, is of
simple construction. A central indirectly-heated cathode is made of lanthanum hexaboride. This
lies along the axis of a vertical cylinder that is the anode. A noble gas, usually argon, is
introduced into the source. The cylinder contains a solenoid that produces an axial magnetic
field. The crossed electric and magnetic fields make the electrons move in cycloids with the
usual increase in path length and degree of ionization. so that an intense plasma is produced in
the source. The fields do not confine the plasma axially and so it escapes from the source into the
chamber. There the electrons, that are very mobile, escape preferentially to the chamber structure
leaving the plasma charged positively without the need for isolated substrate holders. The
deposition sources are thermal, usually electron beam, and they emit evaporant into the plasma
Macleod: Optical Coatings Page 189 Design through Manufacture

where it gains energy and is partially ionized. The evaporant then condenses on the growing film
with additional energy, as in ion plating, and is bombarded simultaneously by ions from the
plasma as in ion-assisted deposition. For reactive processes, the reacting gas is not fed into the
source but into the plasma as it leaves the source. A ring-shower-shaped inlet tube is positioned
just above the aperture of the source for this purpose.

Figure 13-9. Schematic diagram of the Advanced Plasma Source.

13.5 Recent Developments


The most important feature of a manufacturing system for optical coatings is consistency. A
particularly weak part of the manufacturing process is the link between process stages. This
usually involves human intervention. Substrates are frequently carried manually from the
cleaning operation to the jig loading, and from there to the coating machine. The machine stands
open while the substrates are loaded. It is then closed and the coating operation begins by
evacuating the chamber, thermally soaking the substrates and, eventually, coating them. Then
they are cooled, removed, and routed to inspection. Recently, this situation has begun to change.
A particularly important change is that the coating chamber is being isolated from the
surrounding atmosphere except for maintenance and cleaning. A gate valve lies between the
coating chamber proper, and a loading chamber. Only when the loading chamber is evacuated
and the substrates in their jig are ready for coating does the gate vale open and permit entry of
the substrates. In some cases there is an intermediate transfer chamber in between the loading
and coating chambers, which adds additional isolation. Robots frequently are involved in the
transfer.
There is also a move towards integrating several functions in one machine. This is particularly
suited to large scale production of similar components, such as ophthalmic lenses or opto-
Macleod: Optical Coatings Page 190 Design through Manufacture

electronic components. The primary objective in all this is not to reduce the number of personnel
but to assure a stability of the total process that is impossible to achieve with relatively
uncontrolled human links.
The configuration of the coating machines is also changing. In 1975, Schiller and
colleagues[51] proposed a new style of thin-film deposition. Their films were intended for other
than optical purposes, but their idea has had great influence also on optical coatings. Instead of a
continuous process of simultaneous deposition and bombardment they proposed a sequential
process in which a small amount of metallic material would be deposited and then treated by
bombardment to be followed by a further small amount of metal with subsequent bombardment.
In this way the total film would be made up of small increments in thickness, each thin enough
so that the effect of the bombardment would reach through the incremental thickness. This
turned out to give films of very high density. Implementation of the process implied a cylindrical
substrate carrier rotating about a horizontal axis with an evaporation source below it and a
bombardment source off to the side. The cylinder then was rotated sufficiently rapidly to assure
the correct increment of film thickness each time a substrate passed the evaporation source. That
this configuration can be extended to optical applications is clear.
The move to sputtering implies that no longer do the sources need to be situated at the base of
the coating chamber and emitting upwards. Sputtering sources can be mounted vertically in the
wall of the chamber. This configuration is becoming quite common.
This configuration is used in a process is known as Metamode, short for metal mode [the
subject of an issued patent[52]]. Intermittent deposition of a metal film takes place as the
substrates, carried on the outside of a vertical cylindrical drum, rotate past the vertical linear
magnetron sources. Further rotation of the substrates takes them past an ion source where they
are bombarded by a beam of reactive ions.
A quite recent development of some importance is known as Radical Assisted Sputtering or
RAS[53]. Here the configuration is also of a vertical drum rotating past, this time, dual
magnetron sources but the interaction is arranged to be with atomic oxygen. There is essentially
no bombardment. The reactivity of the atomic oxygen is so high that bombardment to assure
implantation is not necessary and, in fact, is intentionally suppressed. The films are consequently
of very high quality.
Some machines have adopted a configuration in which the substrate carrier is a flat disk of
large diameter carrying a ring of substrates near its periphery and rotating about an axis normal
to it and at its center. This ring of substrates rotates past a magnetron sputtering source and past a
treatment stage. It has the advantage that the disk can be horizontal, so that gravity retains the
substrates, facilitating robotic loading and unloading.
It is likely that the trend towards integrated processes and robotic handling will continue.

13.6 Chemical Vapor Deposition


Physical vapor deposition processes are those most often used for the production of optical
coatings. However in the electronic device field, chemical vapor deposition is the principal
method for thin film deposition and there is increasing interest in it for optical purposes, usually
with regard to very special requirements.
Macleod: Optical Coatings Page 191 Design through Manufacture

Chemical vapor deposition[36, 54] differs from physical vapor deposition in that the film
material is produced by a reaction amongst components of the vapor that surrounds the
substrates. The reaction may be induced by the temperature of the substrates themselves, when
the process is the classical thermal chemical vapor deposition, or, and this is more usual in the
optical field, it may be a plasma-induced process.
Usually the components, the reactants or precursors, will be introduced into a carrier gas that
is permitted to flow through the system. This ensures a constant supply of the reactants to the
growing interface and allows sufficient dilution so that the reaction is not so fast as to
overwhelm the film growth.
In this classical form of chemical vapor deposition great problems are created by reactions
that are too efficient. A reaction that proceeds rapidly tends to produce a film that is poorly
packed and poorly adherent. The accurate term snow is often used to describe it. The reactions
must, therefore, be quite weak and this means that impurities that have strong reactions can play
havoc with the process. Such effects limit the possible range of processes severely.
Because of all the difficulties, the classical thermal chemical vapor deposition process is not
often used for optical coatings. Instead, pulsed processes have been largely adopted. Material
added to a thin film is assimilated provided it is not immobilized by material deposited over it
before it has had time to relax into favorable positions. The problem is not really the favorability
of the reaction but rather the large amount of material that arrives in a given time. Earlier
material is buried under the weight of later material and cannot relax to a state of equilibrium,
and what is referred to as snow, loosely packed and exceedingly weak material, is the result. If
an efficient reaction can be made to deliver material at a correct rate then the film will be dense.
It is the overall rate of deposition that determines the microstructure. Pulsing the reaction gives
the control of rate that is required.
The pulsing can most conveniently be achieved when a plasma-assisted process is involved
and Pulsed RF Plasma-Enhance Chemical Vapor Deposition is the process that is attracting
greatest interest at the present time. This can be used for plastic substrates[55] because the need
for heating is avoided in the plasma induced processes. Detailed discussion is beyond the scope
of this course.

13.7 Other Techniques


There are many other techniques for the deposition of optical coatings. Probably the most
important of these is the sol-gel process[56]. The name sol-gel refers to those processes that
involve a solution that undergoes a transition of the sol-gel type, that is, a solution is transformed
into a gel. The common form of the sol-gel process starts with a metal alkoxide. This
organometallic compound is hydrolyzed when it is mixed with water in an appropriate mutual
solvent. The solution is usually made slightly acidic to control the rates of reaction and to help
the formation of a polymeric material with linear molecules. The result is a gradual transition to
an oxide polymer with liquid-filled pores. This gel can be deposited over the surface of an
optical component by dipping. The coating is now heat treated to remove the liquid in the pores
and to densify it. the higher the temperature to which it is raised, the denser is the film. By
treating the gel film at temperatures as high as 1000°C complete densification is achieved.
Macleod: Optical Coatings Page 192 Design through Manufacture

Lower temperatures give partial densification but already by 600°C the film is largely
impermeable.
Typical materials are TEOS (tetraethylorthosilicate, Si(OC2H5)4) for eventual films consisting
of silica, and titanium tetraethoxide (Ti(OC2H5)4) for films of titanium oxide. These materials are
dissolved in ethanol and then hydrolyzed by adding a little distilled water. In the case of the
titanium compound, the rate of hydrolyzation is much faster and so nitric acid is also added to
control the transformation and also the solution is made rather weaker.
The sol-gel process received an enormous boost when it was discovered that sol-gel coated
antireflection coatings had exceptionally high laser damage threshold. For regular coatings,
however, and especially for multilayers, the process is far from straightforward and so it has
never competed successfully with the vacuum processes.

13.8 Chapter 13 References


1. Heitmann, W, Reactive evaporation in ionized gases. Applied Optics, 1971. 10: p. 2414-
2418.
2. Ebert, J, Activated reactive evaporation. Proceedings of the Society of Photo-Optical
Instrumentation Engineers, 1982. 325: p. 29-38.
3. Hacman, D, Optische Messung der Substrat-Temperatur in der Vakuumaufdampftechnik.
Optik, 1968. 28: p. 115-125.
4. Lalezari, R, G Rempe, R J Thompson, and H J Kimble. Measurement of ultralow losses
in dielectric mirrors. in Topical Meeting on Optical Interference Coatings. 1992. Tucson:
Optical Society of America. p. 331-333.
5. Bodemann, A, M Reichling, N Kaiser, and E Welsch, Photothermal microscopy of
defects and laser damage morphology in Al2O3/SiO2 dielectric mirror coatings for
248nm. Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1993.
2114: p. 405-414.
6. Tench, R J, R Chow, and M R Kozlowski, Characterization of defect geometries in
multilayer optical coatings. Proceedings of the Society of Photo-Optical Instrumentation
Engineers, 1993. 2114: p. 415-425.
7. Cox, J T and G Hass, Antireflection coatings for germanium and silicon in the infrared.
Journal of the Optical Society of America, 1958. 48: p. 677-680.
8. Bangert, H and H Pfefferkorn, Condensation and stability of ZnS thin films on glass
substrates. Applied Optics, 1980. 19: p. 3878-3879.
9. Hass, G, J B Ramsay, and R Thun, Optical properties and structure of cerium dioxide
films. Journal of the Optical Society of America, 1958. 48: p. 324-327.
10. Pulker, H K, G Paesold, and E Ritter, Refractive indices of TiO2 films produced by
reactive evaporation of various titanium-oxide phases. Applied Optics, 1976. 15: p.
2986-2991.
Macleod: Optical Coatings Page 193 Design through Manufacture

11. Apfel, J H. The preparation of optical coatings for fusion lasers. in International
Conference on Metallurgical Coatings. 1980. San Diego.

12. Ritter, E, Zür Kentnis des SiO und Si2O3-Phase in dünnen Schichten. Optica Acta, 1962.
9: p. 197-202.
13. Bradford, A P and G Hass, Increasing the far-ultra-violet reflectance of silicon oxide
protected aluminium mirrors by ultraviolet irradiation. Journal of the Optical Society of
America, 1963. 53: p. 1096-1100.
14. Bradford, A P, G Hass, M McFarland, and E Ritter, Effect of ultraviolet irradiation on
the optical properties of silicon oxide films. Applied Optics, 1965. 4: p. 971-976.
15. Ebert, Johannes, Reactive evaporation and plasma processes for thin film optical
coatings. Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1988.
1019: p. 220-227.
16. Hass, G, J B Ramsay, and R Thun, Optical properties of various evaporated rare earth
oxides and fluorides. Journal of the Optical Society of America, 1959. 49: p. 116-120.
17. Smith, D and P W Baumeister, Refractive index of some oxide and fluoride coating
materials. Applied Optics, 1979. 18: p. 111-115.
18. Stetter, F, R Esselborn, N Harder, M Friz, and P Tolles, New materials for optical thin
films. Applied Optics, 1976. 15(10): p. 2315-2317.
19. Lehan, J P, Y Mao, B G Bovard, and H A Macleod, Optical and microstructural
properties of hafnium dioxide thin films. Thin Solid Films, 1991. 203(2): p. 227-250.
20. Baumeister, P W and O Arnon, Use of hafnium dioxide in mutilayer dielectric reflectors
for the near uv. Applied Optics, 1977. 16: p. 439-444.
21. Lubezky, I, E Ceren, and Z Klein, Silver mirrors protected with Yttria for the 0.5 to
14µm region. Applied Optics, 1980. 19: p. 1895.
22. Cox, J Thomas and G Hass, Protected Al mirrors with high reflectance in the 8-12-µm
region from normal to high angles of incidence. Applied Optics, 1978. 17(14): p. 2125-
2126.
23. Cox, J T, G Hass, and W R Hunter, Infrared reflectance of silicon oxide and magnesium
fluoride protected aluminum mirrors at various angles of incidence from 8µm to 12µm.
Applied Optics, 1975. 14: p. 1247-1250.
24. Hall, J F and W F C Ferguson, Optical properties of cadmium sulphide and zinc sulphide
from 0.6 micron to 14 micron. Journal of the Optical Society of America, 1955. 45: p.
714-718.
25. Hass, G and C D Salzberg, Optical properties of silicon monoxide in the wavelength
region from 0.24 to 14.0 microns. Journal of the Optical Society of America, 1954. 44: p.
181-187.
26. Ritchie, F S, Multilayer filters for the infrared region 10-100 microns. 1970. PhD Thesis,
University of Reading.
Macleod: Optical Coatings Page 194 Design through Manufacture

27. Evans, C S, R Hunneman, and J S Seeley, Optical thickness changes in freshly deposited
layers of lead telluride. Journal of Physics D, 1976. 9: p. 321-328.

28. Jacobsson, R and J O Martensson, Evaporated inhomogeneous thin films. Applied


Optics, 1966. 5: p. 29-34.

29. Fujiwara, S, Refractive indices of evaporated cerium dioxide - cerium fluoride films.
Journal of the Optical Society of America, 1963. 53: p. 880.
30. Fujiwara, S, Refractive indices of evaporated cerium fluoride - zinc sulphide films.
Journal of the Optical Society of America, 1963. 53: p. 1317-1318.
31. Yadava, V N, S K Sharma, and K L Chopra, Optical dispersion of homogeneously mixed
ZnS-MgF2 films. Thin Solid Films, 1974. 22: p. 57-66.
32. Feldman, A, E N Farabaugh, W K Haller, D M Sanders, and R A Stempniak, Modifying
structure and properties of optical films by coevaporation. Journal of Vacuum Science
and Technology, A, 1986. 4: p. 2969-2974.
33. Feldman, A and E N Farabaugh. Densification of zirconia films by coevaporation with
silica. in OM85, Basic Properties of Optical Materials. 1985: Special Publication 697,
National Bureau of Standards. p. 122-125.
34. Butterfield, A W, The optical properties of GexSe1-x thin films. Thin Solid Films, 1974.
23: p. 191-194.
35. Jacobsson, R, Inhomogeneous and coevaporated homogeneous films for optical
applications. Physics of Thin Films, 1975. 8: p. 51-98.
36. Vossen, J L and W Kern, Thin Film Processes. 1978, New York, San Francisco and
London: Academic Press.
37. Szczyrbowski, J and G Teschner. Reactive sputtering of SiO2 layers onto large-scale
substrate using an AC twin-magnetron cathode. in 38th Annual Technical Conference.
1995. Chicago: Society of Vacuum Coaters. p. 389-394.
38. Scobey, Michael A. Optical Corporation of America. Low pressure reactive magnetron
sputtering apparatus and method. 1996. USA. Patent 5,525,199.
39. Fulton, Michael L, Applications of ion-assisted-deposition using a gridless end-Hall ion
source for volume manufacturing of thin-film optical filters. Proceedings of the Society
of Photo-Optical Instrumentation Engineers, 1994. 2253: p. 374-393.
40. Müller, K-H, Monte Carlo calculation for structural modifications in ion-assisted thin
film deposition due to thermal spikes. Journal of Vacuum Science and Technology, A,
1986. 4: p. 184-188.
41. Müller, K-H, Models for microstructure evolution during optical thin film growth.
Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1988. 821: p. 36-
44.
42. Targove, J D and H A Macleod, Verification of momentum transfer as the dominant
densifying mechanism in ion-assisted deposition. Applied Optics, 1988. 27(18): p. 3779-
3781.
Macleod: Optical Coatings Page 195 Design through Manufacture

43. Martin, P J, H A Macleod, R P Netterfield, C G Pacey, and W G Sainty, Ion-beam-


assisted deposition of thin films. Applied Optics, 1983. 22: p. 178-184.

44. Messerly, Michael J, Ion-beam analysis of optical coatings. 1987. PhD Dissertation,
University of Arizona.
45. Sainty, W G, R P Netterfield, and P J Martin, Protective dielectric coatings produced by
ion-assisted deposition. Applied Optics, 1984. 23: p. 1116-1119.

46. Hwangbo, C K, L J Lingg, J P Lehan, H A Macleod, J L Makous, and S Y Kim, Ion-


assisted deposition of thermally evaporated Ag and Al films. Applied Optics, 1989.
28(14): p. 2769-2778.
47. Hwangbo, C K, Linda J Lingg, J P Lehan, H A Macleod, and F Suits, Reactive ion-
assisted deposition of aluminum oxynitride thin films. Applied Optics, 1989. 28: p. 2779-
2784.
48. Matl, K, W Klug, and A Zöller, Ion-assisted deposition with a new plasma source.
Materials Science and Engineering, 1991. A140: p. 523-527.
49. Pongratz, S and A Zöller, Plasma ion-assisted evaporative deposition of surface layers.
Annual Review of Materials Science, 1992. 22: p. 279-295.
50. Zöller, A, S Beißwenger, R Götzelmann, and K Matl, Plasma ion assisted deposition: A
novel technique for the production of optical coatings. Proceedings of the Society of
Photo-Optical Instrumentation Engineers, 1994. 2253: p. 394-402.
51. Schiller, S, U Heisig, and G Goedicke, Alternating ion plating - A method of high-rate
ion vapor deposition. Journal of Vacuum Science and Technology, 1975. 12(4): p. 858-
864.
52. Scobey, Michael A, Richard I Seddon, James W Seeser, R Russell Austin, Paul M
LeFebvre, and Barry W Manley. Optical Coating Laboratory, Inc. Magnetron sputtering
apparatus and process. 1989. USA. Patent 4,851,095.
53. Matsumoto, Shigeharu, Kazuo Kikuchi, Masafumi Yamasaki, Qi Tang, and Shigetaro
Ogura. Shincron Co., Ltd. Method of forming a thin film of a composite metal compound
and the apparatus for carrying out the method. 2001. USA. Patent 6,207,536 B1.
54. Vossen, John L and Werner Kern, Thin Film Processes II. 1991, San Diego: Academic
Press.
55. Möhl, W, U Lange, and V Pacquet, Optical coatings on plastic lenses by PICVD-
technique. Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1994.
2253: p. 486-491.
56. Thomas, Ian M, Sol-gel coatings for high power laser optics: past present and future.
Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1993. 2114: p.
232-243.
Macleod: Optical Coatings Page 197 Design through Manufacture

14 MICROSTRUCTURE
One of the most significant features of optical thin films is the way in which their properties
and behavior differ from those of identical materials in bulk form. This is, of course, also true for
thin films in areas other than optics. Almost always, the performance of the film is poorer than
that of the corresponding bulk material. Refractive index is usually lower, although, very
occasionally, for some semiconductor materials it can be slightly higher, losses greater,
durability less, and stability inferior. There is also a sensitivity to deposition conditions,
especially substrate temperature. Heitmann[1] has studied the influence of parameters, such as
the residual gas pressure within the plant and the rate of deposition, on the refractive indices of
cryolite and thorium fluoride. Raising the residual gas (nitrogen) pressure from 4x10-6torr
(5.3x10-6mb) in one case, and 2x10-6torr (2.6x10-6mb) in another, to 2x10-5torr (2.6x10-5mb) had
no measurable effect, within the accuracy of the experiment (±0.1% for thorium fluoride and
±0.3% for cryolite) while a further increase in residual pressure to 2x10-4torr (2.6x10-4mb) gave
a drop in index of 1.5% for cryolite, and 1.4% for thorium fluoride. At this higher pressure, the
mean free path of the nitrogen molecules was less than the distance between boat and substrate,
and the decrease in refractive index was probably caused by increased porosity of the layers.
This tends to confirm that the mean free path of the residual gas molecules should be kept longer
than the source-substrate distance, but that any further increases in mean free path beyond this,
have little effect. Heitmann concluded that the mean free path of the molecules is the important
parameter, not the ratio of the numbers of evaporant molecules to residual gas molecules
impinging on the substrate in unit time, which appeared to have no effect on refractive index. He
also found that changes in the rate of deposition, from a quarterwave in 0.5 minutes (measured at
632.8nm) to a quarterwave in 1.5 minutes, caused a decrease in refractive index of 0.6% in both
cases, but that a further decrease to a quarterwave in 5 minutes produced only slight variations.
Heitmann's results are probably best interpreted in terms of slight changes in film structure,
induced by the variations in deposition conditions. Layer structure is, in fact, the most significant
factor in determining the properties of optical thin films and the way in which they differ from
the same material in bulk form. During the past ten years or so, there has been an increasing
interest in the structure of, and structural effects in, optical thin films.
From many electron microscopical studies[2, 3] it is now well known that optical thin film
materials condense with a columnar structure when thermally evaporated in the normal way. For
coatings produced for visible region applications, the radius of the columns is often in the range
20 to 60nm according to the evaporation conditions and layer thickness. For infrared purposes
where layers are very much thicker the columnar diameter can be much greater and similarly
they can be rather smaller for ultraviolet applications. An important implication of the columnar
structure is that there are voids in the films in between the columns and that there is a very large
so called inner surface area around the columns that must be added to the already large outer
surface area. The term packing density is used to express the incidence of voids within a film.
Packing density is defined as:

Volume of the solid part of the film


p= (14.1)
Total volume of the film ( solid + voids )
Macleod: Optical Coatings Page 198 Design through Manufacture

The packing density (which might be more correctly described as packing coefficient)
depends both on the particular material and on the deposition conditions. It usually increases
with substrate temperature but even at considerably elevated temperatures it is frequently less
than unity. This is especially the case with the refractory oxides. Figure 14-2 shows the variation
of packing density of MgF2 with temperature and indicates that compact films demand substrate
temperatures of around 300°C[4]. This is a well-known effect with magnesium fluoride. Packing
density also varies with film thickness, increasing in some films and decreasing in others.

Figure 14-1. Columnar structure in a film of zinc sulfide[5].


The columnar microstructure fits well with the Movchan-Demchishin zone model of film
growth[6]. Here the important parameter is the ratio of substrate temperature to evaporant
melting point. If the ratio is less than 0.45 (for zirconia for example with its melting point of
2700°C, or 2973K, the ratio for a substrate temperature of 300°C, 573K, is around 0.2; for titania
the ratio is around 0.27) then the microstructure is a pronounced columnar one. The same zone
model applies also to sputtered films but the pressure of the gas in the process affects the result.
Thornton[7, 8] modified the Movchan-Demchishin model to fit sputtering results with the
addition of background gas pressure as an additional axis, but the broad details of the columnar
structure remain. Messier[9] has extended the zone model to include the effects of layer
thickness. Thick layers tend to have larger columns than thin ones and this has always been
puzzling if the columns are considered to be the same size throughout the film. Messier
introduced the idea of a hierarchy in columnar structure. He examined large numbers of films
with thicknesses ranging from a few Angstroms to many microns using a variety of techniques.
It appears that the basic growth unit is some 25 - 40Å in diameter and all films appear to begin
with units of this dimension. As the films grow the basic features extend out from the substrate
surface and either become merged into larger units, the forces of attraction will pull these units
together, or some units grow at the expense of others. In some cases etching of the films has
revealed a fibrous structure running through the columns and this is probably the basic dendritic
growth that still retains some of its independent features even though the dendrites are gathered
into the larger diameter columns.
Macleod: Optical Coatings Page 199 Design through Manufacture

Figure 14-2. The measured variation of p with T for MgF2.

Figure 14-3. The zone model of thin film microstructure [After Movchan and Demchishin[6]]
The microstructure of the films has actually several levels, the columnar structure representing
one level. We will discuss film growth and attempt to reach an understanding of the reasons for
this microstructure.

14.1 Origin of Microstructure


14.1.1 Columnar Microstructure
The conditions under which thin films are grown are very far from equilibrium. The sources
are at very high temperatures and so emit vapor at an enormous rate in an attempt to reach
equilibrium with the vapor. But the structure of the coating plant and the substrates are all at
Macleod: Optical Coatings Page 200 Design through Manufacture

much lower temperature and have very great area compared with the source. The are capable of
sucking the vapor out of the vapor phase at a much higher rate than it can ever be produced by
the sources. The result is that we have two processes that are very far from equilibrium, the
production of the vapor and its condensation. The fraction of the arriving material that is
reemitted from the growing films is very small (except in a few exceptional cases) and it is a
good assumption that the sticking coefficient of arriving material is unity. Arriving material
travels in essentially a straight line from the source and we can think of it in the same way as
illumination from a source of light. As the film grows it rapidly starts to exhibit height variations
on the surface which then cast shadows which the vapor atoms or molecules cannot enter
directly. They must first land on an accessible part of the surface and then migrate over the
surface into the shadow. This migration is not normally effective in improving the growth in the
shadow areas, which are emphasized to the extent that a columnar structure forms.
Much improvement in our understanding of thin film microstructure comes from two-
dimensional models of film growth in which mechanical ideas play an important role. In two
dimensions we can imagine an arriving atom as a circular disk which will land on the film at a
random site and will then try to find a position of minimum energy by a process of thermally
activated hopping from place to place that may take it into the shadow of a column. A useful
source of information on such modeling is the report of a 1987 conference[10].

Figure 14-4. The structure of a bilayer where the mobility of the part nearer the substrate (the substrate is
considered to be at the foot of the film) was higher during deposition that the outer part. The boundary
between the films can be seen roughly half way between substrate and outer surface and the reduction in
packing density in the outer part is clear. The films were modeled for oblique deposition of the vapor.
The thermally activated hopping is primarily a surface effect and cannot take place at
reasonable temperatures in the bulk when all the sites are completely filled with similar atoms or
molecules. Thus there is great movement of material on the surface of the growing film but as
additional material arrives, the existing material is buried and immobilized. Müller[11,
12]demonstrated the dominance of surface diffusion in film growth in a model that contained the
thermal activation of all molecules in the simulated film. The greater the substrate temperature,
the greater the probability of hopping and so the greater the surface diffusion and the greater the
packing density. An enormous volume of calculation was involved in the inclusion of the entire
set of molecules in the thermal activation in each cycle of the model. Müller's model
demonstrated clearly the dependence of packing density on substrate temperature. Packing
Macleod: Optical Coatings Page 201 Design through Manufacture

density increases with substrate temperature but it is a little more complicated than that because
the rate of deposition is also involved.
The columnar structure is reduced and the packing density increased with increased mobility
of the deposited molecules. But by mobility we must understand the total travel of the molecule.
This is a random walk process and so a very large number of hops are required in order for the
molecule to move a significant distance. All else being equal the higher the substrate
temperature, the greater will be the probability of hopping and the further will be the movement
of the molecules. But the hopping ceases when the molecule is buried and the time available
before it is immobilized in this way is a function of the rate of deposition. The faster the rate, the
less the total movement. Increased deposition rate therefore operated against increased substrate
temperature.

Figure 14-5. Sargent's model[13] results showing the variation of the packing density of nickel as a function
both of rate of deposition and of substrate temperature. This is an elaboration of some earlier results by
Müller that did not show a rate dependence.
Müller's demonstration that surface effects were dominant eased the requirements on
modeling and permitted much larger models to be constructed[13]. It was possible to make the
simplifying assumption that any buried atom is unable to move and, further, to alter the sequence
of calculation so that each arriving molecule would be permitted to move to its final position
before the arrival of a subsequent molecule. The results of a model of this kind are shown in
Figure 14-4 and Figure 14-5. The mobility of the arriving molecules can be varied to show the
transition between Zone 1 and Zone 2 structure.
A qualitative representation is indicated in Figure 14-6.
Given that slow deposition leads to higher packing density than fast, why is it that so many
workers prefer to deposit materials as quickly as possible? The probable reason for this is that
there is always residual contamination present in the background of the coating chamber and this
contamination arrives at a fixed rate on the surface of the growing film. The contamination levels
are, therefore, proportional to the inverse of the rate of deposition. The optimum, therefore is
probably given be a rate fast enough to make the contamination levels acceptable but slow
enough to attain a good microstructure. Clearly the greater the background level of
contamination, the faster will be the optimum deposition rate.
Macleod: Optical Coatings Page 202 Design through Manufacture

In the bulk material beneath the surface, there may be voids. Thermally activated hopping can
still take place on the inner surface of these voids and this hopping can, permit some migration
of the voids, rather in the way a hole moves through the filled band of a semiconductor. The
migration of the voids is slow compared with the film growth and so the migration is relatively
unimportant during film deposition but, as Müller has shown, becomes of major importance
during post deposition thermal treatment such as baking or annealing. If the bonding between the
film and the adjacent material is poorer than in the film itself then there is a definite tendency for
the voids to congregate at the film boundary with the consequent deterioration of adhesion. Thus,
baking or annealing of a film system will not usually improve adhesion but will make adhesion
failures more probable. In their migration the voids may touch. There is then a great energetic
advantage in amalgamating into a larger single void. This process implies that smaller voids tend
to disappear during annealing or baking and are replaced by a smaller number of larger voids.
This combining of voids does not change the packing density but it does tend to make the voids
more open and accessible. The improvements in stability that are found to accompany baking are
likely to be principally due to this improved accessibility which reduces the time to reach
equilibrium with the surroundings. When the environment is reasonably stable this gives an
apparent improvement in the stability of the coating but it may not reduce the total changes that
can take place when the environment alters.

Increasing
packing
Deposition density
rate

Substrate
temperature
Figure 14-6. The qualitative influence of substrate temperature and deposition rate on packing density.

14.1.2 Crystalline Microstructure


A second level of microstructure in thin films is their crystalline state. This is less well
understood but considerable progress has been made. Optical thin films are deposited from vapor
that has been derived from sources at comparatively very high temperature. The substrates on
which the films grow, are at relatively very low temperature. There is therefore a great lack of
equilibrium between growing film and arriving vapor. The film material is rapidly cooled or
quenched, and this not only influences the formation of the columnar microstructure but it also
affects the crystalline order. The material that is condensing will attempt to reach the equilibrium
form appropriate to the temperature of the substrate, but the correct rearrangement of the
molecules will take a certain time, and the film will tend to pass through the higher temperature
forms during this rearrangement. If the rate of cooling is greater than the rate of crystallization,
Macleod: Optical Coatings Page 203 Design through Manufacture

then a higher temperature form will be frozen into the layer. The very rapid cooling rate
normally existing in thin films implies the presence of quite high temperature forms and there are
often mixtures of phases. This explains an, at first sight, curious behavior of thin films.
Frequently there is an inversion in the crystalline structure in that at low substrate temperatures a
predominance of high-temperature crystalline forms are found, whereas at high substrate
temperatures, more low temperature material appears to form. The low substrate temperature
leads to a higher quench rate and the rest follows. Amorphous forms, corresponding to a quite
high temperature, can often be frozen by very rapid cooling, and are enhanced by a higher
temperature of the arriving species. For example sputtering, where additional kinetic energy is
possessed by the arriving molecules, often gives amorphous films. The low voltage ion-plating
technique, again with high incident energy, appears virtually invariably to give amorphous films.
The high temperature forms are often only metastable and may change their structure at quite
low temperatures leading to problems of various kinds. Some films deposited in amorphous form
by sputtering may sometimes be induced to recrystallize, in a manner described as explosive, by
a slight mechanical disturbance, such as a scratch, or by laser irradiation[14].

Table 14-1. Samarium Fluoride (SmF3)[15]

Normal high temperature form Hexagonal


Normal low temperature form Orthorhombic
Thermal Substrate temperature Hexagonal (111)
evaporation of 100°C
Substrate temperature Orthorhombic (111) with
≥200°C some hexagonal
Ion assisted Substrate temperature Hexagonal (110) with some
deposition 100°C (111)
Higher bombardment Hexagonal (110) with
at substrate temperature appearance of new peak
100°C SmF2(111)?

Samarium fluoride has two principal crystalline forms, a hexagonal high-temperature form
and an orthorhombic low temperature form. Table 14-1 shows the results of thermal evaporation
and ion-assisted deposition which both lead to this apparently inverted structure[15]. Zirconia
has three principal structures, monoclinic, tetragonal and cubic in ascending temperature.
Klinger and Carniglia[16] found that very thin zirconia shows a cubic structure, but becomes
monoclinic when thicker than a quarterwave at 600nm. This behavior can be explained by a
lower rate of quenching when the film is thicker and less thermally conducting. Alumina,
normally amorphous in thin film form, can recrystallize in the electron microscope when
subjected to the electron bombardment necessary for viewing[17]. Amorphous zirconia, which
can occur when films are very thin, has been shown to exhibit similar behavior[18].
Thin films, therefore, are complicated mixtures of different crystalline phases, some being
high-temperature metastable states. Such behavior is clearly very material and process dependent
and each specific system requires individual study. What is a good structure for one application
Macleod: Optical Coatings Page 204 Design through Manufacture

may not be so for another. The low scattering of the amorphous phases make them attractive for
certain applications, but their high-temperature or high-flux behavior may not be as satisfactory.
Much more needs to be done in attempting to improve our understanding.

14.2 Effects of Microstructure


The microstructure of the layers has a profound effect on almost all of their properties. This
account is a very abbreviated one of some of the more important effects.

14.2.1 Optical Properties


It is clear that the microstructure of the layers will have great influence on the optical
properties. Since the film is made up of solid parts and voids, we can expect that the refractive
index will be reduced below that for a film made up entirely of solid material. Many attempts
have been made at deriving a relationship between packing density and refractive index. An
expression frequently used for the index of a composite film based on nothing more than an
interpolation between two limits, is that of Kinosita and Nishibori[19]

n f = (1 − p )nv + pns (14.2)

where nf is the index of the composite film, ns is the index of the solid material of the film, that is
the columns, and nv is the index of the voids in the film. This expression is reasonably useful for
low index films but less so for high index films where the detailed form of the microstructure
becomes very important.
An isolated column of dielectric material of refractive index rather higher than its
surroundings, which is placed in an electric field, becomes polarized. The charge sheets that
form on the outer surface tend to reduce the field within the material and force it into the
surroundings. A film made up of isolated columns with low refractive index material
surrounding them, exhibits this effect in all columns so that the net refractive index of the
composite film is reduced. An expression for the refractive index of an array of separated
cylindrical columns is due to Bragg and Pippard[20] and can be written:

2 (1 − p )nv4 + (1 + p )nv2 ns2


n =
f (14.3)
(1 + p )nv2 + (1 − p )ns2

where we use p for the packing density, ns for the index of the solid part of the film and nv for
the index of the void material surrounding the solid columns. It is also possible to visualize a
microstructure where the effect of the surface charge sheets is completely suppressed such as
will an intimate mixture of two materials, or simply a columnar structure with very high packing
density. In such cases an expression for refractive index of the composite film is:

n 2f = pns2 + (1 − p )nv2 (14.4)

These two expressions, the Bragg and Pippard of (14.3) and the mixture of (14.4) may be
thought of as the likely limiting values for the index of a structured thin film. Curves are shown
Macleod: Optical Coatings Page 205 Design through Manufacture

in Figure 14-7 along with some measured results for zirconia[21, 22]. The results for one series
of measurements show a shift from the mixture law at high packing density to the Bragg and
Pippard at low packing density, indicating a pronounced columnar structure. The other series
shows a typical mixture result down to quite low packing densities. Here the columnar form is
not so pronounced. This illustrates the sensitivity of film optical constants to their microstructure
and to their deposition conditions. We shall see later that there are additional effects involved in
determining the optical constants of the films.

Figure 14-7. Measured zirconia refractive index as a function of packing density. Mao results measured in air
and compared with the refractive index of a mixture of solid and water-filled voids. Martin results measured
in vacuo and compared with the Bragg and Pippard model with vacuum or air-filled voids
The columnar nature of the films strongly suggests that they should be birefringent with the
principal axes of the dielectric tensor along the columnar direction (n3), and normal to the
columns both in the plane of incidence of the vapor (n1) and normal to the plane of vapor
incidence (n2). For films that are evaporated at normal incidence we would expect the
birefringence to be uniaxial with optic axis normal to the film surface and not detectable by
purely normal incidence measurements. Normal incidence measurements would, however,
permit us to infer the presence of birefringence in obliquely deposited films where we would
expect biaxial birefringence with the optic axes in the plane of incidence of the vapor during film
growth. Because obliquely deposited columns are less closely spaced in the plane of vapor
incidence than normal to the plane, we would expect the birefringence usually to be
characterized by a maximum index along the columns and by a minimum normal to the columns
in the plane of incidence. The optic axes for a biaxial film would therefore be in the plane of
vapor incidence.
Some measurements, at a wavelength around 630nm, of anisotropic behavior in zirconium
oxide deposited at 30° and at 65° and on titanium oxide deposited at 30°[23] are summarized in
the following table.
Macleod: Optical Coatings Page 206 Design through Manufacture

Material α β n1 n2 n3

Zirconium oxide 30° 16.1° 1.948 1.969 2.033


Zirconium oxide 65° 47.0° 1.502 1.575 1.788
Titanium oxide 30° 16.1° 2.437 2.452 2.552
Note:
α is the angle of vapor incidence
β is the column orientation angle, both measured with respect to the substrate
normal,
n1 corresponds to E in plane of vapor incidence normal to column axis
n2 corresponds to E normal to vapor incidence plane - s-polarization
n3 corresponds to E along column axis.

Anisotropy of optical constants has also been observed in metal films. The anisotropy may
originate in the formation of ellipsoidal grains during early stages of film growth. In the case of
aluminum it appears that oxygen has a considerable influence and it is likely that a thin sheath of
aluminum oxide can form around the aluminum columns permitting essentially the same kind of
behavior as in dielectric columns. Thin metal films evaporated obliquely are sometimes used as
polarizers.

Figure 14-8. A double peak is clearly visible in this single-cavity narrowband filter. This is not a moisture-
induced effect but is a consequence of residual birefringence at normal incidence in the titanium oxide/silicon
oxide layers of this filter[24].
Macleod: Optical Coatings Page 207 Design through Manufacture

Thin film optical coatings ideally should not exhibit any anisotropy at normal incidence. If the
films exhibit the form birefringence of a columnar structure then as long as the columns are
arranged normal to the interfaces, the optic axis should also be normal to the interfaces and no
polarization sensitivity should be detectable. Narrowband filters, however, have by their nature
exceedingly sharp characteristics and such features are particularly sensitive to small changes in
the properties of the layers. Figure 14-8 shows a narrowband filter of single cavity design but
with a double peak. The filter was constructed by a process of thermal evaporation and the
characteristic was measured at normal incidence with essentially unpolarized light. The effect
here is not due to moisture adsorption but to residual birefringence at normal incidence in the
materials of the filter. This example shows a quite large effect but lesser effects are not
uncommon in filters of very narrow width and have presented problems in communication
systems. For such applications it is becoming normal to use processes that tend to amorphous
structures that avoid the form birefringence responsible for this effect. The energetic processes,
are especially suitable for this and are gradually becoming the processes of choice for such
filters.

