Sei sulla pagina 1di 169

Total internal

reflection
Fig. 1: Underwater plants in an aquarium, and their
inverted images (top) formed by total internal
reflection in the water-air surface.

Total Internal Reflection (TIR) is the


phenomenon that makes the water-to-air
surface in a fish-tank look like a perfectly
silvered mirror when viewed from below
the water level (Fig. 1). Technically, TIR is
the total reflection of a wave incident at a
sufficiently oblique angle on the interface
between two media, of which the second
("external") medium is transparent to such
waves but has a higher wave velocity than
the first ("internal") medium. TIR occurs
not only with electromagnetic waves such
as light waves and microwaves, but also
with other types of waves, including sound
and water waves. In the case of a narrow
train of waves, such as a laser beam, we
tend to speak of the total internal
reflection of a "ray" (Fig. 2).
Fig. 2: Repeated total internal reflection of a
405 nm laser beam between the front and back
surfaces of a glass pane. The color of the laser light
itself is deep violet; but its wavelength is short enough
to cause fluorescence in the glass, which re-radiates
greenish light in all directions, rendering the zigzag
beam visible.

Refraction is generally accompanied by


partial reflection. When a wavetrain is
refracted from a medium of lower
propagation speed (higher
refractive index) to a medium of higher
propagation speed (lower refractive index),
the angle of refraction (between the
refracted ray and the normal to the
refracting interface) is greater than the
angle of incidence (between the incident
ray and the normal to the interface).
Hence, as the angle of incidence
approaches a certain limit, called the
critical angle, the angle of refraction
approaches 90°, at which the refracted ray
becomes tangential to the interface. As
the angle of incidence increases beyond
the critical angle, the conditions of
refraction can no longer be satisfied; so
we have no refracted ray, and the partial
reflection becomes total. In an isotropic
medium such as air, water, or glass, the ray
direction is simply the direction normal to
the wavefront.

If the internal and external media are


isotropic with refractive indices n1 and n2
respectively, the critical angle is given by
, and is defined if
n2 ≤ n1.  For example, for visible light, the
critical angle is about 49° for incidence
from water to air, and about 42° for
incidence from common glass to air.

Details of the mechanism of TIR give rise


to more subtle phenomena. While total
reflection, by definition, involves absolutely
no continuing transfer of power across the
interface, the external medium carries a
so-called evanescent wave, which travels
along the interface with an amplitude that
falls off exponentially with distance from
the interface. The "total" reflection is
indeed total if the external medium is
lossless (perfectly transparent),
continuous, and of infinite extent, but can
be conspicuously less than total if the
evanescent wave is absorbed by a lossy
external medium ("attenuated total
reflectance"), or diverted by the outer
boundary of the external medium or by
objects embedded in that medium
("frustrated" TIR). Unlike partial reflection
between transparent media, total internal
reflection is accompanied by a non-trivial
phase shift (not just zero or 180°) for each
component of polarization (normal or
parallel to the plane of incidence), and the
shifts vary with the angle of incidence. The
explanation of this effect by Augustin-Jean
Fresnel, in 1823, added to the evidence in
favor of the wave theory of light.

The phase shifts are utilized by Fresnel's


invention, the Fresnel rhomb, to modify
polarization. The efficiency of the
reflection is exploited by optical fibers
(used in telecommunications cables and
in image-forming fiberscopes), and by
reflective prisms, such as erecting prisms
for binoculars.

Optical description

Fig. 3: Total internal reflection of light in a semicircular


acrylic block.

Although total internal reflection can occur


with any kind of wave that can be said to
have oblique incidence, including (e.g.)
microwaves[1] and sound waves,[2]  it is
most familiar in the case of light waves.

Total internal reflection of light can be


demonstrated using a semicircular-
cylindrical block of common glass or
acrylic glass. In Fig. 3, a "ray box" projects
a narrow beam of light (a "ray") radially
inward. The semicircular cross-section of
the glass allows the incoming ray to
remain perpendicular to the curved portion
of the air/glass surface, and thence to
continue in a straight line towards the flat
part of the surface, although its angle with
the flat part varies.
Where the ray meets the flat glass-to-air
interface, the angle between the ray and
the normal to the interface is called the
angle of incidence.[3] If this angle is
sufficiently small, the ray is partly reflected
but mostly transmitted, and the
transmitted portion is refracted away from
the normal, so that the angle of refraction
(between the refracted ray and the normal
to the interface) is greater than the angle
of incidence. For the moment, let us call
the angle of incidence θi and the angle of
refraction θt (where t is for transmitted,
reserving r for reflected). As θi increases
and approaches a certain "critical angle",
denoted by θc (or sometimes θcr), the
angle of refraction approaches 90° (that is,
the refracted ray approaches a tangent to
the interface), and the refracted ray
becomes fainter while the reflected ray
becomes brighter.[4] As θi increases
beyond θc, the refracted ray disappears
and only the reflected ray remains, so that
all of the energy of the incident ray is
reflected; this is total internal reflection
(TIR). In brief:

If  θi < θc , the incident ray is split, being


partly reflected and partly refracted;
If  θi > θc , the incident ray suffers total
internal reflection (TIR); none of it is
transmitted.
Critical angle
The critical angle is the smallest angle of
incidence that yields total reflection.[5] For
light waves and other
electromagnetic waves in isotropic media,
there is a well-known formula for the
critical angle in terms of the refractive
indices. For some other types of waves, it
is more convenient to think in terms of
propagation velocities rather than
refractive indices. The latter approach is
more direct and more general, and will
therefore be discussed first.
Fig. 4: Refraction of a wavefront (red) from a medium

with lower normal velocity v1 to a medium with higher


normal velocity v2. The incident and refracted
segments of the wavefront meet in a common line L
(seen "end-on"), which travels at velocity u.

When a wavefront is refracted from one


medium to another, the incident
(incoming) and refracted (outgoing)
portions of the wavefront meet at a
common line on the refracting surface
(interface). Let this line, denoted by L,
move at velocity u across the surface,[6][7]
where u is measured normal to L (Fig. 4).
Let the incident and refracted wavefronts
propagate with normal velocities and
(respectively), and let them make the
dihedral angles θ1 and θ2 (respectively)
with the interface. From the geometry,
is the component of u in the direction
normal to the incident wave, so that
. Similarly, .
Solving each equation for 1/u and
equating the results, we obtain the general
law of refraction for waves:

.
 
 

 
 
(1)
But the dihedral angle between two planes
is also the angle between their normals.
So θ1 is the angle between the normal to
the incident wavefront and the normal to
the interface, while θ2 is the angle between
the normal to the refracted wavefront and
the normal to the interface; and Eq. (1)
tells us that the sines of these angles are
in the same ratio as the respective
velocities.[8]

This result has the form of "Snell's law",


except that we have not yet said that the
ratio of velocities is constant, nor
identified θ1 and θ2 with the angles of
incidence and refraction (called θi and θt
above). However, if we now suppose that
the media are isotropic, two further
conclusions follow: first, the two velocities,
and hence their ratio, are independent of
their directions; and second, the wave-
normal directions coincide with the ray
directions, so that θ1 and θ2 coincide with
the angles of incidence and refraction as
defined above.[Note 1]

Fig 5: Behavior of a ray incident from a medium of


Fig. 5: Behavior of a ray incident from a medium of
higher refractive index n1 to a medium of lower
refractive index n2 ,  at increasing angles of
incidence.[Note 2]

Fig. 6: The angle of refraction for grazing incidence


from air to water  is the critical angle for incidence
from water to air.

Obviously the angle of refraction cannot


exceed 90°. In the limiting case, we put
θ2 = 90° and θ1 = θc in Eq. (1), and solve
for the critical angle:


 

 
 

(2)

In deriving this result, we retain the


assumption of isotropic media in order to
identify θ1 and θ2 with the angles of
incidence and refraction.[Note 3]

For electromagnetic waves, and especially


for light, it is customary to express the
above results in terms of refractive indices.
The refractive index of a medium with
normal velocity is defined as
, where c is the speed of light in
a vacuum.[9]  Hence .  Similarly,
.  Making these substitutions
in Eqs. (1) and (2), we obtain

 
 

 
 

(3)

and

.  
 

 
 

(4)

Eq. (3) is the law of refraction for general


media, in terms of refractive indices,
provided that θ1 and θ2 are taken as the
dihedral angles; but if the media are
isotopic, then n1 and n2 become
independent of direction while θ1 and θ2
may be taken as the angles of incidence
and refraction for the rays, and Eq. (4)
follows. So, for isotropic media, Eqs. (3) 
and (4) together describe the behavior in
Fig. 5.

According to Eq. (4), for incidence from


water (n1 ≈ 1.333) to air (n2 ≈ 1), we have
θc ≈ 48.6°, whereas for incidence from
common glass or acrylic (n1 ≈ 1.50) to air
(n2 ≈ 1), we have θc ≈ 41.8°.

The arcsin function yielding θc is defined


only if n2 ≤ n1  .  Hence, for
isotropic media, total internal reflection
cannot occur if the second medium has a
higher refractive index (lower normal
velocity) than the first. For example, there
cannot be TIR for incidence from air to
water; rather, the critical angle for
incidence from water to air is the angle of
refraction at grazing incidence from air to
water (Fig. 6).[10]

The medium with the higher refractive


index is commonly described as optically
denser, and the one with the lower
refractive index as optically rarer.[11] Hence
it is said that total internal reflection is
possible for "dense-to-rare" incidence, but
not for "rare-to-dense" incidence.