14.2.2 Mechanical Properties


Thin films exhibit mechanical properties that differ in a number of respects from bulk
material. They are almost invariably in a state of some considerable strain with consequent strain
energy and stress. Some of this strain results from differential contraction of the film and
substrate on cooling from the deposition temperature but there is an intrinsic component of stress
that arises directly from the microstructure. The bonds that hold the film material together link
partially across the gaps between the columns and attempt to draw the columns together. This
force is resisted by the substrate and there is a resultant tensile stress. The energetic processes
beat the films down and so contribute to a compressive stress that in the case of ion plating and
sputtering can be very high. In ion-assisted deposition the bombardment can be finely adjusted
and so it is possible to vary the stress from tensile right through to compressive in the materials
like titania that are considered to be "well behaved. " There are materials, such as zirconia that
exhibit rather stranger behavior sometimes increased tensile stress in the energetic processes.
As we move from the optical properties to the mechanical properties of thin film, our
knowledge and understanding lessen. We tend to talk in the thin-film field as though stress were
an attribute of a particular film, some materials being compressive and some tensile, for
example. The stress in the film is a complicated function of the various phases and the way in
which they are packed together, different crystallites with perhaps different stress, in turn
dependent on deposition conditions, and on the surrounding material that, for equilibrium, must
support the total stress. We pay a great deal of attention to the reaction forces at the substrate,
which tend to distort it, and will often look for material combinations and deposition conditions
for multilayer coatings that will combine to minimize the substrate forces. The low shear loading
across the substrate-multilayer interface does imply that adhesion failures there are less likely.
However, a net low stress in a multilayer does not necessarily mean low shear stress at all
interfaces, nor low tensile or compressive stress in all layers. A set of parallel springs can be in
equilibrium and yet the individual springs can be alternately in significant tension and
compression. A film system can similarly exhibit dangerously high stress at certain interfaces
even though the total stress is low. It is not always at the substrate that adhesion failures occur.
Macleod: Optical Coatings Page 208 Design through Manufacture

The nature of the forces that attach thin films to each other and to the substrate can range from
van der Waals forces, when there is no chemical affinity between the materials, to strong
chemical bonding such as ionic or covalent bonds. Van der Waals forces are the weakest of these
but even they are of considerable magnitude. Quite pessimistic assumptions can lead us to a
figure of around 109Nm-2 as the force of adhesion for a van der Waals bonded film. Hydrogen
bonding and other very much stronger chemical bonds lead to greatly increased figures.
Jacobsson and Kruse[25], using a direct pull technique, measured a value of around one
twentieth of this figure for the adhesion of films such as zinc sulfide, cryolite and silver. This is
still an impressive force (greater than 3 tons/sq in). Yet, in spite of these high forces, adhesion
failures can and do occur more frequently than we could wish. Adhesive systems, such as Scotch
tape, on the other hand, where the adhesive forces are only a small fraction of the forces
involved at thin-film interfaces, seem much less prone to failure.
The problem is related to the type of adhesion failure which is commonly found in thin-film
systems, that of peeling. The peeling may be a simple lifting of the coating, or blistering, but it is
always characterized by progressive delamination rather than by the simultaneous detachment of
a large area. In peel failures, much more than in any other type of failure, the range of the forces
is as important as their magnitude. Indeed it is much more meaningful to talk of work of
adhesion rather than adhesive force. Interatomic and intermolecular forces are all short range
forces, and in the case of our van der Waals forces, we can assign a range of around 0.2nm. The
work which would be involved in removing unit area of the coating, known as the work of
adhesion, would then be

(109)x(0.2x10-9) = 0.2 Jm-2

and we can look on this work as creating a surface energy of 0.1 Jm-2 on each of the fresh
surfaces which are exposed when the coating is removed. The force required to peel unit width
of the coating from the surface is then found by equating the work done by the force to the work
of adhesion, that is to the total energy. In this particular example, we have F=0.2Nm-1,
equivalent to around 200mg force for a 1cm width.
The performance of Scotch tape, which has a very low adhesive force by comparison, is due
to the very much greater range over which the forces operate by stretching of the adhesive.
These aspects of peeling are brought out clearly in an excellent paper by Orowan[26].
A thin film with van der Waals bonding, therefore, presents little resistance to peeling and it is
instructive to consider whether the level of internal stress which is possible in the film, might be
high enough to cause a spontaneous adhesion failure. A modest estimate of stress might be 108Pa
(Chopra[27] suggests ten times this figure) and, in passing, we can note that the tension in a 2µm
thick film, due to this stress, would be 2x102Nm-1 which we can compare with the figure of
0.2Nm-1 for the stripping force for a van der Waals bonded film. Once again, however, we need
to consider the energy involved. It is difficult to estimate the strain energy without a figure for
Young's modulus, and few measurements have been carried out on dielectric thin films. There is
a figure of 6.6x1011dynes cm-2 i.e. 6.6x1010Pa which has been obtained for SiO2 films[27] and
there are some recent measurements by Ledger and Bastien[28] of ZnSe and ThF4 ranging from
3.9x10-5 to 6.8x105kg cm-2, i.e. 3.8x1010 to 6.7x1010Pa. If we assume a figure of 5x1010Pa for
Macleod: Optical Coatings Page 209 Design through Manufacture

Young's modulus and zero for Poisson's ratio then, for a biaxial stress of 108Pa we find a strain
energy of

(108)2/(5x1010) = 2x105 Jm-3

and our 2µm thick film will have an energy of 0.4Jm-2 which is twice what is required to
balance the surface energies which will be created on failure of adhesion.
But the situation is still more complicated than this. For an adhesion failure to be initiated, the
shear forces acting at the interface must exceed the bond strength. The forces depend on the
elastic constants of film and substrate and on the nature of the bonds themselves. If the substrate
and the bonds are infinitely stiff, for instance, the entire tension in the film will be supported by
the outermost bonds which will therefore experience an enormous shear stress. If on the other
hand, the substrate and the bonds have some resilience, then the load will be shared over a
greater area, and the forces experienced by any one bond will be much less. Even with tensions
as high as the figure of 2x102Nm-1, the forces carried by the bonds can still be within the bond
strength. Indeed a rough estimate based on the characteristics of our notional 2µm thick film
suggests that the van der Waals bonds will just be able to survive. Thus, for adhesion to fail, a
stress concentrator may be needed to trigger the failure although once the failure has progressed
sufficiently far, it may itself act as a stress concentrator. Next, during the peeling process,
sufficient strain energy must be available for conversion into the energy of the freshly exposed
surfaces. The calculation which we have carried out so far, just gives the total amount of energy
in the system, and the amount which can actually be used for conversion into surface energy
depends entirely on whether or not other energy dissipating mechanisms are present. Plastic
deformation of the film material or the substrate, or crack generation away from the adhesion
failure are two such mechanisms which are likely to operate in thin film systems. If these take up
too great a fraction of the strain energy, then the adhesion failure will not progress. Effects like
these are probably of importance in all stable thin-film coatings, but it is not possible to say
much more at this stage because quantitative measurements are lacking. Clearly adhesion
failures are complex and difficult problems which have much in common with the propagation of
fractures in engineering structures. There is no doubt, however, that the risk of adhesion failure
is reduced if use can be made of the considerable increase in bond energy associated with the
much stronger chemical bonds and these are certainly involved in all useful thin-film systems.
The surface energies produced by the breaking of chemical bonds can be as high as 5Jm-2.
How do adhesion tests fit into this picture? There are two tests in particular, which are used to
a very great extent in thin films. The first is the Scotch tape test and the second is the abrasion
test. The Scotch tape test appears, at first sight, to be a straightforward peel adhesion test, but it
must be remembered that the actual adhesive force of the tape is small and because the adhesive
stretches, the force is applied over a considerable area of the film. Unless, therefore, there is
some defect which can act as a stress concentrator so that film peeling can begin, the film is
unlikely to be affected, even when the adhesion is low. Once peeling is started, it will tend to
continue, with the force applied as tension in the detached portion of the film until this part of
the film breaks. This type of behavior can be seen when a film in which adhesion is poor, but not
zero, is tested. Such a coating will often pass the straightforward test, but then, if it is scribed
with a diamond into a series of squares, a repeat test will often lift out complete squares.
Abrasion testing is much more complex, adhesion playing only a part, with film structure,
Macleod: Optical Coatings Page 210 Design through Manufacture

hardness and friction all involved. External stress is applied locally through the abrading
particles and film defects are important. Once detachment of a coating begins then peeling can
and usually does take place.

14.2.3 Moisture Adsorption


When films are exposed to the atmosphere they adsorb atmospheric moisture[29-31]. At low
values of relative humidity the principal effect is the formation of a layer, essentially a
monolayer, over the entire inner surface, that is the surface of the columnar grains. The quantity
of moisture adsorbed in this way has a significant effect on the optical properties. The immediate
result is an increase in film refractive index, since the moisture has a higher index than the space
it takes up. The optical thickness, which is the product of the physical thickness and refractive
index, will also be increased. At higher values of relative humidity, the pores in the films
actually fill with liquid water by capillary condensation, increasing still further the refractive
index. If the relative humidity falls, the liquid water tends to desorb, although there may be
considerable hysteresis. The initial monolayer, however, is very tightly bound and resists
desorption unless the material is heated to a very high temperature usually in excess of 600°C.

Figure 14-9. Variation of the index of refraction with humidity of a magnesium fluoride film. (After Ogura[4]
and Macleod[32])
In multilayers, the behavior is even more complicated than in single layers because the
interfaces between the layers obstruct the pores. Even a temporary halt in the evaporation
without altering material can cause trapped pores[4]. This obstruction slows down the rate of
moisture penetration. Now the moisture tends to enter at isolated penetration sites passing
through to deeper layers and spreading laterally. This alters the optical properties of the layers in
circular patches that gradually expand with time. The rate of spreading depends on the relative
humidity and on the characteristics of the particular materials. It may take a matter of minutes or
of months, even years to reach equilibrium. The penetration sites represent defects in the
coatings and may coincide with dust particles or asperities, pits or scratches in the substrate.
They may also be caused by poor substrate cleaning. Experiments aimed at establishing the
origins of the penetration sites show that virtually any departure from perfection can result in a
penetration pore. When coatings are exposed to a reduced relative humidity, the moisture tends
to desorb although there is usually considerable hysteresis.
Macleod: Optical Coatings Page 211 Design through Manufacture

The way in which the characteristics of the coating are affected by the change in optical
properties as the moisture penetrates, depends on the particular design. The wet areas of
narrowband Fabry-Perot filters (that is single cavity filters) simply exhibit a shift towards longer
wavelengths as the depth of penetration increases. The wet patches can be observed by
illuminating the filter with monochromatic light and have been studied in detail be
Richmond[31, 33] and by Lee[30]. Depending on the relationship between the wavelengths of
the illuminating light and the peak of the filter, the patches will appear as dark against a bright
background or as bright against a dark background. Multiple-cavity filters suffer a detuning of
the cavities that is often enough to cause a complete disappearance of the passband, and later,
once the penetration is complete, the passband reappears but at a longer wavelength. The area of
the filter in use will normally be great enough to include a number of moisture penetration
patches. At an early stage of adsorption the wet patches will be small and surrounded by
relatively large areas of dry material that will gradually reduce in size as the adsorption
proceeds. The overall characteristic of the filter will be a combination of the characteristics of
the individual elements of area, which depend both on the stage reached in the adsorption
process and on the filter designs. Narrowband Fabry-Perot filters have passbands shifted more or
less towards longer wavelengths, often broadened and curiously misshapen. Multiple-cavity
filters, on the other hand, show less initial drift but suffer a greater reduction in peak
transmittance that varies considerably over the area of the filter. At a late stage of adsorption the
passband starts to reappear at rather longer wavelengths in areas where it was either missing or
much reduced.

Figure 14-10. A sketch of the equipment for viewing moisture penetration patterns in optical coatings. (After
Macleod and Richmond[31])
Macleod: Optical Coatings Page 212 Design through Manufacture

Figure 14-11. Water adsorption in a narrowband filter constructed from zinc sulfide and cryolite. The picture
was take at a wavelength of 484nm and a relative humidity of 46%, six days after coating. The patch size at
this stage is around 100µm. (After Lee[34].

Figure 14-12. The same area of the filter of Figure 14-11 but at a wavelength of 507nm.

Figure 14-13. the filter of Figure 14-11 eight days after coating. The relative humidity has risen to 50%. The
wavelength is 485nm and the patches are now around 250µm in diameter.
Macleod: Optical Coatings Page 213 Design through Manufacture

Figure 14-14. The filter of Figure 14-13 at a wavelength of 508nm. The patches are the source of the mottled
appearance that coatings sometimes exhibit on exposure to the atmosphere after manufacture.
Heating the filters causes rapid desorption. In the absence of desorption the characteristic
would simply move due to the change in optical thickness. The sense of the shift and its
magnitude depend on the particular materials but for zinc sulfide and cryolite filters, for
example, it is towards longer wavelengths. The desorption of moisture, however, complicates the
shift and, in the case of zinc sulfide and cryolite, can even reverse its sign. There can also be
considerable hysteresis that varies with the cycle time. In such a case the peak wavelength of the
filter is virtually unpredictable[35].

14.2.4 Moisture and Mechanical Properties


Unfortunately the process is still further complicated by additional effects that the moisture
has on layer properties. The film mechanical properties are also affected because the moisture
blocks off the bonds across the gaps between the columns. The tensile stress falls and may even
sometimes become compressive. Because of the weakening of the forces acting between the
columns, the film durability falls.
Furthermore, since the coating of water that forms over all surfaces reduces the surface energy
- in the case of the high energy solids such as the refractory oxides the reduction can be an order
of magnitude - it reduces the work that must be done to cause cracking or delamination. This
means a great increase in the density of stress relief cracking and a greatly increased tendency to
delamination. Such effects are commonly observed after the humidity stage of environmental
testing. Adhesion failures in thin-film systems are peel adhesion failures characterized by a
progressive delamination rather than a sudden detachment of a large area. In this type of failure,
as with stress relief cracking, the work required to cause the failure to propagate is principally
the surface energy that has to be generated for the freshly exposed surfaces together with any
energy dissipated in plastic deformation. Thus the presence of moisture in the films, which
reduces the surface energy, can contribute to a substantial reduction of the work required in the
propagation of an adhesion failure and therefore to a greatly increased probability of such a
failure. Note that these moisture-induced effects do not require that the materials should be water
soluble. In such failures, the strain energy in itself can often be great enough to supply the
energy required. All that is then necessary is a stress concentrator to initiate the failure and this
can coincide with a penetration pore where the presence of moisture makes the delamination
Macleod: Optical Coatings Page 214 Design through Manufacture

more likely. Blistering is a peel adhesion failure in which there is a compressive stress in the
layers so that when they lift off the substrate they tend to form an unbroken dome. At high values
of relative humidity, multilayers of cryolite and zinc sulfide exhibit a moisture uptake that is
much greater than can be explained by the simple filling of the voids of the films. Here the extra
moisture congregates at the boundaries between layers where it causes peel adhesion failures that
form small blisters. The blisters gradually grow in area until the films separate completely and
the filter is destroyed.

Figure 14-15. Behavior of the passband of an unprotected narrowband filter of single cavity construction as a
function of time after production. (After Macleod and Richmond[31])

14.3 Chapter 14 References


[1] W. Heitmann, “The influence of various parameters on the refractive index of evaporated
dielectric thin films,” Applied Optics, vol. 7, pp. 1541-1543, 1968.
[2] P. H. Lissberger and J. M. Pearson, “The performance and structural properties of
multilayer optical filters,” Thin Solid Films, vol. 34, pp. 349-355, 1976.
[3] K. H. Guenther and H. K. Pulker, “Electron microscopical investigations of cross
sections of optical thin films,” Applied Optics, vol. 15, pp. 2992-2997, 1976.
[4] S. Ogura, “Some features of the behaviour of optical thin films,” PhD Thesis, Newcastle
upon Tyne Polytechnic, 1975.
[5] I. M. Reid, H. A. Macleod, E. Henderson, and M. J. Carter, “The ion plating of optical
thin films for the infrared,” in Proc International Conference on Ion Plating and Allied
Macleod: Optical Coatings Page 215 Design through Manufacture

Techniques (IPAT 79), London, July 1979. Edinburgh: CEP Consultants Ltd, 1979, pp.
55-62.
[6] B. A. Movchan and A. V. Demchishin, “Study of the structure and properties of thick
vacuum condensates of nickel, titanium, tungsten, aluminium oxide and zirconium
dioxide,” Fiz Metal Metalloved, vol. 28, pp. 653-660, 1969.
[7] J. A. Thornton, “Influence of apparatus geometry and deposition conditions on the
structure and topography of thick sputtered coatings,” Journal of Vacuum Science and
Technology, vol. 11, pp. 666-670, 1974.
[8] J. A. Thornton, “The microstructure of sputter-deposited coatings,” Journal of Vacuum
Science and Technology, A, vol. 4, pp. 3059-3065, 1986.
[9] R. Messier, “Towards quantification of thin film morphology,” Journal of Vacuum
Science and Technology, A, vol. 4, pp. 490-495, 1986.
[10] M. R. Jacobson, “Modeling of Optical Thin Films,” in Proc SPIE, vol. 821, 1988, pp.
233.
[11] K.-H. Müller, “Models for microstructure evolution during optical thin film growth,”
Proceedings of the Society of Photo-Optical Instrumentation Engineers, vol. 821, pp. 36-
44, 1988.
[12] K.-H. Müller, “Dependence of thin-film microstructure on deposition rate by means of a
computer simulation,” Journal of Applied Physics, vol. 58, pp. 2573-2575, 1985.
[13] R. B. Sargent, “Surface diffusion: A computer study of its effects on thin film
morphology,” PhD Dissertation, University of Arizona, 1989.
[14] R. Messier, T. Takamori, and R. Roy, “Observations on the "explosive" crystallization of
non-crystalline Ge,” Solid State Communications, vol. 16, pp. 311-314, 1975.
[15] L. J. Lingg, “Lanthanide trifluoride thin films: structure, composition and optical
properties,” PhD Dissertation, University of Arizona, 1990.
[16] R. E. Klinger and C. K. Carniglia, “Optical and crystalline inhomogeneity in evaporated
zirconia films,” Applied Optics, vol. 24, pp. 3184-3187, 1985.
[17] J. D. Targove, “The ion-assisted deposition of optical thin films,” PhD Dissertation,
University of Arizona, 1987.
[18] C. Boulesteix and M. Lottiaux, “Behavior of zirconia film in electron microscope,” ,
1987.
[19] K. Kinosita and M. Nishibori, “Porosity of MgF2 films - evaluation based on changes in
refractive index due to adsorption of vapors,” Journal of Vacuum Science and
Technology, vol. 6, pp. 730-733, 1969.
[20] W. L. Bragg and A. B. Pippard, “The form birefringence of macromolecules,” Acta
Crystallographica, vol. 6, pp. 865-867, 1953.
[21] P. J. Martin, R. P. Netterfield, and W. G. Sainty, “The modification of the optical and
structural properties of dielectric ZrO2 films by ion-assisted deposition,” Journal of
Macleod: Optical Coatings Page 216 Design through Manufacture

Applied Physics, vol. 55, pp. 235-241, 1984.

[22] Y. Mao, “Environmental and thermo-mechanical stability of thin films for optical
applications,” PhD Dissertation, University of Arizona, 1990.
[23] F. Horowitz, “Structure-induced optical anisotropy in thin films,” PhD Dissertation,
University of Arizona, 1983.
[24] J. J. Wharton, “Microstructure related properties of optical thin films,” PhD Dissertation,
University of Arizona, 1984.
[25] R. Jacobsson and B. Kruse, “Measurement of adhesion of thin evaporated films on glass
substrates by means of the direct pull method,” Thin Solid Films, vol. 15, pp. 71-77,
1973.
[26] E. Orowan, “The physical basis of adhesion,” Journal of the Franklin Institute, vol. 290,
pp. 493-512, 1970.
[27] K. L. Chopra, Thin Film Phenomena. New York and London: McGraw Hill Book
Company, 1969.
[28] A. J. Glass and A. H. Guenther, “Laser induced damage in optical materials: ninth ASTM
symposium,” Applied Optics, vol. 17, pp. 2386-2411, 1978.
[29] H. Koch, “Optische Untersuchungen zur Wasserdampfsorption in Aufdampfschichten
(inbesondere in MgF2 Schichten),” Physica Status Solidi, vol. 12, pp. 533-543, 1965.
[30] C. C. Lee, “Moisture adsorption and optical instability in thin film coatings,” PhD
Dissertation, University of Arizona, 1983.
[31] H. A. Macleod and D. Richmond, “Moisture penetration patterns in thin films,” Thin
Solid Films, vol. 37, pp. 163-169, 1976.
[32] H. A. Macleod, “The monitoring of thin films for optical purposes,” Vacuum, vol. 27, pp.
383-390, 1977.
[33] D. Richmond, “Thin film narrow band optical filters,” PhD Thesis, Newcastle upon Tyne
Polytechnic, 1976.
[34] C. C. Lee, H. A. Macleod, and R. Potoff, “Moisture adsorption in optical coating (A),”
Journal of the Optical Society of America, vol. 72, pp. 1732, 1982.
[35] P. Roche, L. Bertrand, and E. Pelletier, “Influence of temperature on the optical
properties of narrowband optical filters,” Optica Acta, vol. 21, pp. 927-946, 1974.
Macleod: Optical Coatings Page 217 Design through Manufacture

15 MONITORING, ERRORS AND TOLERANCES


In order to discuss the question of tolerances and accuracy we need to include some
discussion of actual monitoring methods.
Given suitable materials, clean substrates, and plant with substrate-holder geometry to give
the required distribution accuracy, the main problem which remains is that of controlling the
deposition of the layers so that they have the characteristics required by the coating or filter
design. Of course, many properties are required, but refractive index and optical thickness are
the most important. For normal production there is no really satisfactory way, at present, of
measuring the refractive index of that portion of a film which is actually being deposited. Such
measurements can be made later, but for closed loop control, dynamic measurements are
required. Normal practice, therefore, is simply to control, as far as possible, those deposition
parameters that would affect refractive index so that the index produced for any given material is
consistent. Measurements are made of the index and the value usually obtained is used in the
coating design. This procedure is, admittedly, far from ideal and is used simply because, at the
present time, there is no better way. Great improvements have been achieved by use of the
energetic processes because the high packing density helps to stabilize the optical constants, and
hence stabilize the process.
Film thickness can more readily be measured and, therefore, controlled. The simplest systems
display a signal to a plant operator who is responsible for interpreting it and assessing the correct
instant to terminate deposition. At the other end of the scale, there are completely automatic
systems in which operator judgment plays no part and in which even operator intervention is
rarely required.
There are many ways in which the thickness can be measured. All that is necessary is to find a
parameter that varies in a suitable fashion with thickness and to devise a way of monitoring this
parameter during deposition. Thus, parameters such as mass, electrical resistance, optical
density, reflectance and transmittance have all been used. Of all the methods, those most
frequently used involve either optical measurements of reflectance or transmittance or the
measurement of total deposited mass by the quartz crystal microbalance.
The question of the best method for the monitoring of thin films is, of course, inseparable
from that of how accurately the layers must be controlled. This second question is a surprisingly
difficult one to answer. Indeed, it is impossible to separate the two questions: the tolerances
which can be allowed and the method used for monitoring are closely related and one cannot be
considered in depth independently of the other.
For convenience, however, we will consider some of the commoner arrangements for
monitoring, including only the most rudimentary ideas of accuracy. and then, at a later stage,
consider the question of tolerances along with some of the more advanced ideas of monitoring
and its various classifications.
Most of the material will deal with batch processing where thickness control can be directly
exercised and is usually simply a matter of taking more or less time to deposit a particular layer.
Continuous processes present quite a different challenge because the line moves at a given rate
and so the time variable is denied to the thickness control system. There it is a matter of tight
Macleod: Optical Coatings Page 218 Design through Manufacture

control of all possible variables. Sputtering is inherently a much more stable process from the
point of view of rates of deposition and stability of material properties and so continuous
processes tend to use sputtering with very tight control over power, pressure, gas composition
and so on. Thickness is often measured but this is used as a process check rather than a control
variable

15.1 Optical Monitoring Techniques


Optical monitoring systems in batch processes consist essentially of a light source
illuminating a test substrate which may or may not be one of the components in the batch, and a
detector analyzing the reflected or transmitted light. From the results of that analysis, the
evaporation of the layer is stopped as far as possible at the correct point. Particularly in thermal
evaporation, so that the layer may be stopped as sharply as possible, the plant is fitted with a
shutter which can be inserted in front of the vapor sources and is a much more satisfactory
method than merely turning off the supply to the thermal sources that always take a finite time to
stop emitting.
Almost all the early workers in the field used the eye as the detector, and the thicknesses of
the films were determined by assessing their color appearance in white light. In many cases they
were concerned with simple coatings, such as single-layer antireflection coatings for the visible
region, which are fairly resistant to errors. When the antireflecting layer is of the correct
thickness for visible light, the color reflected from the surface in white light has a magenta tint
because the reflectance is lowest in the green. The visual method is quite adequate for this
purpose and is still being widely used. A very clear account of the method is given by Mary
Banning[1], who compiled Table 15-1.
Table 15-1. The colors reflected from films during growth.
Color change for Color change for Optical thickness for
ZnS Na3AlF6 green light
Bluish white Yellow
White Magenta λ/4
Yellow Blue
Magenta White λ/2
Blue Yellow
Greenish white Magenta 3λ/4
Yellow Blue
Magenta Greenish white λ
Blue Yellow
Green Magenta 5λ/4
Macleod: Optical Coatings Page 219 Design through Manufacture

In the production of other types of filter where the errors of the visual method would be too
large, other techniques must be used. An early paper by Polster[2] describes a photoelectric
method which is basically the same as that used most often today. The reflectance, or
transmittance, at one wavelength for a dielectric film with low absorption, varies in a cyclic
manner as the thickness increases. The curve shape is similar to a sine wave, although, for the
higher indices, slightly flattened at high reflectance low transmittance peaks. The turning values
correspond to those wavelengths for which the optical thickness of the film is an integral number
of quarter wavelengths, the reflectance being equal to that of the substrate when the number is
even and a maximum amount removed from the reflectance of the substrate when the number is
odd.

Reflectance versus Optical Thickness


40

30
Reflectance (%)

20

10

0
0.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5 5.0
Quarterwaves
Figure 15-1. The variation of reflectance as a function of layer optical thickness for a number of different
materials deposited on a glass surface of index 1.52. From lowest to highest curve the indices run, 1.38, 1.45,
1.52, 1.65, 1.9, 2.1, 2.4.
Figure 15-1 illustrates the behavior of films of different values of refractive index. This
affords the means for measurement. If the detector in the system is made highly selective, for
example, by putting a narrow filter in front of it, then the measured reflectance or transmittance
will vary in this cyclic way, and the film may be monitored to an integral number of quarter-
waves by counting the number of turning points passed through in the course of the deposition.
The wavelength selecting device may be an interference filter or, with greater flexibility, an
adjustable prism or grating monochromator. A word of warning is appropriate. The output of an
incandescent source varies enormously with wavelength. This means that the output may be
much greater at a wavelength other than that selected by the system. The rejection of the
wavelength selector must, therefore, be exceedingly high otherwise stray light may be present.
Stray light at a different wavelength from that assumed in the signal predictions can play havoc
with the form of the signal and, hence, with the thickness controlled. Even if the filters that are
being produced are straightforward of undemanding performance, the performance of the
monitoring system should be of a very high order. Single cavity narrow band filters, for example,
Macleod: Optical Coatings Page 220 Design through Manufacture

are quite unsuitable for this purpose. Their degree of rejection is insufficient. Multiple cavity
filters, if interference filters are to be used, are absolutely necessary.
If the control is to be manual then the output of the detector should be displayed on a chart
recorder; this makes it easier to determine the turning values. In fact it is almost impossible to
control the process from an instantaneous meter reading. With this arrangement, a trained
operator can usually terminate the layers to an accuracy on the monitoring substrate of around 5-
10%, depending on the index of the film. Of course, as we shall see, this does not necessarily
mean that the actual thickness of the filters in the batch will be as accurate. Other sources of
error operate to introduce differences between the monitor and the batch.

RunSheet1 from layer 1 on chip glass


100

80
Reflectance (%)

60

40

20

0
0 400 800 1200 1600 2000 2400

Distance from Chip (nm)

Figure 15-2. The variation of reflectance at the reference wavelength as a quarterwave stack is deposited. The
horizontal scale is of total physical thickness. The layers have indices of 1.45 and 2.4 and the reference and
monitoring wavelength is 1000nm. Note the very poor contrast as the multilayer grows beyond seven layers.
Consider the deposition of a high-reflectance multilayer stack where all the layers are quarter-
waves. Let the monitoring wavelength be the wavelength for which all the layers are one-
quarter-wavelength thick. The reflectance of the test piece will vary as shown in Figure 15-2.
The example shown is typical of a stack for the visible or near infrared regions. The reflectance
can be seen to increase during the deposition of the first layer, which is of high index, to a
maximum where the deposition is terminated. During the second layer the reflectance falls to a
minimum where the second layer is terminated. The third layer increases the reflectance once
again and the fourth layer reduces it. This behavior is superimposed on a trend towards a
reflectance of 100% so that the variable part of the signal becomes a gradually smaller part of the
total. This puts a limit on the number of layers which can be monitored in reflectance in this way
to around four to six, when a fresh monitoring substrate must be inserted. In transmission
monitoring, the variable part of the signal remains a sufficiently large part of the whole but the
problem now is that the overall trend of the signal is towards zero, so that eventually it will
become too small in comparison with the noise in the system. With good optics and a sensitive
and stable detector the number of layers which may be dealt with in this way is around twenty-
one. At this stage the noise usually becomes too great.
Macleod: Optical Coatings Page 221 Design through Manufacture

MONITOR3: Transmittance
100

80

Transmittance (%) 60

40

20

0
950 960 970 980 990 1000 1010 1020 1030 1040 1050
Wavelength (nm)
Figure 15-3. The theoretical performance curve of the narrowband filter of design:
Air | HLHLHLH LL HLHLHLH | Glass with nH=2.4 and nL=1.45 and λ0=1000nm.

RunSheet1 from layer 1 on chip glassT


100

80
Transmittance (%)

60

40

20

0
0 400 800 1200 1600 2000 2400

Distance from Chip (nm)

Figure 15-4. Transmittance monitoring curves in calculated for the filter of Figure 15-3 with system
bandwidths of 1nm, light curve, and 10nm, dark curve. The 10nm bandwidth, similar to that of the filter itself
shows a gradually reducing contrast in the later layers of the filter culminating in a spurious turning value in
the final layer, with a signal reversal at the final termination point. The shape of this final layer is sometimes
called a dogleg.
In narrowband filter production there is an advantage in monitoring all layers on the same
substrate, preferably one of the filters under construction. Since this filter will be rotating during
deposition, great mechanical and optical stability is required. Here it is important that the
bandwidth of the monitoring system should not be greater than that of the filter that is under
construction. If it does happen to be greater then, as the constructed filter nears the end of its
Macleod: Optical Coatings Page 222 Design through Manufacture

deposition the effect of a decreasing halfwidth may be greater than that of increasing
transmittance and the transmitted monitoring signal may fall rather than rise. There are also
some spurious effects that can occur. Some of these are illustrated in Figure 15-4 where the
calculated monitoring traces for a correctly deposited single-cavity narrow band filter are shown
for different bandwidths of the monitoring system. Spurious extrema and reversed signals
together with much reduced sensitivity can all be seen towards the end of the process.
To improve the signal-to-noise ratio it is usual to chop the light before it enters the plant,
partly because the evaporation process produces a great deal of light during the heating of the
boats, but mainly because at the signal levels encountered, the electronic noise without some
filtering would be impossibly great. The chopper should be placed immediately after the source
of light but before the plant, and the filter should be inserted after the plant. This arrangement
reduces the stray light to a greater extent than would placing either the filter before the plant or
the chopper after it. It is, of course, always advisable to limit as far as possible the total light
incident on the detector, partly because unchopped radiation can push the detector into a non-
linear region and partly because it can cause damage to the device especially if it is a
photomultiplier. If a filter rather than a monochromator is used. then we stress once again that
great care should be taken to ensure that the sidebands are particularly well suppressed.
Detectors and sources have characteristics which vary considerably with wavelength, and if the
monitoring wavelength lies in a rather insensitive region compared with the peak sensitivity,
then small leaks in the more sensitive region, which might not be very noticeable in the
characteristic curve of the filter, can cause considerable difficulties from stray light, even giving
spurious signals of similar or greater magnitude than the true signal. Prism or grating
monochromators are often safer for this work, besides being considerably more flexible.
Care needs to be taken also with the quality of the monitoring light. We have considered the
bandwidth of the source but there are other factors equally or even more important. Particularly
insidious effects are produced by stray light outside the immediate vicinity of the nominal
monitoring wavelength. Figure 15-5 shows the effect on the monitoring of a simple quarterwave
stack of light at one half of the monitoring wavelength. In this case the amount of light at the
shorter wavelength is one tenth of that at the correct wavelength. This may appear to be a rather
large error but bear in mind that the output curve of sources varies enormously with wavelength
and a small leak in a filter may turn into an appreciable amount of energy compared with the
fundamental. In Figure 15-5 the second harmonic has caused the curves for the later high-index
layers in the stack to be quite badly perturbed with a spurious turning point in their centers and a
reversed turning point at their terminations. The low-index layers have their contrast almost
completely washed out. This is a fairly severe example but it is an indication of the kinds of
errors that can occur. This particular component is a quarterwave stack but in narrow band filters
the effect can make the monitoring of the cavity layers almost impossible. There is often
dispersion of the indices of the layers. This tends to make the effects of the second order light
slightly lopsided, particularly in cavity layers. Second-order light is a special danger when
diffraction grating monochromators are being used for producing the monitoring beam. Good
order-sorting filters are essential. More complicated stray light problems can occur when filters
are being used as monochromators. Leaks can occur both at longer and at shorter wavelengths
than the nominal peak of the filter. These leaks can be particularly difficult to detect with grating
spectrometers which are often intended for the examination of samples with some transmission.
Macleod: Optical Coatings Page 223 Design through Manufacture

Strange monitoring signals are usually a sign that there is something that needs to be corrected in
the monitoring system.

RunSheet1 from layer 1 on chip GlassT


100
Transmittance (Signal Units)
80

60

40

20

0
0 400 800 1200 1600 2000 2400
Distance from Chip (nm)
Figure 15-5. The monitoring trace of a simple quarterwave stack with ideal monochromatic light (light line)
and light containing 10% of energy from the second order, i.e. one half of the monitoring wavelength. The
spurious peaks and reversed turning values characteristic of this kind of error can be clearly seen.
We can investigate the errors likely to arise in this type of monitoring under ideal
circumstances as follows: Suppose that in the monitoring of a single quarter-wave layer there is
an error γ in the value of reflectance at the termination point. This will give rise to a
corresponding error ϕ in the phase thickness of the layer δ where

δ = (π 2 ) − ϕ (15.1)

Because of the nature of the characteristic reflectance curve of the single layer, the error in
phase thickness will be rather greater in proportion than the original error in reflectance. The
admittance of the layer will be given by the characteristic matrix:

⎡ cos δ ( i sin δ )
y⎤ ⎡ 1 ⎤
(15.2)
⎢ ⎥⎢ ⎥
⎣iy sin δ cos δ ⎦ ⎣ ysub ⎦

where

cos δ = sin ϕ and sin δ = cos ϕ

This gives

sin ϕ + i ( ysub cos ϕ ) y


Y= (15.3)
iy cos ϕ + ysub sin ϕ
Macleod: Optical Coatings Page 224 Design through Manufacture

where the symbols have their usual meaning. Introducing the approximations for sin ϕ and cos ϕ
up to and including powers of the second order, we have

ϕ + i ( ysub y ) (1 − ϕ 2 2 )
Y= (15.4)
iy (1 − ϕ 2 2 ) + ysubϕ

and the reflectance of the monitor in vacuo will be given by

( ysub − 1) ϕ + i ( y − ysub y ) (1 − ϕ 2 2)
2

R= (15.5)
( ysub + 1) ϕ + i ( y + ysub y ) (1 − ϕ 2 2)

which simplifies to

⎛ ⎞
( y − ysub y ) ⎜1 + ( ysub y 2 ) 4 ysub 2 ⎟
2 2
+ 1 − y 2 − ysub
2

R= 2 ⎜
ϕ ⎟ (15.6)
( y + ysub y ) ⎜
⎝ ( y − ysub y )
2 2 2 2

The values of γ, the reflectance error, and ϕ are related as follows

γ=
(y 2
sub + 1 − y 2 − ysub
2
y 2 ) 4 ysub
ϕ 2 = σϕ 2 (15.7)
(y 2
−y 2
sub y )
2 2

since the first factor in equation (15.6) is just the reflectance when γ and ϕ are both zero.
Now, in most cases it will not be possible to determine the reflectance at the turning value to
better than 1% of the true value. In many cases, especially where there is noise, it will not be
possible even to do as well as this. However, assuming this value for γ, the expression for the
error in the layer thickness becomes

±0.01 = σϕ 2 (15.8)

where the sign ± is taken to agree with σϕ2 and depends on whether or not the turning value is a
maximum or a minimum. If the error is expressed in terms of a quarter-wave thickness which is
equivalent to π/2 radians, the expression becomes

ϕ 0.1
Error = = (15.9)
π 2 π 2σ 12

A typical case is the monitoring of a quarter wave of zinc sulfide on a glass substrate where
y=2.35 and ysub=1.52. Substituting these values in expression (15.7) and using it in (15.9), we
find the fractional error in the quarter wave as 0.08. This is a colossal error compared with the
original error in reflectance, and illustrates the basic lack of accuracy inherent in this method.
Macleod: Optical Coatings Page 225 Design through Manufacture

In the infrared, it is often possible to use wavelengths for monitoring which are shorter than
the wavelengths of the desired filter peaks by a factor of perhaps two or even four. This
improves the basic accuracy by the same factor. For layers similar to that considered above, the
errors would then be 0.04 or 0.02. These errors are on the limit of permissible errors, and it is
clear that this simple system of monitoring is not really adequate for any but the simplest of
designs.
What makes the method particularly difficult to apply is that it is only the portion of the signal
before the turning point that is available to the operator, who has therefore to anticipate the
turning value, and the fact that trained plant operators can achieve the theoretical figures for
accuracy says much for their skill.
An alternative method, inherently more accurate, involves the termination of the layer at a
point remote from a turning value where the signal changes much more rapidly. This consists of
the prediction of the reflectance of the monitoring substrate when the layer is of the correct
thickness and then the termination of the deposition at that point. One disadvantage is that the
reflectance of the monitor, or the transmittance, is not an easy quantity to measure absolutely,
because of calibration drifts during the process due partly to such causes as the gradual coating
of the plant windows - almost impossible to avoid. Another is that whereas with turning value
monitoring it is often possible to use just one single monitor, on which all the layers can be
deposited, so that it becomes an exact replica of the other filters in the batch, in this alternative
method the prediction of the reflectances used as termination values is very difficult if only one
monitor is used, because small errors in early layers affect the shape of the curve for later layers.
Some of these difficulties may be avoided by using a separate monitor for each and every
layer. To avoid the errors due to any shift in calibration which may occur in changing from one
monitor to the next or in the coating of the plant windows, it is wise if at all possible to choose
the parameters of the system so that the layer is thicker than a quarter-wave at the monitoring
wavelength.
This ensures that the termination point of the layer is beyond at least the first turning value,
which can therefore be used as a calibration check. It will also be found necessary to set up the
reflectance scale for each fresh monitoring substrate and the initial uncoated reflectance which
will be known accurately can be used for this. Because a large number of monitor glasses is
required, special monitor changers have been designed and are commercially available, which
will accommodate stacks of 40 or so glasses.
The principal objection which most workers almost instinctively feel towards this system is
that no longer is the monitor an exact replica of the batch of filters. This is to some extent a valid
objection. The layer which is being deposited on an otherwise uncoated substrate is condensing
on top of what may be quite a different structure from the partially finished filters of the batch.
Behrndt and Doughty[3] noticed a definite measurable difference between layers which are
deposited on top of an already existing structure and those deposited on fresh substrates. They
compared the deposition of zinc sulfide shown by a crystal monitor (this special type of monitor
will be discussed shortly), which already had a number of layers on it, with the layer going down
on a fresh glass substrate, and found that the layer began to grow on the crystal immediately the
source was uncovered, but that the optical monitor took some time to register any deposition.
The difference could amount to several tens of nanometres before the rates became equal. This,
Macleod: Optical Coatings Page 226 Design through Manufacture

they decided, was due to the finite time for nuclei to form on the fresh glass surface and the
rather small probability of sticking of the zinc sulfide until the nuclei were well and truly
formed. Once the film started to grow, all the molecules reaching the surface would stick. On the
crystal where a film already existed, not necessarily of zinc sulfide, nucleation sites were already
there and the film started to grow immediately. The sticking coefficient of a material on a fresh
monitor surface falls with rising vapor pressure, and zinc sulfide has a particularly large vapor
pressure. Similar trouble was not experienced with thorium fluoride, which has a much lower
vapor pressure. Behrndt and Doughty found that the problem could be solved by providing
nucleation sites on the clean monitor slides by precoating them with thorium fluoride, which has
a refractive index very close to that of glass. Some 20 nm or so of thorium fluoride was found to
be sufficient and did not affect the monitoring of zinc sulfide deposited on top. This effect
becomes greater the greater the surface temperature of the monitor. By changing the type of
evaporation source to an electron beam unit, which produced less radiant heat for the same
evaporation rate, it was found possible to operate at monitor temperatures low enough to cause
the effect to disappear. The same authors also mention an effect which is often observed in thin-
film optics. Thick substrates tend to have layers condensing on them which are thicker than those
on thin substrates in the same or similar positions in the plant. In the particular case cited, the
thin substrates were around 0.040 inch thick while the thick ones were around half an inch thick.
The difference in coating thickness was sufficient to shift the reflectance turning values by some
40-50 nm at 632.8 nm. This was shown, qualitatively, to be due to the difference in temperature
between the two substrates. The thicker substrates took longer to heat up than the thin ones. The
heating in this particular case was almost entirely due to radiation from the sources and, again
when electron-beam sources were introduced, the effect was considerably reduced.
The accuracy of the monitoring process can be improved greatly if a system devised by
Giacomo and Jacquinot[4], and known usually as the maximètre. is employed. This involves the
measurement of the derivative of the reflectance versus wavelength curve of the monitor. At
points where the reflectance is a turning value, the derivative of the reflectance with respect to
wavelength is zero and is rapidly changing from a positive to a negative value in the case of a
maximum and vice versa in the case of a minimum. The original apparatus consisted of a
monochromator with a small vibrating mirror before the slits on the exit side so that a small
spectral interval was scanned sinusoidally. The output signal from the detector consisted of a
steady DC component, representing the mean reflectance, or transmittance, over the interval, a
component of the same frequency as the scanning mirror representing the first derivative of the
reflectance against wavelength, a component of twice the scanning frequency, representing the
second derivative of the reflectance, and so on. A slight complication is the variation in
sensitivity of the system with wavelength that appears as a change in the reflectance signal and
hence the derivative, unless it is compensated. In their arrangement, Giacomo and Jacquinot
produced an intermediate image of the spectrum within the monochromator, and a razor blade
positioned along it made a linear correction to the intensity over a sufficiently wide region and
was found to be accurate enough. A more usual technique today would be to make a correction
electronically. The accuracy claimed for this system is a few tenths of a nanometre, typically 0.2-
0.3 nm, and this is certainly achieved. We emphasize that in order for this accuracy to be
meaningful, the layers themselves must be sufficiently stable, especially when exposed to the
atmosphere.
Macleod: Optical Coatings Page 227 Design through Manufacture

A method, similar in some respects, but with some definite advantages in interpretation, has
been devised by Lissberger and Ring[5]. It consists of measuring the reflectance or transmittance
at two wavelengths and finding the difference. In the original system, a monochromator was
used, containing a chopping system that switched the output of the monochromator from one
wavelength to another and back again. The AC signal from the detector was a measure of the
difference. Since the two wavelengths could be placed virtually anywhere within the region of
sensitivity of the detector, the method had greater flexibility than the Giacomo and Jacquinot
system. Greatest contrast in the two reflectance signals as a layer was being deposited could be
obtained by placing the two wavelengths at the points of greatest opposite slope in the
characteristic of the thin-film structure at the appropriate stage. When the signals at the two
wavelengths were equal, the output of the system passed through a null, and, if displayed on a
chart recorder, made detection of the terminal point of a particular layer, usually indicated by the
null, particularly easy to detect.
These techniques were developed at a time when electronics and computers were far less
advanced that today. Their construction reflected these limitations. But the fundamental
principles are still completely sound and the ideas inherent in these systems have since been
extended to broad spectral and the advances in detectors and in electronics and data analysis
have made them practical. Many of the advanced systems have been developed in industry and
frequently details have not been published. Those that have been written up rarely include
detailed descriptions of the precise way in which they are employed. Usually the technique
involves a comparison between the spectral characteristic which is actually obtained at any
instant, and that which is required at the instant of termination of the particular layer. In the
earlier systems this was carried out visually by displaying both curves on a cathode ray tube.
This works well when there is a close match between predicted and measured performance but
frequently errors in earlier layers, and changes in the characteristics of layers from what is
expected, cause the actual curves to differ to a greater or lesser extent from the predictions. In
these circumstances, there can be great difficulty in assessing visually the correct moment to
terminate a layer. The most recent systems, therefore, are usually linked to a computer which
calculates continuously a figure of merit. The variation of this figure of merit is not unlike that of
a normal single wavelength monitoring signal but the termination must be at the minimum value.
The signal can be displayed to a plant operator but more often is used in the completely
automatic termination of layers.
Details of scanning monochromator systems have been published by a number of authors. An
early description of such a system is that of Hiraga et al[6], where the scanning was carried out
by a rotating helical slit assembly, again a design forced by the state of the art in detectors and
electronics at the time of the invention.
Pelletier and his colleagues in Marseille[7, 8] developed two such systems with considerable
success. The first uses a stepping motor to rotate a grating and scan the system over a wide
wavelength region. The second uses a holographic grating with a flat spectrum plane in which is
situated a silicon photodiode array detector which can be scanned electronically. Such
approaches are being extended to commercial systems.
These are the basic principles of optical monitoring but the development of monitoring
techniques is really inseparable from the study of tolerances and errors. For convenience, and
because the two subjects were really considered separately in the early stages, tolerances are
Macleod: Optical Coatings Page 228 Design through Manufacture

considered under a different heading, but much about the most recent ideas and developments in
monitoring processes will be found there.

15.2 The Quartz Crystal Monitor


The normal modes of mechanical vibration of a quartz crystal have very high Q and can be
transformed into electrical signals by the piezoelectric properties of the quartz and vice versa.
The crystal acts, therefore, as a very efficient tuned circuit that can be coupled into an electrical
oscillator by adding appropriate electrodes. Any disturbance of its mechanical properties will
cause a change in its resonant frequency. Such a disturbance might be an alteration of the
temperature of the crystal or its mass. The principle of monitoring by the quartz crystal
microbalance (as it is called) is to expose the crystal to the evaporant stream and to measure the
change in frequency as the film deposits on its face and changes the total mass. In some
arrangements the resonant frequency of the crystal is compared with that of a standard (usually
outside the plant) and the difference frequency is measured, in others the number of vibrations in
a given time interval is measured digitally. Usually the frequency shift will be converted
internally into a measure of film thickness using film constants fed in by the operator. Since the
signal from the quartz crystal monitor changes constantly in the same direction it can be more
simply used in automatic systems than optical signals.
The mechanical vibrational modes of a slice of quartz crystal are very complicated. It has
been found possible to limit the possible modes and the coupling between them by cutting the
slice with respect to the axes of the crystal in a particular way, by proportioning the dimensions
of the slice correctly and by supporting the crystal in its holder in the correct way. Quartz crystal
vibrational modes also vary with temperature, some having positive temperature coefficient and
some negative, and the slice can be cut so that modes which have opposite temperature
dependence are intentionally coupled with the resultant effect of a resonant frequency
independent of temperature over a limited temperature range. The usual cut of crystal which is
used in thin-film monitors is the AT cut. This is cut from a slice which was oriented so that it
contained the x-axis of the crystal and was at an angle of 35°15' to the z-axis. The mode of
vibration is a high-frequency shear mode and the temperature coefficient is small over the range
-40°C to +90°C, of the order of ±10-6 °C-1 or slightly greater. The coefficient changes sign
several times throughout the range so that the total fractional change in frequency over the
complete range is only around 5×10-5 Usually the frequency chosen is around 5 MHz although
the range could be anything from 0.5 MHz to 50 or 100 MHz.
As the thickness of the evaporant builds up, the frequency of the crystal falls and the reduction
in frequency is proportional both to the square of the resonant frequency and to the mass of the
film deposited. In a typical arrangement the measurement of mass thickness can be carried out to
an accuracy of around 2%, adequate for most optical filters.
Unfortunately, the sensitivity of the crystal decreases with increasing build up of mass and the
total amount of material which can be deposited before the crystal must be cleaned is limited.
With existing crystals this makes them less useful for multilayer work, especially in the infrared,
where often a single crystal can not accommodate a complete filter. One way round this problem
is to place a screen over the filter which cuts down the material reaching it to a fraction of that
reaching the substrates in the batch. This, of course, reduces the accuracy. Alternatively more
crystals must be used during a deposition and there are multiple crystal heads available.
Macleod: Optical Coatings Page 229 Design through Manufacture

Because the crystal measures mass and not optical thickness, it must be calibrated separately
for each material used and, in fact, the differences in mechanical properties between layer
materials makes it advisable to use a different dedicated crystal for each material. Otherwise the
total mass already deposited may not be a good indication of current calibration.
The temperature of the crystal should not be allowed to rise above the end of the temperature-
insensitive region and this usually implies some kind of cooling, usually by passing water
through the crystal head. This implies that it will not always be possible to keep the crystal
sensor at the same temperature as the other substrates in the coating plant.
The operation of the crystal is not sensitive to alignment in the same way as an optical system.
This is a great advantage but it brings with it a temptation that must be resisted. Because it is
easy to place the crystal virtually anywhere in the chamber, this is exactly what sometimes
happens and crystals can even be situated outside the periphery of the coating jig. Care should be
taken to place the crystal in a position where its signal represents as closely as possible what is
actually being deposited on the substrates. If it is unduly subject to variations in source
uniformity then it is unlikely that there will be a good correspondence between its output and the
actually deposited thickness it is controlling.
Although the mass measured by the quartz crystal monitor is somewhat removed from optical
thickness, provided care is taken in positioning and operating the monitor correctly, the results of
process control by quartz crystal are very satisfactory. The two great advantages of simplicity of
interpretation of signal and flexibility in installation outweigh the disadvantages. The accuracy is
not very different from that achievable by an optical system in single-layer monitoring and so it
is suitable for a wide range of coatings. Narrowband filters and other coatings whose successful
production depends on error compensation processes that are naturally present in the optical
monitoring are better produced using optical techniques.

15.3 Tolerances
The question of how accurately we must control the thickness of layers in the deposition of a
given multilayer is surprisingly difficult to answer and has attracted a great deal of attention over
the years.
One of the earliest approaches to the assessment of errors permissible in multilayers was
devised by Heavens[9] who used an approximate method based a matrix method His method,
useful mainly when calculations must be performed manually, consisted of a technique for
recalculating fairly simply the performance of a multilayer with a small error in thickness in one
of the layers. He showed that the final reflectance of a quarter-wave stack is scarcely affected by
a 5% error in any one of the layers.
Lissberger[10] developed a method for calculating the performance of a multilayer involving
the reflectances at the interfaces. In multilayers made up of quarter-waves, the expressions took
on a fairly simple form which permitted the effects of small errors, in any or all of the layers, on
the phase change caused in the light reflected by the multilayer to be estimated. Lissberger's
results, applied to the all-dielectric Fabry-Perot, or single-cavity, filter, show that the most
critical layer is the spacer. The layers on either side of the spacer layer are next most sensitive,
and the remainder of the layers progressively less sensitive the further they are from the spacer.
Macleod: Optical Coatings Page 230 Design through Manufacture

Giacomo, Baumeister and Jenkins[11] examined the effects on the performance of


narrowband filters of local variations in thickness, or 'roughness', of the films. This involved the
study of the influence of thickness variations in any layer on the peak frequency of the complete
filter. The treatment was similar in some respects to that of Lissberger. For the conventional
Fabry-Perot filter, layers at the center had the greatest effect. If all layers were assumed equally
rough, the design least affected by roughness would have all the layers of equal sensitivity and
attempts were made to find such a design. A phase-dispersion filter gave rather better results
than the simple Fabry-Perot, but still fell short of ideal.
Baumeister[12] introduced the concept of sensitivity of filter performance to changes in the
thickness of any particular layer. The method involved the plotting of sensitivity curves over the
whole range of useful performance of a filter, curves which indicated the magnitude of
performance changes due to errors in any one layer. His conclusions concerning a quarter-wave
stack were that the central layer is the most sensitive and the outermost layers least sensitive. An
interesting feature of these sensitivity curves for the quarter-wave stack is that the sensitivity is
greatest nearest the edge wavelength. This is confirmed in practice with edge filters where errors
usually produce more pronounced dips near the edge of the transmission zone than appear in the
theoretical design.
Smiley and Stuart[13] adopted a different approach using an analogue computer. There were
some difficulties involved in devising an analogue computer, but, once constructed, it possessed
the advantage at the time that any of the parameters of the thin-film assembly could be easily
varied. A particular filter which they examined was: Air | 4H L 4H | Air with nH=5.00 and
nL=1.54. Errors in one of the 4H layers and in the L layer were investigated separately. They
found that errors greater than 1% in one 4H layer had a serious effect, errors of 5%, for example,
caused a drop in peak transmittance to 70% and errors of 10% a drop to 50% together with
considerable degradation in the shape of the pass band. Errors of up to 10 % in the L layer had
virtually no effect on either the shape of the pass band or on the peak transmittance. The design
of filter is that of a very simple multiple cavity design where the 4H layers represent the cavities
and the L layer is the coupling layer in between. This result is therefore consistent with the other
results showing the increased sensitivity of the spacer layer to independent thickness errors.
An investigation was performed by Heather Liddell as part of a study reported by Smith and
Seeley[14] into some effects of errors in the monitoring of infrared Fabry-Perot filters of
designs:

Air | HLHL HH LHLHL | Substrate

and

Air | HL HH LHL | Substrate.

A computer program to calculate the reflectance of a multilayer at any stage during deposition
was used. Monitoring was assumed to be at or near a frequency of four times the peak frequency
(i.e. a quarter of the desired peak wavelength) of the completed filter. It was shown that, if all
Macleod: Optical Coatings Page 231 Design through Manufacture

layers were monitored on one single substrate, then, provided the form of the reflectance curve
during deposition was predicted, and it was possible to terminate layers at reflectances other than
turning values, there could be an advantage in choosing a monitoring frequency slightly removed
from four times peak frequency. If no corrections were made for previous errors, then a distinct
tendency for errors to accumulate in even-order monitoring (that is monitoring frequency an
even integer times peak frequency) was noted.
The major problem in tolerancing is that real errors cannot be treated as small. That is to say
that first order approximations are usually unrealistic because the error in one layer interacts
non-linearly with the errors in other layers. First order approximations effectively treat them
separately and can not include the interactions. This makes many of the analytical studies of
error sensitivity less useful.
In recent years the most satisfactory approach for dealing with the effects of errors and the
magnitude of permissible tolerances has been found to be the use of Monte Carlo techniques. In
this method, the performance of the filter is calculated, first with no errors and then a number of
times with errors introduced in all the layers. In the original form of the technique introduced by
Ritchie[15] and still that most often used, the errors are thickness errors and completely random
and uncorrelated. They belong to the same infinite population, taken as normal with prescribed
mean and standard deviation. The performance curves of the filter without errors and of the
various runs with errors are calculated.
Although statistical analyses of the results can be made, it is almost always sufficient simply
to plot the various performance curves together, when visual assessment of the effects of errors
of the appropriate magnitude can be made.
The method really provides a set of traces which reproduce, as far as possible, what would
actually be achieved in a succession of real production batches. The characteristics of the infinite
normal population can be varied and the procedure repeated. It is sufficient to calculate some
eight or perhaps ten curves for a set of error parameters. The level of error at which a satisfactory
process yield would be achieved can then readily be determined. Although the errors are usually
drawn from a normal population, the type of population has little effect on the nature of the
results. Normal distributions are convenient to program, and since there is no strong reason for
not using them and because errors made up of a number of uncorrelated effects are well
represented by normal distributions, most error analyses employ them.
Figure 15-6 shows some examples of plots where the errors are simple independent thickness
errors of zero mean. From these and similar results we find that the errors which can be tolerated
in a longwave-pass filter are normally of standard deviation between 2% and 5 %, in a
shortwave-pass filter around 2.5 %, and in an antireflection coating such as the quarter-half-
quarter around 2% to 3 %. Of course all depends on the performance deviation from theoretical
that can be tolerated.
Macleod: Optical Coatings Page 232 Design through Manufacture

Design3: Transmittance
100

80
Transmittance (%)

60

40

20

0
400 500 600 700 800 900 1000 1100 1200

Wavelength (nm)

Figure 15-6. The theoretical performance of a 17-layer longwave pass filter with no errors.

Design3: Transmittance
100

80
Transmittance (%)

60

40

20

0
400 500 600 700 800 900 1000 1100 1200

Wavelength (nm)

Figure 15-6b. The performance of the filter of Figure 15-6a with thickness errors of 2% standard deviation.
Macleod: Optical Coatings Page 233 Design through Manufacture

Design3: Transmittance
100

80
Transmittance (%)

60

40

20

0
400 500 600 700 800 900 1000 1100 1200

Wavelength (nm)

Figure 15-6c. The performance of the filter of Figure 15-6a with random thickness errors of 5% standard
deviation.

MONITOR5: Reflectance
5

4
Reflectance (%)

0
400 450 500 550 600 650 700
Wavelength (nm)
Figure 15-7. An antireflection coating of the quarter-half-quarter type on glass showing the effect of random
thickness errors of 2% standard deviation. The intense black line is the performance without errors. There are
ten error curves in the diagram. Some are obscured by the theoretical performance.
Macleod: Optical Coatings Page 234 Design through Manufacture

Design3: Transmittance Design3: Transmittance


100 100

80 80
Transmittance (%)

Transmittance (%)
60 60

40 40

20 20

0 0
960 980 1000 1020 1040 960 980 1000 1020 1040
Wavelength (nm) Wavelength (nm)
Figure 15-8. Random thickness errors of 0.5% standard deviation (right hand plot) are marginal for this two-
cavity filter of design Air | (HL)6 (LH)6 L (HL)6 (LH)6 | Glass. nH=2.4 and nL=1.45. The left-hand curve
shows the filter characteristic with no errors.
In a multiple cavity filter of the type in Figure 15-8 the permissible errors are not greater than
0.5% while, for narrower filters or filters with greater number of cavities, the tolerances must be
tighter. In fact, a rough guide is that the permissible standard deviation is not greater than the
halfwidth of the filter. In a Fabry-Perot or single-cavity filter the main effect of random errors is
a peak wavelength shift, the shape of the pass band being scarcely affected even by errors as
large as 10 %. The standard deviation of the scatter in peak wavelength is usually slightly less
than the standard deviation of the layer thickness errors so that some averaging process is
operating, although the orders of magnitude are the same. Figure 15-9 and Figure 15-10 illustrate
this in the case of a narrow single-cavity filter.

Transmittance (%)
100

80

60

40

20

0
1548 1549 1550 1551 1552
Wavelength (nm)

Figure 15-9. A narrow-band filter of single-cavity construction with design: Air | L (HL)^11 HH (LH)^11 |
Glass with H - Ta2O5 and L - SiO2.
A system of monitoring in which the thickness errors in different layers are uncorrelated
requires that each layer should be controlled independently of the others. In this type of
monitoring, therefore, we cannot expect high precision in the centering of narrowband Fabry-
Macleod: Optical Coatings Page 235 Design through Manufacture

Perot filters and we foresee great difficulties in being able to produce narrowband multiple-
cavity filters at all.

Transmittance (%)
100

80

60

40

20

0
1548 1549 1550 1551 1552
Wavelength (nm)

Figure 15-10. The single-cavity narrow-band filter with random thickness errors of 0.1%.
This monitoring arrangement is what we have called indirect. Systems where each layer is
controlled on a separate monitoring chip are of this type. There are difficulties with monitoring
of low-index layers on a fresh glass substrate because of the small changes in transmittance or
reflectance, and so the monitoring chips are usually changed after a low-index layer and before a
high index, two or four layers per chip being normal. Sometimes these layers will be monitored
to turning values. More frequently what is sometimes called level monitoring will be used. Here
the layer reflectance or transmittance signal is terminated at a point removed from the turning
value where the signal is still changing, leading to an inherently greater accuracy. This approach
involves what is really an absolute measurement of reflectance or transmittance, and so the
termination point is frequently chosen to be after a turning value rather than before, so that the
extremum can be used as a calibration. This usually implies a shorter wavelength for monitoring
or the introduction of a geometrical difference between batch and monitor, placing the monitor
nearer the source or placing masks in front of the batch.
Narrowband filters are not normally monitored in this way. Instead, all the layers are
monitored on the same substrate, usually the actual filter being produced, a system known as
direct monitoring. At the peak wavelength of the filter, the layers should all be quarter-waves or
half-waves, and so we can expect a signal which reaches an extremum at each termination point.
The accuracy cannot therefore be particularly high for any individual layer and, at first sight, it
would appear that the achievable accuracy should be far short of what must be required. Since
each layer is being deposited over all previous layers on the monitor substrate, then there is an
interaction between the errors in any layer and those in the previous layers not included in the
tolerancing calculation described above. We require a technique which models the actual process
as far as possible. Each layer can be considered as deposited over the multilayer which precedes
it, rather than on a completely fresh substrate. The results of such a simulation are shown in
Figure 15-11, (see also Macleod[16]) which demonstrates the powerful error compensation
mechanism that has been found to exist. The compensation has also been independently and
simultaneously confirmed by Pelletier and his colleagues[17]. Its nature is perhaps best
explained by the use of an admittance diagram.
Macleod: Optical Coatings Page 236 Design through Manufacture

Transmittance (%)
100

80

60

40

20

0
1548 1549 1550 1551 1552
Wavelength (nm)

Figure 15-11. Ten simulated runs of the single-cavity narrow-band filter with noise of 0.01% standard
deviation over the monitoring signal. Monitoring wavelength 1550nm

20

10 Layer number
10 20 30 40 50
0

-10

-20

-30

-40
Error (%)

Figure 15-12. The magnitude of the errors actually committed is illustrated by those calculated from one of
the designs in Figure 15-11.
Macleod: Optical Coatings Page 237 Design through Manufacture

1.0 Transmittance (Signal Units)


Re(Admittance) 95
0.5
0.5 1.0 1.5 2.0 2.5 3.0 3.5 90
0.0 85
Error
-0.5 80

-1.0 75
70
-1.5 0 100 200 300 400 500
Im(Admittance) Distance from Chip (nm)

Figure 15-13. The admittance locus of the first two layers of the filter of Figure 15-9 when there is an
overshoot in the first layer of around 1/8th wave optical thickness.
Figure 15-13 shows such a diagram drawn for two quarterwaves. Since both the isoreflectance
contours and the individual layer loci are circles centered on the real axis, the turning values
must always occur at the intersections of the loci with the real axis regardless of what has been
deposited earlier. At the termination point of each layer there is the possibility of restoring the
phase to zero or π. As far as any individual layer is concerned, it is principally the over- or under
shoot of the previous layer that affects it. If the previous layer is too thick, the current one will
tend to be thinner to compensate and vice versa. Of course it is impossible completely to cancel
the effects of an error in a layer. The process is actually transforming the thickness errors into
errors in reflectance at each stage since the loci will be slightly displaced from their theoretical
position. This is not a serious error, As can be guessed from the shape of the diagram, the
reflectance error is a second-order effect. Since the phase is self-corrected each time a layer is
deposited, the peak wavelength of the filter will remain at the desired value, that of the
monitoring wavelength. The remaining error, the residual one in reflectance, is then translated
into changes in peak transmittance and halfwidth. Since the reflectance change is always a
reduction, the bandwidth of an actual filter is invariably wider than theoretical. The peak
transmittance falls to the extent that the reflectances on either side of the cavity layer are
unbalanced. This is usually quite small and the reduction in peak transmittance is generally much
less important that the increase in bandwidth.
In this monitoring arrangement, thickness errors in any individual layer are a combination of a
compensation of the error in the previous layer together with the error committed in the layer
itself. The magnitude of the thickness errors can be quite misleading in interpreting whether or
not the filter can be made successfully. In Figure 15-11, for example, thickness errors of the
order of 50% occur in some layers and yet all the filter characteristics are useful ones.
The important characteristic is actually the error in reflectance or transmittance in determining
the turning values, and it is possible to develop theoretical expressions which relate the
reflectance or transmittance errors to the reduction in performance of the final filter[16]. This
analysis includes an assessment of the sensitivity of each layer to errors which indicate those
layers where the greatest care in monitoring should be exercised. These can be different from the
thickness sensitivity of Lissberger[10] already mentioned.
Macleod: Optical Coatings Page 238 Design through Manufacture

With high-index spacer layers, greatest sensitivity is found in the low-index layers following
the spacer, while with low-index spacers, the spacer itself has the highest sensitivity. A feature of
this analysis is that it demonstrates that for any particular error magnitude, there is a point where
improved halfwidth does not result from an increase in the number of layers because the effect of
errors is increasing more rapidly than the theoretical decrease in bandwidth. Then it is necessary
to move to second- and higher-order spacers if decreased bandwidth is to result. This
corresponds to what is found in practice. The error analysis also demonstrates that high-index
spacers are to be preferred over low-index. We have already seen that high-index spacers give
decreased angular sensitivity and greater tuning range.
Formulae which permit the calculation of the errors in reflectance, in halfwidth and in peak
transmittance as a function of the magnitude of the random errors in determining the turning
values exist[16], but for most purposes a computer simulation will suffice. It should be noted that
the compensation is effective only for the first order. Second-order monitoring, that is
monitoring at the wavelength for which the layers are all half-waves, is not effective in
preserving the peak wavelength. We can understand this because the admittance diagram is quite
different and so the compensation is of a different nature. Likewise, third-order monitoring is not
as effective as first-order, and, although the scatter in peak wavelength is less than that obtained
with second order monitoring, it is, nevertheless, quite large.
Multiple cavity filters are similar in behavior but there are some complications. The coupling
layers in between the various cavity sections of the filter turn out to be particularly sensitive to
errors in a rather peculiar way.
There are some cases where the errors in the layers are better completely decoupled from each
other. Antireflection coatings tend of be of this type. The use of precoatings in antireflection
coatings to help to decouple the errors has been studied[18].

15.4 Multiple Cavity filters


Multiple-cavity filters are particularly important in telecommunications applications as
wavelength division multiplexing beamsplitters. This section is written primarily with this
application in mind and the examples are all drawn from Dense Wavelength Division
Multiplexing (DWDM) applications.
Simple Monte Carlo modeling, Figure 15-14, shows that the allowable tolerances for random
thickness errors in a typical three-cavity filter for channel spacing of 200GHz are around
0.003%. This is several orders of magnitude beyond the capacity of current thickness control
techniques. Fortunately the already discussed inherent error compensation in direct first-order
turning-value monitoring, where all layers are controlled at the peak wavelength and on the filter
under construction, makes the production of such filters possible. For the direct monitoring to be
possible the spectral bandwidth of the monitoring beam must be considerably less than the
bandwidth of the filter under construction. A factor of 0.3 is usually adequate. In this study we
use a monitoring spectral bandwidth of 0.01nm.
Macleod: Optical Coatings Page 239 Design through Manufacture

Wavelength (nm)

1548 1549 1550 1551 1552


0.0
-0.2
-0.4
-0.6
-0.8
-1.0
log(Transmittance) (dB)

Figure 15-14. The effects of random thickness errors of 0.003% standard deviation. The results of ten
different sets of errors drawn from the same normal infinite population are shown. Errors of 0.004% gave
performance outside specification.

Relative Sensitivity
200
Cavity Cavity Cavity
150
Coupling Coupling
100

50

0
10 20 30 40 50 60 70 80 90
Layer Number

Figure 15-15. The error sensitivity of the layers of the three-cavity filter shows that the most sensitive layers
in the filter are not the cavities but the layers following the cavities. The coupling layers are also insensitive
but as seen later they present more special problems.
We have already mentioned the idea of sensitivity to monitoring errors for each layers in
terms of the proportional shift along the real axis of the terminating admittance of the filter as a
function of the reflectance or transmittance error in terminating the individual layer[16]. Such an
analysis is illustrated in Figure 15-15 for the three-cavity filter. The most sensitive layers in the
system are those following the cavity layers. The reason lies partly in the rather flat extremum
exhibited by these layers illustrated by the simulation in Figure 15-16. The coupling layers, that
is the layers midway between each pair of cavity layers, present a special case as we shall see.
Macleod: Optical Coatings Page 240 Design through Manufacture

Cavity layer

Figure 15-16. A simulated monitoring record for first-order turning-value monitoring of three layers in a
narrow-band filter. The flat extremum of the layer following the high-index cavity layer is largely responsible
for the greater error sensitivity of the layers following the cavities. Note the much sharper extrema of both the
layer before the cavity and the cavity itself making them rather less sensitive to error.
The study of the monitoring problem of the filters is assisted by Monte Carlo models of the
process. In the results quoted here the simulation was very simple, involving a required signal
reversal of a specified magnitude before recognition of a turning point. The final two matching
layers of non quarter-wave thickness were controlled by level rather than extremum. Results in
the presence of signal noise of 0.05% in transmittance with a prescribed signal change of 0.2%,
not particularly demanding figures, give the characteristics in Figure 15-17. The characteristics
are close to acceptable with the exception of the misshapen peak. This type of distortion of peak
shape can usually be attributed to the coupling layers.

Wavelength (nm) Wavelength (nm)


1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
0 0.0
-10 -0.2
-0.4
-20
-0.6
-30 -0.8
-40 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 15-17. Signal noise of standard deviation 0.05% absolute, together with prescribed signal reversals of
0.2%, give these results. The misshapen top of the characteristics means that they just fail the specification.
The monitoring curve for the layers around one of the coupling layers is illustrated in Figure
15-18. The admittance locus, Figure 15-19, helps us to understand the problem. The layer before
the coupling layer and the coupling layer have loci on the same side of the real axis so that a
negative error in the first layer will be compensated by a negative error in the coupling layer, and
vice versa. Because of the very different sizes of the loci, the coupling layer will also suffer from
a greatly magnified error. Good transmittance of the final filter at the monitoring wavelength is
assured, however, because of the low error sensitivity of the coupling layer implying that the
admittance error suffered by the complete filter will be small. Unfortunately, the changed
coupling layer thickness together with the wrongly compensated error in the neighboring layer
moves the adjacent cavities either further apart or closer together. It is this that causes the
distortion of the pass-band shape[19].
Macleod: Optical Coatings Page 241 Design through Manufacture

Transmittance (Signal Units)


62
95
90
85
Coupling
80
64 63 60
75 61
70
7000 7200 7400 7600 7800 8000 8200
Distance from Chip (nm)

Figure 15-18. Monitoring curve calculated for the region around one of the coupling layers of the filter. Note
the shallow characteristic of the coupling layer that has the same sense of terminating extremum as one of the
adjacent layers.

This feature
Im(Admittance) gives incorrect
1.0
61 compensation 0.1
Re(Admittance)

0.5 62 1.5 1.6


0.0

0.0
1.5 2.0 2.5 3.0 3.5 -0.1
-0.5 Re(Admittance)
63 -0.2
-1.0 Coupling Im(Admittance)

Figure 15-19. The admittance locus of the coupling layer together with the layers before and after shows that
there is an unavoidable feature that leads to a thickness error compensation of the incorrect sense. Also the
small size of the coupling layer locus im-plies an enormous magnification of the error.
Performance envelopes[20] for the coupling layer help us to confirm that it is indeed a
coupling layer error. The lower, minimum, envelope is a rather narrow spike with very high peak
transmittance. The upper, maximum, envelope is similar in shape to the filter characteristic but
rather broader. In Figure 15-20 these envelopes are superimposed over the error curves of Figure
15-17 and the closeness of fit suggests strongly that the coupling layers are responsible for the
problem.
Macleod: Optical Coatings Page 242 Design through Manufacture

Wavelength (nm) Wavelength (nm)

1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
0 0.0

-10 -0.2
Upper -0.4
-20
-0.6
-30 Lower -0.8
-40 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 15-20. The envelopes of the coupling layer are shown on the left. On the right they are superimposed
over the results of the monitoring simulation. Except for a very small excursion over the top of the upper
envelope the curves are entirely within the envelopes suggesting strongly that the pass-band shape distortion
is due to the coupling layers.
There is another problem associated with the coupling layers. Errors may cause a movement
of the termination of the previous layer along the real axis. If this termination, which represents
the coupling layer start, moves to the admittance of the coupling layer – and only a small
movement is required – then the coupling layer signal becomes completely flat and impossible
for monitoring. A slightly larger error reverses the coupling layer signal.
Because of these problems a common practice is to control these layers separately. Simple
timing is sometimes used but this is more successful with stable sputtering processes than with
thermal sources. Separate optical monitoring is possible but rather complicated. Quartz crystal
monitoring that can readily achieve an accuracy of 2% in layer thickness is usually adequate and
is recommended. The results of monitoring the coupling layers separately are shown in Figure
15-21. Other parameters are as Figure 15-20. The characteristics are now virtually theoretical.

Wavelength (nm) Wavelength (nm)


1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
0 0.0
-10 -0.2
-0.4
-20
-0.6
-30 -0.8
-40 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 15-21. The monitoring of the coupling layers separately by quartz crystal microbalance with random
errors of 2% standard deviation gives results almost indistinguishable from theoretical performance.

15.4.1 Uniformity
We have so far neglected uniformity of deposition. To achieve uniformity of layer thickness,
substrates are rotated about the center of the machine above offset sources. Because the
Macleod: Optical Coatings Page 243 Design through Manufacture

monitoring must be direct, that is on the actual filter under construction, planetary systems can
not be used. In fact they are unsuitable also because of the problem to be discussed.
There are two aspects of any possible lack of uniformity. First there may be a radial variation
of thickness. This can be common to both high and low-index sources or it may vary slightly
from one source to another. Satisfactory filters of the current design require differential variation
of not more than 5% for quite satisfactory pass band shape. The peak wavelength is displaced
somewhat depending on the uniformity of both sources but this is not entirely unwelcome
because it permits achievement of a range of peak wavelengths in a single deposition cycle. A
relative variation of 10% gives a pass-band distortion slightly more than tolerable. The second
aspect of uniformity errors, uniformity around the filter at constant radius, is much more serious.
Such uniformity is assured by the smoothing of the variation of deposition rate around the track.
However, it depends on a complete number of turns during deposition of each layer. A fractional
turn over an offset source introduces errors varying over the filter surface. Source fluctuations
must also be slow compared with the time for a cycle. These errors are random, and random
errors, we know, must be less than 0.003%. Let us assume a maximum error for a partial turn of
25% of the thickness deposited in a complete rotation cycle. Then for errors not greater than
0.003% we must have a total number of revolutions for a single layer of 25/0.003, that is 8300
revolutions. If a layer deposition occupies 10 minutes this implies a rotational speed of at least
830 revolutions per minute. Such a high rotational speed is absolutely necessary and is one of the
factors that distinguishes the production of these filters from virtually all other thin film coatings.
Add to this the fact that the monitoring beam must pass through the filter and must not suffer
from rotational noise. This implies very stable rotation and very accurately parallel sides for the
substrate, which is usually rather thick for increased stability. The high compressive stress,
characteristic of the energetic processes, must not be permitted to bend the substrate significantly
during deposition. The rear surface of the filter may be antireflected during deposition to avoid
the very narrow fringes characteristic of accurate parallel-sided substrates.
It is common to reduce the thickness of the substrates after filter deposition. Then they may,
depending on the mode of use, be once again antireflected on the rear surface. They are then be
diced into small units, frequently around 1.4mm square, using techniques common in the
semiconductor industry.
The strict requirements for rotational speed may seem to indicate that sample and hold
techniques for monitoring where the optical properties are sampled once per revolution might
not work well. However, the rotational errors are random while the monitoring errors are subject
to compensation. The intervals between optical samples are not subject to the same restrictions
as the uniformity errors.

15.5 Other Possibilities


Pelletier and his colleagues[21] have studied theoretically the behavior of the maximètre types
of monitoring systems in the production of narrowband filters. They conclude that, as we would
expect, the accuracy of the system in the production of single layers is very much better than a
single-wavelength system. In the monitoring of narrowband filters all on one substrate there is a
compensation process operating like the turning value method but it is more complex in
operation. For very small errors in most layers the system works adequately, but for large errors
in most layers or small errors in certain critical layers, the errors accumulate in such a way as to
Macleod: Optical Coatings Page 244 Design through Manufacture

cause a drastic broadening of the bandwidth of a Fabry-Perot filter or complete collapse of a


multiple-cavity filter. Pelletier has introduced two concepts to describe this behavior. Accuracy
represents the error that will be committed in any particular layer without reference to the
multilayer system as a whole. Stability represents the way in which the errors accumulate as the
multilayer deposition proceeds. The accuracy of the maximètre is excellent and greater than in
the turning value method, but the stability in the control of narrowband filters is very poor and it
can easily become completely unstable. Subsidiary measurements are therefore required to
ensure stability if advantage is to be taken of the very great accuracy that is possible.
Narrowband filters and their monitoring systems have been surveyed in Macleod[22].
The concepts of accuracy and stability and the discovery that the one does not ensure the other
imply that different measurements may be necessary to ensure that both are simultaneously
assured.
All of these results show most strongly that the detailed study of tolerances can not be
separated from the study of the monitoring process itself, and, to a large extent from the study of
the deposition process. Later work on tolerances, therefore, includes much on the monitoring
process and for convenience we include it here rather than in the section on monitoring.
Elaborations of the straightforward monitoring procedure range from the simple to the
complex.
The traditional approach to the production of an optical coating is to perform a test run and
then to attempt the correction of the errors that occurred in a subsequent run, and so on. Klug and
colleagues[23] reported the results of an investigation of monitoring of antireflection coatings
where a process of inverse synthesis after the deposition of a coating in one single production
test run allowed the automatic determination of all actual refractive indices achieved and
thicknesses deposited. It was found that one test run was sufficient for the completely
satisfactory correction of all the errors. Subsequent production runs then showed exceptionally
small variations.
So far, the techniques discussed up to this point are ones where the monitoring plan is
established at the start of the process and then is followed to the conclusion of the coating. Any
error correction is automatically part of the monitoring procedure. However the obvious question
is whether some correction during production could be applied. The objective is not really to
eliminate the test run, it is very unlikely that manufacturers could be persuaded to dispense with
it, but to improve still further the quality of production.
A straightforward implementation of this idea is to provide a set of rules for a machine
operator so that the termination conditions for a layer can be made to depend on the error
committed in the termination of the previous layer. Willey[24] has introduced the term steering
for this procedure. It is rather like the steering of a motor vehicle when sometimes the sense of
response to the wheel must be reversed. In the case of the multiple-cavity filters it is clear that
the sense of the correction should be reversed for the condition in Figure 15-19. Other types of
filter can be more complicated. Accurately positioned edge filters are good candidates for this
approach. Steering also lends itself to automatic implementation.
All the complications that exist at one single wavelength suggest that it might be useful to
make a more comprehensive set of measurements of coating performance for monitoring
purposes. This leads to a system of monitoring in which simultaneous measurements are made at
Macleod: Optical Coatings Page 245 Design through Manufacture

a large number of wavelengths over a wide spectral region and a merit function representing the
difference between actual and desired signals is computed. The merit function can then be used
as a monitoring signal and layer deposition terminated when the merit function reaches a
minimum. Although perfect deposition should ensure a minimum of zero in the figure of merit,
inevitable errors in layer index and homogeneity perturb the result and the minimum in practice
is rarely exactly zero. The accuracy and stability of such a broadband system in the monitoring
of certain components such as beam splitters has been investigated by computer simulation[25,
26] and it is clear that correction of committed errors by alterations in the subsequent layers is a
powerful and viable technique. Sullivan and Dobrowolski[27] give a detailed account of the
results from a very comprehensive computer program both for simulating and controlling the
wide-band monitoring of optical coatings. The achievements are impressive.
Developments in computer control of processes are such that we can confidently expect such
systems to become fairly common. At the present time, however, the traditional single-
wavelength technique is still that most commonly used in production because in many
application it is completely satisfactory.
Quartz crystal monitoring, in which the mass rather than optical thickness is measured, seems
unlikely to possess powerful compensation. Yet simulation of a simple broadband system for
antireflection coatings comparing optical monitoring with quartz crystal have given results
which indicate that the quartz crystal is in no way inferior[28]. The relative merits of quartz
crystal and optical monitoring form a subject of constant debate and published results for quartz
crystal are impressive[29, 30]. It is clear that narrowband filters, if they are to be controlled in
peak wavelength, do require direct optical monitoring, but quartz crystal monitoring is suitable
for most other filter types. The general opinion, based to some extent on instinct, is that quartz
crystal monitoring is most suitable for production of successive batches of identical components.
For single runs of varying coating types, optical monitoring appears normally to be preferred.
Optical monitoring is also preferred in applications such as filters for the far infrared, where very
large thicknesses of materials are deposited in each coating run.