Everyday examples
Fig. 7: Total internal reflection by the water's surface at

the shallow end of a swimming pool. The broad


bubble-like apparition between the swimmer and her
reflection  is merely a disturbance of the reflecting
surface. Some of the space above the water level can
be seen through "Snell's window" at the top of the
frame.

When standing beside an aquarium with


one's eyes below the water level, one is
likely to see fish or submerged objects
reflected in the water-air surface (Fig. 1).
The brightness of the reflected image —
just as bright as the "direct" view — can be
startling.

A similar effect can be observed by


opening one's eyes while swimming just
below the water's surface. If the water is
calm, the surface outside the critical angle
(measured from the vertical) appears
mirror-like, reflecting objects below. The
region above the water cannot be seen
except overhead, where the hemispherical
field of view is compressed into a conical
field known as Snell's window, whose
angular diameter is twice the critical angle
(cf. Fig. 6).[12]  The field of view above the
water is theoretically 180° across, but
seems less because as we look closer to
the horizon, the vertical dimension is more
strongly compressed by the refraction;
e.g., by Eq. (3), for air-to-water incident
angles of 90°, 80°, and 70°, the
corresponding angles of refraction are
48.6° (θcr in Fig. 6), 47.6°, and 44.8°,
indicating that the image of a point 20°
above the horizon is 3.8° from the edge of
Snell's window while the image of a point
10° above the horizon is only 1° from the
edge.[13]
Fig. 7, for example, is a photograph taken
near the bottom of the shallow end of a
swimming pool. What looks like a broad
horizontal stripe on the right-hand wall
consists of the lower edges of a row of
orange tiles, and their reflections; this
marks the water level, which can then be
traced across the other wall. The swimmer
has disturbed the surface above her,
scrambling the lower half of her reflection,
and distorting the reflection of the ladder
(to the right). But most of the surface is
still calm, giving a clear reflection of the
tiled bottom of the pool. The space above
the water is not visible except at the top of
the frame, where the handles of the ladder
are just discernible above the edge of
Snell's window.

Fig. 8: A round "brilliant"- cut diamond.

The critical angle influences the angles at


which gemstones are cut. The round
"brilliant" cut, for example, is designed to
refract light incident on the front facets,
reflect it twice by TIR off the back facets,
and transmit it out again through the front
facets, so that the stone looks bright.
Diamond (Fig. 8) is especially suitable for
this treatment, because its high refractive
index (about 2.42) and consequently small
critical angle (about 24.5°) yield the
desired behavior over a wide range of
viewing angles.[14] Cheaper materials that
are similarly amenable to this treatment
include cubic zirconia (index ≈ 2.15) and
moissanite (non-isotropic, hence
doubly refractive, with an index ranging
from about 2.65 to 2.69,[Note 4] depending
on direction and polarization); both of
these are therefore popular as
diamond simulants.
Related phenomena
Evanescent wave (qualitative
explanation)

Mathematically, waves are described in


terms of time-varying fields, a "field" being
a function of location in space. A
propagating wave requires an "effort" field
and a "flow" field, the latter being a vector
(if we are working in two or three
dimensions). The product of effort and
flow is related to power (see System
equivalence). For example, for sound
waves in a non-viscous fluid, we might
take the effort field as the pressure
(a scalar), and the flow field as the fluid
velocity (a vector). The product of these
two is intensity (power per unit
area).[15][Note 5] For electromagnetic waves,
we shall take the effort field as the
electric field  E , and the flow field as the
magnetizing field  H. Both of these are
vectors, and their vector product is again
the intensity (see Poynting vector).[16]

When a wave in (say) medium 1 is


reflected off the interface between
medium 1 and medium 2, the flow field in
medium 1 is the vector sum of the flow
fields due to the incident and reflected
waves.[Note 6]  If the reflection is oblique,
the incident and reflected fields are not in
opposite directions and therefore cannot
cancel out at the interface; even if the
reflection is total, either the normal
component or the tangential component
of the combined field (as a function of
location and time) must be non-zero
adjacent to the interface. Furthermore, the
physical laws governing the fields will
generally imply that one of the two
components is continuous across the
interface (that is, it does not suddenly
change as we cross the interface); for
example, for electromagnetic waves, one
of the interface conditions is that the
tangential component of H is continuous
if there is no surface current.[17] Hence,
even if the reflection is total, there must be
some penetration of the flow field into
medium 2; and this, in combination with
the laws relating the effort and flow fields,
implies that there will also be some
penetration of the effort field. The same
continuity condition implies that the
variation ("waviness") of the field in
medium 2 will be synchronized with that of
the incident and reflected waves in
medium 1.
Fig. 9: Depiction of an incident sinusoidal plane wave
(bottom) and the associated evanescent wave (top),
under conditions of total internal reflection. The
reflected wave is not shown.

But, if the reflection is total, the spatial


penetration of the fields into medium 2
must be limited somehow, or else the total
extent and hence the total energy of those
fields would continue to increase, draining
power from medium 1. Total reflection of a
continuing wavetrain permits some energy
to be stored in medium 2, but does not
permit a continuing transfer of power from
medium 1 to medium 2.

Thus, using mostly qualitative reasoning,


we can conclude that total internal
reflection must be accompanied by a
wavelike field in the "external" medium,
traveling along the interface in
synchronism with the incident and
reflected waves, but with some sort of
limited spatial penetration into the
"external" medium; such a field may be
called an evanescent wave.
Fig. 9 shows the basic idea. The incident
wave is assumed to be plane and
sinusoidal. The reflected wave, for
simplicity, is not shown. The evanescent
wave travels to the right in lock-step with
the incident and reflected waves, but its
amplitude falls off with increasing
distance from the interface.

(Two features of the evanescent wave in


Fig. 9 are to be explained later: first, that
the evanescent wave crests are
perpendicular to the interface; and second,
that the evanescent wave is slightly ahead
of the incident wave.)
Frustrated TIR

If reflection is to be total, there must be no


diversion of the evanescent wave.
Suppose, for example, that
electromagnetic waves incident from
glass to air at a certain angle of incidence
are subject to TIR. And suppose that we
have a third medium whose refractive
index is sufficient high that, if the third
medium were to replace the second (air),
we would get a standard transmitted
wavetrain for the same angle of incidence.
Then, if the third medium is brought within
a few wavelengths of the first, where the
evanescent wave has significant
amplitude, the evanescent wave is
effectively refracted into the third medium,
giving non-zero transmission into the third
medium, and therefore less than total
reflection back into the first medium.[18] As
the amplitude of the evanescent wave
decays across the air gap, the transmitted
waves is attenuated, so that there is less
transmission, and therefore more
reflection, than there would be with no gap;
but as long as there is some transmission,
the reflection is less than total. This
phenomenon is called frustrated total
internal reflection, abbreviated "frustrated
TIR" or "FTIR".
Fig. 10: Disembodied fingerprints visible from the
inside of a glass of water, due to frustrated TIR.

Frustrated TIR can be observed by looking


into the top of a glass of water held in
one's hand (Fig. 10). If the glass is held
loosely, contact may not be sufficiently
close and widespread to produce a
noticeable effect. But if it is held more
tightly, the ridges of one's fingerprints
interact strongly with the evanescent
waves, allowing the ridges to be seen
through the otherwise totally reflecting
glass-air surface.[19]

The same effect can be demonstrated


with microwaves, using paraffin wax as the
"internal" medium. In this case the
permitted gap width might be (e.g.) 1 cm
or several cm, which is easily observable
and adjustable.[1][20]

The term frustrated TIR also applies to the


case in which the evanescent wave is
scattered by an object sufficiently close to
the reflecting interface. This effect,
together with the strong dependence of
the amount of scattered light on the
distance from the interface, is exploited in
total internal reflection microscopy.[21]

The mechanism of FTIR is called


evanescent-wave coupling, and is
somewhat analogous to quantum
tunneling. Due to the wave nature of
matter, an electron has a non-zero
probability of "tunneling" through a barrier,
even if classical mechanics would say that
its energy is insufficient.[18][19] Similarly,
due to the wave nature of light, a photon
has a non-zero probability of crossing a
gap, even if ray optics would say that its
approach is too oblique.
Another reason why internal reflection may
be less than total, even beyond the critical
angle, is that the external medium may be
"lossy" (less than perfectly transparent). In
that case, the external medium will absorb
energy from the evanescent wave, so that
the maintenance of the evanescent wave
will draw power from the incident wave.
The consequent less-than-total reflection
is called attenuated total reflectance (ATR).
This effect, and especially the frequency-
dependence of the absorption, can be
used to study the composition of an
unknown external medium.[22]

Derivation of evanescent wave


In a uniform plane sinusoidal
electromagnetic wave, the electric field  E
has the form

 
 

 
 
(5)

where Ek is the (constant) complex


amplitude vector,  i is the imaginary unit,  k
is the wave vector (whose magnitude k is
the angular wavenumber),  r is the
position vector,  ω is the angular
frequency,  t is time, and it is understood
that the real part of the expression is the
physical field.[Note 7] The magnetizing field 
H has the same form with the same k and
ω. The value of the expression is
unchanged if the position r varies in a
direction normal to k; hence k is normal to
the wavefronts.

If ℓ is the component of r in the direction


of k, the field (5) can be written
.  If the argument of is
to be constant,  ℓ must increase at the
velocity known as the phase
velocity.[23] This in turn is equal to
where c is the phase velocity in the
reference medium (taken as a vacuum)
and n is the local refractive index w.r.t. the
reference medium. Solving for k gives
i.e.