15.6 Chapter 15 References


1. Banning, Mary, Practical methods of making and using multilayer filters. Journal of the
Optical Society of America, 1947. 37: p. 792-792.
2. Polster, H D, A symmetrical all-dielectric interference filter. Journal of the Optical
Society of America, 1952. 42: p. 21-25.
3. Behrndt, K H and D W Doughty, Fabrication of multilayer dielectric films. Journal of
Vacuum Science and Technology, 1966. 3: p. 264-272.
4. Giacomo, P and P Jacquinot, Localisation précise d'un maximum ou d'un minimum de
transmission en fonction de la longeur d'onde. Application à la préparation des couches
minces. Journal de Physique et le Radium, 1952. 13(Suppl to No 2): p. 59A-64A.
5. Lissberger, P H and J Ring, Improved methods for producing interference filters. Optica
Acta, 1955. 2: p. 42-46.
Macleod: Optical Coatings Page 246 Design through Manufacture

6. Hiraga, R, N Sugawara, S Ogura, and S Amano, Measurement of spectral characteristics


of optical thin film by rapid scanning spectrophotometer. Japanese Journal of Applied
Physics, 1974(Suppl 2, Pt 1): p. 689-692.
7. Borgogno, J P, P Bousquet, F Flory, B Lazarides, E Pelletier, and P Roche,
Inhomogeneity in films: limitation of the accuracy of optical monitoring of thin films.
Applied Optics, 1981. 20: p. 90-94.
8. Flory, F, B Schmitt, E Pelletier, and H A Macleod, Interpretation of wide band scans of
growing optical thin films in terms of layer microstructure. Proceedings of the Society of
Photo-Optical Instrumentation Engineers, 1983. 401: p. 109-116.
9. Heavens, O S, All-dielectric high-reflecting layers. Journal of the Optical Society of
America, 1954. 44: p. 371-373.
10. Lissberger, P H, Properties of all-dielectric filters. I. A new method of calculation.
Journal of the Optical Society of America, 1959. 49: p. 121-125.
11. Giacomo, P, P W Baumeister, and F A Jenkins, On the limiting bandwidth of interference
filters. Proceedings of the Physical Society, 1959. 73: p. 480-489.
12. Baumeister, P W, Methods of altering the characteristics of a multilayer stack. Journal of
the Optical Society of America, 1962. 52: p. 1149-1152.
13. Smiley, V N and F E Stuart, Analysis of infrared interference filters by means of an
analog computer. Journal of the Optical Society of America, 1963. 53: p. 1078-1083.
14. Smith, S D and J S Seeley, Multilayer filters for the region 0.8 to 100 microns. 1968, Air
Force Cambridge Research Laboratories.
15. Ritchie, F S, Multilayer filters for the infrared region 10-100 microns. 1970. PhD Thesis,
University of Reading.
16. Macleod, H A, Turning value monitoring of narrow-band all-dielectric thin-film optical
filters. Optica Acta, 1972. 19: p. 1-28.
17. Bousquet, P, A Fornier, R Kowalczyk, E Pelletier, and P Roche, Optical filters:
monitoring process allowing the auto-correction of thickness errors. Thin Solid Films,
1972. 13: p. 285-290.
18. Macleod, H A and E Pelletier, Error compensation mechanisms in some thin-film
monitoring systems. Optica Acta, 1977. 24: p. 907-930.
19. Macleod, H A and D Richmond, The effect of errors in the optical monitoring of narrow-
band all-dielectric thin film optical filters. Optica Acta, 1974. 21: p. 429-443.
20. Macleod, Angus and Christopher Clark. Envelopes in optical coating design. in 43rd
Annual Technical Conference Proceedings. 2000. Denver: Society of Vacuum Coaters. p.
197-202.
21. Pelletier, E, R Kowalczyk, and A Fornier, Influence du procédé de contrôle sur les
tolérances de réalisation des filtres interférentiels à bande étroite. Optica Acta, 1973. 20:
p. 509-526.
Macleod: Optical Coatings Page 247 Design through Manufacture

22. Macleod, H A, Thin film narrow band optical filters. Thin Solid Films, 1976. 34: p. 335-
342.
23. Klug, W, M Boos, R Herrmann, and H Schwieker, Inverse synthesis for analyzing the
variations of spectral properties of optical multilayers from different coating runs.
Proceedings of the Society of Photo-Optical Instrumentation Engineers, 1986. 652: p.
118-123.
24. Willey, Ronald R, Practical Design and Production of Optical Thin Films. Optical
Engineering, ed. B. Thompson. 1996, New York: Marcel Dekker Inc.
25. Vidal, B, A Fornier, and E Pelletier, Wideband optical monitoring of nonquarterwave
multilayer filters. Applied Optics, 1979. 18: p. 3851-3856.
26. Sullivan, Brian T and J A Dobrowolski, Deposition error compensation for optical
multilayer coatings. I. Theoretical description. Applied Optics, 1992. 31(19): p. 3821-
3835.
27. Sullivan, Brian T and J A Dobrowolski, Deposition error compensation for optical
multilayer coatings. II. Experimental results - sputtering system. Applied Optics, 1993.
32(13): p. 2351-2360.
28. Macleod, H A, Monitoring of optical coatings. Applied Optics, 1981. 20: p. 82-89.
29. Laan, C J van der and H J Frankena, Monitoring of optical thin films using a quartz
crystal monitor. Vacuum, 1977. 27: p. 391-397.
30. Pulker, H K, Coating production: new ideas at a time of demand. Optical Spectra, 1978.
12(8): p. 43-46.
Macleod: Optical Coatings Page 249 Design through Manufacture

16 OPTICAL CONSTANT DERIVATION


From the point of view of coating design, the most important optical properties of thin films
are their optical constants. The measurement of these is not always a straightforward matter. We
have seen that the microstructure of the films is rather far from that of similar bulk material and
this microstructure not only perturbs the optical constants but also leads to complications such as
birefringence and inhomogeneity.
Derivation techniques for optical constants depend on many factors including
• accurate measurements
• a reliable model for the films
• a method of selecting the appropriate solution from the multiple ones obtainable.
It is particularly important to note that we never measure optical constants directly. We
measure aspects of thin film behavior that are then interpreted in terms of a model. The model
includes optical constants. If the model is to be used essentially to recalculate the input data then
the details of the model are to a large extent unimportant. It is when the parameters of the model
are applied to systems under circumstances some way removed from those corresponding to the
input data that problems occur. Then, if the real film does not correspond to the model, large
calculation errors may appear.
In the design of optical coatings the thin film model that is used is homogeneous and isotropic
and may include absorption. An inhomogeneous layer is treated as a combination of
homogeneous sublayers.
The parameters that describe such a film are n, the refractive index, k, the extinction
coefficient and d, the physical thickness.
Real thin films can be inhomogeneous, anisotropic, absorbing and scattering. Films that depart
too far from the ideal model are usually considered "badly behaved" and are avoided. Most films
are anisotropic but the optic axis is usually normal to the film surface and the effects of
anisotropy may be suppressed by making measurements at normal incidence only. Multilayer
coatings naturally have a large anisotropy with optic axis normal to their surface and this is
included in the use of the isotropic film model. Inhomogeneity, also common in films, is more
difficult to suppress. Films that scatter appreciably are usually rejected
Many methods for optical constant extraction exist and a useful account is given by
Heavens[1]. Measurement of the optical constants of thin films is also included by Liddell[2].
Here we shall be concerned with just a few methods that are frequently used.
Given the optical constants and thicknesses of any series of thin films on a substrate, the
calculation of the optical properties is straightforward. The inverse problem, that of calculating
the optical constants and thicknesses of even a single thin film, given the measured optical
properties, is much more difficult and there is no general analytical solution to the problem of
inverting the equations. For an ideal thin film there are three parameters involved, n, k and d, the
real and imaginary parts of refractive index and the geometrical thickness, respectively. Both n
and k vary with wavelength, which increases the complexity. The traditional methods of
Macleod: Optical Coatings Page 250 Design through Manufacture

measuring optical constants, therefore, rely on special limiting cases that have straightforward
solutions.
Perhaps the simplest case of all is represented by a quarterwave of material on a substrate,
both of which are lossless and dispersionless, that is, k is zero and n is constant with wavelength.
The reflectance is given by

F1− n
R=G
2
/ nsub I 2

H1+ n
f
2
f / nsub JK (16.1)

where nf is the index of the film, nsub that of the substrate and the incident medium is assumed to
have an index of unity. Then nf is given by

nf = n 1/ 2 FG 1 ± R IJ
1/ 2 1/ 2

(16.2)
sub
H1∓ R K1/ 2

where the refractive index of the substrate, nsub is known. The measurement of reflectance must
be reasonably accurate. If, for instance, the refractive index is around 2.3, with a substrate of
glass, then the reflectance should be measured to around one third of a percent (absolute ∆R of
0.003) for a refractive index measurement accurate in the second decimal place. It is sometimes
claimed that this method gives a more accurate value for refractive index than the original
measure of reflectance since the square root of R is used in the calculation. This may be so, but
the value obtained for refractive index will most likely be used in the subsequent calculation of
the reflectance of a coating, and therefore the computed figure can be only as good as the
original measurement of reflectance. In the absence of dispersion, the curve of reflectance versus
wavelength of the film will be similar to that in Figure 16-1. The extrema correspond to integral
numbers of quarterwaves, even numbers being halfwave absentees and giving reflectance equal
to that of the uncoated substrate, and odd corresponding to the quarterwave of equations (16.1).
It is easy to pick out those values of reflectance which correspond to the quarterwave condition.

Figure 16-1 The reflectance of a dispersionless, homogeneous, lossless thin film.


The technique can be adapted to give results in the presence of slight dispersion. The maxima
in Figure 16-1 will now no longer be at the same heights but, provided the index of the substrate
is known throughout the range, the heights of the maxima can be used to calculate values for
Macleod: Optical Coatings Page 251 Design through Manufacture

film index at the corresponding wavelengths. Interpolation can then be used to construct a graph
of refractive index against wavelength.
This simple method yields results which are usually sufficiently accurate for design purposes.
If, however, the dispersion is somewhat greater, or if rather more accurate results are required,
then the slightly more involved formulae given by Hass, Ramsay and Thun[3] must be applied. It
is still assumed that absorption is negligible. In a the curve of reflectance or transmittance of a
thin film possessing dispersion the extrema corresponding to odd quarterwaves are displaced in
wavelength from the true quarterwave values while the halfwave extrema are unchanged. The
shift is due to dispersion. In the absence of absorption, the extrema of R, T 1/R and 1/T must all
coincide. If we assume that the incident medium has index unity then

4
T=
iLMN1 − cosFGH 4π nλ d IJ OP (16.3)
nsub + 2 + n −1
sub + 0.5n−1
sub dn 2
f −1− n 2
sub +n 2
sub n −2
f
f f

KQ
Since the extrema of T and 1/T coincide, the positions of the extrema can be found in terms of
d/λ by differentiating the expression for 1/T and equating it to zero.

4π n f d f
iLMN FG IJ OP
−1
1 nsub + 2 + nsub
T
=
4
+
1
8nsub
n 2f − 1 − nsub
2
d 2
+ nsub n −f 2 1 − cos
λ H KQ (16.4)

so that

0=
b g
d 1/ T
b g
d d/λ

= 0.25n′ d n f
−1
sub iFGH
n f − nsub n −f 3 1 − cos
4π n f d f
λ
IJ
K (16.5)

d
−1
+0.5π nsub −1
n 2f − nsub − nsub + nsub n −f 2 n f + n′f iFGH df IJ sin 4π n d
λK λ
f f

b g
where n ′f = dn f / d d / λ That the equation is satisfied exactly at all halfwave positions can be
seen since both sind4π n d / λ i and
f f d
1 − cos 4π n f d f / λ i are zero. At wavelengths
corresponding to odd quarterwaves a shift does occur and this can be determined by
manipulating the equation into

2π n f d f
2
c
n5f − 1 + nsub h
n3f + nsub
2
nf Fn + d I
tan
λ
= −2π 4
n −n
f
2
sub
GH n′ λ JK
f

f
f
(16.6)

This equation has too many unknowns to solve directly for nf. A method of successive
approximations is usually the best approach. It is important that the dispersion of the substrate
Macleod: Optical Coatings Page 252 Design through Manufacture

should be considerably less than that of the film otherwise it will need to be taken into account
also.
If absorption is present then the above equation cannot be used. When absorption is heavy then
the extinction coefficient can be calculated from

1− R
= exp
FG
4π k f d f IJ
T λH K (16.7)

If something better is required then we can write

ψ=
T
=
Re nsub b g (16.8)
1 − R Re BC∗ c h
2π k f d f
If we write β = then we can show that near the extrema,
λ
1
ψ=
Fn
1+ G
n I (16.9)

Hn
s

f
+ f β
ns JK
λ 1− R − T
so that kf = ⋅ (16.10)
d i d
2π d f nsub / n f + n f / nsub i T

In the methods discussed so far, we have been assuming that the thickness of the film is
unknown, except as it can be deduced from the measurements of reflectance and transmittance,
and the extrema have been the principal indicator of film thickness. However, it is possible
accurately to measure film thickness in other ways, such as multiple beam interferometry, or
electron microscopy, or by using a stylus step-measuring instrument. Once there is an
independent accurate measure of physical thickness, the problem of calculating the optical
constants becomes much simpler. The most frequently used technique of this type, was devised
by Hadley[1]. Since two optical constants, nf and kf, are involved at each wavelength, two
parameters must be measured, and these can most conveniently be R and T. In the ideal form of
the technique, if now a value of nf is assumed, then by trial and error one value of kf can be
found, which, together with the known geometrical thickness and the assumed nf, yields the
correct measured value of R, and then a second value of kf that similarly yields the correct value
of T. A different value of nf will give two further values of kf, and so on. Proceeding thus, we can
plot two curves of kf against nf, one corresponding to the T-values and the other to the R-values,
and, where they intersect, we have the correct values of nf and kf for the film. The angle of
intersection of the curves gives an indication of the precision of the result. Hadley, at a time
when such calculations were exceedingly cumbersome, produced a book of curves giving the
reflectance and transmittance of films as a function of the ratio of geometrical thickness to
wavelength, with nf and kf as parameters, which greatly speeded up the process. Nowadays, the
method can be readily programmed and precision estimates incorporated. This method can be
Macleod: Optical Coatings Page 253 Design through Manufacture

applied to any thickness of film, not just at the extrema, although maximum precision is
achieved, as we might expect, near optical thicknesses of odd quarterwaves, while, at halfwave
optical thicknesses, it is unable to yield any results. As with many other techniques, it suffers
from multiple solutions particularly when the films are thick and in practice a range of
wavelengths is employed, which adds an element of redundancy and helps to eliminate some of
the less probable solutions. Hadley's method involves simple iteration and does not require any
very powerful computing facilities. Even in the absence of Hadley's precalculated curves, it can
be accommodated on a programmable calculator of modest capacity. It does, however, involve
the additional measurement of film thickness, which is of a different character to the
measurements of R and T.
There is a problem, however, with virtually all techniques that make independent
measurements of thickness. Unless the thickness is very accurately determined and the model
used for the thin film is well chosen, the values of optical constants that are derived may have
quite serious errors. The source of the difficulty is that the extrema of the reflectance or
transmittance curves are essentially fixed in position by the value of n and d. There is only a very
small influence on the part of k. Should the value for d be incorrect then there is no way in which
a correct choice of n can satisfy both the value and the position of the extremum. What happens,
then, is that the extremum position is assured by an apparent dispersion, usually enormous and
quite false, and the values of n are seriously in error, sometimes showing abrupt gaps in the
curve. The situation is often worse at the halfwave points than it is at the quarterwave ones but
even in between the extrema there are clear errors in level which tend to be alternately too high
and then too low in between successive extremum pairs. A technique that has been used to avoid
this difficulty is to permit some small variation of d around the measured value and to search for
a value that removes to the greatest extent the incorrect features of the variation of n.
A different, computer-based, approach, developed by Pelletier and his colleagues in
Marseille[4], retains the measurement of R and T, but, instead of an independent measure of film
thickness, adds the measurement of R', the reflectance of the film from the substrate side. Now
we have three parameters to calculate at each wavelength and three measurements and it might
appear possible that all three could be calculated by a process of iteration, just like the Hadley
method, but the precision possible to attain in such a process is poor and it would break down
completely in the absence of absorption. To overcome this difficulty, the Marseille method uses
the fact that the geometrical thickness of the film does not vary with wavelength, and therefore,
if information over a spectral region is used, there will be sufficient redundancy to permit an
accurate estimate of geometrical thickness. Then once the thickness has been determined, a
computer method akin to refinement finds accurate values of the optical constants nf and kf over
the whole wavelength region. For dielectric layers of use in optical coatings, kf will usually be
small, and often negligible, over at least part of the region and a preliminary calculation
involving an approximate value of nf is able to yield a value for geometrical thickness, which in
most cases is sufficiently accurate for the subsequent determination of the optical constants.
Given the thickness, R and T, as we have seen should in fact be sufficient to determine nf and kf.
But this would mean discarding the extra information in R', and so the determination of the
optical constants uses successive approximations to minimize a figure of merit consisting of a
weighted sum of the squares of the differences between measured T, R and R' and the calculated
values of the same quantities using the assumed values of nf and kf. Although seldom necessary,
Macleod: Optical Coatings Page 254 Design through Manufacture

the new values of the optical constants can then be used in an improved estimate of the
geometrical thickness, and the optical constants recalculated. For an estimate of precision, the
changes in nf and kf to change the values of T, R and R' by a prescribed amount, usually 0.3%,
are calculated. Invariably, there are regions around the wavelengths for which the film is an
integral number of halfwaves thick, where the errors are greater than can be accepted and results
in these regions are rejected. In practice the films are deposited over half of a substrate, slightly
wedged to eliminate the effects of multiple reflections, and measurements are made of R and R'
and T and T' on both coated and uncoated portions of the substrate. This permits the optical
constants of the substrate to be estimated, and the redundancy in the measurements of T and T',
the transmittance measured in the opposite direction, gives a check on the stability of the
apparatus. A very large number of different dielectric thin-film materials have been measured in
this way.
A particularly useful and straightforward family of techniques are known as envelope
methods. The envelope method was first described in detail by Manifacier, Gasiot and Fillard[5]
and was later elaborated by Swanepoel[6]. Provided the absorption in a thin film is small then
the transmittance at the quarter and halfwave points is a fairly simple function of nf, kf and df.
Unfortunately, the transmittances at these points for one single film can only be measured for
different wavelengths. The optical constants of the film are functions of wavelength and an
iterative process involving interpolation is necessary to extract their values. In their method,
therefore, Manifacier, Gasiot and Fillard begin by interpolating the actual values of transmittance
by drawing two envelope curves around the transmittance characteristic for the film. These
envelope curves are then supposed to mark the loci of quarterwave and halfwave points
assuming that the thickness of the film were to vary by a small amount. This gives at each
wavelength point two values of transmittance corresponding to the two envelopes and therefore
to the transmittances that a films of thickness an integral number of halfwaves or of an odd
number of quarterwaves would have at that particular wavelength. These transmittances are
denoted by Tmax and Tmin respectively for a film of high index on a substrate of lower index. For
such a film we can write

α=
b
C1 1 − Tmax / Tmin g
1/ 2

(16.11)
C 1 + bT g
1/ 2
1 max / Tmin

where

d i
α = exp −4π k f d f / λ
4π k f d / λ = m π bquarter or halfwave thicknessg
f

C = dn + n idn + n i
1 f 0 sub f
(16.12)
C = dn − n idn − n i
1 f 0 sub f

T = 16n n n α / bC + C α g 2 2
max 0 sub f 1 2

T = 16n n n α / bC − C α g 2 2
max 0 sub f 1 2
Macleod: Optical Coatings Page 255 Design through Manufacture

Then from (8.12) and (8.14), if we define N as

n02 + nsub
2
T −T
N= + 2n0ns max min (16.13)
2 Tmax Tmin

nf is given by

c
n f = N + N 2 − n02 ns2 h
1/ 2 1/ 2
(16.14)

Once nf has been determined, equation (16.11) can be used to find a value for α. The
thickness df can then be found from the wavelengths corresponding to the various extrema and
the extinction coefficient kf from the values of df and α. The method has the advantage of
explicit expressions for the various quantities, which makes it easily implemented on machines
as small as programmable calculators. Unfortunately, as with many of the other techniques, the
results can suffer from appreciable errors in the presence of inhomogeneity.
Computers bring the advantage that we no longer need to devise methods of optical constant
measurement with the principal objective of ease of calculation. Instead, methods can be chosen
simply on the basis of precision of results, regardless of the complexity of the analytical
techniques that are required. This is the approach advocated by Hansen[7], who has developed a
reflectance attachment making it possible to measure the reflectance of a thin film for virtually
any angle of incidence and plane of polarization, the particular measurements carried out being
chosen to suit each individual film.
For rapid, straightforward measurement of refractive index, a method due to Abelès[8] is
particularly useful. It depends on the fact that the reflectance for p-polarization is the same for
substrate and film at an angle of incidence that depends only on the indices of film and incident
medium and not at all on either substrate index or film thickness, except, of course, that layers
that are a halfwave thick at the appropriate angle of incidence and wavelength will give a
reflectance equal to the uncoated substrate regardless of index. It is fairly easy to use Snell's Law
and the expressions for p-admittances to give

tan θ 0 = n f / n0 (16.15)

The measurement of index reduces to the measurement of the angle θ0 at which the
reflectances are equal. Heavens[1] shows that the greatest accuracy of measurement is, once
again, obtained when the layer is an odd number of quarterwaves thick at the appropriate angle
of incidence. This is because there is then the greatest difference in the reflectances of the coated
and uncoated substrate for a given angular misalignment from the ideal. It is possible to achieve
an accuracy of around 0.002 in refractive index provided the film and substrate indices are
within 0.3 of each other, but not equal. Hacskaylo[9] has developed an improved method based
on the Abelès technique. It involves incident light that is plane polarized with the plane of
polarization almost but not quite parallel to the plane of incidence. The reflected light is passed
through an analyzer and the analyzer angle, for which the reflected light from the uncoated
substrate and from the film-coated substrate are equal, is plotted against the angle of incidence.
Macleod: Optical Coatings Page 256 Design through Manufacture

A very sharp zero at the angle satisfying the Abelès condition is obtained, which permits
accuracies of 0.0002 to 0.0006 in the measurement of indices in the range 1.2 to 2.3. It is not
necessary for the film index to be close to the substrate index.
Unfortunately, the behavior of real thin films is often more complicated than we have been
assuming. They are frequently inhomogeneous, that is, their refractive index varies throughout
their thickness. They tend also to be anisotropic, although little work has been done on this
aspect of their behavior, but the possibility should be borne in mind when considering which
methods to use for index determination.
Provided that the variation of index throughout the film is either a smooth increase or a
smooth decrease, so that there are no extrema within the thickness of the film, the highest and
lowest values being at the film boundaries, then we can use a very simple technique to determine
the difference in behavior at the quarterwave and halfwave points, which would be obtained with
an inhomogeneous film. We assume that the film is absorption-free and that its properties can be
calculated by a multiple-beam approach, which considers the amplitude reflection and
transmission coefficients at the boundary only. We assume that the index of that part of the film
next to the substrate is nb and that next to the surrounding medium is na. The corresponding
admittances are yb and ya. The only reflections that take place are assumed to be at either of the
two interfaces. There is one further complication, also indicated in the figure, before we can sum
the multiple beams to arrive at transmittance and reflectance. A beam propagating from the outer
surface of the film to the inner, is assumed to suffer no loss by reflection and, therefore, the
irradiance is unaltered. Since irradiance is proportional to the square of the electric amplitude
times admittance, a beam that is of amplitude Ea just inside interface a, will have amplitude
(ya/yb)Eb at interface b. The correction will be reversed in traveling from b back to a. This is in
addition to any phase changes. The inverse correction applies to magnetic amplitudes. Since the
correction cancels out for each double pass it does not affect the result for resultant reflectance
but it must be taken into account when the multiple beams are being summed for the calculation
of transmittance. The characteristic matrix for the layer is then given by[10]

LMb y / y g 1/ 2
cos δ
i sin δ OP
MM
b a
by y g P
a b
1/ 2
(16.16)
N ib y y g
a b
1/ 2
sin δ b y / y g cos δ PQ
a b
1/ 2

Now we consider cases where the layer is either an odd number of quarterwaves or an integral
number of halfwaves. We apply the expression (16.16) in the normal way and find the well
known relations

Fy −y
R=G 0 a yb / ysub IJ 2

for a quarter wave (16.17)


Hy +y 0 a yb / ysub K
and
Fy −y
R=G 0 a ysub / yb IJ 2

for a half wave (16.18)


Hy +y 0 a ysub / yb K
Macleod: Optical Coatings Page 257 Design through Manufacture

The expression for a quarterwave layer is indistinguishable from that of a homogeneous layer of
admittance (yayb)1/2: and so it is impossible to detect the presence of inhomogeneity from the
quarterwave result. The halfwave expression is quite different. Here the layer is no longer an
absentee layer and cannot therefore be represented by an equivalent homogeneous layer. The
shifting of the reflectance of the halfwave points from the level of the uncoated substrate in
absorption-free layers is a sure sign of inhomogeneity and can be used to measure it.
The Hadley method of deriving the optical constants, takes no account of inhomogeneity. Any
inhomogeneity, therefore, introduces errors. The Marseille method, however, includes halfwave
points and therefore has sufficient information to accommodate inhomogeneity. The matrix
expression is a good approximation when the inhomogeneity is not too large and when the
admittances ya and yb are significantly different from those of substrate and incident medium. To
avoid any difficulties due to the model, the Marseille group actually uses a model for the layer
consisting of ten homogeneous sublayers with linearly varying values of n but identical values of
k and thickness d. The halfwave points still give the principal information on the degree of
inhomogeneity. They are also affected by the extinction coefficient k and this has also to be
taken into account. One halfwave point within the region of measurement can be used to give a
measure of inhomogeneity that is assumed constant over the rest of the region. Several halfwave
points can yield values of inhomogeneity that can be fitted to a Cauchy expression, that is an
expression of the form

∆n B C
= A+ 2 + 4 (16.19)
n λ λ

The envelope method has also been extended[11] to deal with inhomogeneous films using the
inhomogeneous matrix expression for the calculations. The extinction coefficient k, as in the
Marseille method, is assumed constant through the film.
A further technique that is covered in the section on oblique incidence is that of ellipsometry.
Ellipsometry avoids the need for absolute measurements by using, at oblique incidence, the s-
response to calibrate the p-response, and vice versa. the light reflected from a simple surface at
oblique incidence is, in the general case, elliptically polarized and measurement of the ellipticity
and orientation yield the two ellipsometric parameters, ψ and ∆, where

ρp
tan ψ =
ρs
d
and ∆ = ϕ p − ϕ s i (16.20)

and the ellipsometric sign convention for ϕp is used, which differs by 180° from the thin film
convention used everywhere else in these notes.
The optical constants of a simple surface can be derived analytically from the ellipsometric
parameters but a surface coated with a thin film has a minimum of three unknown parameters, n,
k and d. Additional measurements are therefore needed and these can involve additional angles
of incidence, additional wavelengths, or both. There is more information in the section on
oblique incidence where ellipsometry is discussed.
Macleod: Optical Coatings Page 258 Design through Manufacture

16.1 Chapter 16 References


[1] O. S. Heavens, “Measurement of optical constants of thin films,” in Physics of Thin
Films, vol. 2, G. Hass and R. E. Thun, Eds. New York and London: Academic Press,
1964, pp. 193-238.
[2] H. M. Liddell, Computer-aided techniques for the design of multilayer filters. Bristol:
Adam Hilger Ltd, 1981.
[3] G. Hass, J. B. Ramsay, and R. Thun, “Optical properties of various evaporated rare earth
oxides and fluorides,” Journal of the Optical Society of America, vol. 49, pp. 116-120,
1959.
[4] E. Pelletier, P. Roche, and B. Vidal, “Détermination automatique des constantes optiques
et de l'épaisseur de couches minces: application aux couches diélectriques,” Nouvelle
Revue d'Optique, vol. 7, pp. 353-362, 1976.
[5] J. C. Manifacier, J. Gasiot, and J. P. Fillard, “A simple method for the determination of
the optical constants n, k and the thickness of a weakly absorbing thin film,” Journal of
Physics E, vol. 9, pp. 1002-1004, 1976.
[6] R. Swanepoel, “Determination of the thickness and optical constants of amorphous
silicon,” Journal of Physics E, vol. 16, pp. 1214-1222, 1983.
[7] W. Hansen, “Optical characterization of thin films: theory,” Journal of the Optical
Society of America, vol. 63, pp. 793-802, 1973.
[8] F. Abelès, “La détermination de l'indice et de l'épaisseur des couches minces
transparentes,” Journal de Physique et le Radium, vol. 11, pp. 310-314, 1950.
[9] M. Hacskaylo, “Determination of the refractive index of thin dielectric films,” Journal of
the Optical Society of America, vol. 54, pp. 198-203, 1964.
[10] F. Abelès, “Recherches sur la propagation des ondes électromagnetiques sinusoidales
dans les milieus stratifiés,” Annales de Physique, 12ième Serie, vol. 5, pp. 706-784, 1950.
[11] D. P. Arndt, R. M. A. Azzam, J. M. Bennett, J. P. Borgogno, C. K. Carniglia, W. E. Case,
J. A. Dobrowolski, D. P. Arndt, U. J. Gibson, T. T. Hart, F. C. Ho, V. A. Hodgkin, W. P.
Klapp, H. A. Macleod, E. Pelletier, M. K. Purvis, D. M. Quinn, D. H. Strome, R.
Swenson, P. A. Temple, and T. F. Thonn, “Multiple determination of the optical
constants of thin-film coating materials,” Applied Optics, vol. 23, pp. 3571-3596, 1984.
Macleod: Optical Coatings Page 259 Design through Manufacture

17 REVERSE ENGINEERING
We frequently extract parameters of individual layers from optical measurements on thin film
systems. In some cases we require these parameters for future predictions of performance of new
coating designs. In other cases we are interested in analyzing the makeup of an existing coating.
The derivation of such fundamental information about layers from measurements of overall
performance is frequently called reverse engineering. A frequent application of reverse
engineering is in the determination of errors that have been committed in the attempted, but less
than successful, manufacture of a given coating. We exclude completely from this discussion the
situation where the subject is a coating produced elsewhere, perhaps by a competitor. We assume
that we understand very well what we attempted to produce
The procedure is not unlike that of solving any puzzle. No two cases are alike. However, there
are some general rules and there are some pitfalls to be avoided. This chapter deals with these.

17.1 Random and Systematic


When something has gone wrong we usually want to understand not simply what is wrong
with the specific coating but also how to put it right.
Problems can exhibit two principal and different aspects of behavior, random and systematic.
Random problems are ones that fluctuate. A wavelength may be too long in one run and too
short in the next. An antireflection coating may have a green color in one run and a red color in
the next. Systematic problems are ones where the coating performance is biased in some way
from what is desired. For example, the wavelength of an important feature might be always a
certain amount too long, or a peak might appear in a characteristic where there should be none.
Of course most problems exhibit features that fall into both categories.
It is very important that the random and systematic natures of the problem should be identified
before any steps are taken to remedy it. Curing a random problem implies a greater degree of
control. Curing a systematic problem implies making a systematic change to one or more control
parameters.
The application of a systematic change to a random problem will simply make matters worse.
Random problems should always be solved before tackling systematic ones.
Imagine the situation where we require a feature at 500nm. The process fluctuates from
450nm to 550nm. We make a measurement of a product that gives us a value of 530nm. We
correct the monitoring to move this 30nm towards shorter wavelengths, Our fluctuation is now
from 420nm to 520nm, rather worse than before. To cure this problem we need to identify the
source of the fluctuation and put that right before we make any changes to our monitoring
wavelength.
First and foremost, therefore, is that we identify the random element of our problem and we
put them right. Once the problem has been converted into a systematic one, the solution is more
straightforward.
A random problem can not be identified from one result only.
Macleod: Optical Coatings Page 260 Design through Manufacture

An antireflection coating that was expected to be green in color is actually blue. The next
attempt produces a red color. Without that second run we could not have deduced that there is a
random element in the problem.
We should use all the information that is available to us. Here good records are of immense
importance. The pumping performance of the machine, the batch of materials, the time between
cleaning of substrates and coating. The operator in charge of the machine may have noticed
something. Only if we keep good records do we have access to this information. Reliable records
also imply reliable instrumentation. Unreliable instrumentation is likely to be a contributor to the
problem.
The most common source of random problems in production is probably a small water leak.
Almost all deposition chambers are supplied with large quantities of cooling water. A small
pinhole leak in a water line is self-sealing. A small block of ice forms because of the evaporation
cooling and this inhibits further evaporation. Pumping performance at this stage is unaffected.
During deposition, however, when the sources are operating, the added heat can melt the ice plug
so that suddenly water vapor is added to the internal atmosphere of the machine with, sometimes,
disastrous consequences. Common symptoms are unexpected variations in optical performance
and, frequently, environmental durability. When there is a random problem, especially in a
process that has been previously satisfactory, think water leak. A residual gas analyzer can help
in this respect but finding the actual leak can be a laborious matter of simply replacing seals and
suspect pipe.
The only way of avoiding random problems is good housekeeping. This includes strict
attention to procedures so that consistency is assured, machines in good condition, sensible
maintenance schedules, avoidance of panic (end of the month rush). A tidy coating laboratory
where casual visitors are discouraged is a good start. Clean-room conditions are not always
strictly necessary from a purely technical point of view but their beneficial psychological effect
on personnel can be enormous. Also casual traffic is completely discouraged by the need to don
special clothing.
Random problems are better avoided than solved.

17.2 Systematic Problems in General


The process of reverse engineering is allied to the process of design. We must derive a system
of layers having a predicted performance that is equal to that measured. However we must also
be sure that the structure we derive is sufficiently close to that of the actual component for useful
deductions to be made.
Two major problems present themselves. The first is that the model films may not be able to
reproduce sufficiently well the behavior of the real films. Rather more serious is the second
problem that the measurements themselves may be in error, and we shall look much more closely
at this one.
Measurements are most frequently made using spectrometers. Although the specifications for
photometric accuracy in high-performance instruments are normally in the range from 0.1% to
0.25% considerable attention to detail, and especially calibration, is required to achieve such
figures. Routine measurements are likely to be much less accurate.
Macleod: Optical Coatings Page 261 Design through Manufacture

17.3 Single layers


In the optical coating field we have become used to the fact that our thin films possess neither
the stability nor the exact optical constants of the corresponding bulk material. We make
measurements of the optical constants, n and k, of thin films in a number of ways but most often
using a spectrometer to measure the reflectance and transmittance of a film thick enough to
exhibit several fringes within the range of measured wavelength. These results are then adjusted
as measured information on actual coatings becomes available. In all of this the measured data
are fundamental. But what if they are in error?
To simplify the discussion we consider normal incidence measurements only.
We frequently describe the process as the measurement of n, k and d but in fact we never
make such measurements on thin films. The process is really the comparison of measures of film
behavior with calculations of similar behavior of a model of a thin film. The model involves
certain parameters that we label as n, k and d, and we adjust these parameters until we achieve
the best fit possible between measured and calculated behavior. The values of the parameters at
that stage are then taken as the “measured” values of n, k and d. The model of the film is crucial.
If the model behaves as the real thin film then we can rely on the extracted parameters as
representative of the real film. If the model behavior does not match that of the real film then the
extracted parameters will be deficient.
It is useful to employ the term stability in connection with optical constants. Normally
because of the extraction process, whatever the qualities of the model, when the extracted
parameters are used to recalculate the measured values, the agreement will be good. In this case
any deficiencies of the model are compensated by deficiencies in the parameter values. Such
deficiencies in the parameter values, however, become uncompensated when they are used in
order to predict behavior other than under the exact conditions of the original measurements.
Parameters are said to be stable when they are effective in the accurate prediction of behavior
and unstable when they are not. In virtually all cases it is desirable to have stable parameters.
Apart from the model there is also the possibility of errors in the actual measurement. These
errors, too, can have effects on the extracted optical parameters that are out of all proportion to
their magnitudes. The effect of such errors depends both on the model and on the techniques
used in the extraction process.
The simplest of the possible models is that of a homogeneous, absorption-free film on a
substrate that is also homogeneous and free of absorption. The rear surface of the substrate is
uncoated and parallel to the front, coated, surface.
To calculate the reflectance or transmittance of this system we need three parameters, the
geometrical thickness, d/λ, the refractive index, n, of the film and the refractive index, nsub, of
the substrate. Since there is no absorption, the transmittance and reflectance will sum to give
100% and so there would be no additional information derived from measuring transmittance as
well as reflectance. Thus there is insufficient information in any single measurement for it to be
possible to extract a value of index for the film unless some of the parameters are already known.
Measurements over a range of wavelengths help because the physical thickness, d, of the film
does not change with wavelength but this redundancy in physical thickness is insufficient for a
complete solution.
Macleod: Optical Coatings Page 262 Design through Manufacture

In order to make progress, we need more information and it is principally in this respect that
the various methods differ. We can consider two extremes. In the first case we know both the
index of the substrate and the physical thickness of the film, that is we lack information about the
film index only. In the second case we have no information whatsoever about the parameters of
substrate and film except that they fit the model.
The situation is made more difficult by the presence of the rear surface of the substrate. How
this is dealt with depends on the particular technique. In this discussion the rear surface is an
unnecessary complication and so we imagine either that the inversion has been accomplished or
that the measurement has been performed on wedged substrates so that the effect of the rear
surface has been or can readily be eliminated.

Indices 2.0, 1.9, 1.8, 1.7, 1.6, 1.5, 1.4, 1.3


10
9
8
Reflectance (%)

7
6
5
4
3
2
1
0
0 10 20 30 40 50
Layer Thickness (nm)
Figure 17-1. The reflectance at 1000nm of a thin layer of index varying from 2.0 (top curve) to 1.3 (bottom
curve) at intervals of 0.1. The substrate has an index of 1.52 and the incident medium of 1.00.
Very thin films that are less than a quarterwave in thickness are notoriously difficult to
characterize. Figure 17-1 shows the variation of reflectance as a function of physical thickness
for absorption-free films ranging in refractive index from 1.3 to 2.0 on a substrate of index 1.52.
The difference in reflectance is especially small for the lower values of refractive index. Even
with an accurate measure of film thickness the precision with which the index can be determined
is exceedingly poor. Only for accurate thickness measurements of around 50nm or greater and
indices superior to 1.7 would it be possible to make a reasonable determination of index in the
presence of reflectance errors not greater than 0.5%. If absorption is known to be present then
the task becomes virtually impossible. Of course for thin metallic films, where the absorption is
very high, techniques based on ellipsometry [74] or surface plasmon resonances [75] become
feasible.
Once the data contains at least one extremum the extraction of the necessary parameters
becomes more straightforward. At any given wavelength the reflectance is determined by three
parameters, and, in principle, if any two are known then the third can be calculated. This is the
basis of one of the two extremes that we are considering. We assume that we know the index of
the substrate and the thickness of the film.
Macleod: Optical Coatings Page 263 Design through Manufacture

The variation of reflectance with index of refraction for a given film geometrical thickness,
d/λ, and substrate refractive index, nsub, is shown in Figure 17-2. It is immediately clear that
there are multiple solutions and the greater the film geometrical thickness, the greater is the
number of possible solutions. The thin horizontal line indicates a measured value of reflectance.
All points of intersection between this line and the curve that corresponds to the geometric
thickness of the film are potential solutions. The multiple solution problem is usually overcome
by predicting a value for the index of the film from other available information such as the index
of similar bulk material or other aspects of film behavior and the adopted solution is that closest
to the predicted value. Figure 17-2 actually includes two curves, a bold one that corresponds to a
geometric thickness of 1.6 and a thin line representing one that is some 6% thinner. Should the
incorrect value be used rather than the correct one then the value of index derived for the film
will clearly be in error. It is also possible that the value of reflectance may be in error. Provided
the point of intersection simply moves up or down the appropriate curve of R against n the
extracted value of n should not be too remote from the correct value.
Now let us imagine that the measured value of R is close to the top or bottom of the
appropriate cycle. There will be two values of n that are close together and it will be difficult to
distinguish between them. Let the value of R be slightly in error, just enough so that there is no
longer an intersection with the correct peak. The allowable values of n will now be removed
quite far from the correct value such that even that nearest to the estimate is seriously in error.
This behavior results in the appearance of abrupt and usually serious discontinuities in the
curve of extracted n against wavelength with values that are usually clearly unrealistic. The
technique in this raw form is certainly far from stable. The usual way of dealing with this
problem is to smooth the results by forcing them to fit a normal dispersion formula such as the
Cauchy

bg
n λ = A + B λ2 + C λ4 (17.1)

This is reasonably effective provided the region over which the smoothing is carried out is
large, that is there are at least several fringes within it. A good example of the usefulness of such
smoothing is given by Pelletier and colleagues [76].
Macleod: Optical Coatings Page 264 Design through Manufacture
Geometrical thickness 1.6 (thick), 1.5 (thin)
0.6

0.5

0.4

Reflectance 0.3

0.2

0.1

0.0
1 2 3 4 5
Film index, n

Figure 17-2. The variation of reflectance (scale of 0 to 1) as a function of film index, n, for a thin film of
geometrical thickness 1.6 (bold line) and of 1.5 (narrow line). The substrate index is 3.0. The thin horizontal
line marks a reflectance of 0.15. Each intersection between that line and either of the curves is a potential
solution.
An alternative and inherently very stable technique actually uses less information about the
film and substrate combination [77-79]. Figure 17-2 demonstrates quite clearly that the two sets
of fringes denoted by the two curves share a common pair of envelopes. We can draw these
envelopes separately as in Figure 17-3. Now the measured reflectance will cut, at most two
values only of n and, except in the small region around the point where the film is a perfect
antireflection coating for the substrate, there will be virtually no uncertainty in the selection of
the correct solution. How is the reflectance of the envelope measured? This is the point where
there is some subjective input. The reflectance is found by drawing the envelopes on the curves
of film R, or T, against λ. At first sight it may seem that this is an operation subject to significant
errors but in fact errors, even when the envelopes are sketched manually, are usually small and
do not perturb the measured values of n significantly. Their effect lacks the violent excursions of
the first method. The stability of the method is high especially because of the absence of multiple
solutions in most cases. The envelope corresponding to half-wave film thicknesses should
coincide with the horizontal envelope in Figure 17-3. If it does not, it usefully indicates the
presence of absorption, inhomogeneity or an error of measurement. If we have no knowledge of
the substrate then this level can be used to give us that information.
Thus there are good viable techniques that minimize the effects of measurement inaccuracies
in the extraction of film parameters in very simple systems. There can be problems when the film
is rather more complex in its real behavior than can be accommodated by the model. Typical of
such films are those where the index varies cyclically throughout the thickness.
Macleod: Optical Coatings Page 265 Design through Manufacture

Envelopes
0.6

0.5

0.4
Reflectance
0.3

0.2

0.1

0.0
1 2 3 4 5
Film index, n
Figure 17-3. The envelopes of the curves in Figure 17-2. The horizontal line is associated with half-wave
thicknesses of the films, the alternate curve with quarter-wave thicknesses. This removes almost all the
ambiguity associated with the solution for n.

17.4 Problems of the model in n and k extraction


17.4.1 Measurements
The measurements that we consider to be available in this study consist of the normal
incidence reflectance, R, and transmittance, T, the ellipsometric parameters ψ and ∆, and the
physical thickness, d, of the film. These measurements are considered to be made after
deposition of the film and represent the full knowledge that we have of the film and its
properties. We assume also that we do have full details of the incident medium and the substrate.
Macleod: Optical Coatings Page 266 Design through Manufacture

17.4.2 Single homogeneous dielectric layers

Homogeneous, dielectric
40
2.50
Reflectance (%) 30

20

10

0 1.35
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Optical)
Figure 17-4. The reflectance in air of a dielectric, homogeneous thin film as a function of optical thickness
with index of refraction as a parameter varying from 1.35 (lowest reflectance) to 2.50 (highest reflectance) at
intervals of 0.05. The substrate index is 1.52.
The simplest film is one that is dielectric, that is, free of loss, and homogeneous and that can
be represented by a thickness and refractive index. In such a film the transmittance and
reflectance have the sum of unity and so measurement of one while the other is available yields
no extra information.
The variation of reflectance for a homogeneous dielectric film as a function of optical
thickness is well-known and is shown in Figure 17-4 where the range of possible refractive
indices has been limited to 1.35 to 2.50 corresponding roughly to those available in the visible
and near infrared regions. Clearly, if we know the substrate and incident medium indices then,
except for the region where the optical thickness is near an integral number of halfwaves,
knowledge of the optical thickness and the reflectance of the film gives an unambiguous value of
film refractive index. Unfortunately we are rarely in this situation.
A separate measurement of layer thickness will usually involve physical rather than optical
thickness. This means that the horizontal scale of the diagram is altered in inverse proportion to
the refractive index. The various curves of Figure 17-4 are therefore smeared out across each
other to give Figure 17-5.
Macleod: Optical Coatings Page 267 Design through Manufacture

Homogeneous. n = 1.35 to 2.50.


40
2.50
30

Reflectance (%)
20

10

0 1.35
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-5. The reflectance in air of a dielectric, homogeneous thin film as a function of geometrical
thickness (d/λ) with index of refraction as a parameter varying from 1.35 (lowest reflectance) to 2.50 (highest
reflectance) at intervals of 0.05. The substrate index is 1.52.
Now we see that there are only limited regions where there is a clear unambiguous
relationship between thickness, index and reflectance. The wider the range of possible indices
the narrower are the available regions. Elsewhere there are multiple solutions and the possibility
of jumping from one solution to another with discontinuities in the estimated refractive index.
The situation is even worse when a high index substrate is used in the measurement of a film
with lower refractive index. This is shown in Figure 17-6. Here the region of unique solution is
very small. Thus, even with this very simple model where the films are characterized by only
two parameters and one of these is known independently there are great problems in finding a
valid index of refraction. In fact measurement of a film index that is lower than that of the
substrate, especially if it is close to the value consistent with a reasonable antireflection coating,
should be avoided whenever possible.
Macleod: Optical Coatings Page 268 Design through Manufacture

Homogeneous. n = 1.35 to 3.00. nsub = 3.00


40

30

Reflectance (%)
20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-6. The reflectance in air of a dielectric homogeneous thin film as a function of geometrical
thickness. The substrate index is 3.0 and the film index ranges from 1.35 to 3.00 at intervals of 0.05. There
are only very limited regions where unambiguous solutions may be obtained.