 
 

 
 

(6)
where is the wavenumber in a
vacuum.[24][Note 8]

From (5), the electric field in the "external"


medium has the form

 
 

 
 
(7)

where kt is the wave vector for the


transmitted wave (we assume isotropic
media, but the transmitted wave is not yet
assumed to be evanescent).
Fig. 11: Incident, reflected, and transmitted wave
vectors (ki , kr , and kt ), for incidence from a medium
with higher refractive index n1 to a medium with lower
refractive index n2. The red arrows are perpendicular to

the wave vectors and therefore parallel to the


respective wavefronts.

In Cartesian coordinates (x, y,z), let the


region y < 0 have refractive index n1 , and
let the region y > 0 have refractive index
n2. Then the xz plane is the interface, and
the y axis is normal to the interface
(Fig. 11). Let i and j (in bold roman type)
be the unit vectors in the x and y
directions, respectively. Let the plane of
incidence (containing the incident wave-
normal and the normal to the interface) be
the xy plane (the plane of the page), with
the angle of incidence θi measured from j
towards i. Let the angle of refraction,
measured in the same sense, be θt  (t for
transmitted, reserving r for reflected).

From (6), the transmitted wave vector kt


has magnitude n2k0. Hence, from the
geometry,
where the last step uses Snell's law.
Taking the dot product with the position
vector, we get

so that Eq. (7) becomes

 
 

 
 
(8)

In the case of TIR, the angle θt does not


exist in the usual sense. But we can still
interpret (8) for the transmitted
(evanescent) wave, by allowing cos θt to
be complex. This becomes necessary
when we write cos θt in terms of sin θt ,
and thence in terms of sin θi using Snell's
law:

For θi greater than the critical angle, the


value under the square-root symbol is
negative, so that[25]

.  
 

 
 
(9)

To determine which sign is applicable, we


substitute (9) into (8), obtaining

 
 

 
 

(10)
where the undetermined sign is the
opposite of that in (9). For an evanescent
transmitted wave — that is, one whose
amplitude decays as y increases — the
undetermined sign in (10) must be minus,
so the undetermined sign in (9) must be
plus.[Note 9]

With the correct sign, the result (10) can


be abbreviated

 
 

 
 
(11)

where

 
 

 
 

(12)
and k0 is the wavenumber in a vacuum,
i.e.  .

So the evanescent wave is a plane


sinewave traveling in the x direction, with
an amplitude that decays exponentially in
the y direction (cf. Fig. 9). It is evident that
the energy stored in this wave likewise
travels in the x direction and does not
cross the interface. Hence the Poynting
vector generally has a component in the x
direction, but its y component averages to
zero (although its instantaneous y
component is not identically zero).[26][27]
Fig. 12: Penetration depth of the evanescent wave
(in wavelengths) vs. angle of incidence, for various
values of the relative refractive index (internal
w.r.t. external)

Eq. (11) indicates that the amplitude of the


evanescent wave falls off by a factor e as
the coordinate y (measured from the
interface) increases by the distance
commonly called the
"penetration depth" of the evanescent
wave. Taking reciprocals of the first
equation of (12), we find that the
penetration depth is[27]

where λ0 is the wavelength in a vacuum,


i.e.  .[28]  Dividing the numerator and
denominator by n2 yields

where is the wavelength in


the second (external) medium. Hence we
can plot d in units of λ2 , as a function of
the angle of incidence, for various values
of (Fig. 12).  As θi decreases
towards the critical angle, the denominator
approaches zero, so that d increases
without limit — as is to be expected,
because as soon as θi is less than critical,
uniform plane waves are permitted in the
external medium. As θi approaches 90°
(grazing incidence),  d approaches a
minimum

For incidence from water to air, or


common glass to air,  dmin is not much
λ2
different from 2π .  But d is larger at
smaller angles of incidence (Fig. 12), and
the amplitude may still be significant at
distances of several times d; for example,
because e−4.6 is just greater than 0.01, the
evanescent wave amplitude within a
distance 4.6d  of the interface is at least
1% of its value at the interface. Hence,
speaking loosely, we tend to say that the
evanescent wave amplitude is significant
within "a few wavelengths" of the
interface.

Phase shifts

Between 1817 and 1823, Augustin-Jean


Fresnel discovered that total internal
reflection is accompanied by a non-trivial
phase shift (that is, a phase shift that is
not restricted to 0° or 180°), as the Fresnel
reflection coefficient acquires a non-zero
imaginary part.[29] We shall now explain
this effect for electromagnetic waves in
the case of linear, homogeneous, isotropic,
non-magnetic media. The phase shift
turns out to be an advance, which grows
as the incidence angle increases beyond
the critical angle, but which depends on
the polarization of the incident wave.

In equations (5), (7), (8), (10), and (11), we


advance the phase by the angle ϕ if we
replace ωt by ωt+ϕ  (that is, if we replace
−ωt by −ωt−ϕ),  with the result that the
(complex) field is multiplied by e−iϕ. So a
phase advance is equivalent to
multiplication by a complex constant with
a negative argument. This becomes more
obvious when (e.g.) the field (5) is factored
as where the last factor
contains the time-dependence.[Note 10]

To represent the polarization of the


incident, reflected, or transmitted wave,
the electric field adjacent to an interface
can be resolved into two perpendicular
components, known as the s and p
components, which are parallel to the
surface and the plane of incidence,
respectively; in other words, the s and p
components are respectively square and
parallel to the plane of incidence.[Note 11]

For each component of polarization, the


incident, reflected, or transmitted electric
field (E in Eq. (5)) has a certain direction,
and can be represented by its (complex)
scalar component in that direction. The
reflection or transmission coefficient can
then be defined as a ratio of complex
components at the same point, or at
infinitesimally separated points on
opposite sides of the interface. But, in
order to fix the signs of the coefficients, we
must choose positive senses for the
"directions". For the s components, the
obvious choice is to say that the positive
directions of the incident, reflected, and
transmitted fields are all the same (e.g.,
the z direction in Fig. 11). For the p
components, this article adopts the
convention that the positive directions of
the incident, reflected, and transmitted
fields are inclined towards the same
medium (that is, towards the same side of
the interface, e.g. like the red arrows in
Fig. 11).[Note 12]  But the reader should be
warned that some books use a different
convention for the p components, causing
a different sign in the resulting formula for
the reflection coefficient.[30]
For the s polarization, let the reflection and
transmission coefficients be rs and ts
respectively. For the p polarization, let the
corresponding coefficients be rp and tp .
Then, for linear, homogeneous, isotropic,
non-magnetic media, the coefficients are
given by:[31]

 
 

 
 

(13)

 
 

 
 

(14)

 
 

 
 

(15)

.  
 

 
 

(16)
(For a derivation of the above, see  Fresnel
equations  § Theory.)

Now we suppose that the transmitted


wave is evanescent. With the correct sign
(+), substituting (9) into (13) gives

where

that is, n is the index of the "internal"


medium relative to the "external" one, or
the index of the internal medium if the
external one is a vacuum.[Note 13]  So the
magnitude of rs is 1, and the argument of
rs is

which gives a phase advance of [32]


 

 
 

(17)

Making the same substitution in (14), we


find that ts has the same denominator as
rs with a positive real numerator (instead
of a complex conjugate numerator) and
therefore has half the argument of rs , so
that the phase advance of the evanescent
wave is half that of the reflected wave.
With the same choice of sign,[Note 14]
substituting (9) into (15) gives

whose magnitude is 1, and whose


argument is

which gives a phase advance of [32]

.
 
 

 
 

(18)
Making the same substitution in (16), we
again find that the phase advance of the
evanescent wave is half that of the
reflected wave.

Equations (17) and (18) apply when


θc ≤ θi < 90°, where θi is the angle of
incidence and θc is the critical angle
arcsin (1/n).  These equations show that

each phase advance is zero at the


critical angle (for which the numerator is
zero);
each phase advance approaches 180°
as θi → 90°; and
δp > δs at intermediate values of θi
(because the factor n is in the
numerator of (18) and the denominator
of (17)).[33]

For θi ≤ θc , the reflection coefficients are


given by equations (13) and (15), and are
real, so that the phase shift is either 0° (if
the coefficient is positive) or 180° (if the
coefficient is negative).

In (13), if we put
(Snell's law) and multiply the numerator
and denominator by n1  sin θt ,
1
we obtain [34][35]
 
 

 
 
(19)

which is positive for all angles of incidence


with a transmitted ray (since θt > θi),
giving a phase shift δs of zero.

If we do likewise with (15), the result is


easily shown to be equivalent to [36][37]

 
 

 
 
(20)

which is negative for small angles (that is,


near normal incidence), but changes sign
at Brewster's angle, where  θi and θt are
complementary. Thus the phase shift δp is
180° for small θi but switches to 0° at
Brewster's angle. Combining the
complementarity with Snell's law yields
θi = arctan (1/n) as Brewster's angle for
dense-to-rare incidence.[Note 15]

(Equations (19) and (20) are known as


Fresnel's sine law and Fresnel's tangent
law.[38] Both reduce to 0/0 at normal
incidence, but yield the correct results in
the limit as θi → 0. That they have
opposite signs as we approach normal
incidence is an obvious disadvantage of
the sign convention used in this article; the
corresponding advantage is that they have
the same signs at grazing incidence.)
Fig. 13: Phase advance at "internal" reflections for
refractive indices of 1.55, 1.5, and 1.45 ("internal"
relative to "external"). Beyond the critical angle, the
p (red) and s (blue) polarizations undergo unequal
phase shifts on total internal reflection; the
macroscopically observable difference between these
shifts is plotted in black.
That completes the information needed to
plot δs and δp for all angles of incidence.
This is done in Fig. 13,[32] with δp in red
and δs in blue, for three refractive indices.
On the angle-of-incidence scale (horizontal
axis), Brewster's angle is where δp (red)
falls from 180° to 0°, and the critical angle
is where both δp and δs (red and blue)
start to rise again. To the left of the critical
angle is the region of partial reflection,
where both reflection coefficients are real
(phase 0° or 180°) with magnitudes less
than 1. To the right of the critical angle is
the region of total reflection, where both
reflection coefficients are complex with
magnitudes equal to 1. In that region, the
black curves show the phase advance of
the p component relative to the
s component:[39]

It can be seen that a refractive index of


1.45 is not enough to give a 45° phase
difference, whereas a refractive index of
1.5 is enough (by a slim margin) to give a
45° phase difference at two angles of
incidence: about 50.2° and 53.3°.