Reflectance Psi. 70deg.


40
35
Reflectance Psi (deg)

30
1.35
25
20
15
10
5
2.50
0
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-7. The ellipsometric parameter ψ at 70°as a function of geometrical thickness (d/λ) with index of
refraction as a parameter varying from 1.35 (lowest reflectance) to 2.50 (highest reflectance) at intervals of
0.05. The substrate index is 1.52.
Macleod: Optical Coatings Page 269 Design through Manufacture

Reflectance Delta. 70deg.


180
135

Reflectance Delta (deg)


90
1.35
45
0
-45
-90
-135
2.5
-180
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-8. The ellipsometric parameter ∆ at 70°as a function of geometrical thickness (d/λ) with index of
refraction as a parameter varying from 1.35 to 2.50 at intervals of 0.05. The substrate index is 1.52.
Note that the difficulties exist because the parameter is going to be used in predictive
calculations. If it were simply to be used to recalculate the measured results then any of the
multiple values could be chosen with no influence on the result.
The same problems exist with ellipsometric measurements, Figure 17-7 and Figure 17-8.
There are several ways in which we can solve this problem. The most useful ones depend on
the availability of measurements over a range of wavelengths. The parameter against which the
performance is plotted is the geometric layer thickness, d/λ. Thus if measurements are made over
a sufficiently wide range of wavelength then there will be at least some measurements that are in
the unambiguous regions. The model will normally have only a few adjustable parameters and
the unambiguous values can be used to extract the parameters of the model. Usually the
parameter values will be refined by refinement. A Cauchy model is frequently used.
A different approach is to use the envelopes of the measured performance for the extraction of
index. This has the advantage that no physical model of the variation of index with wavelength is
required.
Real thin films are, unfortunately, still more complicated.

17.4.3 Absorbing Films


Absorption is another parameter that must be taken into account. When absorption is present
the sum of reflectance and transmittance is no longer unity. However, the absorption also affects
reflectance and transmittance quite differently. As an example we take a film of refractive index
2.00 on a substrate of index 1.52, corresponding to glass.
Macleod: Optical Coatings Page 270 Design through Manufacture

Effect of k = 0.005 on R
40
35
30

Reflectance (%)
k=0
25
20
15
10
5 k=0.005
0
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-9. The reflectance of an absorption-free film with index 2.00 (grey curve) compared with the
reflectance of a similar film with extinction coefficient 0.005 (black curve).

Efect of k = 0.005 on T
100
95
k=0
Transmittance (%)

90
85
80
75
k=0.005
70
65
60
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-10. The transmittance of an absorption-free film with index 2.00 (grey curve) compared with the
transmittance of a similar film with extinction coefficient 0.005 (black curve).
Absorption is usually expressed in terms of the extinction coefficient, k. The relationship
between k and the absorption coefficient α is α = 4πk/λ. It is clear from Figure 17-9 and Figure
17-10 that reflectance is hardly affected even by quite significant absorption and therefore does
not represent a good vehicle for the determination of extinction coefficient. Transmittance is
much better. The same is true of reflective ellipsometric measurements. They are not nearly as
sensitive to absorption as transmission measurements.

17.4.4 Inhomogeneous Films


In inhomogeneous films there is a variation of refractive index through the thickness of the
film. There may also be a variation of extinction coefficient but such variation is exceedingly
Macleod: Optical Coatings Page 271 Design through Manufacture

difficult to detect, even when it is large. We therefore concentrate on a refractive index variation
and, in the first instance, assume that there is no absorption so that the extinction coefficient is
zero.
If there is no absorption then the reflectance and transmittance records are completely
complementary. Thus the effect of inhomogeneity is equally detectable from transmittance or
reflectance measurements. However, since reflectance is insensitive to absorption, reflectance is
a better vehicle for the detection of inhomogeneity than transmittance.
Figure 17-11 shows the effect of inhomogeneity in a layer with index rising from substrate to
outer surface. This has displaced the turning values of reflectance from the value of the uncoated
substrate and is the sole indicator of inhomogeneity because the transmittance curves show a
reduced transmittance at the halfwave values that could be consistent with absorption. There is
similar displacement but in the opposite sense for a layer with index falling from that of the
substrate. In the second case the halfwave transmittance is increased and can not be interpreted
as absorption.

Inhomogeneous layer
40
35
30
Reflectance (%)

25
20
15
10
5
0
0.0 0.2 0.4 0.6 0.8 1.0
Layer Thickness (Geometric)
Figure 17-11. The reflectance of a layer with linear variation of refractive index from 1.905 next to the
substrate to 2.00 at the outer edge. The horizontal line is the substrate reflectance. The inhomogeneity in the
layer lifts its turning values away from the substrate reflectance at the halfwave points. There is little other
indication of inhomogeneity.

17.4.5 Incorrect Models


Incorrect models are more likely to exist in the interpretation of transmittance measurements
than reflectance.
We can look at a simple case. We construct a film with thickness 1000nm and index varying
linearly from 2.00 at the outer surface to 1.905 at the surface of the substrate. We now attempt to
interpret the transmittance of this film using a homogeneous absorbing film model.
We know that the total thickness of the film is 1000nm. But this still leaves us with two
adjustable parameters, n and k, but only the measurement of T at each wavelength. We know the
wavelengths that correspond to the extrema and we know exactly the thickness of the film. All
Macleod: Optical Coatings Page 272 Design through Manufacture

that is then necessary to be able to assign an index to each extremum is the order number of the
extremum. Trial and error will usually permit the choice of the correct series of order numbers
for the fringes. In this case we can quickly determine that the most appropriate solution is a
dispersionless index of 1.952. However, there is then no way that we can fit a reasonable value
of k to this. The value oscillates with the pitch of the fringes. If we try to arrange a reasonable
variation for k then the index shows serious discontinuities. This is typical of techniques that rely
too heavily on an exact value of layer thickness.
We therefore adopt the envelope approach that reduces dependence on thickness and also
introduces two effectively measured values at each wavelength. These are the maximum and
minimum envelopes of transmittance. The stability inherent in the envelope measurements has
already been discussed.

Extinction Coefficient
0.0015
Extinction Coefficient

0.0012

0.0009

0.0006

0.0003

0.0000
400 600 800 1000 1200 1400 1600
Wavelength
Figure 17-12. The extinction coefficient derived from fitting the transmittance of an inhomogeneous and
nonabsorbing layer with a model of a homogeneous absorbing layer.
The results give a film of index 1.93, and a thickness of 1011.6nm and an extinction
coefficient that increases linearly with wavelength, Figure 17-12. The fit between the
transmittance calculated from these derived results and the original inhomogeneous performance
is so good that no difference can effectively be determined, Figure 17-13.
Macleod: Optical Coatings Page 273 Design through Manufacture

Transmittance comparison
100

95

Transmittance (%)
90

85

80

75

70
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-13. The fit between the original inhomogeneous results (grey curve) and those calculated from the
derived index and extinction coefficient (black curve). The agreement is so good that no difference can be
detected.

Reflectance comparison
30

25
Reflectance (%)

20

15

10

0
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-14. Comparison between the original inhomogeneous reflectance (grey curve) and that calculated
from the extracted homogeneous constants (black curve).
This illustrates the danger in assuming that the extracted constants must be correct if their use
reproduces exactly the input data. Calculation of the reflectance now shows a serious difference,
Figure 17-14.
Now let us look at a more complex example. Figure 17-15 shows a refractive index profile
through a film. The variation consists of a gradually increasing index, on which is superimposed
a variation that is close to cyclic. The total thickness of the film is 1005nm. There is no
dispersion and no absorption. The transmittance of this film on a dispersionless substrate of
index 1.52 corresponding roughly to glass is in Figure 17-16. A similar variation of index
through the film, although usually less regular, often occurs in practice when the control of a
Macleod: Optical Coatings Page 274 Design through Manufacture

process parameter is deficient, or when manual adjustments to deposition rate are frequently
made. The appearance of the fringes is usually a good guide. If there is some irregularity in terms
of wavelength then a complicated inhomogeneity should be suspected.
Because of their considerable stability, envelope methods were used to extract n and k for the
film. The models consisted of an inhomogeneous nonabsorbing film with a linear variation of
index and a homogeneous absorbing film.
The results derived using the inhomogeneous model are shown in Figure 17-17. The total
thickness was estimated at 1005.9nm, a satisfactory figure. Also the degree of inhomogeneity
corresponds well with that of the original film. However, around the region where the fringe
pattern is irregular there is a pronounced disturbance in index, the inhomogeneity actually being
reversed over the greater part of the region. In fact the index exhibits the type of variation
associated with a resonance, and, of course, the cyclic variation of index through the film could
be thought of in that way.

Inhomogeneous layer
2.2
2.0
Refractive Index

1.8
1.6
1.4
1.2
1.0
0.8
-200 0 200 400 600 800 1000 1200
Physical Distance from Medium
Figure 17-15. Index variation through an inhomogeneous film. The variation consists of an almost linear
increase with cyclic variation superimposed. Total thickness is 1005nm.
Macleod: Optical Coatings Page 275 Design through Manufacture

Inhomogeneous layer
100

95

Transmittance (%)
90

85

80

75

70
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-16. Transmittance of the film of Figure 17-15. Dispersion and absorption were both assumed zero.
The rear surface of the substrate is assumed uncoated. The fringes around 800nm are misshapen.

Inhomogeneous model
2.05

Outer
2.00
Refractive Index

Mean
1.95

Inner
1.90

1.85
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-17. The refractive indices extracted from the results in Figure 17-16 using a model of a linearly
inhomogeneous film. The curves, from top to bottom at the left-hand side are outer index, mean index and
inner index. The value of thickness derived for the film was 1005.9nm, which should be compared with the
1005nm of the original film.
The results using the homogeneous but absorbing model are no better. The transmittance of
the coated substrate actually exceeded that of the uncoated substrate over a very limited region
near 800nm. Since this would have yielded a negative value of k, that particular point was
discounted. The results are shown in Figure 17-18 and Figure 17-19. The thickness of the film
was determined as 1017.2nm, a rather poorer result than that due to the inhomogeneous model.
Macleod: Optical Coatings Page 276 Design through Manufacture

Homogeneous model
2.05

2.00

Refractive Index
1.95

1.90

1.85
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-18. Refractive index derived for the transmittance of Figure 17-16 using a model of a
homogeneous and absorbing thin film. The thickness was determined as 1017.2nm while that of the original
film was 1005nm. The results are close to, but not exactly the same as, the mean values in Figure 17-17.

Homogeneous model
0.0012
Extinction Coefficient

0.0009

0.0006

0.0003

0.0000
400 600 800 1000 1200 1400 1600
Wavelength (nm)
Figure 17-19. The extinction coefficient extracted from the results of Figure 17-16 using the
homogeneous and absorbing model. The original film was completely free of absorption.

17.5 Multilayers
A field of considerable interest, particularly at the present time, is the extraction of film
properties from performance measurements on multilayers. This is a much less well organized
field. The objective is in most cases a very attractive one, to discover where an attempt at the
production of a given coating has failed so that a subsequent attempt may be more successful.
There is no theoretical body of knowledge. Clearly with no information about the structure of
the layer system in the coating the matching of a sequence of layers to a measured performance
Macleod: Optical Coatings Page 277 Design through Manufacture

is similar to an open-ended process of design. There are likely to be many solutions. But in the
normal way we know the intended structure. It is simply that the real coating has deviated from it
at some point. Provided the deviation is small might we then use a process of refinement to alter
a starting design slightly so that its performance becomes that of the perturbed design? Provided
we remain in the same local minimum of the function of merit then we might argue that the
probability should be high of a coincidence between the actual structure and the calculated
perturbed one. In fact we present shortly a demonstration that this is essentially correct.
Unfortunately we can rely on the measurements only to the extent of their accuracy. It is
impossible to measure transmittance and reflectance precisely. Thus there will be some deviation
between the actual intrinsic performance of the structure under study and the measurements
available to us. Such errors can have a major effect on the reverse engineering process.
Let us begin with a very simple and rather crude example involving a four-layer high-
performance antireflection coating for the visible region. This is made up of magnesium fluoride
and zirconium dioxide with a structure that can be written:

Air | L 2.13H 0.33L 0.27H | Glass

with reference wavelength 510nm. The performance of the perfect design is shown in Figure
17-20 where we are again ignoring the influence of the rear surface to simplify the calculations.
Note that there is dispersion included in the materials although it is very small in the magnesium
fluoride.

Air | L 2.13H 0.33L 0.27H | Glass


5

4
Reflectance (%)

0
400 450 500 550 600 650 700
Wavelength (nm)
Figure 17-20. The initial perfect performance of the antireflection coating.
Let us now measure the performance over an extended region, 300nm to 1000nm, but let the
measurement be in error by +0.5%. Although this may seem a large error, it is not unusual
especially when the real measurement may involve also the rear surface of the substrate. We now
attempt to see what changes we must make in the design so that it will match this performance
measurement.
Macleod: Optical Coatings Page 278 Design through Manufacture

The measured performance is taken as the target for refinement. At first we try to change only
thicknesses of the layers but this turns out to be ineffective. Next we try to include the possibility
of a change in the refractive indices of the materials. We do this by varying the packing density
so that the form of the dispersion is not greatly altered but initially retain the same packing
density for both layers of the same material. This, too, is not very successful. Finally we permit
the thicknesses and the packing densities of the layers to change independently. This achieves
the very good result shown in Figure 17-21. The target, supposed measured, performance is
shown as crosses while the calculated performance from the altered design is shown as a bold
continuous line.
The final, altered design is compared with the starting, ideal design in Figure 17-22. We see
little change in the thicknesses but an inhomogeneity in refractive index that shows values a little
elevated towards the substrate surface and somewhat reduced towards the outer surface. This
might well be thought a real effect because it corresponds to what we know about film growth
but it is a complete artefact.

Targets (crosses) compared with final performance


30

25
Reflectance (%)

20

15

10

0
300 400 500 600 700 800 900 1000
Wavelength (nm)
Figure 17-21. The match between the target performance and the design performance is excellent once both
thicknesses and refractive indices of the layers are permitted to vary.
Macleod: Optical Coatings Page 279 Design through Manufacture

Index profile of final (bold) and starting (grey) designs


2.4

2.1

Refractive Index 1.8

1.5

1.2

0.9
-0.2 0.0 0.2 0.4 0.6 0.8 1.0 1.2
Optical Distance from Medium
Figure 17-22. The final design (bold line) compared with the starting, ideal design (fine line). The
thicknesses of the layers are scarcely altered but there is a definite inhomogeneity in the structure showing a
high index near the substrate and a reduced index towards the front of the system.
A second example concerns a long-wave-pass filter consisting of a core that is a quarter-wave
stack surrounded by a three-layer matching system on either side. There are 31 layers in all and
the design is illustrated in Figure 17-23 as a plot of refractive index against optical thickness.
This system is perturbed by altering the thickness of layer 2, that is the layer next to the front
layer, from 0.3λ0 to 0.2λ0, where λ0 is 500nm.

Starting design of LWP filter


2.7

2.4
Refractive Index

2.1

1.8

1.5

1.2

0.9
-2 0 2 4 6 8 10
Optical Distance from Medium
Figure 17-23. The design of a long-wave-pass filter using silicon dioxide and titanium dioxide. The core is a
quarter-wave stack and it is surrounded by a three-layer matching system on either side.
Refinement with the perfect coating as starting design and targets consisting of the exact
calculated transmittance from 400nm to 1000nm at intervals of 5nm yields exactly the perturbed
design. This supports the idea that the correct result should be obtained provided the refinement
does not move to a completely new minimum of the merit function.
Macleod: Optical Coatings Page 280 Design through Manufacture

Now the calculated performance, including the effect of the change in layer 2, is transformed
into a faulty measured performance by multiplying all results by a factor of 0.99. The targets are
as before the “measured” performance from 400nm to 1000nm at intervals of 5nm. A careful
process of refinement in which the constraints are gradually opened up, until all layer
thicknesses and all layer indices (through packing density) are permitted to take part in the
refinement, results in the performance shown in Figure 17-24. Again the targets are shown as
crosses and the calculated performance as a bold line. The agreement is impressive. The
resulting design, however, is quite remote from the known structure of Figure 17-23 with only
the thickness of the second layer altered. It is shown in Figure 17-25 and exhibits large thickness
variations in the matching system next to the substrate together with a pronounced cyclic
variation of index through the structure that applies both to high and low-index layers. Again
virtually all of these effects appear plausible but are complete artefacts.

Final transmittance compared with targets


100

80
Transmittance (%)

60

40

20

0
400 500 600 700 800 900 1000
Wavelength (nm)
Figure 17-24. Comparison of final performance with target “measured” performance.
Macleod: Optical Coatings Page 281 Design through Manufacture

Index profile of final design


2.7
2.4

Refractive Index
2.1
1.8
1.5
1.2
0.9
-2 0 2 4 6 8
Optical Distance from Medium
Figure 17-25. The final design. Note the differences between this design and that of Figure 17-23. The
thicknesses of the layers next to the substrate are quite different and there is a cyclic variation of index
through the structure.

17.6 Implications
What are the implications of all of this? Clearly we need to be very cautious when trying to
extract information about particular parameters from sample measurements. It is a model fitting
process and defects in either the model or the measurements or both will immediately lead to
incorrect extracted parameters. These parameters may be quite sensitive functions of the
measured values or of the model and the errors in them may be very much larger than the
original errors in measurement or model. There is enormous danger in attempting to draw serious
conclusions from the results of one experiment and one measurement. In the case of single layers
measurements on several samples, preferably of different thicknesses and different coating
operations should be included. Results should be viewed with great suspicion until confidence
can be assured.

17.7 Conclusion
We must solve any random problems before we can tackle systematic ones. There is a clear
danger in assigning too much weight to a process of reverse engineering when only one result is
available. Inevitable errors in measurement have a major effect on the extracted parameters of
the coating. For single layers we can apply straightforward smoothing processes that can solve
many of the problems. We do not yet have anything equivalent for multilayers.
We must make use of all the information available to us.
Macleod: Optical Coatings Page 282 Design through Manufacture

17.8 References
1. Aspnes, D E and H G Craighead, Multiple determination of the optical constants of thin-
film coating materials: a Rh sequel. Applied Optics, 1986. 25: p. 1299-1310.

2. Hwangbo, C K, L J Lingg, J P Lehan, H A Macleod, J L Makous, and S Y Kim, Ion-


assisted deposition of thermally evaporated Ag and Al films. Applied Optics, 1989.
28(14): p. 2769-2778.
3. Borgogno, J P, B Lazarides, and E Pelletier, Automatic determination of the optical
constants of inhomogeneous thin films. Applied Optics, 1982. 21: p. 4020-4029.
4. Manifacier, J C, J Gasiot, and J P Fillard, A simple method for the determination of the
optical constants n, k and the thickness of a weakly absorbing thin film. Journal of
Physics E, 1976. 9: p. 1002-1004.
5. Swanepoel, R, Determination of the thickness and optical constants of amorphous
silicon. Journal of Physics E, 1983. 16(12): p. 1214-1222.
6. Arndt, D P, R M A Azzam, J M Bennett, J P Borgogno, C K Carniglia, W E Case, J A
Dobrowolski, D P Arndt, U J Gibson, T Tuttle Hart, F C Ho, V A Hodgkin, W P Klapp,
H A Macleod, E Pelletier, M K Purvis, D M Quinn, D H Strome, R Swenson, P A
Temple, et al., Multiple determination of the optical constants of thin-film coating
materials. Applied Optics, 1984. 23: p. 3571-3596.
Macleod: Optical Coatings Page 283 Design through Manufacture

18 COATINGS IN SYSTEMS
The combination of electromagnetic beams in interference phenomena is a linear process. The
electric and magnetic field vectors are simply added. In a general case this could represent a
rather involved procedure but because the process is linear, we can decompose any interference
problem into a set of operations on fundamental components chosen for their simplicity. As long
as the components can be reassembled to yield the composite fields they can be chosen in any
way we wish. There are great advantages in choosing infinite, plane, harmonic, linearly
polarized waves as the fundamental components and we use a process of Fourier decomposition
to set up these components. We use the term spectrum to describe the complete set of
components. We find it very convenient to use the spectral components in our design and
calculation. Except in very special cases involving very short optical pulses, we are normally
dealing with steady-state phenomena and we do not have to be concerned about the relative
phases of the different spectral components. The final combination into a resultant light beam is
then a simple summation of the component irradiances, taking no account of any relative
properties such as phase. We quite often carry out this summation only when we are dealing with
the optical system as a whole and as far as coating design is concerned it is usual to express
performance in terms of the interaction with the fundamental spectral component. A
transmittance or reflectance curve, for example, records the effect of the coating on plane,
harmonic, linearly polarized spectral components. However, the performance is derived from
interference effects and these are not simply dependent on wavelength but vary also with other
parameters, such as angle of incidence or even temperature. When, however, short pulses are
concerned we can no longer simply add irradiances to derive the performance of the component.
Amplitude and phase of the spectral components are both important and we shall look at some of
the implications.
A single plane, harmonic, linearly polarized spectral component such as we use in our
calculations and assume in our performance never exists in practice. The light in a measuring
instrument or application may sometimes approach it but it is much more usually it must be
represented as a distribution of such components. And then the filter or coating itself often has
performance that departs from the ideal theoretical model.
In this section we consider a few consequences of these complications.

18.1 General
An important question that should always be asked about a filter is "where does the unwanted
energy go?" Most filters exhibit the various modifications of the input in Figure 18-1. For
example an absorption filter to eliminate short wavelengths will reflect residual light
(redirection), will convert some light (fluorescence), will likely scatter some light, and will emit
thermal radiation - all unwanted. If this light is accepted back into the system, performance will
suffer. Unfortunately, the techniques for actually measuring the components and verifying their
performance are designed to avoid perturbation by the very same defects that can upset
instrumental performance.
Macleod: Optical Coatings Page 284 Design through Manufacture

Figure 18-1. Filters may discard the rejected light in different ways. In most cases the ideal would be to have
the unwanted light absorbed with no other consequence. This is rarely the case. Some of the possibilities are
shown.
Figure 18-2 shows a simple instrument with the important stops. The image of the aperture
stop in source space is the entrance pupil and in receiver space the exit pupil. Note that the
coating and filter performance cannot be separated from the details of stops and pupils. These are
part of the overall system design and we shall return to some aspects of this point. “Where does
the unwanted light go?” Can it be scattered back into the acceptance zone of the system? Baffles
can help to prevent return of unwanted energy and their location and shape is a specialized task.
Cooled infrared detectors need to be surrounded by cooled enclosures to avoid unwanted
thermal radiation. Filters within the field of view of the detector, necessarily so because the light
they transmit must reach the detector, can be a major source of thermal radiation unless they are
cooled. It is convenient to place them immediately in front of the detector so that they are cooled
with it. Then, unfortunately, the aperture subtended at the detector is effectively very wide and
scattered light, intended to be rejected, becomes a problem. This scattered light is rarely
measured by normal spectrometers often designed to measure the spectral properties of
scattering samples without problems. A similar difficulty can occur with wedged filters in front
of array detectors for simple scanning of spectral regions. Serious crosstalk can be caused by
scattered light.
Macleod: Optical Coatings Page 285 Design through Manufacture

Figure 18-2. Sketch of a simple instrument showing the stops. The image of the aperture stop in source space
is the entrance pupil and in receiver space the exit pupil.
It is very important to remember that although performance is system dependent, an optical
filter is usually specified with regard to standard (and that usually means ideal) conditions.
Entrance and exit pupil are at infinity (light ideally collimated) and scattered, redirected,
converted and emitted light components are assumed lost to the system. A real system will rarely
have this arrangement. Therefore the performance may not correspond to the standard specified
performance of the filter.
In the remainder of this chapter we will be mainly concerned with the ideal properties of
coatings and filters. That is we will consider their properties as being those calculated by the
usual thin-film calculation techniques and the surfaces on which they are deposited of
sufficiently high optical quality so that the behavior is specular. Specular means that the light
redirected by the surface by reflection or transmission remains in the plane of incidence with the
angle of reflection equal to the angle of incidence, and the angle of refraction related to it by
Snell’s Law.

18.2 Reflection from a rejection filter

Figure 18-3. Transmittance (calculated) of a germanium filter with antireflection coatings.


Filters that transmit the selected light often operate by a mixture of absorption and reflection.
Sometimes they also consist of several different components combined. For example,
Macleod: Optical Coatings Page 286 Design through Manufacture

semiconductors are intrinsically longwave pass absorption filters. Photons of energy greater than
the gap between valence and conduction bands of the electrons are absorbed by transferring
energy to electrons in the valence band to move them into the conduction band. High
transmittance implies intrinsic semiconductors of high resistivity. Gallium arsenide, silicon,
germanium, indium arsenide, indium antimonide are all useful. Note that these semiconductors
have high refractive index so must be antireflected in the pass region. They are rarely
antireflected on purpose in the absorbing region and so a large amount of the rejected light may
be actually reflected rather than absorbed.

Figure 18-4. Reflectance (calculated) of the filter of Figure 18-4.


The reflected light is now traveling in a new direction. Depending on the design of the
instrument this light may be returned, eventually, to the very filter that reflected it. Figure 18-5
shows how it may be that light might arrive back at the filter but at a new angle of incidence.
The performance of the filter at this changed angle will differ, possibly significantly, from that
expected at the design angle. We illustrate this with a long-wave pass filter for the visible region.

Figure 18-5. The short-wave light reflected by this long-wave pass filter may be scattered back at oblique
incidence and then, because of the angular shift, may be transmitted by the filter rather than simply rejected.
Macleod: Optical Coatings Page 287 Design through Manufacture

Figure 18-6. Performance at normal incidence of an edge filter consisting of a quarterwave core and 8-layer
matching systems on either side of the core totalling 49 layers in all. Materials are TiO2 and SiO2.

Front

Back

Figure 18-7. The reflectance of the edge filter where the short wave sidebands have been suppressed by an
OG 570 absorption filter used as substrate rather than transparent glass. The reflectance is significant in the
rejection region and depends on the orientation of the filter.
The shortwave sidebands of the edge filter in Figure 18-6 can be suppressed by the use of an
absorbing glass filter, OG 570 [1] in this case. Here in Figure 18-7 it replaces the glass substrate
of Figure 18-6. The reflectance of the filter depends now very much on the orientation but in
either case there is considerable reflectance between the edge of the absorption filter and that of
the dielectric long-wave pass element. Light that is reflected can return to the filter at an oblique
angle of incidence. The characteristic of the filter moves to shorter wavelengths as it is tilted.
Therefore there can be serious transmission of the unwanted light. Error! Reference source not
found. illustrates the problem.
Macleod: Optical Coatings Page 288 Design through Manufacture

Figure 18-8. The product of the normal incidence reflectance with the 30º transmittance characteristic shows
the danger zone to the shortwave side of the filter edge where leakage can occur. The zone broadens as the
angle of incidence of the returning light increases.

18.3 Spectral width of measurement


A typical device, a spectrometer, for assessing the spectral performance of a filter or coating,
probes it with a beam of light that has, as nearly as possible, the characteristics of a
monochromatic beam, and measures the response. As the wavelength or frequency of the beam
are varied the response varies, and the record of this variation is taken as the measured
performance of the filter. The two principal deficiencies of the probe beam are a finite spectral
width and a finite range of angles of incidence. We shall consider the range of angles of
incidence shortly but in this section concentrate on the spectral width, usually referred to as the
bandwidth of the spectrometer.
Unless the spectrometer has been very badly designed the spectral distribution of the probe
beam will be a narrow, continuous band of wavelengths clustered around a mean that is
nominally the measuring wavelength of the instrument. During the process the measuring
wavelength is varied, either continuously or, sometimes, in small steps.
In mathematical terms the result of the measurement is the convolution of the spectral
distribution of the source and the spectral performance of the coating. The result depends on the
spectral features in the coating performance. As long as the bandwidth of the spectrometer is
much smaller that the width of any coating feature then the measured performance and actual
performance will be similar. But how narrow must the bandwidth be? A good rule of thumb is
that the bandwidth of the spectrometer should be not greater than one third of the width of any
spectral features in the performance of the coating to be measured.
We can take as our first example a narrowband filter such as might be employed in a
telecommunication applications, Figure 18-9. The width of the filter passband is 0.6nm. It is
clear that the degradation in measured performance is slight when the measuring width is one
third of the filter width (i.e. 0.2nm) but large when it is twice (i.e. 1.2nm).
Macleod: Optical Coatings Page 289 Design through Manufacture

Transmittance (%) 0.2nm


100

80
0.6nm
60

40
1.2nm
20

0
1549.0 1549.5 1550.0 1550.5 1551.0
Wavelength (nm)

Figure 18-9. A narrowband filter (bandwidth of 0.6nm) scanned with varying spectral bandwidths as shown.

Reflectance (%)
25

20

15

10

0
400 600 800 1000 1200 1400 1600 1800 2000
Wavelength (nm)

Figure 18-10. The grey curve in the background represents the fringes calculated for a 50µm foil of
polyester.(n = 1.65). The black curve represents the result of scanning the fringes with a spectral bandwidth
of 5nm.
This result is straightforward but rather more complicated results can be obtained in scanning
closely packed fringes. Figure 18-10 represents the results of scanning the fringes produced by a
50µm thick polyester foil using a bandwidth of 5nm. The curious beat pattern is quite typical.
The beat nodes occur when the scanning bandwidth is exactly equal to one, two, three and so on,
fringe periods. The fringe pattern is said to have variable visibility. The fringe visibility record
contains information about the thickness of the foil that may be easier to interpret than the record
of the individual fringes. Allied techniques are often used in astronomy to determine stellar
diameters.

18.4 Tilted and cone response


Coating and filter calculations are normally made for perfectly collimated light of infinitely
narrow spectral width. Light sources must have finite dimensions and this implies that any beam
that is generated must contain a range of propagation angles. The greater this range the greater
the energy that can be transported. An angular range of zero is accompanied by zero irradiance.
Macleod: Optical Coatings Page 290 Design through Manufacture

We can consider the light source to be a circular part of a sphere with the coating at the sphere
center, the sphere being so large that the area of the coating is vanishingly small in comparison.
Now if we draw all possible rays from the source to the coating they form a solid cone. If all
possible rays have equal irradiance then we describe the light source as Lambertian and the
illuminating cone as a Lambertian cone. A different type of cone, a Gaussian cone, will be
considered shortly.
In an instrument, and this includes the spectrometer used to measure the component, the
illumination conditions are not ideal. An instrument can be considered largely as a succession of
apertures. In a well-designed instrument correct illumination of the various apertures is assured.
The simple sketch in Figure 18-11 shows a pair of successive plane apertures in such a system.
In most cases the illuminating aperture A1, where A1 is both designating symbol and aperture
area, has a characteristic of even illumination in all directions. The total power that leaves it in
any relevant direction is proportional to the product of its total area A1 and the cosine of the
angle between the direction and the normal to the aperture. A plane source that has this property
is known as Lambertian. A true Lambertian luminous surface looks equally bright from all
directions. Most normal surfaces are of this type. The exit hole from a black body cavity is a
Lambertian source. The second aperture, A2, the receiver, is also of this type, although in reverse.
If both A1 and A2 are Lambertian then the total power transferred between the two apertures is
given by

A1 A2
L⋅ = LA1Ω 2 = LA2Ω1 (18.1)
r2

where L is the radiance, that is the radiant power (watts) per unit solid angle (steradian) per unit
surface area (square metre) subtended in the direction of the radiation. (L used to be called the
brightness of the source). Note that each aperture area is coupled with the solid angle subtended
at it by the other aperture. It is easy to mix these up. In a lossless system, radiance is conserved.
This product thus represents the maximum that can be transferred from one aperture to the other.
The well-designed instrument makes sure that the AΩ product is conserved through the
instrument – in other words each aperture should be of the appropriate size and optical elements
should be employed with sufficient power to avoid vignetting. Then the throughput is given by
the product of the radiance of the source, the area of the source and the solid angle subtended at
it by the first optical aperture, usually the clear aperture of the first optical element.
Macleod: Optical Coatings Page 291 Design through Manufacture

A1 Ω2 Ω1 A2

Figure 18-11. Two successive apertures in an instrument illustrating the AΩ product. The A represents the
area of the aperture and the Ω the solid angle subtended at the other.
The properties of thin film optical coatings are calculated for collimated light. This is
equivalent to shrinking the appropriate solid angle until its magnitude is zero. Clearly this would
represent zero throughput, and render operation of the instrument impossible. We need,
therefore, to consider the effect of solid angles that are finite.
We know that the effect of tilting an optical interference coating is to shift the characteristic to
a shorter wavelength. There are some other effects as well but this is enough for us for the
moment. The phase thickness of a layer at oblique incidence is given by

2πnd cosϑ
δ= (18.2)
λ

where ϑ is the angle of propagation in the particular layer, related to the angle of incidence, ϑ0,
in the incident medium by Snell’s Law. The various films in a multilayer dielectric coating are
not all of the same refractive index and therefore the change in phase thickness varies. For many
coatings, however, the effect of tilt is so close to a simple translation in wavelength that we can
approximately model it by using an effective index. This has been shown to work well for
narrowband filters. The effective index depends to some extent on the particular design but for
rough calculations of somewhat better than order of magnitude a good starting approximation for
a two-material coating is the geometric mean of the indices involved. This should not be
stretched too far. As the tilt angle moves further and further from the design angle of incidence
the characteristic of the coating becomes more and more distorted. At oblique incidence, too,
there is always the polarization splitting that complicates matters further.
Let us denote the effective index by ne. Then the shift in characteristic will be given by:

2πne d cosϑ 2πne d


= (18.3)
λ λ0

λ = λ 0 − ∆λ (18.4)
Macleod: Optical Coatings Page 292 Design through Manufacture
LM Fn I 2 OP
∆λ = 1 − 1 − G sin ϑ J
MN PQ ⋅ λ
0
(18.5)
Hn Ke
0 0

which becomes, if we neglect third-order effects,

n02ϑ 20
∆λ = ⋅ λ0 (18.6)
2ne2

In this expression ϑ0 is given in radians. With ϑ0in degrees the relationship becomes

n02ϑ 20
∆λ = 15
. × 10 −4
⋅ λ0 (18.7)
ne2

A typical curve calculated from the accurate expression, for the pair of materials 2.40 and 1.45,
geometric mean √(1.45×2.40) = 1.865, is shown in Figure 18-12.

Angle-induced shift at 1000nm


30
Wavelength shift (nm)

25

20

15

10

0
0 5 10 15 20 25 30
Angle of incidence (deg)
Figure 18-12. Calculation of the angle-induced shift for a combination of 1.45 and 2.40 material and λ0 of
1000nm.
A graph like this allows some simple deductions about the sensitivity of coatings to tilts and
to illumination in the form of cones. In an instrument, the finite solid angle will have a maximum
and a minimum angle of incidence on the coating. If the illumination cone is at normal incidence
then the minimum will be zero. Two lines can be marked on Figure 18-12 corresponding to the
maximum and minimum angle of incidence. This will give immediately a good idea of the effect
on the particular coating. For example, a cone that has semiangle 5° will have little effect, other
than a slight spectral shift, on a feature that is some 30nm in width. Steep sides would suffer not
more than a 1 to 2nm increase in the distance from top to bottom, small compared with the basic
characteristic. 10° represents something a little more serious. Here we have a 5nm increase in the
slope of the edges.
This is shown clearly in Figure 18-13 where a multiple cavity filter of design:
Macleod: Optical Coatings Page 293 Design through Manufacture

Air | (HL)3 H (HL)7 H HH(HL)7 HHH (HL)7 H (HL)3 H | glass (18.8)

made up of material of indices, 2.20 and 1.45, not significantly far from those of Figure 18-12, is
shown. The effects of an illuminating cone of semi-angle 5° and then 10° are compared with the
ideal curve. The most important feature of the figure is the shift in effective center wavelength of
the filter. These cones are normally incident, yet the peak wavelength has shifted to a shorter
wavelength. Note that spectrometers often have apertures corresponding to f/6.3 equivalent to a
cone with semi-angle just under 5°.

0, 5, 10 deg cones
100

80
Transmittance (%)

60

40

20

0
985 990 995 1000 1005 1010
Wavelength (nm)
Figure 18-13 Calculated response of multiple-cavity filter illuminated by cones of semi-angle, 0° (black,
right-hand characteristic), 5°, (grey, center characteristic) and 10° (grey, left-hand characteristic).

Reflectance (%)
25

20

15

10

0
400 600 800 1000 1200 1400 1600 1800 2000
Wavelength (nm)

Figure 18-14. The fringes of Figure 18-10 measured with zero bandwidth but an aperture of f/6.3 or a cone
semi-angle of 4.5º.
The use of a Lambertian cone as source for performance measurement is not dissimilar to the
use of a finite bandwidth. In fact both are attributes of a normal spectrometer. Figure 18-14
shows the fringes of Figure 18-10 measured using a Lambertian cone of illumination of semi-
Macleod: Optical Coatings Page 294 Design through Manufacture

angle 4.5º (or an aperture of f/6.3). The effect is clearly of the same nature as the limited
bandwidth in Figure 18-10, but somewhat less.
⎡ ϑ2 ⎤
I = I 0 exp ⎢ −2 2 ⎥
⎣ ϑ0 ⎦

ϑ0

4λ Beam
d0 = waist
π n ( 2ϑ0 )

Figure 18-15. The details of a Gaussian beam. The narrowest part of the beam is called the beam waist. At a
distance from the beam waist the beam has a circular wavefront converging on the point at the center of the
waist. The irradiance at angle ϑ from the beam axis is given by exp(-2ϑ2/ϑ02) times that at the center where
ϑ0 is the beam divergence and marks the angle where the irradiance is 1/e2 times that at the center. A
Lambertian cone would have constant irradiance up to angle ϑ0 and zero irradiance beyond it.
Rather like the infinite plane wave, a light ray, that is a beam of light that is so limited in
lateral dimensions that it appears simply as a line, cannot exist. Light rays are used in system and
lens design but they are theoretical entities that represent unattainable limits. The nearest
practical equivalent of the light ray is the Gaussian beam. Gaussian beams appear superficially
like light rays but they are rather more like a cone that is focused on a spot called the beam
waist. The Gaussian beam has an irradiance that falls off from the central axis as a Gaussian
function. The principal details of such a beam are shown in Figure 18-15.

Divergence (degree)
6

4
λ = 1550nm
3

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Spot size (mm)

Figure 18-16. The relationship between spot size, that is beam waist diameter, with beam divergence of a
Gaussian beam. The wavelength is assumed to be 1550nm.
Macleod: Optical Coatings Page 295 Design through Manufacture

Transmittance (%)
100

80

60

40
10° 5° 0°
20

0
1540 1542 1544 1546 1548 1550 1552
Wavelength (nm)

Figure 18-17. The performance of the filter of Figure 18-9 measured at 0º, 5º and 10º with light forming a
Lambertian cone of 1º semi-angle. The curves in the background are the ideal performance curves of the
filter for perfect monochromatic plane waves.

Transmittance (%)
100

80

60

40
10° 5° 0°
20

0
1540 1542 1544 1546 1548 1550 1552
Wavelength (nm)

Figure 18-18. The filter of Figure 18-9 measured at 0º, 5º and 10º with a 1º semi-angle divergence Gaussian
beam. Background curves are ideal performance.
In practice many beams behave like the ideal Gaussian and in particular the properties of the
beam link illuminated spot size to beam divergence, Figure 18-16. The spot size has recently
become an issue in communication filters. Figure 18-16 shows that as long as the spot size is
larger than 60µm that the beam divergence should be less than 1º.
How dangerous is a beam divergence of 1º? Figure 18-17 and Figure 18-18 show the effects
of scanning a narrowband filter similar to that of Figure 18-9 with Lambertian and Gaussian
beams of 1º divergence. There is not a great deal of difference between the Lambertian and
Gaussian cases (although they will exhibit significant differences for much greater divergences).
Changes in performance with tilt angle follow a cosine law and so the effect becomes much
worse as the angle of incidence of the illuminating cone gets worse and that is very clear from
the diagrams. Strictly there are also rather small variations with the polarization of the incident
light. Here we are assuming that the incident light is unpolarized.
In most communication applications involving narrowband filters the spot size is arranged to
be larger than 500µm. This implies that for small angles of incidence of a Gaussian beam the
performance degradation should be vanishingly small.
Macleod: Optical Coatings Page 296 Design through Manufacture

NA 0.0

NA 0.5

Figure 18-19. An extended zone reflector in collimated light (Numerical Aperture zero) and in light of cone
semi-angle 30º (Numerical Aperture 0.5). Note the steep fall in performance at the red end of the spectrum.
However, it is not just the narrowband filter that can exhibit effects due to illuminating cones.
Figure 18-19 shows the performance of an extended zone reflector when used in a cone of 30º
semi-angle (Numerical Aperture 0.5). There is now a steep drop in performance at the red end of
the spectrum.
Then the edge filter from Figure 18-6 exhibits polarization splitting when used at 30º
incidence. Now at this angle of incidence a cone of only 6º semi-angle (that is a numerical
aperture of just over 0.1) causes quite significant degradation, Figure 18-20.

mean

Figure 18-20. The degradation in the split edge of the long-wave pass filter when a 6º cone rather than a
collimated beam is used to illuminate it.