This 45° relative shift is employed in


Fresnel's invention, now known as the
Fresnel rhomb, in which the angles of
incidence are chosen such that the two
internal reflections cause a total relative
phase shift of 90° between the two
polarizations of an incident wave. This
device performs the same function as a
birefringent quarter-wave plate, but is
more achromatic (that is, the phase shift
of the rhomb is less sensitive to
wavelength). Either device may be used,
for instance, to transform linear
polarization to circular polarization (which
Fresnel also discovered) and vice versa.

In Fig. 13,  δ is computed by a final


subtraction; but there are other ways of
expressing it. Fresnel himself, in 1823,[40]
gave a formula for  cos δ.  Born and Wolf
(1970, p. 50) derive an expression for
tan (δ/2), and find its maximum
analytically.

For TIR of a beam with finite width, the


variation in the phase shift with the angle
of incidence gives rise to the Goos–
Hänchen effect, which is a lateral shift of
the reflected beam within the plane of
incidence.[27][41] This effect applies to
linear polarization in the s or p direction.
The Imbert–Fedorov effect is an analogous
effect for circular or elliptical polarization,
and produces a shift perpendicular to the
plane of incidence.[42]
Applications
Optical fibers exploit total internal
reflection to carry signals over long
distances with little attenuation.[43] They
are used in telecommunication cables, and
in image-forming fiberscopes such as
colonoscopes.[44]

In the catadioptric Fresnel lens, invented


by Augustin-Jean Fresnel for use in
lighthouses, the outer prisms use TIR to
deflect light from the lamp through a
greater angle than would be possible with
purely refractive prisms, but with less
absorption of light (and less risk of
tarnishing) than with conventional
mirrors.[45]

Fig. 14: Porro prisms (labeled 2 & 3) in a pair of


binoculars.

Other reflecting prisms that use TIR


include the following (with some overlap
between the categories):[46]
Image-erecting prisms for binoculars
and spotting scopes include paired
45°-90°-45° Porro prisms (Fig. 14), the
Porro–Abbe prism, the inline Koenig[47]
and Abbe–Koenig prisms, and the
compact inline Schmidt–Pechan prism.
(The last consists of two components,
of which one is a kind of Bauernfeind
prism, which requires a reflective
coating on one of its two reflecting
faces, due to a sub-critical angle of
incidence.) These prisms have the
additional function of folding the optical
path from the objective lens to the
prime focus, reducing the overall length
for a given primary focal length.
A prismatic star diagonal for an
astronomical telescope may consist of
a single Porro prism (configured for a
single reflection, giving a mirror-reversed
image) or an Amici roof prism (which
gives a non-reversed image).
Roof prisms use TIR at two faces
meeting at a sharp 90° angle. This
category includes the Koenig, Abbe–
Koenig, Schmidt–Pechan, and Amici
types (already mentioned), and the roof
pentaprism used in SLR cameras; the
last of these requires a reflective
coating on one non-TIR face.
A prismatic corner reflector uses three
total internal reflections to reverse the
direction of incoming light.
The Dove prism gives an inline view with
mirror-reversal.

Polarizing prisms: Although the Fresnel


rhomb, which converts between linear and
elliptical polarization, is not birefringent
(doubly refractive), there are other kinds of
prisms that combine birefringence with
TIR in such a way that light of a particular
polarization is totally reflected while light
of the orthogonal polarization is at least
partly transmitted. Examples include the
Nicol prism,[48] Glan–Thompson prism,
Glan–Foucault prism (or "Foucault
prism"),[49][50] and Glan–Taylor prism.[51]

Refractometers, which measure refractive


indices, often use the critical angle.[52][53]

Rain sensors for automatic


windscreen/windshield wipers have been
implemented using the principle that total
internal reflection will guide an infrared
beam from a source to a detector if the
outer surface of the windshield is dry, but
any water drops on the surface will divert
some of the light.[54]

Edge-lit LED panels, used (e.g.) for


backlighting of LCD computer monitors,
exploit TIR to confine the LED light to the
acrylic glass pane, except that some of the
light is scattered by etchings on one side
of the pane, giving an approximately
uniform luminous emittance.[55]

Fig. 15: Operation of a "trans-geometry" TIR


fluorescence microscope: (1) objective, (2) emission
beam [signal], (3) immersion oil, (4) cover slip,
(5) specimen, (6) evanescent wave range, (7) excitation
beam, (8) quartz prism.
Total internal reflection microscopy
(TIRM) uses the evanescent wave to
illuminate small objects close to the
reflecting interface. The consequent
scattering of the evanescent wave (a form
of frustrated TIR), makes the objects
appear bright when viewed from the
"external" side.[21] In the total internal
reflection fluorescence microscope
(TIRFM), instead of relying on simple
scattering, we choose an evanescent
wavelength short enough to cause
fluorescence (Fig. 15).[56] The high
sensitivity of the illumination to the
distance from the interface allows
measurement of extremely small
displacements and forces.[57]

A beam-splitter cube uses frustrated TIR


to divide the power of the incoming beam
between the transmitted and reflected
beams.[18]

Optical modulation can be accomplished


by means of frustrated TIR with a variable
gap.[58] As the transmission coefficient is
highly sensitive to the gap width (the
function being approximately exponential
until the gap is almost closed), this
technique can achieve a large dynamic
range.
Optical fingerprinting devices have used
frustrated TIR to record images of
persons' fingerprints without the use of ink
(cf. Fig. 11).[59]

Gait analysis can be performed by using


frustrated TIR with a high-speed camera,
to capture and analyze footprints.[60]

A gonioscope, used in optometry and


ophthalmology for the diagnosis of
glaucoma, suppresses TIR in order to look
into the angle between the iris and the
cornea. This view is usually blocked by TIR
at the cornea-air interface. The
gonioscope replaces the air with a higher-
index medium, allowing transmission at
oblique incidence, typically followed by
reflection in a "mirror", which itself may be
implemented using TIR.[61][62]

History
Discovery

The surprisingly comprehensive and


largely correct explanations of the rainbow
by Theodoric of Freiberg (written 1304–
1310) and Kamāl al-Dīn al-Fārisī (1309),[63]
although sometimes mentioned in
connection with total internal reflection
(TIR), are of dubious relevance because
the internal reflection of sunlight in a
spherical raindrop is not total.[Note 16] But,
according to Carl Benjamin Boyer,
Theodoric's treatise on the rainbow also
classified optical phenomena under five
causes, the last of which was "a total
reflection at the boundary of two
transparent media".[64] Theodoric's work
was forgotten until it was rediscovered by
Giovanni Battista Venturi in 1814.[65]
Johannes Kepler (1571–1630).

Theodoric having fallen into obscurity, the


discovery of TIR was generally attributed
to Johannes Kepler, who published his
findings in his Dioptrice in 1611. Although
Kepler failed to find the true law of
refraction, he showed by experiment that
for air-to-glass incidence, the incident and
refracted rays rotated in the same sense
about the point of incidence, and that as
the angle of incidence varied through ±90°,
the angle of refraction (as we now call it)
varied through ±42°. He was also aware
that the incident and refracted rays were
interchangeable. But these observations
did not cover the case of a ray incident
from glass to air at an angle beyond 42°,
and Kepler promptly concluded that such a
ray could only be reflected.[66]

René Descartes rediscovered the law of


refraction and published it in his Dioptrique
of 1637. In the same work he mentioned
the senses of rotation of the incident and
refracted rays and the condition of TIR. But
he neglected to discuss the limiting case,
and consequently failed give an
expression for the critical angle, although
he could easily have done so.[67]

Huygens and Newton: Rival


explanations

Christiaan Huygens, in his Treatise on Light


(1690), paid much attention to the
threshold at which the incident ray is
"unable to penetrate into the other
transparent substance".[68] Although he
gave neither a name nor an algebraic
expression for the critical angle, he gave
numerical examples for glass-to-air and
water-to-air incidence, noted the large
change in the angle of refraction for a
small change in the angle of incidence
near the critical angle, and cited this as the
cause of the rapid increase in brightness
of the reflected ray as the refracted ray
approaches the tangent to the
interface.[69] Huygens' insight is confirmed
by modern theory: in Eqs. (13) and (15)
above, there is nothing to say that the
reflection coefficients increase
exceptionally steeply as θt approaches
90°, except that, according to Snell's law, 
θt itself is an increasingly steep function of
θi.

Christiaan Huygens (1629–1695).


Huygens offered an explanation of TIR
within the same framework as his
explanations of the laws of rectilinear
propagation, reflection, ordinary refraction,
and even the extraordinary refraction of
"Iceland crystal" (calcite). That framework
rested on two premises: first, every point
crossed by a propagating wavefront
becomes a source of secondary
wavefronts ("Huygens' principle"); and
second, given an initial wavefront, any
subsequent position of the wavefront is
the envelope (common tangent surface) of
all the secondary wavefronts emitted from
the initial position. All cases of reflection
or refraction by a surface are then
explained simply by considering the
secondary waves emitted from that
surface. In the case of refraction from a
medium of slower propagation to a
medium of faster propagation, there is a
certain obliquity of incidence beyond
which it is impossible for the secondary
wavefronts to form a common tangent in
the second medium;[70] this is what we
now call the critical angle. As the incident
wavefront approaches this critical
obliquity, the refracted wavefront becomes
concentrated against the refracting
surface, augmenting the secondary waves
that produce the reflection back into the
first medium.[71]

Huygens' system even accommodated


partial reflection at the interface between
different media, albeit vaguely, by analogy
with the laws of collisions between
particles of different sizes.[72] However, as
long as the wave theory continued to
assume longitudinal waves, it had no
chance of accommodating polarization,
hence no chance of explaining the
polarization-dependence of extraordinary
refraction,[73] or of the partial reflection
coefficient, or of the phase shift in TIR.
Isaac Newton (1642/3–1726/7).

Isaac Newton rejected the wave


explanation of rectilinear propagation,
believing that if light consisted of waves, it
would "bend and spread every way" into
the shadows.[74] His corpuscular theory of
light explained rectilinear propagation
more simply, and it accounted for the
ordinary laws of refraction and reflection,
including TIR, on the hypothesis that the
corpuscles of light were subject to a force
acting perpendicular to the interface.[75] In
this model, for dense-to-rare incidence, the
force was an attraction back towards the
denser medium, and the critical angle was
the angle of incidence at which the normal
velocity of the approaching corpuscle was
just enough to reach the far side of the
force field; at more oblique incidence, the
corpuscle would be turned back.[76]
Newton gave what amounts to a formula
for the critical angle, albeit in words: "as
the Sines are which measure the
Refraction, so is the Sine of Incidence at
which the total Reflexion begins, to the
Radius of the Circle".[77]
Newton went beyond Huygens in two
ways. First, not surprisingly, Newton
pointed out the relationship between TIR
and dispersion: when a beam of white light
approaches a glass-to-air interface at
increasing obliquity, the most strongly-
refracted rays (violet) are the first to be
"taken out" by "total Reflexion", followed by
the less-refracted rays.[78] Second, he
observed that total reflection could be
frustrated (as we now say) by laying
together two prisms, one plane and the
other slightly convex; and he explained this
simply by noting that the corpuscles would
be attracted not only to the first prism, but
also to the second.[79]
In two other ways, however, Newton's
system was less coherent. First, his
explanation of partial reflection depended
not only on the supposed forces of
attraction between corpuscles and media,
but also on the more nebulous hypothesis
of "Fits of easy Reflexion" and "Fits of easy
Transmission".[80] Second, although his
corpuscles could conceivably have "sides"
or "poles", whose orientations could
conceivably determine whether the
corpuscles suffered ordinary or
extraordinary refraction in "Island-
Crystal",[81] his geometric description of
the extraordinary refraction[82] was
theoretically unsupported[83] and
empirically inaccurate.[84]

Laplace, Malus, and attenuated total


reflectance (ATR)

William Hyde Wollaston, in the first of a


pair of papers read to the Royal Society of
London in 1802,[53] reported his invention
of a refractometer based on the critical
angle of incidence from an internal
medium of known "refractive power"
(refractive index) to an external medium
whose index was to be measured.[85] With
this device, Wollaston measured the
"refractive powers" of numerous materials,
some of which were too opaque to permit
direct measurement of an angle of
refraction. Translations of his papers were
published in France in 1803, and
apparently came to the attention of Pierre-
Simon Laplace.[86]

Pierre-Simon Laplace (1749–1827).


According to Laplace's elaboration of
Newton's theory of refraction, a corpuscle
incident on a plane interface between two
homogeneous isotropic media was
subject to a force field that was
symmetrical about the interface. If both
media were transparent, total reflection
would occur if the corpuscle were turned
back before it exited the field in the
second medium. But if the second
medium were opaque, reflection would not
be total unless the corpuscle were turned
back before it left the first medium; this
required a larger critical angle than the one
given by Snell's law, and consequently
impugned the validity of Wollaston's
method for opaque media.[87] Laplace
combined the two cases into a single
formula for the relative refractive index in
terms of the critical angle (minimum angle
of incidence for TIR). The formula
contained a parameter which took one
value for a transparent external medium
and another value for an opaque external
medium. Laplace's theory further
predicted a relationship between refractive
index and density for a given
substance.[88]
Étienne-Louis Malus (1775–1812).

In 1807, Laplace's theory was tested


experimentally by his protégé, Étienne-
Louis Malus. Taking Laplace's formula for
the refractive index as given, and using it
to measure the refractive index of bees'
wax in the liquid (transparent) state and
the solid (opaque) state at various
temperatures (hence various densities),
Malus verified Laplace's relationship
between refractive index and density.[89][90]

But Laplace's theory implied that if the


angle of incidence exceeded his modified
critical angle, the reflection would be total
even if the external medium was
absorbent. Clearly this was wrong: in
Eqs. (12) above, there is no threshold
value of the angle θi beyond which κ
becomes infinite; so the penetration depth
of the evanescent wave (1/κ) is always
non-zero, and the external medium, if it is
at all lossy, will attenuate the reflection. As
to why Malus apparently observed such an
angle for opaque wax, we must infer that
there was a certain angle beyond which
the attenuation of the reflection was so
small that ATR was visually
indistinguishable from TIR.[91]

Fresnel and the phase shift

Fresnel came to the study of total internal


reflection through his research on
polarization. In 1811, François Arago
discovered that polarized light was
apparently "depolarized" in an orientation-
dependent and color-dependent manner
when passed through a slice of doubly-
refractive crystal: the emerging light
showed colors when viewed through an
analyzer (second polarizer). Chromatic
polarization, as this phenomenon came to
be called, was more thoroughly
investigated in 1812 by Jean-Baptiste Biot.
In 1813, Biot established that one case
studied by Arago, namely quartz cut
perpendicular to its optic axis, was
actually a gradual rotation of the plane of
polarization with distance.[92]
Augustin-Jean Fresnel (1788–1827).

In 1816, Fresnel offered his first attempt at


a wave-based theory of chromatic
polarization. Without (yet) explicitly
invoking transverse waves, his theory
treated the light as consisting of two
perpendicularly polarized components.[93]
In 1817 he noticed that plane-polarized
light seemed to be partly depolarized by
total internal reflection, if initially polarized
at an acute angle to the plane of
incidence.[94] By including total internal
reflection in a chromatic-polarization
experiment, he found that the apparently
depolarized light was a mixture of
components polarized parallel and
perpendicular to the plane of incidence,
and that the total reflection introduced a
phase difference between them.[95]
Choosing an appropriate angle of
incidence (not yet exactly specified) gave
a phase difference of 1/8 of a cycle. Two
such reflections from the "parallel faces"
of "two coupled prisms" gave a phase
difference of 1/4 of a cycle. In that case, if
the light was initially polarized at 45° to the
plane of incidence and reflection, it
appeared to be completely depolarized
after the two reflections. These findings
were reported in a memoir submitted and
read to the French Academy of Sciences in
November 1817.[96]

In 1821, Fresnel derived formulae


equivalent to his sine and tangent laws
(Eqs. (19) and (20), above)  by modeling
light waves as transverse elastic waves
with vibrations perpendicular to what had
previously been called the plane of
polarization.[97][Note 17] He promptly
confirmed by experiment that the
equations correctly predicted the direction
of polarization of the reflected beam when
the incident beam was polarized at 45° to
the plane of incidence, for light incident
from air onto glass or water.[98] The
experimental confirmation was reported in
a "postscript" to the work in which Fresnel
expounded his mature theory of chromatic
polarization, based on transverse
waves.[99] Details of the derivation were
given later, in a memoir read to the
Academy in January 1823.[100] The
derivation combined conservation of
energy with continuity of the tangential
vibration at the interface, but failed to
allow for any condition on the normal
component of vibration.[101]

Meanwhile, in a memoir submitted in


December 1822,[102] Fresnel coined the
terms linear polarization, circular
polarization, and elliptical polarization.[103]
For circular polarization, the two
perpendicular components were a quarter-
cycle (±90°) out of phase.

The new terminology was useful in the


memoir of January 1823,[100] containing
the detailed derivations of the sine and
tangent laws: in that same memoir,
Fresnel found that for angles of incidence
greater than the critical angle, the resulting
reflection coefficients were complex with
unit magnitude. Noting that the magnitude
represented the amplitude ratio as usual,
he guessed that the argument represented
the phase shift, and verified the hypothesis
by experiment.[104] The verification
involved

calculating the angle of incidence that


would introduce a total phase difference
of 90° between the s and p components,
for various numbers of total internal
reflections at that angle (generally there
were two solutions),
subjecting light to that number of total
internal reflections at that angle of
incidence, with an initial linear
polarization at 45° to the plane of
incidence, and
checking that the final polarization was
circular.[105]
This procedure was necessary because,
with the technology of the time, one could
not measure the s and p phase-shifts
directly, and one could not measure an
arbitrary degree of ellipticality of
polarization, such as might be caused by
the difference between the phase shifts.
But one could verify that the polarization
was circular, because the brightness of the
light was then insensitive to the
orientation of the analyzer.