18.5 Multiple Reflections and Stray Light


The effect of the rear surface of a substrate has already been discussed in the chapters on
theoretical techniques. Beams reflected back and forth between the two surfaces can have a large
Macleod: Optical Coatings Page 297 Design through Manufacture

effect on the light that is transmitted and reflected by the component. The same is true of a series
of components as in a lens or other optical system. When the surfaces are sufficiently wedged
then the reflected light is directed out of the aperture of the receiver and it has no effect. Thus
there are two extreme conditions, surfaces sufficiently wedged to avoid multiple reflection
effects and completely parallel so that all multiply reflected beams are collected by the receiver.
Multiple reflections increase the apparent reflectance and transmittance of the system. The
effect is least when the surface reflectance is low and the transmittance high. Light that would
otherwise be lost, is collected by the receivers. The light representing the increase is strictly stray
light. It has followed different paths from that of the desired primary beam. If the system is an
imaging one the extra light will tend to arrive in the focal plane of the instrument in the wrong
place. It may form a veiling glare or it may form ghost images. If the system is not imaging one
then it can increase the amount of light that is received in a spectral region where it should have
been rejected.
The difference between the transmittance and reflectance when the surfaces are completely
wedged and completely parallel is a measure of the stray light that can be present. Two examples
will suffice.
The transmittances are given by:

1
Tparallel =
FG 1 1
+ −1
IJ (18.9)

H Ta Tb K
Twedged = TaTb (18.10)

The first example is a rejection filter where two highly reflecting coated surfaces are placed in
series. There are no losses associated with either coated surface. The transmittance of each
coating is 1%. The wedged case gives an overall transmittance of 0.01%. The parallel case gives
a transmittance of 0.5%. This represents 50 times the desired light level.
The second example is a lens with some 10 elements and 20 surfaces. Here we imagine that
each surface has the same reflectance. The ratio of stray light to useful light as a function of the
single surface reflectance is shown in Figure 18-21. The stray light may be a much more
demanding parameter in determining coating performance requirements than overall
transmittance, Figure 18-22. Figure 18-23 shows the amount of stray light that can occur in a
five-element lens (ten surfaces) through this effect.;
Macleod: Optical Coatings Page 298 Design through Manufacture

Stray light - 10 elements - 20 surfaces


1.2

1.0

Stray/Useful (%)
0.8

0.6

0.4

0.2

0.0
0.0 0.2 0.4 0.6 0.8 1.0
Surface Reflectance (%)
Figure 18-21. A system consists of 10 elements each with two surfaces. The potential stray light as a fraction
of the useful light is plotted against the residual single-surface reflectance.

Transmittance - 10 elements - 20 surfaces


100
98
Transmittance (%)

96
94
92
90
88
86
84
0.0 0.2 0.4 0.6 0.8 1.0
Surface Reflectance (%)
Figure 18-22. The useful transmittance of the 10-element 20-surface system of Figure 18-21.
Macleod: Optical Coatings Page 299 Design through Manufacture

Stray Light (%)


2.0

1.5

1.0 Single-layer

0.5
Four-layer
0.0
400 500 600 700
Wavelength (nm)

Figure 18-23. Calculated stray light in a five-element lens when a single-layer antireflection coating is used
on all surfaces and when a four-layer coating of superior performance is used. For calculation purposes the
elements are supposed to have indices of refraction similar to Schott BK7 glass (around 1.52).
Multiple beams reflected back and forth between coatings can have other major effects on
performance. The magnitude is greatest when the reflectance of the two coatings is high. Figure
18-24 shows the effect on the rejection of a pair of notch filters.

Wavelength (nm)
400 500 600 700 800 900
0

-10
A
-20
B

-30
C

-40
log(Transmittance) (dB)

Figure 18-24. The effect of placing two notch filters in series. A represents one single notch filter. In B the
filters are arranged so that the light is reflected back and forth between the two filters. In C the filters are
arranged slightly wedged so that reflected light is directed outside the aperture of the receiver.

18.6 Effect of cement


Filters are frequently cemented into optical assemblies. Especially if the cementing operation
is carried out by a customer after filter delivery, the filters may have been measured in air to
verify their performance. This is sometimes the case in telecommunication applications, for
example. This later cementing operation can bring some problems.
Macleod: Optical Coatings Page 300 Design through Manufacture

Transmittance (%) Transmittance (%)


100 B 100
A
A
90 90
B
80 80

70 70

60 60
650 700 750 800 850 900 650 700 750 800 850 900
Wavelength (nm) Wavelength (nm)

Figure 18-25a. Performance in an incident Figure 18-25b. Performance of the same two
medium of air of two similar edge filters. edge filters but this time the incident medium is
cement of index 1.55.

A simple transparent surface can have only a real refractive index, otherwise the material will
be absorbing. A material of index 1.55 will match exactly a cement of index 1.55 to give zero
reflectance. In air this surface of index 1.55 will yield a reflectance of 4.65%. A surface of index
1/1.55, i.e. 0.645 will also give a reflectance of 4.65% in air but the reflectance in cement will be
17%. For a simple transparent surface, of course, an index of 0.645 is not possible but this is not
the case when a surface is covered by a coating. Then even a lossless system can have not only
an index less than unity but a complex index. A coating that presents a reflectance of 4.65% in
air can have any reflectance from zero to 17% when the incident medium is changed to cement.
Specifications will often call for a transmittance to be typical of what will give the correct
performance when cemented. This is really no guarantee of the cemented performance and the
only way to be sure, is to measure the filter after it has been cemented. This can obviously create
difficulties and so an alternative that is sometimes used is to require a representative filter from a
simultaneously manufactured batch to be tested either cemented or with an incident medium of a
suitable index matching fluid.
Figure 18-25 illustrates the problem in the case of a long-wave-pass filter.

18.7 Temperature
Temperature changes can have significant effects on the performance of optical coatings. For
example, plastic substrates include the possibility of dissipating strain energy in deformation, a
mechanism that is missing in the brittle inorganic materials considered shortly. This energy
dissipation can reduce compressive strain during temperature cycling and hence increase
susceptibility to tensile cracking.
The performance of coatings changes with temperature. Until recently, temperature
coefficients were mainly derived by empirical measurements on actual filters. Only a very few
Macleod: Optical Coatings Page 301 Design through Manufacture

studies attempted to create models of behavior. Recently this has changed for rigid, brittle
substrates with the publication of a study by Takahashi[2].

Poisson’s ratio Strain birefringence Substrate-induced


thickness change change in index differential strain

Thickness change by Temperature-induced


thermal expansion change in index
Figure 18-26. Diagram illustrating the details of the Takahashi model for the effects of temperature on the
optical performance of a coating.
The Takahashi model assumes that the substrate is dominant and changes in substrate
dimensions by temperature changes are transmitted unchanged to the layers of the filter. These
layers also suffer temperature changes of the same magnitude as the substrate but their
dimensional changes may differ. Differential expansion or contraction in the plane of the film
causes tensile strain. There are four effects that determine the change in performance, Figure
18-26.
1. The layers of the filter show a change in refractive index due to the temperature change.
2. The physical thicknesses of the layers change because of thermal expansion or
contraction.
3. The tensile strain in the film plane causes a change in refractive index through strain
birefringence.
4. The tensile strain in the film plane is translated by Poisson’s ratio into a change of film
physical thickness.
This model has been very effective in predicting temperature coefficients of narrowband
filters for telecommunications, Figure 18-27. Unfortunately there is as yet no corresponding
model for systems on less rigid substrates.

Transmittance (%) BK 7
100

S 7003 80 Silica
60

S 7006 40

20

0
1548 1549 1550 1551 1552 1553 1554
Wavelength (nm)
Ambient
Figure 18-27. Predicted performance of a narrowband filter of the same type as in Figure 18-9 on different
substrates when the temperature is increased by 100ºC. The shift of the filter on S 7003 glass is negligible.
[BK 7 is a general purpose glass made by Schott. S 7003 and S 7006 are special high-expansion coefficient
substrate materials manufactured by Schott specifically for telecommunication filters.]
Macleod: Optical Coatings Page 302 Design through Manufacture

18.8 Transient response


It is only recently that we have become interested in the transient response of certain coatings.
Ultrafast pulses of light have now reached widths of just a few femtoseconds where they are
shorter than the time constant of many of the components that reflect or transmit them. The
situation is also becoming true in the telecom field where, although the pulse lengths are
measured in picoseconds nevertheless the coatings are so complex and involve so many layers
that their time constant is also in the picosecond region.
Fortunately the interaction between coatings and these short pulses is essentially linear. This
means that we can continue to decompose the incident light into spectral components but now
we can no longer simply consider the power response of a coating. We now need to consider the
amplitude and phase response.
The amplitude response is reasonably straightforward. If it is anything other than unity over
the width of the pulse spectrum the pulse shape will suffer. We therefore assume that the design
of the component has achieved this level either in transmission or in reflection. Then we are
principally concerned with the phase terms. As long as all the spectral components of the pulse
retain a coincidence of phase at some point then, at that point, they will simply reconstitute the
pulse from which they were derived.
We can take the parameters of the carrier as our reference. Then it is convenient to expand the
phase shift, ϕ, either in reflection or transmission as a Taylor series in terms of ∆ω, the
difference in phase between that of the particular component and the carrier. The term involving
the first derivative of ϕ with respect to ω can be balanced by altering the time. This represents a
time delay or time advance in the pulse position. Terms pulse distortion. The first three
derivatives are those most usually considered:


Group Delay (GD ) = − (units of time)

d 2ϕ
Group Delay Dispersion (GDD ) = − (units of time 2 ) (18.11)
dω 2

d 3ϕ
Third Order Dispersion (TOD ) = − (units of time3 )
dω 3

Note the negative signs. Group Delay Dispersion, in particular, broadens and chirps the pulse,
that is changes the frequency of the carrier gradually through its length. It can be shown for a
Gaussian pulse that if 2µ is the width of the pulse at the 1/e power points then the altered width
of the pulse will be given by
1
⎡ GDD 2 ⎤ 2
τ new = τ old ⎢1 + µ 4 ⎥ (18.12)
⎣ ⎦

As long as GDD is less than µ2 there is little effect but the effect increases rapidly with GDD if it
should exceed µ2.
Macleod: Optical Coatings Page 303 Design through Manufacture

Figure 18-28. The electric field distribution in the extended zone high reflectance coating at awavelength of
560nm. The incident irradiance is 1Wm-2. Note the considerable magnification of the electric field in the
structure at the front (left) of the coating.

Figure 18-29. The Group Delay Dispersion of the extended zone reflector showing the high values associated
with the resonances in the structure.
We can imagine the use of our extended zone high reflectance coating of Figure 18-19 as a
suitable reflector for laser pulses. there are two related problems. The first is that at longer
wavelengths in the high-reflectance zone the high reflectance is created by the structure towards
the rear of the coating. The structure towards the front assures the high reflectance at the short-
wave end and at longer wavelengths is simply acting to transmit the light to and from the
underlying reflector. However, there are interference effects within this outer structure and they
can act to increase the standing electric field. In Figure 18-28 we see a considerable
magnification of the field at 560nm. This implies a reduction in the resistance to high power that
is in itself serious. But the high field is also evidence of a resonant structure which immediately
increases the Group Delay at that wavelength. The derivative of the Group Delay gives the
Group Delay Dispersion and this shows, in Figure 18-29, the characteristic reverse-peaked
structure. In fact there are several such resonances. A 10 femtosecond pulse with a carrier of
500nm has a spectral width of approximately 100nm. The square of the half pulse width is,
Macleod: Optical Coatings Page 304 Design through Manufacture

therefore, 25 fs2. The extended-zone high reflectance coating is clearly not a reasonable coating
for use with such pulses. Reflecting coatings with limited values of Group Delay Dispersion
require special design.

18.9 Contamination Sensitivity


Little has been done on the sensitivity of coatings to contamination. In many cases, the
application demands that the coating be exposed to the external, and often poorely controlled,
environment. Material can deposit on the coating from the environment and affect the optical
properties. It is the electric field that interacts with materials and, therefore, interaction is
maximized where the electric field is highest. The theory is straightforward and is discussed in a
different chapter. Antireflection coatings are designed to match the incident medium and so there
is no, or very little, standing field in the incident medium and the electric field at the front
surface is essentially the incident field. This means that all antireflection coatings have virtually
identical sensitivity to contamination and little can be done to modify that without introducing
higher residual reflectance. Reflecting coatings, on the other hand, produce a large standing
wave in the incident medium and hence the value of the field at the front surface can vary
between close to zero and twice that in the incident wave depending on the phase shift there.

18.10 Other Effects


There are many other effects that must be included when considering the employment rather
than simply the design of optical coatings. Some of these are related to defects in the coatings
themselves, for example inhomogeneities, scattering, reduced packing density and consequent
moisture adsorption, thermal nonlinearities, and so on. Space prevents their inclusion here.

18.11 References
1. OG 570 is one of a series of filter glasses with long wave pass characteristics made by
Schott. For more information see http://www.schott.com/ or http://www.schott.de/
2. Takashashi, Haruo, Temperature stability of thin-film narrow-bandpass filters produced
by ion-assisted deposition. Applied Optics, 1995. 34(4): p. 667-675.
3. Palik, E D, ed. Handbook of Optical Constants of Solids. 1985, Academic Press Inc.
Macleod: Optical Coatings Page 305 Design through Manufacture

19 ABSORPTION AND CONTAMINATION


SENSITIVITY
In spite of careful design, checked by equally careful calculation, there can be surprises in the
eventual performance of a coating, particularly when it is included in an optical system. There
are many reasons for this but a major cause is the interaction, beyond simple interference,
between the electromagnetic field of the light and the materials of the coating. At optical
frequencies the effect of the magnetic field is vanishingly small and the interactions are all
through the electric field. In linear systems absorption and scattering are the principal effects.
Both of these have magnitudes that are proportional to the square of the electric field amplitude.
The distribution of electric field through a multilayer is therefore of some importance and it is
relatively simple to derive. Its interpretation in terms of absorption and scattering needs some
subsidiary information. Scattering is dealt with in a separate chapter. This chapter is concerned
mainly with losses due to absorption.
The optical properties of any material are determined largely by the electrons and their
interaction with electromagnetic disturbances. Any optical material is made up of atoms or
molecules consisting of heavy positively charged masses surrounded by negatively charged
electrons. These electrons are light and mobile compared with the heavy positively charged
nuclei. An electric field can exerts a force on a charged particle even while it is stationary, but a
magnetic field can interact only when the charged particle moves and for any significant
interaction, the particle must be moving at an appreciable fraction of the speed of light. At the
very high frequencies of optical waves the magnetic interaction is virtually zero, so that any
interaction is entirely through the electric field. Where the electric field amplitude is high the
potential for interaction is high. When thin film optical coatings are illuminated by light,
standing wave patterns form which can exhibit considerable variations in electric field amplitude
both in terms of wavelength and position within the coating. We have already discussed simple
techniques for assessing these amplitude variations. From them deductions about losses can be
made, sometimes with surprising results.

Figure 19-1. Lines of constant electric field amplitude for dielectric materials in the admittance diagram.. The
figures are in volt/metre if the transmitted irradiance is 1watt/sq metre.
We recall that contours of constant electric field are therefore lines, normal to the real axis in
the admittance diagram. If we put Y the admittance at any point in a multilayer in free space units
then:
Macleod: Optical Coatings Page 306 Design through Manufacture
T ⋅ Iinc
E = 27.46 volt/metre (19.1)
bg
Re Y

19.1 Absorption Of Light


A very thin slice of absorbing material is embedded in a multilayer. What can we say about the
absorption of this slice?
The result is contained in the expression:

⎡ i sin δ ⎤
⎡ E ′ ⎤ ⎢ cos δ ⎡E⎤
⎢ H ′⎥ = ⎢ y ⎥⎢ ⎥ (19.2)
⎣ ⎦ iy sin δ ⎥ H
⎣⎢ cos δ ⎦⎥ ⎣ ⎦

where the input and exit irradiances are given by

Iin =
1
2
c h 1
Re E ′ ⋅ H ′∗ and Iexit = Re E ⋅ H ∗
2
c h (19.3)

The irradiance lost by absorption in the layer is the difference between these two quantities. Now
let the layer be extremely thin. Since the layer is absorbing δ is given by

δ=
b
2π n − ik d g = α − iβ (19.4)
λ

Equation (19.4) defines the quantities α and β. By extremely thin, we mean that d/λ should be
sufficiently small to make both α and β vanishingly small, whatever the size of either n or k.
Then,

⎡ i sin (α − i β ) ⎤
⎡ E ′ ⎤ ⎢ cos (α − i β ) ⎥⎡E⎤
⎢ H ′⎥ = ⎢ y ⎥ ⎢H ⎥
⎣ ⎦ ⎢iy sin α − i β
⎣ ( ) cos (α − i β ) ⎥⎦ ⎣ ⎦
⎡ i (α − i β ) ⎤
⎢ 1 ⎡E⎤
=⎢ ( n − ik ) Y ⎥⎥ ⎢ ⎥ (19.5)
H
⎢i (α − i β )( n − ik ) Y 1 ⎥⎣ ⎦
⎣ ⎦
⎡ i (α − i β ) H ⎤
⎢ E+ ⎥
=⎢ ( n − ik ) Y ⎥
⎢i (α − i β )( n − ik ) YE + H ⎥
⎣ ⎦

where we are including terms up to the first order only in α and β.


Macleod: Optical Coatings Page 307 Design through Manufacture

The irradiance at the entrance to this thin layer will then be given by

1 ⎡ ⎪⎧ i (α − i β ) H ⎪⎫ ∗⎤
I in = Re ⎢ ⎨ E + ⎬ ⋅ {i (α − i β )( n − ik ) YE + H } ⎥
2 ⎣⎢ ⎪⎩ ( n − ik ) Y ⎪⎭ ⎦⎥
1
= Re ⎡ E ⋅ H ∗ + E ⋅ {−i (α + i β )( n + ik ) YE ∗ }⎤⎦
2 ⎣
1 ⎡ i (α − i β ) H .H ⎤

+ Re ⎢ ⎥
2 ⎣ ( n − ik ) Y ⎦

(19.6)

The second of the two terms in (19.6) simplifies to

1 ⎡ i (α − i β ) H .H ⎤ 1 ⎡ i (α − i β )( n − ik ) H .H ⎤
∗ ∗

Re ⎢ ⎥ = Re ⎢ ⎥
2 ⎣ ( n − ik ) Y ⎦ 2 ⎢
⎣ ( n 2
+ k 2
) Y ⎥⎦

1 ⎡ { β n − α k + i (α n + β k )} H .H ⎤

= Re ⎢ ⎥ (19.7)
2 ⎢
⎣ ( n 2
+ k 2
) Y ⎥⎦

1 ⎡ ( β n − α k ) H .H ⎤

= ⎢ ⎥
2 ⎢ ( n2 + k 2 ) Y ⎥
⎣ ⎦

2πkd 2πnd
But βn − αk = n− k=0 (19.8)
λ λ

The first term gives

1
I in = Re ⎡ E ⋅ H ∗ + E ⋅ {−i (α + i β )( n + ik ) YE ∗ }⎤⎦
2 ⎣
(19.9)
1 1
= Re ⎡⎣ E ⋅ H ∗ ⎤⎦ + ⎡⎣(α k + β n ) YE ⋅ E ∗ ⎤⎦
2 2

4πnkd
where αk + βn = and E ⋅ E ∗ = E 2 (19.10)
λ

The irradiance that has been absorbed is therefore given by the difference between the irradiance
incident on the thickness element, Iin, and that emerging on the exit side, Iexit. and this is

2πnkd
Iabsorbed = ⋅ Y ⋅E2 (19.11)
λ
Macleod: Optical Coatings Page 308 Design through Manufacture

The magnitude of the absorbed energy is directly proportional to the produce of n and k. Both
must be nonzero for absorption to occur. The absorption will be small both for a metal with
vanishingly small n and a dielectric with vanishingly small k. The factor involving n and k may
be thought of as a phase thickness multiplied by k or as a quantity β multiplied by n.
What is the contribution to the absorption A of the multilayer?
Potential transmittance, ψ, of any element of a coating system is defined as the ratio of the
output to the input irradiances, the input being the net irradiance rather than the incident.
Potential transmittance has several advantages over transmittance when dealing with absorbing
systems because it completely avoids any problems associated with the mixed Poynting vector in
absorbing media. The potential transmittance of a complete system is simply the product of the
individual potential transmittances.

Iexit
ψ= (19.12)
Iin
ψ system = ψ 1 ⋅ ψ 2 ⋅ ψ 3 ⋅ ψ 4 ⋅ ψ 5 ...ψ q

with the eventual overall transmittance given by

b g
T = 1 − R ⋅ ψ system (19.13)

The potential transmittance of the thin elemental film is given by

Iexit I
ψ= = 1 − absorbed = 1 − A (19.14)
Iin Iin

where A is the potential absorptance. But

Iin =
1
2
bg
Y ⋅ Re Y ⋅ E 2 (19.15)

where Y is given in free space units. Then

2πnkd 2
ψ = 1− A = 1− ⋅ (19.16)
λ bg
Re Y

and

4π nkd 1
A= ⋅ (19.17)
λ Re (Y )

This result allows interpretation of an admittance locus in terms of potential absorption.


To move from potential absorption to absorption is straightforward when the absorption is
confined to a very thin layer, the rest of the multilayer being essentially transparent. Then the
absorption, A, is given by:
Macleod: Optical Coatings Page 309 Design through Manufacture
b g
A = 1− R A (19.18)

If, however, the absorption is distributed through the layer then the calculation is rather more
involved. Normally the absorption would be calculated by the normal matrix expression for the
entire film and the would be completely accurate. We, however, are looking for a way of
estimating the absorption and its variation through a layer given the locus in the admittance
diagram or the electric field distribution. Let us assume that the absorption is rather small. The
layer may be considered as a succession of slices of equal optical thickness and extinction
coefficient and so the first factor in the expression for A is a constant. Each slice has a potential
absorptance that depends on the real part of the optical admittance following.(19.16) Then the
potential transmittance is given by the product of the individual potential transmittances,

ψ = ψ 1 ⋅ ψ 2 ⋅ ψ 3 ⋅ ψ 4 ...
b gb gb
= 1 − A1 ⋅ 1 − A 2 ⋅ 1 − A3 ... g (19.19)
b g
= 1 − A1 + A 2 + A3 + A 4 +... + A1 A 2 +...

Provided the potential absorptances are small enough the product terms can be neglected and
then the total potential absorptance is given by the sum of the individual absorptances,

A = A1 + A 2 + A3 + A 4 +... (19.20)

In terms of an integral this can be written as

2k 2n
A = ∑Aj = ∫ dδ = ∫ dβ (19.21)
j δ Re (Y ) β Re ( Y )

It is already well known that for a metal, where the performance is determined by β and k, the
losses are determined by the residual magnitude of n, while for a dielectric, where performance
is a function of δ and n, the losses are due to a residual magnitude of k. The theory presents
nothing new in this regard but it is always comforting to find the result that we know is correct.
For both metals and dielectrics we can consider the product of n and k to be the determinant of
quality from the point of view of losses.
We note that for the same potential absorptance, the actual absorptance will be finally
determined by the quantity (1-R). This implies straight away that transmitting systems are much
more sensitive, by orders of magnitude, to absorption losses than reflecting systems. We often
have the experience that materials that are perfectly satisfactory in reflectors, let us down in
transmitting systems. In an admittance diagram, for a coating with any given value of
reflectance, the potential absorptance, and the actual absorptance, will be higher the nearer the
appropriate part of the locus is to the imaginary axis.

19.2 Contamination sensitivity


Optical coatings are rarely used in an ideal environment. They are subjected to all kinds of
environmental disturbances ranging from abrasion to high temperature and humidity. These
cause performance degradation that mostly originates in an actual irreversible and usually visible
Macleod: Optical Coatings Page 310 Design through Manufacture

destruction of the layers. However, performance may be degraded in a rather less spectacular
way by the simple acquisition of a contaminant that may have no aggressive effect on the layers
other than a reduction of the level of performance of the coating as a whole. The action of water
vapor that is adsorbed by a process of capillary condensation and causes a spectral shift of the
coating is well known. Here we are concerned with much smaller amounts of absorbing material
such as carbon in the form of sub-molecular over the surface after deposition.
Although there are many tests for the assessment of the resistance of a coating to most
environmental disturbances there is no standard test for the measurement of susceptibility to
contamination. Yet it can be shown that the response of coatings can vary enormously,
depending on many factors including design, wavelength, and even on errors committed during
deposition. The reason may be that, often, careful cleaning will restore the performance but this
does not avoid the degradation in between cleanings, and more frequent cleanings are required
for more susceptible coatings.
Fortunately it is possible to make some predictions of coating response to low levels of
contamination and, especially, to make assessments of comparative sensitivity. Electric field
distribution and potential absorption are the keys to understanding the phenomenon.
We can think of the contamination as a vanishingly thin absorbing layer that is deposited over
the outer surface. The actual absorptance, A, will be just the potential absorptance times (1-R).
Provided the contaminant is sufficiently thin then we can use the same Ye to define both potential
absorptance and the reflectance. Then a little analysis yields:

A = (1- R ) A

=
4π nkd

(1- R ) (19.22)
λ Re (Ye )
4π nkd 4 y0
= ⋅
λ 2
⎡⎣1 + Re (Ye ) ⎤⎦ + ⎡⎣ Im (Ye ) ⎤⎦
2

If we define H by:

4π nkd
H= (19.23)
λ

and replace Ye by x-iz, then we find

A 4 y0
= (19.24)
H ( y0 + x )2 + z 2

H is a measure of the absorption capacity of the film, while A is the actual absorptance. We can
think of A/H as a measure of the sensitivity to absorptance of an optical coating. this sensitivity
is purely a function of the optical admittance of the complete coating. Further, it can be deduced
from (19.24) that contours of constant sensitivity are circles in the admittance plane that are
centered on the point –y0 and exhibit decreasing sensitivity with increasing radius. Figure 19-2
shows some contours for normal incidence in air. We can see immediately that antireflection
Macleod: Optical Coatings Page 311 Design through Manufacture

coatings that terminate at the point 1.0, corresponding to the air incident medium, all have equal
sensitivity of 1.0, regardless of their design. We see also that the maximum possible sensitivity
for air as incident medium is 4.0, that is four times that of the perfect antireflection coating. For
coatings with reflectance other than zero there can be great variation. High reflectance coatings,
consisting of a quarterwave stack have least sensitivity if they terminate as far out on the real
axis as possible. This implies that they should end with a high-index quarterwave layer.

Figure 19-2. Circles of constant contamination sensitivity in the admittance plane calculated for an incident
admittance of 1.0. Greatest sensitivity corresponds to the origin where the value is 4.0.
A particularly interesting case is the extended zone high reflectance coating. A plot of the
reflectance of such a system consisting essentially of two staggered quarterwave stacks, with
some adjustment of the outermost layers and those at the junction of the two stacks, is shown in
Figure 19-3. The contamination sensitivity, compared with a simple quarterwave stack is shown
in Figure 19-4. The sensitivity to contamination of the extended zone reflector reaches values
much more than an order of magnitude greater. It also exhibits considerable variation with
wavelength.
The reason is not difficult to find. Because of the considerable thickness of the extended zone
reflector, its front surface admittance travels around the admittance plane, frequently passing
very close to the origin, where the sensitivity is at its maximum. Another way of understanding
the behavior is that over much of the range the incident light penetrates into the coating altering
the phase change on reflection. As a result the electric field amplitude at the front surface
exhibits considerable variation, reaching values that can be twice the incident amplitude. When
the front surface amplitude is equal to the incident amplitude the contamination sensitivity is 1.0,
and so a field that is double that of the incident wave will yield a sensitivity of 4.0.
Macleod: Optical Coatings Page 312 Design through Manufacture

Figure 19-3. The reflectance of the extended zone reflector described in the text.

Figure 19-4. The contamination sensitivity in air of the extended zone reflector compared with that of a
simple quarterwave stack. The extended zone reflector exhibits a sensitivity that is more than an order of
magnitude greater.
The Essential Macleod Function Enhancement includes a script, Contamination Sensitivity,
that will calculate the contamination sensitivity of any coating
Macleod: Optical Coatings Page 313 Design through Manufacture

20 OPTICAL COATINGS IN COMMUNICATIONS


20.1 Introduction
Demand for communication capacity is increasing constantly. The single-mode fiber, thought
in the 1980’s to be sufficient for years to come, has proved insufficient to handle the traffic due
to the internet. Combining channels in what is known as Wavelength Division Multiplexing is
now increasing capacity. The capacity of a single channel is limited by the bit rate capability of
the fiber. This has risen from 2.5Gb/s to, currently, 10Gb/s, but is soon to be increased to 40Gb/s
and even to 80Gb/s. The primary wavelengths of interest for long haul applications are the C-
band (192-196THz, i.e. 1528-1561nm) and the L-band (185-192THz i.e. 1561-1620nm). For the
purposes of this contribution we assume a wavelength of 1550nm. The wavelengths of the
various channels in the bands have been defined by the International Telecommunications Union
in what is usually known as the ITU Grid. This takes a frequency of 196.1THz (1528.77nm) as
starting point and spaces channel frequencies at intervals of 50GHz from this value. Every fourth
channel, therefore, is spaced at 200GHz and every second at 100GHz. At 1550nm, 50GHz,
100GHz and 200GHz spacings translate into wavelength intervals of 0.4nm, 0.8nm and 1.6nm
respectively. Current emphasis is on 200GHz intervals with 100GHz gradually increasing in
importance and 50GHz largely in the future. Since the channels are carrying information they
have a finite bandwidth. A pulse rate of 10Gb/s can not have a bandwidth less than 5GHz but in
practice is much more. Allowance must also be made for tolerances, influence of temperature,
wavelength variation and other drifts and so the spectral width for a single channel at 200GHz
spacing is frequently specified as around 0.7nm and for 100GHz as 0.3nm. Sometimes the
requirement may be in the form of a formula that includes the temperature coefficient of the
component. More information will be found in Kartalopoulos[1].
There are many areas in telecommunications where optical coatings are required. This
presentation considers just five of these, Wavelength Division Multiplexing, the Separation of
the C and L bands, Gain Equalization Filters, Antireflection Coatings and Dispersion Correction
Filters. A certain level of knowledge of optical coatings is assumed. For more fundamental
details see Macleod[2].

20.2 Wavelength Division Multiplexing


The components for multiplexing and demultiplexing are, in their most general form, four-
port devices. The separation of one channel from the others is known as dropping and the adding
of a fresh channel as adding. These operations are indicated schematically in Figure 20-1.
Typical requirements for such a device are shown in Table 20-1. The requirement for Coherent
Crosstalk is not one that can be met with thin film devices at present so that they must be three-
port drop or add devices only. The specification is such that narrowband transmitting filters are
preferred over narrowband reflectors. In the C and L bands the term dense is used to describe the
close spacing of the channels and the description Dense Wavelength Division Multiplexing
(DWDM) is usually preferred.
Macleod: Optical Coatings Page 314 Design through Manufacture

In Through

Add Drop

Figure 20-1. The four ports of a drop-Add device for wavelength division multiplexing.

Table 20-1. Requirements for a four-port add-drop device.


Parameter Meaning Requirement Interpretation
Loss suffered in
Insertion any channel due to Out-of-band R ≥ 93%
≤0.3dB
loss passage through In-band T ≥ 93%
the device
Leakage of signal
Incoherent
from one channel ≤-20dB Out-of-band T ≤1%
crosstalk
to another
Leakage of signal
Coherent from the Add to
≤-20dB In-band R ≤1%
crosstalk the Drop channel
and vice versa
Variation of
insertion loss
Ripple ≤0.3dB
across channel
width
Performance in dB (decibel) corresponds to 10log10(1/R) or 10log10(1/T)
Materials in DWDM beamsplitters must have exceptionally low loss. Amorphous materials
are preferred over polycrystalline since they avoid grain-boundary scattering. Stability
requirements imply dense materials and so the deposition is by energetic processes involving
supply of momentum to condensing films. Ion Assisted Deposition, Ionized Plasma Assisted
Deposition and Ion Beam Sputtering are currently most often used. The materials must be
suitable for these processes. Tantalum pentoxide (n = 2.10) and silicon dioxide (n = 1.44) are the
two most frequent choices for coatings at 1550nm.
For reasons connected with manufacture, quarter-wave layers are used virtually exclusively in
the design of narrowband filters. Tolerances are such that optical monitoring of deposition is
necessary. Optical monitoring carries with it the possibility of error compensation[3-5] without
which successful construction is impossible. The error compensation relies on the use of quarter-
wave thicknesses. The final matching layers need not be so constrained and some of the more
advanced designs have one or two non quarterwave thicknesses at the start and/or finish of the
coating. It can readily be demonstrated that a single-cavity filter, however it is designed, is
incapable of meeting the specification. Multiple-cavity filters are necessary. A basic structure:
Macleod: Optical Coatings Page 315 Design through Manufacture

Air | 1.214L 0.378H [ (HL)7 6H (LH)7 L (HL)7 6H (LH)7 L (HL)7 6H (LH)7 ] | Glass

(where we are using the usual convention of H and L indicating quarterwave layers of high and
low index respectively) possesses suitable properties for channel spacing of 200GHz (what is
usually known as a 200GHz filter). The part of the structure inside the square brackets is a three-
cavity structure while the final two layers at the air side form a matching structure that is
essentially a V-coat[2].
For improved performance at the expense of increased complexity, greater difficulty in
manufacture and, usually, decreased accommodation of changes in angle of incidence, (the
passband shape is inclined to change) we can make more use of the concept of symmetrical
systems[2]. Figure 20-3 shows the equivalent admittance of three similar structures except the
outermost layers have different orders. The lower orders can be used as matching layers for the
higher. A design based on this and including the usual V-coat matching is:

Air | 1.244L 0.363H (HL)7 H (HL)15 H [ 4H (HL)15 H 4H ]2 (HL)15 H (HL)7 H | Glass

This design has performance shown in Figure 20-4.


The symmetrical systems need not have precisely the same admittance at the center. The
requirement is that the total matching system should have the correct value and dispersion.
Figure 20-5 shows a 100GHz design with low-index cavity layers that uses a two-stage matching
systems that is retains the same order as the central period but relaxes the total number of
quarterwave layers in the symmetrical period. The order and the numeber of layer can both be
changed and even additional halfwave layers can be inserted in the structures with the intention
of slightly modifying the artificial dispersion to improve the quality of the matching. Note the
final V-coat matching in Figure 20-5.

Wavelength (nm) Wavelength (nm)


1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
0 0.0
-0.2
-10
-0.4
-20
-0.6
-30 -0.8
-40 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 20-2. Theoretical performance of the three-cavity filter. The grey lines indicate the specification.
Macleod: Optical Coatings Page 316 Design through Manufacture

E
1500
(HL)15 H
1000
HH (HL)15 H HH
500 HHHH (HL)15 H HHHH

0
1548 1549 1550 1551 1552
Wavelength (nm)

Figure 20-3. The variation of equivalent admittance of three similar structures of increasing order.

Wavelength (nm) Wavelength (nm)


1548 1549 1550 1551 1552 1548 1549 1550 1551 1552
0 0.0
-20 -0.2
-0.4
-40
-0.6
-60 -0.8
-80 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 20-4. The performance of an improved but more complex design detailed in the text.

Wavelength (nm) Wavelength (nm)


1549.0 1549.5 1550.0 1550.5 1551.0 1549.0 1549.5 1550.0 1550.5 1551.0
0 -0.0

-0.2
-10

-0.4
-20
-0.6

-30
-0.8

-40 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 20-5. A 100GHz filter with design:


Air | 1.29L 0.34H (LH)7 L [4L (LH)15 L 4L] [(4L (LH)16 L 4L)2] 4L (LH)15 L 4L (LH)7 L | Glass.
Simple Monte Carlo modeling, Figure 20-6 (left), shows that the allowable tolerances for
random thickness errors in the three-cavity filter are around 0.003%, several orders of magnitude
beyond the capacity of current thickness control techniques and forcing rapid rotation of the
substrate carrier to avoid random errors created by uniformity defects due to partial rotations.
Transmittance and reflectance of growing films exhibit extrema at quarter-wave thicknesses
implying that such layers exhibit least accuracy in their control suggesting at first sight that
direct first-order monitoring might not be appropriate. Fortunately there is a well-known inherent
error compensation in direct first-order turning-value monitoring, where all layers are controlled
Macleod: Optical Coatings Page 317 Design through Manufacture

at the peak wavelength and on the filter under construction. An error in any layer is compensated
by a similar but negative error in the subsequent layer[3-5]. It is this compensation that enables
the successful construction of such multiple-cavity filters, , Figure 20-6 (right).

Wavelength (nm)

1548 1549 1550 1551 1552


0.0
-0.2
-0.4
Wavelength (nm)
-0.6
-0.8 1548 1549 1550 1551 1552
0.0
-1.0
-0.2
log(Transmittance) (dB)
-0.4
-0.6
-0.8
-1.0
log(Transmittance) (dB)

Figure 20-6. Left: The effects of random thickness errors of 0.003% standard deviation. The results of ten
different sets of errors drawn from the same normal infinite population are shown. Errors of 0.004% gave
performance outside specification. Right: Simulated direct turning-value monitoring of the filter with λ0 =
1550nm, signal noise 0.05% and minimum signal change 0.2%, with coupling layers monitored separately by
quartz crystal microbalance with random errors of 2% standard deviation, gives results almost
indistinguishable from theoretical performance.
An admittance diagram of a pair of layers makes the origin of the compensation clear when it
is remembered that the extrema of reflectance and transmittance correspond to the intersections
with the real axis, Figure 20-7. For the direct monitoring to be possible the spectral bandwidth of
the monitoring beam must be considerably less than the bandwidth of the filter under
construction. A factor of 0.3 represents the desirable upper limit. Note that the compensation is
not effective in the correct sense with the coupling layers and it is better to control them
separately. Monitoring is considered in more detail in another section.
Macleod: Optical Coatings Page 318 Design through Manufacture

1.0

0.5 Re(Admittance)

0.5 1.0 1.5 2.0 2.5 3.0 3.5


0.0

Error
-0.5

-1.0

-1.5
Im(Admittance)

Figure 20-7. The admittance diagram of an HL sequence shows the effect of an error of 20% in the first (H)
layer. If the L layer is still terminated at the extremum the resultant error is very small.

20.3 Separating the C and L bands


A more difficult requirement is a filter to separate the C- and L-bands. A wider filter has a less
steep edge for the same number of layers. We therefore choose a specification that transmits all
the C-band 1528-1561nm and reflects the L band 1566-1620nm.
Table 20-2. Specification for a coating to separate the C and L bands.

11528-1561nm ≥-0.3dB ≥93%


Transmittance
1566-1620nm ≤-20dB ≤1%
Using techniques similar to those in the previous sections we arrive at a design:
Air | (HL)2H (HL)6H [HH (HL)3 LL H LL (LH)3 HH]q (HL)6H (HL)2H LH | Glass
where a certain amount of trial and error has been used. This particular design uses high-index
coupling layers (those in the center of the central symmetrical structure). This is intentional. The
result is that the extra V-coat matching must come between the filter and substrate and the LH
layers represent that V-coat. The parameter q represents the number of times the central structure
must be repeated. Two calculations with values for q of 7 and 9 are included in Figure 20-8.
An alternative approach is an edge filter. Figure 20-9 shows an edge filter consisting of a
quarterwave stack of some 103 layers surrounded by 9 and 10-layer matching structures where
the layers are of thicknesses other than quarterwaves. The filter design meets the specification
and has an even better ripple performance than the multiple-cavity designs but it is doubtful if
such a filter could actually be manufactures. Monitoring would represent a major challenge.
Macleod: Optical Coatings Page 319 Design through Manufacture

Wavelength (nm) Wavelength (nm)


1520 1540 1560 1580 1600 1620 1520 1540 1560 1580 1600 1620
0 0.0
-0.2 q=7
-20
-0.4
-0.6
-40 q=9
q=9 q=7 -0.8

-60 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 20-8. The multiple-cavity filter to separate the C and L bands.

Wavelength (nm) Wavelength (nm)

1520 1540 1560 1580 1600 1620 1520 1540 1560 1580 1600 1620
0 0.0
-0.2
-20
-0.4

-40 -0.6
-0.8
-60 -1.0
log(Transmittance) (dB) log(Transmittance) (dB)

Figure 20-9. A short-wave-pass filter solution with design:


Air | L (HL)4 (HL)51H (LH)5 | Glass
The design has 122 layers that compares with 119 for the advanced band-pass filter. Unfortunately it is
difficult to make. Matching layers L and H are not quarterwaves. λ0 = 1747.5nm

20.4 Gain Equalization


Erbium-doped fiber amplifiers (EDFA) have gain that varies with wavelength. A typical curve
using data from Madsen[6] is shown in Figure 20-10. Curves issued by different manufacturers
show considerable variation but have usually the rather limited region of higher gain around
1530. A gain equalization filter can only reduce the gain and thus the desired level of the
effectively flat gain fixes the spectral width of the filter-amplifier combination. For the purposes
of demonstration we choose 2dB. This gives us a specification as shown in Figure 20-11 where
all required points that would demand a transmittance above 100% have been removed.
Macleod: Optical Coatings Page 320 Design through Manufacture

Gain (db)
10

0
1520 1530 1540 1550 1560 1570
-2 Wavelength (nm)
-4

Figure 20-10. Gain of an erbium-doped fiber amplifier after Madsen[6].

Transmittance (%)
100

80

60

40

20

0
1520 1530 1540 1550 1560
Wavelength (nm)

Figure 20-11. The gain equalization specification in terms of transmittance. The reflectance specification
would look exactly the same except the scale would read reflectance.
Some information on what can be expected in the design of such filters can be derived from
the appearance of the curve. The spacing of the fringes, ∆λ, is around 10nm. This means that we
have an order number of λ/∆λ that is very roughly 1550/10 or 155. Each increase in order
demands another full wave in path difference or a halfwave in optical thickness and so we can
expect a total thickness for the filter of something in the region of 155/2, that is 72.5, full waves
maybe a little smaller because the fringe spacing is, in fact, a little greater than 10nm. We can
also foresee a problem in the rather sharp fringe at 1540nm. Simple reflection fringes that are
near 100% have flat tops and narrow bases so a design in reflection will have difficulty with the
region in the close vicinity of 1540nm and will likely require greater than the minimum
Macleod: Optical Coatings Page 321 Design through Manufacture

thickness of material to deal with it. Transmission fringes tend to have narrow tops and broader
bases, and so a filter operating in transmission will tend to be easier to design and to have
thickness nearer the minimum. There is no systematic way of designing such filters at the present
time. Brute force, however, can be quite effective. For the coating illustrated here, the process
used for design was straightforward computer synthesis. Starting materials were chosen to be
silica and tantala. A nine-layer design derived in this way is shown in Figure 20-12 and the final
performance compared with the specification in Figure 20-13. The total thickness of this design
is 63 full waves at 1550nm.
Both filters have some rather thick layers. No constraints were placed on film thickness in the
process of design. Hard constraints on layer thicknesses tend to inhibit refinement and synthesis
processes and render them rather less effective. However, if thick layers represent a problem
then they can be broken down into smaller components although the total thickness of the filter
is unlikely to be much affected. Manufacture is clearly a much greater problem than design.

Figure 20-12. A design of a gain equalization filter this time having only nine layers. Optical thickness in
units of the reference wavelength, λ0=1550nm, are shown.
Macleod: Optical Coatings Page 322 Design through Manufacture

Wavelength (nm)
1520 1530 1540 1550 1560
0
-1
-2
-3
-4
-5
-6
-7
-8
Log(T) (dB)

Figure 20-13. The performance of the 9-layer filter of Figure 20-12 and the specification in terms of dB
shown behind it in grey.

Wavelength (nm)
1520 1530 1540 1550 1560
0
-1
-2
-3
-4
-5
-6
-7
-8
log(Transmittance) (dB)

Figure 20-14. The effect of random absolute errors of 0.0025 full waves in layer optical thickness plotted in
terms of db. The spread in curves is around 0.25db.
The layers in the Gain Equalization Filters are very thick and this is important in assessing
allowable errors. The significance of an error in thickness is largely the difference in phase that it
induces between the interfering beams in the structure. This is more a question of absolute
thickness change rather than proportional thickness change. A 1% error in a layer that is ten
quarterwaves thick, for example, will be ten times as serious as a 1% error in a single
quarterwave because it represents a phase change of ten times 1% of a quarterwave. This means
that the thickness tolerances in terms of proportional errors will be very small for any design
with thick layers. This, however, is not strictly how monitoring systems actually operate. An
optical system, for example, will count extrema and the error in terminating the layer will apply
Macleod: Optical Coatings Page 323 Design through Manufacture

simply to the last fractional thickness beyond the final extremum. A more realistic error analysis
would use the error in the final quarterwave rather than a proportional error. Quartz crystal
monitors, too, will give rather better accuracy for a thick layer than the 2% or so that can be
expected for much thinner layers. But the crystal must be very carefully calibrated and the
tooling factor accurately known so it is less likely to be as effective in this application as optical
monitoring. Figure 20-14 shows the effects of an absolute thickness error of standard deviation
0.0025 full waves. This is equivalent to a 1% error in a quarterwave and represents what might
be achieved by extreme care in optical monitoring. The performance is satisfactory.

20.5 Antireflection Coatings


Antireflection coatings reduce reflectance and increase transmittance. These are usually
identified with crosstalk and insertion loss respectively. The two requirements are related, the
requirements for crosstalk reduction being normally more severe.

Table 20-3. Crosstalk and insertion loss comparison for a typical antireflection coating.
Reflectance Crosstalk Transmittance Insertion loss
0.1% -30dB 99.9% -0.0043dB

We take as a requirement low reflectance, ≤0.1%, over the C and L bands (1528-1620nm).
Theoretically this can be achieved with a single V-coat but Figure 20-15 shows that even with
1% random thickness errors, representing rather better than normal accuracy in coating
deposition, the specification is not quite achieved for both band simultaneously.

Crosstalk Crosstalk
-20 -20
C-Band L-Band C-Band L-Band
-30 -30

-40 -40

-50 -50

-60 -60

-70 -70
1500 1550 1600 1650 1500 1550 1600 1650
Wavelength (nm) Wavelength (nm)

Figure 20-15. V-coat antireflection coating with 2% random thickness variations (left) and 1% (right). Even
1% does not quite meet the specification of the C and L bands.
An alternative design, involving six layers gives better stability against errors. The design is a
double V-coat structure with a partial buffer layer inserted in the outermost low index layer. the
admittance diagram, Figure 20-16 left, should help to explain. This design is rather more
complex than the two-layer one but has the advantage of being rather less sensitive to errors,
Figure 20-16 right.
Macleod: Optical Coatings Page 324 Design through Manufacture

Table 20-4. The design of a six-layer antireflection coating.


Incident Air Massive
Layer 1 SiO2 247nm
Layer Ta2O5 269nm
Layer SiO2 133.7nm
Layer Ta2O5 28nm
Layer SiO2 379nm
Layer Ta2O5 20.2nm
Substrate S 7003 glass (n=1.55) Massive

Im(Admittance) Crosstalk (dB)


-20
0.75
C-Band L-Band
0.50 -30

0.25 -40

0.00 -50
1.00 1.25 1.50 1.75 2.00 2.25 2.50
-0.25 -60
Re(Admittance)
-0.50 -70
1500 1550 1600 1650
-0.75 Wavelength (nm)

Figure 20-16. The six-layer design. Left shows the admittance diagram at 1550nm. Right shows the
theoretical performance (black) together with the effects of random thickness errors of 2% standard
deviation. This meets the specification for both the C and L bands simultaneously.

20.6 Dispersion Compensation


Thin film filters are potentially useful in compensating for the effects of dispersion in
components such as optical fibers[7]. It is always easier to design reflecting compensators
because perfect dielectric layers added to a perfect reflector affect only the phase but not the
magnitude of the reflectance. In practice the reflector is not perfect and the layers not completely
dielectric so that there is usually a small effect on reflectance. Additional layers can always be
added to the reflector to boost its performance.
The important parameters involved are defined as follows:
Macleod: Optical Coatings Page 325 Design through Manufacture


Group Delay GD = − (time)

d 2ϕ
Group Delay Dispersion GDD = − (time2 )
dω 2
d 3ϕ
Third Order Dispersion GD = − (time3 )
dω 3

These affect the envelopes of the pulses, delaying, broadening and distorting them.
In 1964 Gires and Tournois[8] suggested the use of an asymmetric interferometer, operating
in reflectance, for pulse compression. The system can be thought of as a thin-film resonator
deposited over a reflector. The reflected light enters the resonator and takes time to emerge so
that there is a delay. The greatest delay is at the resonant wavelength of the resonator.

Group
Delay

Resonator Reflector

Wavelength
Figure 20-17. Schematic of a dispersion compensator base on the Gires-Tournois[8] structure.

Reflectance Group Delay (ps) Reflectance Group Delay Dispersion (ps^2)


3.0 3 5
Wavelength (nm)
2 1549 1550 1551 1552
0
2.0
1

0 -5
1.0 1549 1550 1551 1552
-1 Wavelength (nm)
-10
0.0 -2
1548 1549 1550 1551 1552
Wavelength (nm) -15
-3
Reflectance Third Order Dispersion (ps^3)

Figure 20-18. The reflectance group delay (left), group delay dispersion (middle) and third-order dispersion
(right) of the dispersion compensator in the text.
The resonator can be quite complex. Here we consider a very simple arrangement.

Air | (HL)6 HH (LH)6 L (HL)7 H | Glass

which uses the usual silicon dioxide and tantalum pentoxide materials. The performance of this
design is shown in Figure 20-18.
Macleod: Optical Coatings Page 326 Design through Manufacture

The dispersive properties are all linear and so the resultant dispersive properties of a series of
reflecting coatings is just the sum of the individual ones. This means we can add the
characteristics of any number of such compensators by cascading them. Here, Figure 20-19, four
of the simple resonators are each offset by 0.5nm to give essentially constant third-order
dispersion over a small range of wavelengths.

Reflectance Third Order Dispersion (ps^3)


20

10

0
1549 1550 1551 1552
Wavelength (nm)
-10

-20

Reflectance Group Delay Dispersion (ps^2)


8

0
1549 1550 1551 1552
-2
Wavelength (nm)
-4

-6

-8

Figure 20-19. Here four of the simple resonators of Figure 20-18 are each offset by 0.5nm to give essentially
constant third-order dispersion when they are cascaded as shown.
Metal layers have been used to eliminate some of the dielectric layers in the underlying
reflectors of the compensating filters. The design principles are, however, not changed by this.

20.7 References
1. Kartalopoulos, Stamatios V, Introduction to DWDM Technology: Data in a Rainbow.
First ed. 2000, New York: IEEE Press.
2. Macleod, H A, Thin-Film Optical Filters. Second ed. 1986, Bristol: Adam Hilger.
3. Macleod, H A, Turning value monitoring of narrow-band all-dielectric thin-film optical
filters. Optica Acta, 1972. 19: p. 1-28.
4. Macleod, H A, Absorption in turning value monitoring of narrow-band thin-film optical
filters. Optica Acta, 1973. 20: p. 493-508.
Macleod: Optical Coatings Page 327 Design through Manufacture

5. Macleod, H A and D Richmond, The effect of errors in the optical monitoring of narrow-
band all-dielectric thin film optical filters. Optica Acta, 1974. 21: p. 429-443.

6. Madsen, Christi K and Jian H Zhao, Optical Filter Design and Analysis. First ed. Wiley
Series in Microwave and Optical Engineering, ed. K. Chang. 1999, New York: John
Wiley & Sons, Inc.
7. Jablonski, Mark, Yuichi Takushima, and Kazuro Kikuchi, The realization of all-pass
filters for third-order dispersion compensation in ultrafast optical fiber transmission
systems. Journal of Lightwave Technology, 2001. 19(8): p. 1194-1205.
8. Gires, F and P Tournois, Interféromètre utilisable pour la compression d'impulsions
lumineuses modulées en fréquence. Contes Rendus de l'Academie de Science, 1964. 258:
p. 6112-6115.
Macleod: Optical Coatings Page 329 Design through Manufacture

21 ANTIREFLECTION COATINGS ON PLASTICS


There is essentially no difference in the optical performance of a coating on plastic compared
with a coating on any other substrate. The principles of design are completely unchanged. The
major problems are connected with mechanical, thermal and chemical behavior. The thin films
are hard, brittle inorganic materials while the substrates are soft and yielding, and they dissipate
energy, properties completely different from those of the films. These differences in properties
are the prime cause of problems.

21.1 Substrate Materials


There is a number of plastic substrate materials in use for spectacle lenses and windows today.
Typical of the thermosetting resins is CR 39, and then there are some essentially similar
materials that have rather higher refractive index permitting thinner (and lighter) lenses for a
given correction. The thermosetting resins once set are essentially unmodified by application of
heat and have good mechanical properties. Thermoplastic resins are also used. They have the
advantage that they can be injection molded. Early thermoplastics like polymethylmethacrylate
(PMMA) had exceedingly poor abrasion resistance but polycarbonate has been found to be more
satisfactory. It has a relatively high softening point of 140°C or higher and it has very high
impact resistance. Cyclo-olefin thermoplastic polymers (COP) with exceedingly good optical
transmittance and refractive indices around 1.52, similar to crown glass, are also beginning to
appear as optical substrates.

21.2 Durability
To improve the abrasion resistance, hard coats, as they are known, are frequently applied. The
most common hard coat is a UV-cured acrylic coating. Polysiloxane is another common coating.
These are strictly polymeric coatings but somewhat more resistant and harder than the very soft
underlying material. Sometimes the hard coat is deliberately made porous so that it can be
stained, often with a dye that absorbs ultraviolet light. This is particularly the case with
polycarbonate materials. A stainable hard coat is considerably softer than a normal one. The
usual hard coat is, therefore, essentially still a polymeric and plastic material. However, new
nanocomposite materials are under development consisting of a matrix of organo silicon
polymeric material with colloidal particles of silica dispersed in it.
Plastics are organic while most optical coatings are inorganic. Plastics are rather soft, can
accommodate large dimensional changes, have high coefficients of thermal expansion and
dissipate strain energy when responding to changes in applied stress. The inorganic materials
used for thin films are quite brittle, they have no sensible mechanisms for dissipating strain
energy other than fracturing so that they are unable to accommodate large dimensional changes
without catastrophic failure. Their thermal coefficients of expansion are comparatively small.
The bonds that hold plastics together are quite weak but the molecules are in the form of long
chains with very many bonds linking with neighboring molecules. They are also tangled, which
effectively adds to the bonding. Their ability to stretch and to dissipate strain energy means that
although the bonds are weak, the range is effectively large and so a large amount of work must
be done to break them. The ability to stretch prevents the stress concentrating effect of isolated
Macleod: Optical Coatings Page 330 Design through Manufacture

defects and so propagating fractures are virtually eliminated. The inorganic molecules are
attached together by much stronger bonds but they are simply from one small unit to the next and
they form a relatively rigid assembly that is unable to support any major distortion without bond
rupture. The bond force is large but the range is small and so the work that has to be done to
break a bond is comparatively small. Stress concentration at defects is common and fractures can
propagate. Films are brittle and much weaker in tension than in compression.
These differences cause adhesion problems because of the different nature of the bonds. The
brittle nature of the films, their weakness in tension, the large thermal expansion coefficients of
the substrate and the ability to deform plastically are of major importance in cracking and
crazing problems.
There is no ideal solution. Various recipes, almost all of them trade secrets, are used to reduce
the magnitude of the problems. Most of the studies have considerable empirical content.
Adhesion is a complicated topic. It might be thought that improved adhesion would result
from stronger bonds but this is not necessarily true. Bond strength is, of course, a factor in
adhesion but it is a rather misleading one. Electrons are involved in bonding and their energy in a
bonded state is lower than when unbonded. To move from a bonded state to an unbonded one, in
other words to break a bond, means that extra energy must be supplied. The supply of energy
will be in the form of work and we can think of this as the work that must be done to break the
bond. The bonds that hold inorganic materials together are very strong but their range is very
small and so the work that has to be done to rupture them is small. When an adhesion failure
takes place, two fresh surfaces are generated. The broken bonds on these surfaces possess their
characteristic increased energy. This is known as surface energy, an older term being surface
tension. To induce the adhesion failure, work equal to the new surface energy must be supplied
and this is known as the work of adhesion. The usual failure mode is that of a peel adhesion
failure where bonds break successively rather than simultaneously and the breaking of each bond
transfers the load to the next which breaks in turn. The failure may be driven by internal strain
energy rather than an external agent. A peel adhesion failure may be accompanied by a
deformation of the materials on either side of the failure. If this deformation dissipates energy
then additional work must be supplied to the failure process. This implies that the work of
adhesion must increase, even when the bonds retain exactly their previous strength. The plastic
substrates exhibit this kind of dissipative behavior but the brittle optical coatings do not.
Plastic surfaces are known as low energy surfaces. This means that the surface energy
associated with the broken bonds is small. The inorganic oxides, nitrides and fluorides typical of
optical coatings have higher surface energy. The work of adhesion in plastics receives a
considerable contribution from the dissipation of energy in deformation but in brittle inorganics
it is primarily from the new surface energy.
Coatings are normally in a state of considerable stress. The net stress in the coating is
supported completely by the interface between coating and substrate. Any lack of adhesion here
or any defects are therefore likely to be converted into an adhesion failure. A compressive stress
in the coating is desirable but this exerts a considerable shear stress across the final interface.
The softness of the substrate and its ability to deform plastically makes it a poor support for
the brittle film system which tends to crack when it is subjected to a concentrated load, such as
an abrading particle. The bending of the coating places the outer part of the curved surface in
Macleod: Optical Coatings Page 331 Design through Manufacture

tension. The films are weak in tension and the cracking then propagates across the film. For a
stationary load the cracks define a set of roughly trapezoidal plates that define an approximately
circular area around the point load. When the abrading particle is moving, the damage in
enhanced because the friction tends to lifts some of the cracked plates from the surface and the
particle then ploughs its way across the coating leaving a shattered track of damage behind it.
This scratch can then act as a stress concentrator enhancing the probability of further abrasion-
induced damage.
If the internal strain energy is tensile this can also cause tensile cracking because the inorganic
thin films are weak in tension but strong in compression. It is therefore desirable that the films
should be compressively strained and the substrate must then support the resulting compressive
stress. A problem with plastic substrates is that they have high coefficients of expansion and the
compressive stress in the film system may become tensile at elevated temperatures. Even if the
stress is not immediately converted to tensile the strain energy may be partially dissipated in the
plastic substrate. Over a number of thermal cycles this effect can accumulate and eventually
cause tensile cracking.

21.3 Surface Treatment


Hard coats are intended to stiffen the substrate surface so that a point load is not entirely
supported by the thin coating. These hard coats are frequently applied to the plastic surface as a
separate process before any antireflection coating. Their details are somewhat proprietary but the
commonest are applied as a liquid that is then cured by exposure to ultraviolet light or by baking,
perhaps both. Acrylic or polysiloxane varnishes are normal but there are other possibilities.
Silica films, sometimes vacuum-deposited, may be used. Direct hardening of the plastic surface
itself, by bombarding it with energetic fragments from a discharge in oxygen or nitrogen is a
further possibility. Nanocomposite varnishes containing colloidal silica particles are a promising
new approach. Note that some scratch resistant coatings, particularly for uncoated lenses, really
act just to reduce the coefficient of friction at the surface to reduce the load imparted by an
abrading object.
The hard coat also acts as a transition layer between the coating and the plastic that appears to
help to reduce the incidence of tensile cracking from temperature cycling. It is usually a micron
or two in thickness. An increase in thickness does lead to improved abrasion resistance, as would
be expected, because the deflection under a point load is reduced, but, for reasons that are not
clear, the thicker hard coat usually appears to have a reduced resistance to crazing on
temperature cycling.
The most common techniques employ a transparent varnish that is applied in thicknesses of
several microns by dipping, spinning or spraying. The varnish is then usually cured by exposure
to ultraviolet light or sometimes simply by heating. Acrylic resins are most commonly used in
the ophthalmic lens industry. [Acrylic resins are thermoplastic materials produced by the
polymerization of methacrylic acid esters (methacrylates) or acrylic acid esters (acrylates).]
Silicone varnishes consist of a polysiloxane polymer in a solvent, usually alcohol. The
characteristic of these materials is a silicon-oxygen backbone. The oxygen and silicon crosslink
under ultraviolet light to give a tough hard coating. Nanocomposite lacquers are now beginning
to be used. These consist of lacquers to which have been added particles of mineral, usually
silica, of size some 10 to 100nm
Macleod: Optical Coatings Page 332 Design through Manufacture

Sometimes called plasma polymerization, plasma enhanced chemical vapor deposition or


PECVD is capable of depositing dense organic layers with stable optical properties over curved
and irregular surfaces with good uniformity. By altering the precursors, that is the starting
material, the index of refraction can be varied so that the substrates can be matched in index and
the fringes of the conventional process avoided. Plasma polymerization is quite unlike normal
polymerization where monomers are linked into chains of repeat units. The plasma is
characterized by energetic electrons that break the reactants into active fragments and these
fragments link with each other to form the deposited film. Some of this combination may take
place in the gaseous phase forming clusters that may deposit on the growing film or may be
broken into fragments again by the plasma. Strong binding occurs so that the deposited film is
tough and hard and dense. It is not strictly polymeric and contains free radicals that may
combine also with any oxygen that is also present. The mechanical properties can range from
plastic to elastic and glass-like. Because the films are insulating, in fact they are used as
capacitor dielectrics in some applications, RF discharges are usual for this process. Speed of
deposition can be very high, up to 1µm per minute although rates of one tenth to one hundredth
of this are more common.
The process has been used for some time in the semiconductor industry to deposit silicon
dioxide. The normal precursor is tetraethoxysilane (TEOS) together with oxygen but the
substrate temperature is usually quite high, at least 250°C, much higher than can be possible for
plastic substrates. When the temperature is reduced to permit coating of plastic substrates, the
film composition becomes much more complicated. Apart from the silicon oxide content they
include, for example, silanol that results from reactions involving residual water vapor. There
are, in fact, many silicone compounds that can be and have been used as precursors in the
PECVD deposition of such silica rich films. The feature that they tend to share is a backbone of
alternate silicon and oxygen atoms. Apart from the tetraethoxysilane already mentioned other
suitable compounds include hexamethyldisiloxane (HMDSO), tetramethoxysilane (TMOS),
methyltrimethoxysilane (MTMOS) and trimethylmethoxysilane (TMMOS). As might be
expected, they are toxic, although their toxicity varies.
Surfaces may also be subjected to a plasma to improve their properties[1]. There are two
major aspects of plasma treatment. Energetic particles, ions, neutrals and electrons bombard the
plastic surface. Also ultraviolet light that is produced in the plasma irradiates the surface. Both
the energetic particles and the ultraviolet photons have sufficient energy to rupture the bonds in
the plastic molecules. The rupturing of the bonds implies the breaking up of the molecules into
functional groups or even free radicals. These functional groups may relink, crosslink, branch or
link with molecules from the plasma, or they may leave the surface altogether. The nature of the
plasma has a major influence on the induced surface properties. Fluorine-containing plasmas
tends to produce fluorocarbon surfaces that are hydrophobic. Oxygen-containing plasmas
produce surfaces that are hydrophilic. Plasma processing using an oxygen plasma is one of the
more effective treatments for improving the adhesion of metals or metal oxides at the surface of
the plastic. Oxygen forms a particularly aggressive plasma including molecular and atomic ions.
Even fluoropolymers to which virtually nothing sticks can be sufficiently modified in their
surface properties that adhesion to them becomes possible. It is a complicated process. Reference
[1] explains why a DC plasma treatment in argon and water vapor gives greatly improved
adhesion of PMMA while a similar treatment using a radio-frequency plasma does not.
Macleod: Optical Coatings Page 333 Design through Manufacture

21.4 Optical Coatings


The hard coat that is frequently deposited over the plastic before the optical coating, should,
theoretically, have no influence on the subsequent optical properties. However the index is
seldom matched exactly to the substrate and the result is the appearance of closely spaced
fringes, usually of quite small amplitude, superimposed on the characteristic of the coating.

Figure 21-1. Four-layer antireflection coating for a substrate identical to glass (n=1.52) where a hard coat of
index 1.55 and thickness 5 micron is interposed between the coating and the substrate.
The sensors in the eye have spectral bandwidth around 100nm and so the fringes due to the hard
coat are not normally visible.
In the ophthalmic field the residual color is an important feature of antireflection coatings.
Although the reflectance is weak, the saturation of the color is high and small variations can be
detected quite readily. It is difficult to suppress the color completely and so the color, which
might be considered a defect, has become a rather positive fashion statement and nowadays is
considered part of the coating design. Variations in the quality of the color are usually
considered to be a greater defect than variations in residual reflectance levels. It is particularly
difficult to maintain the color of a coating that has very low reflectance. If the variations of color
due to random thickness errors plotted in the chromaticity diagram then they tend to cluster
around the white point and the color can vary widely. Even if the cluster does not change in size,
a movement away from the white point will reduce the variation in color. This technique is
frequently practiced in the coating of ophthalmic lenses.
Macleod: Optical Coatings Page 334 Design through Manufacture

Figure 21-2. The four-layer antireflection coating with performance as shown on the left-hand side is
subjected to random errors in thickness of 2% standard deviation. the result is the variation in color shown in
the chromaticity diagram to the right. The color varies from red to blue.

Figure 21-3. The design of the coating of Figure 21-2 has been modified to give a more pronounced green
color. Now the random errors yield a smaller variation in color.
Macleod: Optical Coatings Page 335 Design through Manufacture

Ta2O5

2.5µm

SiO2

Plastic substrate

Figure 21-4. A new type of antireflection coating for plastic invented at the Fraunhofer Institute, Jena. It
consists of a very thick silica layer with thin layers of Ta2O5 interspersed. Because it consists essentially of a
very thick silica layer it is very tough and resistant to abrasion.
Dr Ulrike Schulz of the Fraunhofer Institute, Jena, has invented a new type of antireflection
coating that is largely a very thick layer of silica with inserted extremely thin layers of tantala
and has unprecedentedly high abrasion resistance[2]. The performance of a typical coating of
this new type is shown. Color, which is important for spectacle lenses, can be introduced by
varying the details of the design. The structure of the design can be understood as a series of
symmetrical periods based on silica/tantala/silica. These are used to construct an antireflection
coating of index that gradually reduces from the substrate to the incident medium. However,
instead of, as is normal, basing the design on quarterwave layers, it is based on thicknesses of
three-quarters of a wavelength. This allows the Herpin equivalent admittance to be reduced
below the admittance of silica. The three-quarter-wave approach does not have the same width
as a similar system based on quarterwaves but the width can be great enough to cover the visible
region.
Macleod: Optical Coatings Page 336 Design through Manufacture

The characteristic curve of the coating in Figure 21-4 is neutral in color but adjustment to the
design to give a particular color is straightforward.

21.5 Chapter 21 References


1. Schulz, Ulrike, Peter Munzert, and Norbert Kaiser, Surface modification of PMMA by
DC glow discharge and microwave plasma treatment for the improvement of coating
adhesion. Surface and Coatings Technology, 2001. 142-144: p. 507-511.
2. Schulz, Ulrike, Uwe B Schallenberg, and Norbert Kaiser, Novel antireflection coating
design for plastic optics. Applied Optics, 2002. 41(16): p. 3107-3110.
Macleod: Optical Coatings Page 337 Design through Manufacture

22 BIBLIOGRAPHY
A complete bibliography of primary references would stretch to an enormous length. This list is,
therefore, primarily one of secondary references wherever possible. Primary references are given
usually only where secondary references are difficult to obtain or do not exist.

Anders, H, Dünne Schichten für die Optik (English translation: Thin films in optics, Focal Press,
1967). 1965, Wissenschaftliche Verlagsgesellschaft mbH: Stuttgart.
Bach, Hans and Dieter Krause, eds. Thin Films on Glass. 1997, Springer-Verlag: Berlin,
Heidelberg.
Baumeister, Philip W, Optical Coating Technology. 2004, SPIE Press: Bellingham, Washington,
USA.
Dobrowolski, J A, Optical properties of films and coatings, in Handbook of Optics, M. Bass, et
al., Editors. 1995, McGraw-Hill, Inc: New York. p. 42.1-42.130.
Flory, François R, ed. Thin Films for Optical Systems. 1 ed. Optical Engineering, ed. B.J.
Thomson. Vol. 49. 1995, Marcel Dekker Inc: New York. 585.
Frey, Hartmut and Gerhard Kienel, eds. Dünnschicht Technologie. 1987, VDI-Verlag GmbH:
Düsseldorf.
Furman, Sh A and A V Tikhonravov, Basics of Optics of Multilayer Systems. First ed. Basics of,
ed. J.T.T. Van. 1992, Editions Frontières: Gif-sur-Yvette. 242.
Hartnagel, H L, A L Dawar, A K Jain, and C Jagadish, Semiconducting Transparent Thin Films.
1995, Institute of Physics Publishing: Bristol and Philadelphia. 358.
Heavens, O S, Optical Properties of Thin Solid Films. 1955, Butterworths Scientific
Publications: London.
Hodgkinson, Ian J and Qi hong Wu, Birefringent Thin Films and Polarizing Elements. First ed.
1997, World Scientific Publishing Co Pte Ltd: Singapore.
Holland, L, Vacuum Deposition of Thin Films. 1956, Chapman and Hall: London.
Hummel, Rolf E and Karl H Guenther, eds. Thin Films for Optical Coatings. Handbook of
Optical Properties. Vol. 1. 1995, CRC Press: Boca Raton, Florida. 361.
Jacobson, Michael Ray, ed. Deposition of optical coatings. 1 ed. SPIE Milestone Series, ed. B.J.
Thompson. Vol. MS 6. 1989, SPIE: Bellingham, Washington. 667.
Jacobson, Michael Ray, ed. Design of optical coatings. SPIE Milestone Series, ed. B.J.
Thompson. Vol. MS 26. 1990, SPIE: Bellingham, Washington. 700.
Jacobson, Michael Ray, ed. Characterization of optical coatings. 1 ed. SPIE Milestone Series,
ed. B.J. Thompson. Vol. MS 63. 1992, SPIE: Bellingham, Washington. 711.
Knittl, Z, Optics of Thin Films. 1976, John Wiley and Sons and SNTL: London, New York,
Sydney and Toronto, and Prague.
Macleod: Optical Coatings Page 338 Design through Manufacture

Liddell, Heather M, Computer-aided techniques for the design of multilayer filters. 1981, Adam
Hilger Ltd: Bristol. 194.
Lissberger, P H, Optical applications of dielectric thin films. Reports of Progress in Physics,
1970. 33: p. 197-268.
Macleod, H A, Thin-Film Optical Filters. Third ed. 2001, Institute of Physics Publishing: Bristol
and Philadelphia. 519.
Perilloux, Bruce E, Thin Film Design: Modulated Thickness and Other Stopband Design
Methods. Tutorial Texts in Optical Engineering, ed. J. Arthur R Weeks. Vol. TT57. 2002,
SPIE Optical Engineering Press: Bellingham. 134.
Pulker, Hans K, Coatings on Glass. Thin Films Science and Technology, ed. G. Siddal. 1984,
Elsevier: Amsterdam. 484.
Rancourt, James D, Optical Thin Films: Users' Handbook. Optical and Electro-Optical
Engineering, ed. W.J. Smith and R. Fischer. 1987, Macmillan Publishing Company: New
York. 289.
Thelen, Alfred, Design of Optical Interference Coatings. First ed. McGraw-Hill Optical and
Electro-Optical Engineering Series, ed. R.E. Fischer and W.J. Smith. 1988, McGraw-Hill
Book Company: New York. 255.
Vasicek, A, Optics of Thin Films. 1960, North Holland: Amsterdam.
Macleod: Optical Coatings Page 339 Design through Manufacture

23 CHARACTERISTICS OF THIN-FILM DIELECTRIC


MATERIALS
This list gives some details of the commoner thin film dielectric materials. It is not a definitive
list but is intended to show the wide range of available materials. The metals exhibit enormous
dispersion and so an abbreviated table of values is of little use. For extended tables of the optical
constants of metals consult [1-4]. Surveys of many thin film materials are given by Ritter[5, 6]
and by Palik[2-4]. For a fuller account of the fluorides of the rare earths consult Lingg[7].
In most cases the materials in the table can be deposited by many different processes. Where
thermal evaporation is possible it is the main process listed. Many of the materials with the
principal exception of the fluorides can be sputtered in their dielectric form by either radio
frequency sputtering or neutral ion-beam sputtering. A few materials, the nitrides especially, are
not capable of evaporation or reactive evaporation and require an energetic process such as ion-
assisted deposition.
The optical properties of thin films are very dependent on deposition conditions and other
factors. The values quoted should be interpreted simply as values that were reported at some
time and not as necessarily intrinsic and repeatable properties of the materials.
Macleod: Optical Coatings Page 341 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
1.62 at 0.6µm Can also be produced
Ts=300°C
Aluminium oxide 1.59 at 1.6µm by anodic oxidation
E-beam [9]
(Al2O3) 1.62 at 0.6µm of Al in ammonium
Ts=40°C tartrate solution [8]
1.59 at 1.6µm
E-beam evaporation Index varies
Aluminium
of Al with nitrogen 1.71-1.93 at 350nm continuously as [10, 11]
oxynitride <300nm-6.5µm
ion assist and oxygen 1.65-1.83 at 550nm function of
(AlOxNy)
background composition.
Important to avoid
Antimony trioxide 2.20 at 366nm overheating [12]
Molybdenum boat 300nm-1µm
(Sb2O3) 2.04 at 546nm otherwise
decomposes
Antimony sulphide Brief note [13] p 189 [14, 15]
3.0 at 589nm 500nm-10µm
(Sb2S3)
Tantalum boat.
Beryllium oxide Reactive evaporation 1.82 at 193nm
190nm-infrared Highly toxic [16]
(BeO) of Be metal in 1.72 at 550nm
activated oxygen.
E-beam[17]. Also 2.7 at 600nm Good infrared
(E-beam)
Bismuth oxide reactive sputtering of 2.2 at 9µm material but less
<550nm-12µm [17, 18]
(Bi2O3) bismuth in abrasion resistant
oxygen[18] 2.45 at 550nm (Sputter) than other oxides[17]
Bismuth trifluoride Graphite Knudsen 1.74 at 1µm
260nm-20µm [19]
(BiF3) cell 1.65 at 10µm
Cadmium sulphide Quartz crucible with 2.6 at 600nm Avoid overheating.
600nm-7µm [14, 20]
(CdS) spiral filament in 2.27 at 7µm Filament temperature
Macleod: Optical Coatings Page 342 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
contact with charge must be ≤1025°C
Cadmium telluride
Molybdenum boat 3.05 in near IR [21](brief)
(CdTe)

1.23-1.26 at
(porous)
546nm
Molybdenum or
Calcium fluoride [14, 17, 21,
tantalum boat. E- 150nm-12µm
(CaF2) 22]
beam[17]
1.40 at 600nm
(E-beam)
1.32 at 9µm

2.2 at 550nm Tends to form


Ceric oxide 2.18 at 550nm. Ts=50°C inhomogeneous
Tungsten boat 400nm-16µm [23-26]
(CeO2) 2.42 at 550nm. Ts=350°C layers. Suffers from
2.2 in near IR moisture adsorption.
1.63 at 550nm Hot substrate. Crazes
Cerous fluoride Tungsten boat. E- [17, 21, 23,
1.59 at 2µm 300nm-12µm on cold substrate[23].
(CeF3) beam[17] 24, 27]
1.57 at 9µm (E-beam) High tensile stress
Chiolite Howitzer or tantalum
Similar to cryolite [21]
(5NaF.3AlF3) boat
Chromium oxide 2.242 at 700nm
E-beam <600nm-8µm [17]
(Cr2O3) 2.1 at 8µm
Cryolite Howitzer or tantalum Slightly hygroscopic. [14, 21-23,
1.35 at 550nm <200nm-14µm
(Na3AlF6) boat Soft, easily damaged 28, 29]
Gadolinium fluoride
E-beam 1.55 at 400nm 140nm->12µm [7]
(GdF3)
Macleod: Optical Coatings Page 343 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
Absorption band
Germanium E-beam or graphite 4.25 in IR (usually slightly
1.7-100µm centred at approx [21, 23]
(Ge) boat higher than bulk value)
25µm
2.088 at 350nm
Hafnium dioxide [17, 24, 30,
E-beam 2.00 at 500nm 220nm-12µm
(HfO2) 31]
1.88 at 8µm
Hafnium fluoride 1.57 at 600nm
E-beam <600nm-12µm [17]
(HfF4) 1.46 at 10µm
1.59 at 550nm
Lanthanum fluoride Tungsten boat. E- [17, 23, 24,
1.57 at 2µm 200nm-12µm Heated substrate
(LaF3) beam[17] 27, 32, 33]
1.52 at 9µm (E-beam)
Lanthanum oxide 1.95 at 550nm Hot substrate
Tungsten boat 350nm->2µm [23, 24, 27]
(La2O3) 1.86 at 2µm (∼300°C)
Lead chloride Platinum or 2.3 at 550nm
300nm->14µm [21, 34]
(PbCl2) molybdenum boat 2.0 at 10µm
1.75 at 550nm
Lead fluoride Platinum boat. E- [17, 21, 23,
1.70 at 1µm 240nm->20µm
(PbF2) beam[35] 36]
1.3 at 10µm (E-beam)
Avoid overheating.
Lead telluride
Tantalum boat 5.5 in IR 3.4µm->30µm Hot substrate (see [37-39]
(PbTe)
text)
Lithium fluoride
Tantalum boat 1.36 to 1.37 at 546nm 110nm-7µm [14, 40]
(LiF)
Lutetium fluoride
E-beam 1.51 at 400nm 140nm-12µm [7]
(LuF3)
Macleod: Optical Coatings Page 344 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
Films on heated
[14, 21, 22,
Magnesium fluoride 1.38 at 550nm substrates much more
Tantalum boat 210nm-10µm 24, 30, 41-
(MgF2) 1.35 at 2µm rugged. High tensile
43]
stress.
Magnesium oxide 1.7 at 550nm. Ts=50°C
E-beam 210nm-8µm [44]
(MgO) 1.74 at 550nm. Ts=300°C
1.60 at 550nm
Neodymium fluoride Tungsten boat. E- [17, 23, 24,
1.58 at 2µm 220nm-12µm Hot substrate 300°C
(NdF3) beam[17] 27]
1.60 at 9µm (E-beam)
Hot substrate 300°C.
Neodymium oxide 2.0 at 550nm
Tungsten boat 400->2µm Decomposes at high [23, 27]
(Nd2O3) 1.95 at 2µm
boat temperature
Praseodymium oxide 1.92 at 500nm
Tungsten boat 400->2µm Hot substrate 300°C [27]
(Pr6O11) 1.83 at 2µm
Samarium fluoride
E-beam 1.56 at 400nm 160nm->12µm [7]
(SmF3)
Scandium oxide
E-beam 1.86 at 550nm 350nm-13µm [45]
(Sc2O3)
E-beam with water-
Silicon
cooled hearth. 3.5 in IR 1.1-14µm [23]
(Si)
Sputtering.
[21](brief)
Silicon monoxide Tantalum boat or 2.0 at 550nm Fast evaporation at
500nm-8µm [9, 14, 23,
(SiO) howitzer 1.7 at 6µm low pressure
30, 46]
Disilicon trioxide Tantalum boat or [9, 23, 47-
1.52-1.55 at 550nm 300nm-8µm
(Si2O3) howitzer 52]
Macleod: Optical Coatings Page 345 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
Silicon dioxide E-beam. Mixture in 1.46 at 500nm <200nm-8µm [9, 23, 53,
(Si)2) tungsten boat. 1.445 at 1.6µm (in thin films) 54]
Silicon nitride Low voltage reactive
2.06 at 500nm 320nm-7µm [55]
(Si3N4) ion plating
Sodium fluoride
Tantalum boat 1.34 in visible <250nm-14µm [14](brief)
(NaF)
Strontium fluoride 1.46 at 600nm
E-beam <600nm->12µm [17]
(SrF2) 1.3 at 10µm
Tantalum pentoxide 2.16 at 550nm
E-beam 300nm-10µm [17, 24]
(Ta2O5) 1.95 at 8µm
Tellurium [21, 23, 56,
Tantalum boat 4.9 at 6µm 3.4µm-20µm
(Te) 57]
Reactive evaporation
Can also be produced [14, 23, 47,
Titanium dioxide of TiO, Ti2O3 or 2.2-2.7 at 550nm depending
350nm-12µm by subsequent 52, 53, 58-
(TiO2) Ti3O5 in O2. E-beam on structure
oxidation of Ti film 65]
reactive evaporation.
Thallous chloride
Tantalum boat 2.6 at 12µm Visible->20µm [21, 66]
(TlCl)
Thorium oxide 1.8 at 550nm [21, 23, 67-
E-beam 250nm-15µm Radioactive
(ThO2) 1.75 at 2µm 69]
Radioactive.
Note: Thorium
Thorium fluoride 1.52 at 400nm [21, 23, 67-
Tantalum boat 200nm->15µm oxyfluoride (ThOF2)
(ThF4) 1.51 at 750nm 70]
actually forms ThF4
when evaporated.
Macleod: Optical Coatings Page 346 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
Ytterbium fluoride 1.52 at 600nm
E-beam <600nm-12µm [17]
(YbF3) 1.48 at 10µm
Yttrium oxide 1.82 at 550nm [17, 24, 30,
E-beam 250nm-12µm
(Y2O3) 1.69 at 9µm 71]
Zinc selenide Platinum or tantalum
2.58 at 633nm 600nm->15µm [68]
(ZnSe) boat
[14, 21, 23,
Zinc sulphide Tantalum boat or 2.35 at 550nm
380nm-~25µm 26, 29, 39,
(ZnS) howitzer 2.2 at 2.0µm
41, 69]
Zirconium dioxide 2.1 at 550nm
E-beam 340nm-12µm [17, 24, 46]
(ZrO2) 2.05 at 9.0µm
Substance H1† Tungsten boat or E- Does not melt
2.1 at 550nm 360nm-7µm [72, 73]
(Zirconia/titania) beam completely [72]
Substance H2†
Some weak
(Mixed
E-beam 2.1 at 550nm 400-7µm absorption bands in [72]
Praseodymium and
visible [72]
titanium oxides)
Substance H4†
E-beam with
(Lanthanum and 2.1 at 500nm. Ts=300°C 360nm-7µm [72]
molybdenum liner
titanium oxide)


Substance H1, Substance H2, Substance H4 and Substance M1 are members of the Patinal® series of optical coating materials manufactured by E Merck,
Darmstadt, Germany.
Macleod: Optical Coatings Page 347 Design through Manufacture

Region of
Materials Deposition technique Refractive index Remarks References
transparency
Substance M1†
(Mixed
E-beam 1.71 at 500nm. Ts=300°C 300nm-9µm [72]
Praseodymium and
aluminium oxides)
Macleod: Optical Coatings Page 348 Design through Manufacture

23.1 References
1. Hass, G and L Hadley, Optical constants of metals, in American Institute of Physics
Handbook, D.E. Gray, Editor. 1972, McGraw Hill Book Company: New York and London. p.
6.124-6.156.
2. Palik, E D, ed. Handbook of Optical Constants of Solids. 1985, Academic Press Inc.
3. Palik, Edward D, Handbook of Optical Constants of Solids II. 1991, San Diego and
London: Academic Press.
4. Palik, Edward D, Handbook of Optical Constants of Solids III. 1998, San Diego and
London: Academic Press.
5. Ritter, E, Dielectric film materials for optical applications, in Physics of Thin films, G.
Hass, M.H. Francombe, and R.W. Hoffman, Editors. 1975, Academic Press: New York,
San Fransisco and London. p. 1-49.
6. Ritter, E, Optical film materials and their applications. Applied Optics, 1976. 15(10): p.
2318-2327.
7. Lingg, Linda Jeanne, Lanthanide trifluoride thin films: structure, composition and
optical properties. 1990. PhD Dissertation, University of Arizona.
8. Hass, G, On the preparation of hard oxide films with precisely controlled thickness on
evaporated aluminum mirrors. Journal of the Optical Society of America, 1949. 39: p.
532-540.
9. Cox, J T, G Hass, and J B Ramsay, Improved dielectric films for multilayer coatings and
mirror protection. Journal de Physique, 1964. 25: p. 250-254.
10. Hwangbo, C K, Linda J Lingg, J P Lehan, H A Macleod, and F Suits, Reactive ion-
assisted deposition of aluminum oxynitride thin films. Applied Optics, 1989. 28: p. 2779-
2784.
11. Targove, J D, L J Lingg, J P Lehan, C K Hwangbo, H A Macleod, J A Leavitt, and L C
McIntyre, Jr. Preparation of aluminum nitride and oxynitride thin films by ion-assisted
deposition. in Materials Modification and Growth using Ion Beams Symposium. 1987.
Anaheim, CA, USA: Materials Research Society, Pittsburgh, PA, USA. p. 311-316.
12. Jenkins, F A, Extension du domaine spectral de pouvoir réflecteur élevé des couches
multiples diélectriques. Journal de Physique et le Radium, 1958. 19: p. 301-306.
13. Heavens, O S, J Ring, and S D Smith, Interference filters for the infra-red.
Spectrochimica Acta, 1957. 10: p. 179-194.
14. Heavens, O S, Optical properties of thin films. Reports on Progress in Physics, 1960. 23:
p. 1-65.
15. Billings, S H and M Hyman Jr, The infra-red refractive index and dispersion of
evaporated stibnite thin films. Journal of the Optical Society of America, 1947. 37: p.
119-121.
16. Ebert, J, Activated reactive evaporation. Proceedings of the Society of Photo-Optical
Instrumentation Engineers, 1982. 325: p. 29-38.
Macleod: Optical Coatings Page 349 Design through Manufacture

17. Kruschwitz, Jennifer D Traylor and Walter T Pawlewicz, Optical and durability
properties of infrared transmitting thin films. Applied Optics, 1997. 36(10): p. 2157-
2159.
18. Holland, L and G Siddall, Heat-reflecting windows using gold and bismuth oxide films.
British Journal of Applied Physics, 1958. 9: p. 359-361.
19. Moravec, T J, R A Skogman, and E Bernal G, Optical properties of bismuth trifluoride
thin films. Applied Optics, 1979. 18(1): p. 105-110.
20. Hall, J F and W F C Ferguson, Optical properties of cadmium sulphide and zinc sulphide
from 0.6 micron to 14 micron. Journal of the Optical Society of America, 1955. 45: p.
714-718.
21. Ennos, A E, Stresses developed in optical film coatings. Applied Optics, 1966. 5: p. 51-
61.
22. Heavens, O S and S D Smith, Dielectric thin films. Journal of the Optical Society of
America, 1957. 47: p. 469-472.
23. Ritter, E, Gesichtspunkte bei der Stoffauswahl für dünne Schichten in der Optik.
Zeitschrift für Angewandte Mathematische Physik, 1961. 12: p. 275-276.
24. Smith, D and P W Baumeister, Refractive index of some oxide and fluoride coating
materials. Applied Optics, 1979. 18: p. 111-115.
25. Hass, G, J B Ramsay, and R Thun, Optical properties and structure of cerium dioxide
films. Journal of the Optical Society of America, 1958. 48: p. 324-327.
26. Cox, J T and G Hass, Antireflection coatings for germanium and silicon in the infrared.
Journal of the Optical Society of America, 1958. 48: p. 677-680.
27. Hass, G, J B Ramsay, and R Thun, Optical properties of various evaporated rare earth
oxides and fluorides. Journal of the Optical Society of America, 1959. 49: p. 116-120.
28. Pelletier, E, P Roche, and B Vidal, Détermination automatique des constantes optiques et
de l'épaisseur de couches minces: application aux couches diélectriques. Nouvelle Revue
d'Optique, 1976. 7: p. 353-362.
29. Netterfield, R P, Refractive indices of zinc sulphide and cryolite in multilayer stacks.
Applied Optics, 1976. 15(8): p. 1969-1973.
30. Borgogno, J P, B Lazarides, and E Pelletier, Automatic determination of the optical
constants of inhomogeneous thin films. Applied Optics, 1982. 21: p. 4020-4029.
31. Baumeister, P W and O Arnon, Use of hafnium dioxide in mutilayer dielectric reflectors
for the near uv. Applied Optics, 1977. 16: p. 439-444.
32. Bourg, A, N Barbaroux, and M Bourg, Propriétés optiques et structure de couches
minces de fluorure de lanthane. Optica Acta, 1965: p. 151-160.
33. Targove, J D, J P Lehan, L J Lingg, H A Macleod, J A Leavitt, and L C McIntyre, Ion-
assisted deposition of lanthanum fluoride thin films. Applied Optics, 1987. 26(17): p.
3733-3737.
Macleod: Optical Coatings Page 350 Design through Manufacture

34. Penselin, S and A Steudel, Fabry-Perot-Interferometerverspiegelungen aus


dielektrischen Vielfachschichten. Zeitschrift für Physik, 1955. 142: p. 21-41.