For glass with a refractive index of 1.51,


Fresnel calculated that a 45° phase
difference between the two reflection
coefficients (hence a 90° difference after
two reflections) required an angle of
incidence of 48°37' or 54°37'. He cut a
rhomb to the latter angle and found that it
performed as expected.[106] Thus the
specification of the Fresnel rhomb was
completed. Similarly, Fresnel calculated
and verified the angle of incidence that
would give a 90° phase difference after
three reflections at the same angle, and
four reflections at the same angle. In each
case there were two solutions, and in each
case he reported that the larger angle of
incidence gave an accurate circular
polarization (for an initial linear
polarization at 45° to the plane of
reflection). For the case of three
reflections he also tested the smaller
angle, but found that it gave some
coloration due to the proximity of the
critical angle and its slight dependence on
wavelength. (Compare Fig. 13 above,
which shows that the phase difference δ is
more sensitive to the refractive index for
smaller angles of incidence.)

For added confidence, Fresnel predicted


and verified that four total internal
reflections at 68°27' would give an
accurate circular polarization if two of the
reflections had water as the external
medium while the other two had air, but
not if the reflecting surfaces were all wet
or all dry.[107]

Fresnel's deduction of the phase shift in


TIR is thought to have been the first
occasion on which a physical meaning
was attached to the argument of a
complex number. Although this reasoning
was applied without the benefit of knowing
that light waves were electromagnetic, it
passed the test of experiment, and
survived remarkably intact after James
Clerk Maxwell changed the presumed
nature of the waves.[108] Meanwhile,
Fresnel's success inspired James
MacCullagh and Augustin-Louis Cauchy,
beginning in 1836, to analyze reflection
from metals by using the Fresnel
equations with a complex refractive
index.[109] The imaginary part of the
complex index represents absorption.[110]

The term critical angle, used for


convenience in the above narrative, is
anachronistic: it apparently dates from
1873.[111]

In the 20th century, quantum


electrodynamics reinterpreted the
amplitude of an electromagnetic wave in
terms of the probability of finding a
photon.[112] In this framework, partial
transmission and frustrated TIR concern
the probability of a photon crossing a
boundary, and attenuated total reflectance
concerns the probability of a photon being
absorbed on the other side.

Research into the more subtle aspects of


the phase shift in TIR, including the Goos–
Hänchen and Imbert–Fedorov effects and
their quantum interpretations, has
continued into the 21st century.[42]

See also
Attenuated total reflectance
Evanescent field
Fiberscope
Fresnel equations
Fresnel lens
Fresnel rhomb
Goos–Hänchen effect
Imbert–Fedorov effect
Optical fiber
Polarization (waves)
Snell's window
TIR fluorescence microscope
TIR microscopy
Total "external" reflection

Notes
1. Birefringent media, such as calcite, are
non-isotropic (anisotropic). When we
say that the extraordinary refraction of
a calcite crystal "violates Snell's law",
we mean that Snell's law does not
apply to the extraordinary ray, because
the direction of this ray inside the
crystal generally differs from that of
the associated wave-normal (Huygens,
1690, tr. Thompson, p. 65, Art. 24), and
because the wave-normal speed is
itself dependent on direction. (Note
that the cited passage contains a
translation error: in the phrase
"conjugate with respect to diameters
which are not in the straight line AB",
the word "not" is unsupported by
Huygens' original French , and is
geometrically incorrect.)
2. According to Eqs. (13) and (15),
reflection is total for incidence at the
critical angle. On that basis, Fig. 5
ought to show a fully reflected ray, and
no tangential ray, for incidence at θc.
But, due to diffraction, an incident
beam of finite width cannot have a
single angle of incidence; there must
be some divergence of the beam.
Moreover, the graph of the reflection
coefficient vs. the angle of incidence
becomes vertical at θc (Jenkins &
White, 1976, p. 527), so that a small
divergence of the beam causes a large
loss of reflection. Similarly, near the
critical angle, a small divergence in the
angle of incidence causes a large
divergence in the angle of refraction
(cf. Huygens, 1690, tr. Thompson,
p. 41); the tangential refracted ray
should therefore be taken only as a
limiting case.
3. For non-isotropic media, Eq. (1) still
describes the law of refraction in
terms of wave-normal directions and
speeds, but the range of applicability
of that law is determined by the
constraints on the ray directions
(cf. Buchwald, 1989, p. 29).
4. The quoted range varies because of
different crystal polytypes.
5. Power "per unit area" is appropriate for
fields in three dimensions. In two
dimensions, we might want the
product of effort and flow to be power
per unit length. In one dimension, or in
a lumped-element model, we might
want it to be simply power.
6. We assume that the equations
describing the fields are linear.
7. The above form (5) is typically used by
physicists. Electrical engineers
typically prefer the form
that is, they not only
use j instead of i for the imaginary
unit, but also change the sign of the
exponent, with the result that the
whole expression is replaced by its
complex conjugate, leaving the real
part unchanged. The electrical
engineers' form and the formulae
derived therefrom may be converted to
the physicists' convention by
substituting −i for j (Stratton, 1941,
pp. vii–viii).
8. We assume that there are no Doppler
shifts, so that ω does not change at
interfaces between media.
9. If we correctly convert this to the
electrical engineering convention, we
get −j√ ⋯   on the right-hand side of (9),
which is not the principal square root.
So it is not valid to assume, a priori,
that what mathematicians call the
"principal square root" is the physically
applicable one.
10. In the electrical engineering
convention, the time-dependent factor
is e jωt, so that a phase advance
corresponds to multiplication by a
complex constant with a positive
argument. This article, however, uses
the physics convention, with the time-
dependent factor e−iωt.
11. The s originally comes from the
German senkrecht, meaning
"perpendicular" (to the plane of
incidence). The alternative mnemonics
in the text are perhaps more suitable
for English speakers.
12. In other words, for both polarizations,
this article uses the convention that
the positive directions of the incident,
reflected, and transmitted fields are all
the same for whichever field is normal
to the plane of incidence; this is the E
field for the s polarization, and the H
field for the p polarization.
13. This nomenclature follows Jenkins &
White, 1976, pp. 526–9. Some authors,
however, use the reciprocal refractive
index and therefore obtain different
forms for our Eqs. (17) and (18).
Examples include Born & Wolf [1970,
p. 49, eqs. (60)] and Stratton [1941,
p. 499, eqs. (43)]. Furthermore, Born &
Wolf define δ⊥ and δ∥ as arguments
rather than phase shifts, causing a
change of sign.
14. It is merely fortuitous that the principal
square root turns out to be the correct
one in the present situation, and only
because we use the time-dependent
factor e−iωt.  If we instead used the
electrical engineers' time-dependent
factor e jωt,  choosing the principal
square root would yield the same
argument for the reflection coefficient,
but this would be interpreted as the
opposite phase shift, which would be
wrong. But if we choose the square
root so that the transmitted field is
evanescent, we get the right phase
shift with either time-dependent factor.
15. The more familiar formula  arctan n is
for rare-to-dense incidence. In both
cases, n is the refractive index of the
denser medium relative to the rarer
medium.
16. For an external ray incident on a
spherical raindrop, the refracted ray is
in the plane of the incident ray and the
center of the drop, and the angle of
refraction is less than the critical angle
for water-air incidence; but this angle
of refraction, by the spherical
symmetry, is also the angle of
incidence for the internal reflection,
which is therefore less than total.
Moreover, if that reflection were total,
all subsequent internal reflections
would have the same angle of
incidence (due to the symmetry) and
would also be total, so that the light
would never escape to produce a
visible bow.
17. Hence, where Fresnel says that after
total internal reflection at the
appropriate incidence, the wave
polarized parallel to the plane of
incidence is "behind" by 1/8 of a cycle
(quoted by Buchwald, 1989, p. 381), he
refers to the wave whose plane of
polarization is parallel to the plane of
incidence, i.e. the wave whose
vibration is perpendicular to that plane,
i.e. what we now call the s component.
References
1. R.P. Feynman, R.B. Leighton, and
M. Sands, 1963–2013, The Feynman
Lectures on Physics, California
Institute of Technology, Volume II, 
§ 33-6 .
2. P.P. Antich, J.A. Anderson,
R.B. Ashman, J.E. Dowdey,
J. Gonzales, R.C. Murry, J.E. Zerwekh,
and C.Y.C. Pak (April 1991),
"Measurement of mechanical
properties of bone material in vitro by
ultrasound reflection: Methodology
and comparison with ultrasound
transmission" , Journal of Bone and
Mineral Research, 6 (4): 417–26,
doi:10.1002/jbmr.5650060414 .
3. Jenkins & White, 1976, p. 11.
4. Jenkins & White, 1976, p. 527. (The
refracted beam becomes fainter in
terms of total power, but not
necessarily in terms of visibility,
because the beam also becomes
narrower as it becomes more nearly
tangential.)
5. Jenkins & White, 1976, p. 26.
6. Cf.  Thomas Young in the Quarterly
Review, April 1814, reprinted in
T. Young (ed. G. Peacock),
Miscellaneous Works of the late
Thomas Young, London: J. Murray,
1855, vol. 1, at p. 263 .
7. Cf. Born & Wolf, 1970, pp. 12–13.
8. Cf.  Huygens, 1690, tr. Thompson, p. 
38.
9. Born & Wolf, 1970, p. 13; Jenkins &
White, 1976, pp. 9–10. This definition
uses a vacuum as the "reference
medium". In principle, any isotropic
medium can be chosen as the
reference. For some purposes it is
convenient to choose air, in which the
speed of light is about 0.03% lower
than in a vacuum (cf. Rutten and van
Venrooij, 2002, pp. 10, 352). The
present article, however, chooses a
vacuum.
10. Cf. Jenkins & White, 1976, p. 25.
11. Jenkins & White, 1976, pp. 10, 25.
12. Cf.  D.K. Lynch (1 February 2015),
"Snell's window in wavy water" ,
Applied Optics, 54 (4): B8–B11,
doi:10.1364/AO.54.0000B8 .
13. Huygens (1690, tr. Thompson, p. 41),
for glass-to-air incidence, noted that if
the obliqueness of the incident ray is
only 1° short of critical, the refracted
ray is more than 11° from the tangent.
N.B.: Huygens' definition of the "angle
of incidence" is the complement of the
modern definition.
14. J.R. Graham, "Can you cut a gem
design for tilt brightness?" ,
International Gem Society, accessed
21 March 2019; archived
14 December 2018.
15. 'PJS' (author), "Sound Pressure, Sound
Power, and Sound Intensity: What's the
difference?" Siemens PLM
Community, accessed 10 April 2019;
archived 10 April 2019.
16. Stratton, 1941, pp. 131–7.
17. Stratton, 1941, p. 37.
18. Cf.  Harvard Natural Sciences Lecture
Demonstrations, "Frustrated Total
Internal Reflection" , accessed 9 April
2019; archived 2 August 2018.
19. R. Ehrlich, 1997, Why Toast Lands
Jelly-side Down: Zen and the Art of
Physics Demonstrations, Princeton
University Press, ISBN 978-0-691-
02891-0, p. 182 , accessed 26 March
2019.
20. R. Bowley, 2009, "Total Internal
Reflection" (4-minute video), Sixty
Symbols, Brady Haran for the
University of Nottingham, from 1:25.
21. E.J. Ambrose (24 November 1956), "A
surface contact microscope for the
study of cell movements", Nature, 178
(4543): 1194,
Bibcode:1956Natur.178.1194A ,
doi:10.1038/1781194a0 .
22. Thermo Fisher Scientific, "FTIR Sample
Techniques: Attenuated Total
Reflection (ATR)" , accessed 9 April
2019.
23. Jenkins & White, 1976, p. 228.
24. Born & Wolf, 1970, pp. 16–17,
eqs. (20), (21).
25. Born & Wolf, 1970, p. 47, eq. (54),
where their n is our (not our
).
26. Stratton, 1941, p. 499; Born & Wolf,
1970, p. 48.
27. Laboratory of Cold Atoms Near
Surfaces (Jagiellonian University),
"Evanescent wave properties" ,
accessed 11 April 2019; archived
28 April 2018. (N.B.: This page uses z
for the coordinate normal to the
interface, and the superscripts ⊥ and ∥
for the s ("TE") and p polarizations,
respectively. Pages on this site use the
time-dependent factor e+iωt — that is,
the electrical engineers' time-
dependent factor with the physicists'
symbol for the imaginary unit.)
28. Born & Wolf, 1970, p. 16.
29. Whittaker, 1910, pp. 132, 135–6.
30. One notable authority that uses the
"different" convention (but without
taking it very far) is The Feynman
Lectures on Physics, at Volume I,
eq. (33.8) (for B), and Volume II,
Figs. 33-6 and 33-7.
31. Born & Wolf, 1970, p. 40, eqs. (20), (21),
where the subscript ⊥ corresponds to
s, and ∥ to p.
32. Cf. Jenkins & White, 1976, p. 529.
33. "The phase of the polarization in which
the magnetic field is parallel to the
interface is advanced with respect to
that of the other polarization."
— Fitzpatrick, 2013, p. 140; Fitzpatrick,
2013a; emphasis added.
34. Fresnel, 1866, pp. 773, 789n.
35. Born & Wolf, 1970, p. 40, eqs. (21a);
Hecht, 2002, p. 115, eq. (4.42); Jenkins
& White, 1976, p. 524, eqs. (25a).
36. Fresnel, 1866, pp. 757, 789n.
37. Born & Wolf, 1970, p. 40, eqs. (21a);
Hecht, 2002, p. 115, eq. (4.43); Jenkins
& White, 1976, p. 524, eqs. (25a).
38. Whittaker, 1910, p. 134; Darrigol, 2012,
p.213.
39. Stratton, 1941, p. 500, eq. (44). The
corresponding expression in Born &
Wolf (1970, p. 50) is the other way
around because the terms represent
arguments rather than phase shifts.
40. Buchwald, 1989, pp. 394,453; Fresnel,
1866, pp. 759, 786–7, 790.
41. P.R. Berman, 2012, "Goos-Hänchen
effect" , Scholarpedia 7 (3): 11584, 
§ 2.1, especially eqs. (1) to (3). Note
that Berman's n is the reciprocal of the
n in the present article.
42. K.Y. Bliokh and A. Aiello (January
2013), "Goos–Hänchen and Imbert–
Fedorov beam shifts: An overview",
Journal of Optics, 15 (1): 014001,
arXiv:1210.8236v2, doi:10.1088/2040-
8978/15/1/014001 .
43. Jenkins & White, 1976, pp. 40–42.
44. W.W.H. Rudd (December 1971),
"Fiberoptic Colonoscopy: A Dramatic
Advance in Colon Surgery" (PDF,
4 pages), Canadian Family Physician,
17 (12): 42–5.
45. Levitt, 2013, pp. 79–80.
46. Jenkins & White, 1976, pp. 26–7
(Porro, Dove, 90° Amici, corner
reflector, Lummer-Brodhun);  Born &
Wolf, 1970, pp. 240–41 (Porro,
Koenig), 243–4 (Dove).
47. Born & Wolf, 1970, p. 241.
48. Born & Wolf, 1970, pp. 690–91.
49. R. Nave, "Prisms for Polarization"
(Nicol, Glan–Foucault), Georgia State
University, accessed 27 March 2019;
archived 25 March 2019.
50. Jenkins & White, 1976, pp. 510–11
(Nicol, Glan–Thompson, "Foucault").
51. J.F. Archard; A.M. Taylor (December
1948), "Improved Glan-Foucault prism",
Journal of Scientific Instruments, 25
(12): 407–9,
Bibcode:1948JScI...25..407A ,
doi:10.1088/0950-7671/25/12/304 .
52. Buchwald, 1989, pp. 19–21; Jenkins &
White, 1976, pp. 27–8.
53. W.H. Wollaston, "A method of
examining refractive and dispersive
powers, by prismatic reflection" ,
Philosophical Transactions of the
Royal Society, vol. 92, part II (1802),
pp. 365–80 (read 24 June 1802),
doi:10.1098/rstl.1802.0014 .
54. HELLA GmbH & Co. KGaA, "Rain
sensor & headlight sensor testing -
Repair instructions & fault diagnosis" ,
accessed 9 April 2019; archived
8 April 2019.
55. J. Gourlay, "Making Light Work – Light
Sources for Modern Lighting
Requirements" , LED Professional,
accessed 29 March 2019; archived
12 April 2016.
56. D. Axelrod (April 1981), "Cell-substrate
contacts illuminated by total internal
reflection fluorescence" , Journal of
Cell Biology, 89 (1): 141–5,
doi:10.1083/jcb.89.1.141 ,
PMC 2111781 , PMID 7014571 .
57. D. Axelrod (November 2001), "Total
Internal Reflection Fluorescence
Microscopy in Cell Biology" (PDF),
Traffic, 2 (11): 764–74,
doi:10.1034/j.1600-
0854.2001.21104.x , PMID 11733042 .
58. R.W. Astheimer; G. Falbel; S. Minkowitz
(January 1966), "Infrared modulation
by means of frustrated total internal
reflection", Applied Optics, 5 (1): 87–
91, doi:10.1364/AO.5.000087 ,
PMID 20048791 .
59. N.J. Harrick (1962-3), "Fingerprinting
via total internal reflection" , Philips
Technical Review, 24 (9): 271–4.
60. Noldus Information Technology,
"CatWalk™ XT" , accessed 29 March
2019; archived 25 March 2019.
61. E. Bruce, R. Bendure, S. Krein, and
N. Lighthizer, "Zoom in on
Gonioscopy" , Review of Optometry,
21 September 2016.
62. Glaucoma Associates of Texas,
"Gonioscopy" , accessed 29 March
2019; archived 22 August 2018.
63. H.G. Topdemir (2007), "Kamal al-Din
Al-Farisi's Explanation of the
Rainbow" , Humanity & Social
Sciences Journal, ISSN 1818-4960 ,
2 (1): 75–85.
64. Boyer, 1959, pp. 113, 114, 335.  Boyer
cites J. Würschmidt's edition of
Theodoric's De iride et radialibus
impressionibus, in Beiträge zur
Geschichte der Philosophie des
Mittelalters, vol. 12, nos. 5–6 (1914),
at p. 47.
65. Boyer, 1959, pp. 307, 335.
66. E. Mach (tr. J.S. Anderson & A.F.A. 
Young), The Principles of Physical
Optics: An Historical and Philosophical
Treatment (London: Methuen & Co,
1926), reprinted Mineola, NY: Dover,
2003, pp. 30–32.
67. A.I. Sabra, Theories of Light: From
Descartes to Newton (London:
Oldbourne Book Co., 1967), reprinted
Cambridge University Press, 1981,
pp. 111–12.
68. Huygens, 1690, tr. Thompson, p. 39.
69. Huygens, 1690, tr. Thompson, pp. 40–
41. Notice that Huygens' definition of
the "angle of incidence" is the
complement of the modern definition.
70. Huygens, 1690, tr. Thompson, pp. 39–
40.
71. Huygens, 1690, tr. Thompson, pp. 40–
41.
72. Huygens, 1690, tr. Thompson, pp. 16, 
42.
73. Huygens, 1690, tr. Thompson, pp. 92–
4.
74. Newton, 1730, p. 362.
75. Darrigol, 2012, pp. 93–4, 103.
76. Newton, 1730, pp. 370–71.
77. Newton, 1730, p. 246. Notice that a
"sine" meant the length of a side for a
specified "radius" (hypotenuse),
whereas nowadays we take the radius
as unity or express the sine as a ratio.
78. Newton, 1730, pp. 56–62, 264.
79. Newton, 1730, pp. 371–2.
80. Newton, 1730, p. 281.
81. Newton, 1730, p. 373.
82. Newton, 1730, p. 356.
83. Buchwald, 1980, pp. 327, 331–2.
84. Buchwald, 1980, pp. 335–6, 364;
Buchwald, 1989, pp. 9–10, 13.
85. Buchwald, 1989, pp. 19–21.
86. Buchwald, 1989, p. 28.
87. Darrigol, 2012, pp. 187–8.
88. Buchwald, 1989, p. 30.
89. Buchwald, 1980, pp. 29–31.
90. E. Frankel (May 1976), "Corpuscular
optics and the wave theory of light:
The science and politics of a
revolution in physics", Social Studies
of Science, 6 (2): 141–84, at p. 145.
91. Cf. Buchwald, 1989, p. 30 (quoting
Malus).
92. Darrigol, 2012, pp. 193–6, 290.
93. Darrigol, 2012, p. 206.
94. This effect had been previously
discovered by Brewster, but not yet
adequately reported. See: "On a new
species of moveable polarization" ,
[Quarterly] Journal of Science and the
Arts, vol. 2, no. 3, 1817, p. 213; 
T. Young, "Chromatics", Supplement to
the Fourth, Fifth, and Sixth Editions of
the Encyclopædia Britannica, vol. 3
(first half, issued February 1818),
pp. 141–63, at p. 157 ;  Lloyd, 1834, p. 
368.
95. Darrigol, 2012, p. 207.
96. A. Fresnel, "Mémoire sur les
modifications que la réflexion imprime
à la lumière polarisée" ("Memoir on the
modifications that reflection
impresses on polarized light"), signed
& submitted 10 November 1817, read
24 November 1817; printed in Fresnel,
1866, pp. 441–85, including pp. 452
(rediscovery of depolarization by total
internal reflection), 455 (two
reflections, "coupled prisms",
"parallelepiped in glass"), 467–8
(phase difference per reflection); see
also p. 487, note 1, for the date of
reading.
97. Darrigol, 2012, p. 212.
98. Buchwald, 1989, pp. 390–91; Fresnel,
1866, pp. 646–8.
99. A. Fresnel, "Note sur le calcul des
teintes que la polarisation développe
dans les lames cristallisées" ("Note on
the calculation of hues that
polarization develops in crystalline
laminae"), Annales de Chimie et de
Physique, vol. 17, pp. 102–12 (May
1821), 167–96 (June 1821), 312–16
("Postscript", July 1821); reprinted in
Fresnel, 1866, pp. 609–48.
100. A. Fresnel, "Mémoire sur la loi des
modifications que la réflexion imprime
à la lumière polarisée" ("Memoir on the
law of the modifications that reflection
impresses on polarized light"), read
7 January 1823; reprinted in Fresnel,
1866, pp. 767–99 (full text, published
1831), pp. 753–62 (extract, published
1823). See especially pp. 773 (sine
law), 757 (tangent law), 760–61 and
792–6 (angles of total internal
reflection for given phase differences).
101. Buchwald, 1989, pp. 391–3; Darrigol,
2012, pp. 212–13; Whittaker, 1910,
pp. 133–5.
102. A. Fresnel, "Mémoire sur la double
réfraction que les rayons lumineux
éprouvent en traversant les aiguilles
de cristal de roche suivant les
directions parallèles à l'axe" ("Memoir
on the double refraction that light rays
undergo in traversing the needles of
rock crystal [quartz] in directions
parallel to the axis"), signed &
submitted 9 December 1822; reprinted
in Fresnel, 1866, pp. 731–51 (full text,
published 1825), pp. 719–29 (extract,
published 1823). On the publication
dates, see also Buchwald, 1989,
p. 462, ref. 1822b.
103. Buchwald, 1989, pp. 230–31; Fresnel,
1866, p. 744.
104. Lloyd, 1834, pp. 369–70; Buchwald,
1989, pp. 393–4, 453; Fresnel, 1866,
pp. 781–96.
105. Fresnel, 1866, pp. 760–61, 792–6;
Whewell, 1857, p. 359.
106. Fresnel, 1866, pp. 760–61, 792–3.
107. Fresnel, 1866, pp. 761, 793–6;
Whewell, 1857, p. 359.
108. Bochner, 1963, pp. 198–200.
109. Whittaker, 1910, pp. 177–9.
110. Bochner, 1963, p. 200; Born & Wolf,
1970, p. 613.
111. Merriam-Webster, Inc., "critical angle" ,
accessed 21 April 2019. (No primary
source is given.)
112. R.P. Feynman, 1985 (seventh printing,
1988), QED: The Strange Theory of
Light and Matter, Princeton University
Press, esp. pp. 33, 109–10.