35. Lès, Z, F Lès, and L Gabla, Semitransarent metallic-dielectric mirrors with low
absorption coefficient in the ultra-violet region of the spectrum (3200 - 2400A). Acta
Physica Polonica, 1963. 23: p. 211-214.
36. Carl-Zeiss-Stiftung. Interference filters. 1965. UK. Patent 994,638.
37. Smith, S D and J S Seeley, Multilayer filters for the region 0.8 to 100 microns. 1968, Air
Force Cambridge Research Laboratories.
38. Yen, Yi-Hsun, Ling-Xin Zhu, Wen-Di Zhang, Feng-Shan Zhang, and Shou-Yin Wang,
Study of PbTe optical coatings. Applied Optics, 1984. 23: p. 3597-3601.
39. Ritchie, F S, Multilayer filters for the infrared region 10-100 microns. 1970. PhD Thesis,
University of Reading.
40. Schulz, L G, The structure and growth of evaporation LiF and NaCl films on amorphous
substrates. Journal of Chemical Physics, 1949. 17: p. 1153-1162.
41. Hall, J F, Jr and W F C Ferguson, Dispersion of zinc sulfide and magnesium fluoride
films in the visible spectrum. Journal of the Optical Society of America, 1955. 45: p. 74-
75.
42. Wood, O R, II, H G Craighead, J E Sweeney, and P J Maloney, Vacuum ultraviolet loss
in magnesium fluoride films. Applied Optics, 1984. 23: p. 3644-3649.
43. Hall, J F, Optical properties of magnesium fluoride films in the ultraviolet. Journal of the
Optical Society of America, 1957. 47: p. 662-665.
44. Pulker, H K, Characterization of optical thin films. Applied Optics, 1979. 18(12): p.
1969-1977.
45. Arndt, D P, R M A Azzam, J M Bennett, J P Borgogno, C K Carniglia, W E Case, J A
Dobrowolski, D P Arndt, U J Gibson, T Tuttle Hart, F C Ho, V A Hodgkin, W P Klapp,
H A Macleod, E Pelletier, M K Purvis, D M Quinn, D H Strome, R Swenson, P A
Temple, et al., Multiple determination of the optical constants of thin-film coating
materials. Applied Optics, 1984. 23: p. 3571-3596.
46. Hass, G and C D Salzberg, Optical properties of silicon monoxide in the wavelength
region from 0.24 to 14.0 microns. Journal of the Optical Society of America, 1954. 44: p.
181-187.
47. Improvements in or relating to the manufacture of thin light transmitting layers. 1957.
UK. Patent 775,002.
48. Ritter, E, Zür Kentnis des SiO und Si2O3-Phase in dünnen Schichten. Optica Acta, 1962.
9: p. 197-202.
49. Okamoto, E and Y Hishinuma, Properties of evaporated thin films of Si2O3. Trans 3rd
Int Vac Congress, 1965. 2(2): p. 49-56.
50. Bradford, A P, G Hass, M McFarland, and E Ritter, Effect of ultraviolet irradiation on
the optical properties of silicon oxide films. Applied Optics, 1965. 4: p. 971-976.
Macleod: Optical Coatings Page 351 Design through Manufacture

51. Bradford, A P and G Hass, Increasing the far-ultra-violet reflectance of silicon oxide
protected aluminium mirrors by ultraviolet irradiation. Journal of the Optical Society of
America, 1963. 53: p. 1096-1100.

52. Auwärter, Max. Process for the manufacture of thin films. 1960. USA. Patent 2,920,002.

53. Reichelt, W, Fortschritte in der Herstellung von Oxydschichten für optische und
elektrische Zwecke. Trans 3rd Int Vac Congress, 1965. 2(2): p. 25-29.
54. Libbey-Owens-Ford Glass Company. Method of coating with quartz by thermal
evaporation. 1947. UK. Patent 632,442.
55. Bovard, Bertrand B, Juergen Ramm, Ralf Hora, and Fritz Hanselmann, Silicon nitride
thin films by low voltage reactive ion plating: optical properties and composition.
Applied Optics, 1989. 28(220): p. 4436-4441.
56. Moss, T S, Optical properties of tellurium in the infra-red. Proceedings of the Physical
Society, 1952. 65: p. 62-66.
57. Greenler, R G, Interferometry in the infrared. Journal of the Optical Society of America,
1955. 45: p. 788-791.
58. Hass, G, Preparation, properties and optical applications of thin films of titanium
dioxide. Vacuum, 1952. 2: p. 331-345.
59. Brinsmaid, Doris S, William J Keenan, George J Koch, and William F Parsons. Eastman
Kodak Co. Method of producing titanium dioxide coatings. 1957. USA. Patent 2,784,115.
60. Balzers Patent und Lizenz Anstalt. Improvements in and relating to the oxidation and/or
transparency of thin partly oxidic layers. 1962. UK. Patent 895,879.
61. Pulker, H K, G Paesold, and E Ritter, Refractive indices of TiO2 films produced by
reactive evaporation of various titanium-oxide phases. Applied Optics, 1976. 15: p.
2986-2991.
62. Heitmann, W, Reactive evaporation in ionized gases. Applied Optics, 1971. 10: p. 2414-
2418.
63. Heitmann, W, Properties of evaporated SiO2, SiOxNy, and TiO2, films. Applied Optics,
1971. 10: p. 2685-2689.
64. Heitmann, W, Reaktives Aufdampfen in ionisierten Gasen. Vakuum-Technik, 1972. 21:
p. 1-11.
65. Chiao, Shu-Chung, Bertrand G Bovard, and H A Macleod, Repeatability of the
composition of titanium oxide films produced by evaporation of Ti2O3. Applied Optics,
1998. 37(22): p. 5284-5290.
66. Perkin_Elmer Corporation. Infrared filters. 1961. UK. Patent 970,071.
67. Heitmann, W and E Ritter, Production and properties of vacuum evaporated films of
thorium fluoride. Applied Optics, 1968. 7: p. 307-309.
68. Heitmann, W, Extrem hochreflektierende dielektrische Spiegelschichten mit Zincselenid.
Zeitschrift für Angewandte Physik, 1966. 21: p. 503-508.
Macleod: Optical Coatings Page 352 Design through Manufacture

69. Behrndt, K H and D W Doughty, Fabrication of multilayer dielectric films. Journal of


Vacuum Science and Technology, 1966. 3: p. 264-272.
70. Ledger, A M and R C Bastien, Intrinsic and thermal stress modeling for thin-film
multilayers. 1977, The Perkin Elmer Corporation, Norwalk, Connecticut 06856.

71. Lubezky, I, E Ceren, and Z Klein, Silver mirrors protected with Yttria for the 0.5 to
14µm region. Applied Optics, 1980. 19: p. 1895.

72. Fritz, M, F Koenig, E Merck, and S Feiman, New materials for production of optical
coatings, in 35th Annual Technical Conference Proceedings. 1992, Society of Vacuum
Coaters: Albuquerque, New Mexico. p. 143-147.
73. Stetter, F, R Esselborn, N Harder, M Friz, and P Tolles, New materials for optical thin
films. Applied Optics, 1976. 15(10): p. 2315-2317.
Macleod: Optical Coatings Page 353 Design through Manufacture

24 A QUICK TOUR OF THE PROGRAM


Double click on the Essential Macleod icon and the program will begin to load. The initial
screen is quite blank with four menus, File, Tools, Options, Help.

Select the File menu. It has several sections but the first carries the four principal items, New,
Open..., Open Material..., Open Substrate... and Open Function…. Select New .
The symbol, , at the end of the menu item shows that there is a submenu. The number of
items depends on the configuration but there will be several, the first four being Design, Material,
Optical Constant and Table.
Select Design to open the Design window with a new design and a default design will appear.
The idea of a default rather than a blank design is that all necessary parameters are initialized so
that calculations can proceed without problems. The details of the default can be set by the user
as described later in the manual but, if this is still the first time the package has been used the
default will be a simple single layer coating.
Right at the top of the design window is the heading Design1. This is a default name that has
been given to this design and, if not changed, will be the name of the design file in which the
design will be stored. Next, immediately below the title, is the incident angle. This is the angle,
measured in degrees, between the direction of the incident light and the normal to the coating
surface. Here we have zero entered and so we are at normal incidence.
Macleod: Optical Coatings Page 354 Design through Manufacture

The reference wavelength, λ0, is shown next. This is 510nm. In the initial configuration as the
program is normally supplied, all wavelengths are given in nm. We will discuss wavelength units
in detail later but for the moment we retain nm. The actual design is shown next along with the
incident medium and substrate. The convention used throughout the package is that the incident
medium is uppermost in design tables. The substrate is the emergent medium and is at the foot.
The layers are numbered in turn from incident medium to substrate so that the layer nearest to
the incident medium is numbered one and is listed at the top of the table immediately after the
incident medium. The other layers, if any, follow in order through to the substrate. The columns
are labeled in turn, Layer, Material, Refractive Index, Extinction Coefficient, Optical Thickness.
Layer lists the reference number of the layer or indicates the incident medium or substrate. The
next three columns give information on the various materials, name and optical constants. Since
we are dealing with dispersive materials the values of refractive index and extinction coefficient
are those at the reference wavelength, λ0, (which is 510nm currently). The optical thickness is
next. This is nd/λ0 where n is refractive index, d is physical thickness and λ0 is the reference
wavelength. 0.25, therefore indicates a quarterwave layer at the reference wavelength. [There are
other options for the display of layer thickness and some other layer attributes (packing density,
locked and linked) that are invisible for the moment, all of which we will consider later.] This
design is therefore a quarterwave at 510nm of Na3AlF6 (cryolite) on glass in air and can be
described as an antireflection coating. Let us examine the performance. Select Plot from the
Performance menu. A new window entitled Design1: Transmittance appears. The transmittance
of the coating is shown plotted against wavelength from 400nm to 700nm.
Macleod: Optical Coatings Page 355 Design through Manufacture

The transmittance is everywhere very high. Since this is an antireflection coating, reflectance
may be a more suitable parameter to plot.
The active window is the plot, however, and so the current menu bar refers to it. We must
therefore make the design window active so that the menu bar displays the design menu. To do
this place the mouse cursor over any visible part of the design window and click, or use the
Windows menu. The design window should come to the front showing that it is active.
Before we change the plot, however, let us be sure that two plots may coexist. Select the
Options menu and then General...The Essential Macleod Options dialog box will appear. In the
Windows tab there is a check box labeled Keep Old Plots and Tables Displayed. Check this box if
it is not already checked. Later you may wish to uncheck this box because displayed plots use up
resources.

Now select the Parameters menu and the item Performance... The following dialog window
opens. Note that some parameters apply only to certain choices of horizontal or vertical axis.
Those that do not apply are shown in gray.
Macleod: Optical Coatings Page 356 Design through Manufacture

We want reflectance so we click on the Vertical Axis tab and then set the performance to
Reflectance Magnitude (%). The vertical scale is listed as Maximum Value 100, Minimum Value 0
and Interval for Plot 20. Automatic Scale is checked and so the plot will be re-scaled to fit the
data. Leave all other parameters unchanged and click on the Plot button.

You can activate a cursor on the plot by holding down the <Alt> key and dragging the mouse
over the plot with the left mouse button also pressed. A pair of markers will follow the curve
nearest the mouse pointer and the (x,y) coordinates of the curve are displayed in a box beneath
the plot.
Macleod: Optical Coatings Page 357 Design through Manufacture

The cursor is removed by selecting Clear Cursor in the Edit menu, or by right-clicking the
mouse and selecting Clear Cursor from the menu that appears.
It is a good idea at this stage to remove any unwanted plot that has been generated so far.
Click in a plot to make it active and then click on the close box at the upper right corner. Do not
accept the invitation to save the plot.
Now place the mouse over the performance curve and double-click. A new dialog box
appears. This controls the display of the curve.

To change the thickness of the curve, enter a number into the Width box and click on Close.
This produces the following plot.
Macleod: Optical Coatings Page 358 Design through Manufacture

It is important to realize, however, that editing the plot changes only the parameters of the
plot. The data are unaffected. Thus the data cannot be changed from reflectance to transmittance
just by altering the title of the vertical axis. Reflectance data remain reflectance whatever title is
given to the axis.
We will now save the plot. With the plot as the active window, select the File menu and then
Save As... The Save Plot As dialog box appears. Note that there is a shortcut key, F12, that
moves straight to the dialog box. Against File name type in one. Note that the extension for plot
files will be .npl but that it is not necessary to add any extension. That will be done automatically
by the package. Then select Save. The plot will be saved as file one.npl.
Choose Performance... from the Parameters menu. In the dialog box that appears select the
Vertical Axis tab and then clear the Automatic Scale check box. Choose Transmittance Magnitude
(%) with Maximum Value 100, Minimum Value 95. Set the Interval for Plot to blank. Clear the
Automatic Scale check box. Click on the Plot button. The result is shown below. If the plot
interval is set to blank then the program automatically calculates a reasonable interval. If a
number is inserted in the field then that number will be used regardless of its suitability.
Macleod: Optical Coatings Page 359 Design through Manufacture

Now examine the Window menu. It lists the various windows that have been created. Select
each in turn and see how they come to the front. This is another way of making a window active
and it is particularly useful when a window is completely obscured behind others.
Finally from the Performance menu select Table. The results are displayed in tabular form.

Note that above the heading of the table are entries giving reference wavelength and incident
angle. These entries are for reference only. It is possible to change these entries (the read-only
status of the table must first be canceled) but such changes will have no effect whatsoever on the
actual values of the parameters, neither will they affect the values of the data in the table.
The table lists values of reflectance, transmittance and phase changes on transmission and
reflection at intervals that are determined by Interval for Table under Horizontal Axis in the
Performance Parameters dialog box. The table window may have to be adjusted by stretching
with the mouse in the normal way.
Macleod: Optical Coatings Page 360 Design through Manufacture

This current coating is not a particularly good one because the residual reflectance is too high
for many applications. Also, although it has nothing to do with its optical properties, cryolite is
not a rugged material. We will now examine a coating that is rather broader in its characteristic,
and uses more resistant materials, but which is, at the same time, more complicated. This is the
well-known quarter-half-quarter coating and the version we are going to examine consists of a
quarterwave of magnesium fluoride outermost followed by a halfwave of zirconia, and then next
to the substrate a quarterwave of alumina. The pattern of this coating is a low index quarterwave
next to the incident medium and an intermediate index quarterwave next to the substrate which
form an antireflection coating for the reference wavelength. These layers are separated by a
flattening halfwave of high index, which gives added breadth to the characteristic. The names for
the materials in the materials list are mostly based on the chemical formulae and the materials we
are suggesting have names MgF2, ZrO2 and Al2O3 respectively.
Save the design in a file named one. The name at the top of the design window will change to
one. We want to keep this design. It is, however, still the active one. Any changes we make now
will risk being incorporated in one.dds should an automatic save be initiated. To avoid this we
immediately save the design yet again as two.dds. Press F12 and enter TWO (or two) in the File
Name box and then select Save.
This is really good practice that is worth cultivating. It is all too easy to lose a design by
overwriting it. It is usually impossible to recover a design once it has been overwritten. This
procedure is particularly useful also when many of the parameters remain unchanged. When you
want to start a completely new design, use the New command in the File menu. The new design
will then coexist with the previous one, which will not be automatically closed. Being able to
edit several designs at the same time is a valuable feature of the Multiple Document Interface
(MDI) of the program.
We will now edit the design two.dds.
There is only one layer in the current design and we need three. The first task is to increase
the number of layers to three. Place the cursor over a cell in the row corresponding to the cryolite
layer (Na3AlF6, Layer 1) and click. Now go to the Edit menu and select Insert Layers... In the
resulting dialog box enter 2 as the number of layers to be inserted and then select the OK button.
Now the design will show three similar layers of Na3AlF6. The material for layer 1, that next to
the incident medium should be magnesium fluoride, MgF2 in the materials list. Place the cursor
over the cell in layer 1 containing Na3AlF6 and click. The cell will be selected and a small arrow
will appear at the right. A click on the arrow brings up a small scroll box with the names of the
existing materials in it. MgF2 is a little way down the list. Click on it and immediately the name
of the material for layer 1 will change to MgF2 and the scroll box will close. The insertion of the
extra layers resulted in a thickness of zero for layer 1. This must be changed to the required 0.25.
Place the cursor over the thickness cell for layer 1. Click once to select it. Then type 0.25. Note
that as soon as we selected the thickness cell, deselecting the material, the value of refractive
index changed to match the index of magnesium fluoride at the reference wavelength.
Incidentally, a double click in the thickness cell will place it in insert mode and then anything
typed will be added to what is already there. If that happens then use the Delete key to remove
the unwanted characters. Now we pass to layer 2. Here we need a halfwave of zirconia, ZrO2.
The same process as before permits us to replace the Na3AlF6 material with ZrO2. The zero
thickness must then be changed to 0.5. Finally we change the material for layer 3 to Al2O3. This
Macleod: Optical Coatings Page 361 Design through Manufacture

time, instead of selecting the layers from the list, select the materials cell and type in Al2O3.
Now select any other cell. The entered material will change to Al2O3 denoting that the new
material has been accepted by the program and the Refractive Index cell will display the
appropriate value at the reference wavelength. Materials can be entered either manually or from
the table. Sometimes manual entry can be quicker than looking up the table each time. The
thickness of layer 3 is already 0.25 (if the insertion was performed exactly as described here) and
so it needs no alteration. At this stage the design window should appear as reproduced below:

We added the layers by using the insert layers command. There are several other ways of
performing this task and a particularly useful one is available when the Insert mode is visible in
the bar at the foot of the Essential Macleod window. (If Overstrike is shown then press the
<Insert> key and the mode will change to Insert.) Place the cursor in a cell on one of the layers.
Press <Enter> repeatedly, each time making the cell to the right of the current one active. When
the thickness cell is active, pressing <Enter> will create a new layer. A layer that is unwanted can
be deleted very quickly by clicking in the little square selection box at the extreme left of the
layer and then pressing <Delete>. This method of adding and deleting layers is very useful when
a number of layers is being entered by hand.
Now we plot the performance of this new design. In the Parameters dialog, set the vertical
axis to be Reflectance Magnitude and check the Automatic Scale box. Now click on Plot.
Macleod: Optical Coatings Page 362 Design through Manufacture

Save this plot as plot file two using the same procedure as before, that is selecting the plot by
making the plot window active and then using the Save As... command in the File menu, or,
pressing <F12>.
It would be useful to compare these two plots in the same diagram. To illustrate how this can
be done let us close all the plot and design windows. Now either go to the File menu and select
Open..., or press <Ctrl><F12>. This brings up a dialog box with a list of files to open. A list of
design files is shown but we want the plot files. Click the arrow to the right of the Files of Type:
window and select Plots. The list of designs will disappear to be replaced by the list of plots that
we have saved so far. Select one.npl. This is the plot corresponding to the single layer. The line
thickness was made perhaps rather too large . Let us change that to a fine dotted line. With the
mouse over the curve, double-click the left mouse button. In the dialog box that appears change
the Pattern to be Dotted and then click on Close. This changes the appearance of the curve. Now
we want to add the three layer antireflection coating curve to the same diagram. Return to the
File menu and select Add Line... This time there is no equivalent single key stroke but you can
use <Alt><F> followed by <d>. The Add dialog box appears and this time the file two.npl can be
selected. Both curves are now on the same diagram as shown below. This plot may in turn be
stored for later recall or for adding further curves.
Macleod: Optical Coatings Page 363 Design through Manufacture

It would be interesting to see how the three-layer coating operates at oblique incidence. Open
the design named two. Let us choose an angle of 45°. This can readily be entered in the cell
labeled Incident Angle right at the top of the Design window (or in the Performance Parameters
dialog box). We decide we want to plot both s- and p-polarizations together with the mean. We
therefore select the Parameters menu and Performance... Down at the foot of the resulting dialog
box there is a set of polarization check boxes. Check P, S and Mean. Select Plot either by
clicking the plot button or selecting Plot from the Performance menu and the result is as shown
below.

In the normal configuration of the package, the curves will all be full lines but with different
colors, red for p-polarization, blue for s-polarization and green for mean. In this version of the
plot the s-polarization curve is the full line, the p-polarization the broken line with broad dashes
and the mean the broken line with short dashes. This is so that it can be printed in black and
white if necessary.
Macleod: Optical Coatings Page 364 Design through Manufacture

24.1 A Word about Data Files...


The program generates many different files containing data as it operates. You will want to
preserve various designs that have been entered. Significant plots and tables will be stored.
Specifications for refinement will be created. Text files with pasted-in design and performance
details for incorporation in reports will be saved. The amount of material that can accumulate in
one project is enormous and from time to time you will want to archive this data and start again.
If all of this is preserved in a dedicated folder, then archiving and clearing out is straightforward.
If, however, it is mixed in with program files and with other project files then there can be a real
problem during archiving. The program is arranged so that many different design folders can be
used. These must first be created in the usual way (Windows Explorer is simplest). Try to adopt
a convention in naming the folders so that they have a clear and obvious meaning. This will help
enormously in keeping things organized.

24.2 ...about Materials Databases


Many materials databases can be used. The only limitation is that only one database at a time
may be active. The databases are completely separate entities and can contain duplicate names
without any risk of ambiguity because the design files are linked to the materials database with
which they were created. Thus a separate database with subtle differences in material properties
could be maintained for each coating plant. Two databases, one labeled WET and the other DRY
could permit rapid calculation of the effect of moisture on different filters. Although the coating
design files are linked to their own materials databases, nevertheless it is very easy to change
from one database to another with the same material names. It is simply difficult to do it
accidentally. It is also possible to vary variable units from one database to another. Coatings for
the far infrared are better handled in microns that in nanometres, for example. A materials
database is a collection of material files that are held in a separate folder together with control
files. The collection can be as extensive or narrow as desired. The easiest way to create a
material database is to type the name including path in the Materials Folder window in the
Essential Macleod Options dialog box. Choose General... in the Options menu to activate the
dialog box. If the name that has been entered is a new one then the program will automatically
create the new database. Another way of creating a materials database is simply to copy an
existing one into a new folder. It is most important that the control files should be kept in the
same folder as the material files otherwise the material files cannot be used by the program.
Rather like the design file folders, the names of materials databases are best chosen carefully so
that they convey an idea of their content. IR, VIS and UV, for example convey immediately the
idea of spectral regions and differences in wavelength units. PlantA, PlantB or ClientA, ClientB
are also designations that convey an immediate meaning.

24.3 ...and about Conventions


There are conventions concerning designs and parameters that are dealt with in detail in the
next section but some are so important that we mention them also here. The term, Medium, refers
to the incident medium and Substrate to the emergent medium. Designs are normally written
with the flow of light from left to right or from top to bottom so that a formula representing a
design is always assumed to have the incident medium on the left and the substrate, or emergent
Macleod: Optical Coatings Page 365 Design through Manufacture

medium, on the right. The designs shown in the design window have incident medium at the top
of the table and emergent medium, or substrate, at the foot. However, not all practitioners adhere
to these conventions. They can therefore be changed. Note that in this course we will adhere to
the normal convention.
To change the convention select General in the Options menu and click on the Designs tab.
Display Order can be set to Substrate at Top instead of Medium at Top and Formula Order to
Medium at Right instead of Medium at Left. The numbering in the design document can also be
from top to bottom or the oposite.

Note that throughout the Macleod package there is just one exception to this arrangement and
that is in Runsheet where the order of the layers is necessarily that in which they are deposited,
that is with the layer next to the substrate always first.
Layer thicknesses may be specified in four principal ways. Optical thickness is given by
nd/λ0, geometric by d/λ0, and physical by d where λ0 is the reference wavelength, n, the real part
of refractive index, and d, the physical thickness of the film, that is the thickness that would be
measured by a conventional ruler. Then there is a fourth way headed QWOT that gives the
thickness of the films in quarterwaves at λ0. QWOT stands for QuarterWave Optical Thickness,
and is nd/(λ0/4), that is 4nd/λ0. To switch from FWOT to QWOT use the Optical Thickness
Convention field in the above dialog. To display the various conventions use Display Setup… in
the File menu.
Macleod: Optical Coatings Page 366 Design through Manufacture

The design document with all the thickness conventions checked looks like the following.
Note that the total thickness in the appropriate units is always given at the foot of the particular
column.

Note that a number of standard designs is available in the Designs folder. These have
extensive notes attached.
Macleod: Optical Coatings Page 367 Design through Manufacture

25 EXERCISES
These exercises are intended simply as suggestions because there is nothing worse
than sitting at the computer with nothing to do. Of course the best form of training is to
work on a real problem and if you have one you would like to spend some time on
please do so.
For many of the questions, especially the earlier ones, a nondispersive database of
materials should be used. This allows for easier interpretation and understanding of the
results. Real materials suffer from dispersion, however, and it should always be
included in a final design. Try changing databases. You may find it useful, for example,
to create a fresh database with slightly different indices for the materials to simulate the
changing of a process parameter, or, perhaps shifting from one plant or process to a
different one.
With real problems there is seldom a clear single answer. Rather there is a set of
compromises amongst performance, complexity, cost, time and so on. Many of the
following questions have no unique correct solution and the specifications are quite
loose. Try out different designs and see, for example, how increases in the number of
layers can alter the achievable performance. Use the designs, once they are created, in
computer-aided experiments.
A dispersionless database exists under the name “Non dispersive”. To make this the
current materials database first of all make sure that all document windows except the
main Essential Macleod window are closed. Then select General in the Options menu.
In the Materials Folder box use the drop down menu and look for the Materials Folder
called “Non dispersive.” Once the name is entered, click on OK and this materials
database will be activated. Provided the thicknesses are set to optical and the reference
wavelength to 1000nm then the frequency scale will be dimensionless in terms of g (the
ratio of the reference wavelength to the actual wavelength). The names of the materials
in the database are all simply the values of index/admittance that they possess. Not all
possible values are represented. To create a material with a fresh value use New in the
File menu. In the material window that will then appear enter the new index at
wavelengths 100 and 10000. Then choose Save As and enter the value of
index/admittance against the material name.
Macleod: Optical Coatings Page 368 Design through Manufacture

1 Assume that H, J, L and M are quarterwaves of TiO2 (2.4), HfO2 (2.0),


MgF2 (`1.38), and Al2O3 (1.65), respectively. The reference wavelength
is 510nm and the performance is to be calculated over the region 400nm
to 700nm. Substrates are SiO2 (1.45), glass (1.52), or flint glass (1.65).
Design and calculate the properties of the following antireflection
coatings.
(a) Single layer.
(b) Quarter-quarter.
(c) W-coat (LJJ).
(d) Quarter-half-quarter.

2 Calculate the characteristic curves of


Air | (HL)^q H | Glass
where H and L represent quarterwaves of TiO2 (n=2.4) and SiO2
(n=1.45). Glass is crown glass (n=1.52). The reference wavelength is
700nm and the wavelength range is 400nm to 1200nm. The angle of
incidence is to be both normal and 45 deg. Use q in the range 5 to 11.

3 Plot admittance diagrams at the reference wavelength for the examples


in question 1.

4 This exercise is to be attempted using dispersive materials. Calculate the


performance first at normal incidence and then at 45°.
(a) Calculate the reflectance of aluminum over the range 300nm to
1200nm. Note that the design must include a layer so use Air | Al
(0.01 thick) | Al
(b) Calculate the performance of Air | HLHL | Al with reference
wavelength 510nm where H is TiO2 and L is SiO2.
(c) Attempt to move this design to a longer wavelength to improve the
reflectance of aluminum in the near infrared.

5 Design a V-coat for silica at wavelength 510nm using material of


admittance 1.9 and 1.38. Use the thinner of the two solutions. Insert a
buffer layer in the appropriate place and examine the variation of the
properties over the range 400nm to 900nm as the buffer layer is varied in
thickness.
Macleod: Optical Coatings Page 369 Design through Manufacture

6 A coating consisting of Air | LLL | Glass with L a quarterwave of n=1.38,


and glass of n=1.52, has low R at the reference wavelength and at three
times the reference wavelength. A halfwave H layer (n=1.9) at three
times the reference wavelength will have no influence on the
performance at that wavelength but if it is next to the glass it can have a
major flattening effect. Examine the performance over the region 400nm
to 1800nm of Air | LLLHHHHHH | Glass with reference wavelength of
510nm. There will be some flattening also at 510nm which can be
increased somewhat by shifting the design to Air | LHHHHHHLL | Glass.
Plot this characteristic over the whole range. A further zero of R appears
at an intermediate wavelength. The admittance diagram explains this
zero. On this basis study also the design Air | LHHHHLL | Glass.

7 Design a V-coat for flint glass (n=1.65) using TiO2 and SiO2. Insert
halfwave flattening layers. There are two places where these may be
inserted. Try first one place and then the other. Try the effect of having
both halfwaves present together. In each case refine the performance
over the region 400nm to 700nm to have as low a reflectance as
possible.

8 In refinement of antireflection coatings (as with any other design) it is


important to avoid local minima if possible. The starting design can be
very important. A starting design for an antireflection coating for the
visible region consists of Air | LHHLLHHLL | Glass with MgF2 and TiO2,
say, as the two materials. This design is very near a local minimum and
it is very difficult to move away from it. Experiment with the various
refinement processes and see if the parameters can be adjusted so that
the local minimum may be avoided.

9 Repeat question 8 using as starting design the structure


Air | LHH.1L.1HLL | Glass. This should give much better results and
more quickly. Why?

10 The following design with H of 2.35 and L of 1.38 and glass of 1.52
Air | L(0.3087) H(0.1019) L(0.1413) H(0.0749) L(0.5698) | glass
with reference wavelength 510nm is a good antireflection coating for the
visible region. Use the admittance diagram to see how it works.

11 Redesigns the coating of question 10 to use materials of


index/admittance 1.45 and 2.4
Macleod: Optical Coatings Page 370 Design through Manufacture

12 With no restriction on the number of layers but with only two materials,
Ge and ZnS design an antireflection coating for Ge to cover the region 8
to 12 micron with a reflectance less than 1% if possible.

13 Use Air | LHLH| glass as starting design for two-material AR for 400nm
to 700nm. The substrate should be flint glass (n=1.65).

14 Insert a flattening layer in the beam splitter of design:


Air | LHLH | Glass
with L - 1.38, H - 2.30, such that flatter performance is achieved over the
visible region. Assume that the incidence is normal. Refine the
performance to see if still flatter performance can be achieved.

15 Design a high reflectance coating to have greater than 99% reflectance


over the region 400 to 700nm. Use ZnS and cryolite deposited on glass
with air as the incident medium.

16 Edge filters. Use SiO2 and TiO2 on a glass substrate in air to construct
a longwave pass edge filter with edge at 700nm and good transmission
over the region 730nm to 1000nm. [Start with (0.5HL0.5H)^q then try
various techniques.]

17 Design a longwave pass filter using Ge and ZnS on a germanium


substrate in air. The filter is to have a 50% edge at 8 micron and high
transmittance from 8.5 to 13 micron. Transmittance should be less than
0.05% at wavelengths less than 7.5 micron.

18 This illustrates halfwave holes. Use dispersionless materials for this. A


basic short-wave pass edge filter has design
Air | (0.5LH0.5L)^13 ).5L | glass with L, SiO2 and H, TiO2. The reference
wavelength is 1.1 micron. Set up this design and two others with
(0.4L1.2H0.4L) and (0.6L0.8H0.6L) as the basic periods. The latter two
designs will have halfwave holes at half the reference wavelength. Tilt all
three coatings at 45 deg and see how the halfwave holes vary.
Macleod: Optical Coatings Page 371 Design through Manufacture

19 Design a minus filter. This filter must have 50% points at 650nm and
850nm. Transmittance should be as high as possible from 400nm to
630nm and from 870nm to 1100nm. Transmittance should be very low
from 670nm to 825nm. The filter is to be constructed from TiO2 and
SiO2 on a glass substrate with air as the incident medium.

20 Design a longwave pass edge filter using Ge and SiO on Ge with air as
the incident medium. The edge is to be at 3 micron (50%) and there is to
be high transmittance from 3.2 micron to 5.0 micron with low
transmittance (<0.1%) at wavelengths shorter than 2.8 micron.

21 Design a multiple-cavity bandpass filter for the visible region using ZnS
and cryolite. The incident medium and the substrate are to be glass. The
filter is to have a peak wavelength of 600nm with halfwidth as near 20nm
as possible.

22 Design a narrowband filter for the infrared using germanium and SiO on
sapphire (Al2O3) in air. The 50% points are to be 4.2 and 4.3 micron and
the peak transmittance must be greater than 95%. Transmittance outside
the pass band is to be generally low, but less than 0.01% at 4.5 micron.

23 Design a bandpass filter for the far infrared region. The center
wavelength should be 10 micron and the halfwidth 1 micron. Materials
are ZnS, Ge and substrate Ge with air as the incident medium.

24 Investigate possible bandpass filter designs for a peak wavelength of


250nm using layer materials aluminum oxide and cryolite on a substrate
of silica with incident medium air.

25 Design an induced transmission filter for 600nm. The metal layer(s) are
to be silver which must not have a physical thickness less than 20nm.
The peak wavelength of the filter is to be greater than 87%. Use TiO2
and SiO2 as the layer materials and glass as both substrate and incident
medium.

26 What are the prospects for using aluminum as metal layer and cryolite
and aluminum oxide as layer materials on silica substrates to construct
induced transmission filters for the near UV (say 250nm peak
wavelength)?
Macleod: Optical Coatings Page 372 Design through Manufacture

27 Use dispersive optical constants for this. Design a dielectric-metal-


dielectric filter to have zero reflectance at 550nm using TiO2 and Ag.
Use the maximum thickness of silver and use glass as both incident and
emergent media. Plot the characteristic (T and R) over the region 300nm
to 1500nm. Now use two layers of Ag in a dielectric-metal-dielectric-
metal-dielectric design. Note that it is possible now to use thicker silver
layers than the maximum possible before, but there are some
disadvantages. Go back to the single silver layer and increase the
reflectance of the matching assemblies on either by adding a low
admittance quarterwave. This should allow a thicker silver layer than in
the first design.

28 Find the range of angles of incidence over which the quarter-half-quarter


antireflection coating, using Air|MgF2|TiO2|Al2O3|Glass, gives less than
1% R for both s- and p-polarization over the region 450nm to 650nm.

29 (a) Using TiO2 and SiO2 on SiO2, design an antireflection coating to


give zero R for s-polarization at 600nm. The angle of incidence is
to be 60 deg in air. Over what range of angles of incidence will Rs
be less than or equal to 0.1%?
(b) Investigate the same operation as (a) but for p-polarization.

30 Attempt the design of an antireflection coating for s-polarization at 60


deg angle of incidence. The substrate is to be SiO2 and the incident
medium air. TiO2, SiO2 and Al2O3 are the materials to be used and a
reflectance of less than 1% (Rs) over the region 450nm to 650nm should
be attempted.

31 Plot the p-polarization characteristic of the coating designed in question


30.

32 Design a beam splitter to have Rp=Rs=50% at 45 deg angle of incidence


in air and at a wavelength of 550nm. The substrate is to be glass and
any material with admittance between 1.35 and 2.4 can be assumed to
be available.
Macleod: Optical Coatings Page 373 Design through Manufacture

33 Design a plate polarizer for 1.06 micron. The substrate is SiO2 and the
materials are to be SiO2 and TiO2. The specification is Tp>98% and
Ts<0.1% for 1.06 micron. [Make sure that the angle of incidence is the
Brewster angle for SiO2 in air, to avoid any antireflection coating on the
rear surface of the component when eventually it is manufactured.]

34 Using TiO2 and SiO2 on glass and introducing one other material with
admittance to be determined in the design process, design a
nonpolarizing edge filter of the longwave pass type with an edge at
500nm to have minimum polarization splitting at 45 deg angle of
incidence in air.

35 Using only TiO2 and SiO2, use the edge filter option to design a
nonpolarizing edge filter with the characteristics in question 34. Compare
these two filters.

36 Design a shortwave pass filter with edge at 700nm with other details as
in question 35

37 Design (a) a shortwave pass filter and (b) a longwave pass filter using
ZnS and Ge on Ge to have a 6 micron edge and minimum polarization
splitting at 45 deg incidence in air. Optimize the designs to reduce ripple
if necessary.

You might also like