Bibliography
S. Bochner (June 1963), "The
significance of some basic
mathematical conceptions for physics",
Isis, 54 (2): 179–205;
jstor.org/stable/228537 .
M. Born and E. Wolf, 1970, Principles of
Optics, 4th Ed., Oxford: Pergamon Press.
C.B. Boyer, 1959, The Rainbow: From
Myth to Mathematics, New York:
Thomas Yoseloff.
J.Z. Buchwald (December 1980),
"Experimental investigations of double
refraction from Huygens to Malus",
Archive for History of Exact Sciences,
21 (4): 311–373.
J.Z. Buchwald, 1989, The Rise of the
Wave Theory of Light: Optical Theory and
Experiment in the Early Nineteenth
Century, University of Chicago Press,
ISBN 0-226-07886-8.
O. Darrigol, 2012, A History of Optics:
From Greek Antiquity to the Nineteenth
Century, Oxford, ISBN 978-0-19-964437-
7.
R. Fitzpatrick, 2013, Oscillations and
Waves: An Introduction, Boca Raton, FL:
CRC Press, ISBN 978-1-4665-6608-8.
R. Fitzpatrick, 2013a, "Total Internal
Reflection" , University of Texas at
Austin, accessed 14 March 2018.
A. Fresnel, 1866  (ed.  H. de Senarmont,
E. Verdet, and L. Fresnel), Oeuvres
complètes d'Augustin Fresnel, Paris:
Imprimerie Impériale (3 vols., 1866–70),
vol. 1 (1866) .
E. Hecht, 2002, Optics, 4th Ed., Addison
Wesley, ISBN 0-321-18878-0.
C. Huygens, 1690, Traité de la Lumière
(Leiden: Van der Aa), translated by
S.P. Thompson as Treatise on Light ,
University of Chicago Press, 1912;
Project Gutenberg, 2005. (Cited page
numbers match the 1912 edition and the
Gutenberg HTML edition.)
F.A. Jenkins and H.E. White, 1976,
Fundamentals of Optics, 4th Ed.,
New York: McGraw-Hill, ISBN 0-07-
032330-5.
T.H. Levitt, 2013, A Short Bright Flash:
Augustin Fresnel and the Birth of the
Modern Lighthouse, New York:
W.W. Norton, ISBN 978-0-393-35089-0.
H. Lloyd, 1834, "Report on the progress
and present state of physical optics" ,
Report of the Fourth Meeting of the
British Association for the Advancement
of Science (held at Edinburgh in 1834),
London: J. Murray, 1835, pp. 295–413.
I. Newton, 1730, Opticks: or, a Treatise of
the Reflections, Refractions, Inflections,
and Colours of Light , 4th Ed. (London:
William Innys, 1730; Project Gutenberg,
2010); republished with Foreword by
A. Einstein and Introduction by
E.T. Whittaker (London: George Bell &
Sons, 1931); reprinted with additional
Preface by I.B. Cohen and Analytical
Table of Contents by D.H.D. Roller, 
Mineola, NY: Dover, 1952, 1979 (with
revised preface), 2012. (Cited page
numbers match the Gutenberg HTML
edition and the Dover editions.)
H.G.J. Rutten and M.A.M. van Venrooij,
1988 (fifth printing, 2002), Telescope
Optics: A Comprehensive Manual for
Amateur Astronomers, Richmond, VA:
Willmann-Bell, ISBN 978-0-943396-18-7.
J.A. Stratton, 1941, Electromagnetic
Theory, New York: McGraw-Hill.
W. Whewell, 1857, History of the
Inductive Sciences: From the Earliest to
the Present Time, 3rd Ed., London:
J.W. Parker & Son, vol. 2 .
E.T. Whittaker, 1910, A History of the
Theories of Aether and Electricity: From
the Age of Descartes to the Close of the
Nineteenth Century , London: Longmans,
Green, & Co.

External links

Wikimedia Commons has media


related to Total internal reflection.

Mr. Mangiacapre, "Fluorescence in a
Liquid" (video, 1m28s), uploaded
13 March 2012.  (Fluorescence and TIR
of a violet laser beam in quinine water.)
PhysicsatUVM, "Frustrated Total Internal
Reflection" (video, 37s), uploaded
21 November 2011.  ("A laser beam
undergoes total internal reflection in a
fogged piece of plexiglass...")
SMUPhysics, "Internal Reflection"
(video, 12s), uploaded 20 May 2010. 
(Transition from refraction through
critical angle to TIR in a 45°-90°-45°
prism.)

Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Total_internal_reflection&oldid=913509655"

Last edited on 1 September 2019, at 14:35


Content is available under CC BY-SA 3.0 unless
otherwise noted.

Potrebbero piacerti anche