You are on page 1of 70

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/359272079

Young's interference experiment: Past, present, and


future

Chapter  in  Progress in Optics · March 2022


DOI: 10.1016/bs.po.2022.01.003

CITATION READS

1 202

2 authors:

Greg Gbur Taco D. Visser


University of North Carolina at Charlotte Vrije Universiteit Amsterdam
154 PUBLICATIONS   4,408 CITATIONS    176 PUBLICATIONS   4,323 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Plasmons - optics View project

Coherence View project

All content following this page was uploaded by Taco D. Visser on 03 September 2022.

The user has requested enhancement of the downloaded file.


CHAPTER FOUR

Young’s interference experiment:


Past, present, and future
Greg Gbura and Taco D. Visserb,c
a
Department of Physics and Optical Science, University of North Carolina at Charlotte, Charlotte,
NC, United States
b
Department of Physics and Astronomy, Vrije Universiteit, Amsterdam, The Netherlands
c
Institute of Optics, University of Rochester, Rochester, NY, United States

Contents
1. Introduction 275
2. Life and work of Thomas Young 277
3. Young’s interference experiment 286
4. Weak light 289
5. The wave-particle duality for massive particles 296
6. Optical coherence 304
7. Surface plasmons 314
8. Entanglement and duality 324
9. Multi-order interference and exotic paths 327
10. Singular optics 331
11. Conclusions 337
Acknowledgments 337
References 337

1. Introduction
The experiment that is the subject of this review has been referred to as
“the two-pinhole experiment” or “the double-slit experiment” (Young
himself referred to it as both). It has had a tremendous influence on the his-
tory of science, and it continues to inspire and direct modern researchers to
use it, and its modifications, to probe new areas of physics. In this article we
explore the history of Young’s experiment, its implications, and those
discoveries and experiments that have followed in its wake, even up to
the present day.

Progress in Optics, Volume 67 Copyright # 2022 Elsevier B.V. 275


ISSN 0079-6638 All rights reserved.
https://doi.org/10.1016/bs.po.2022.01.003
276 Greg Gbur and Taco D. Visser

We begin with a short biographical sketch of Thomas Young in which


not only his work in optics is described, but also his other, wide-ranging
interests. This is followed, in Section 3, by a mathematical analysis of the
classic two aperture experiment, for both scalar and vector fields. It was this
seminal experiment that first showed that light is a wave, more precisely a
transverse wave, with a wavelength that can be determined from the posi-
tion of the interference fringes.
Section 4 is concerned with weak light and single-photon experiments.
This leads naturally to a discussion of the wave-particle duality as proposed
by de Broglie, which is the topic of Section 5. We review how this concept
has been investigated using Young’s experiment for ever larger particles.
Also Wheeler’s delayed-choice Gedanken experiment and its eventual phys-
ical realization are described.
In Section 6 we describe the elements of optical coherence theory. The
correlation functions that describe the statistical properties of a wave field are,
in contrast to the rapid oscillations of the field itself, quantities that can be
observed with a two-pinhole setup. The work of Zernike and Wolf showed
how this statistical information can be extracted from the interference process.
It is also discussed how coherence affects both the state of polarization and the
spectrum of the diffracted field
The role of Surface Plasmon Polaritons, or SPPs, in optical transmission
through nanoapertures is the topic of Section 7. We use a simple phenom-
enological model that takes into account the possibility of SPPs that are gen-
erated at one slit being backscattered into a freely propagating field at an
adjacent slit. With this model the influence of SPPs on the intensity, polar-
ization, and state of coherence of the transmitted field is analyzed.
Section 8 deals with the recently discovered connections between a
wavefield’s degree of polarization, fringe visibility, distinguishability, and
entanglement. The equations that are presented apply, albeit with different
meanings, to the quantum domain as well as to the classical domain.
In Section 9 it is discussed how Young’s setup with three slits can be used
to validate Born’s Rule, one of the basic tenets of Quantum Mechanics. This
naturally segues into an examination of nonclassical or “looped” trajectories
in which a particle travels more than once through the slit arrangement.
Section 10 describes the pivotal role that Young’s setup plays in elucidat-
ing the connection between different types of field singularities such as phase
singularities, polarization singularities, and coherence singularities. It also
describes how, in a two- or three-pinhole setup, these different singularities
unfold and can evolve into each other.
Young’s interference experiment 277

2. Life and work of Thomas Young


Thomas Young was born in 1773 into an upper middle class family in
Milverton, in the English county of Somerset. To describe him as a child
prodigy would be somewhat of an understatement: according to one of
his biographers (Peacock, 1855a) he had twice read the Bible before the
age of 4. His early education was a mixture of styles. He was home schooled,
and for a while he attended a boarding school, an experience that he would
later describe as “miserable.” He also taught himself many subjects. A vora-
cious reader and gifted student of European and Oriental languages, he
translated parts of the Bible into 13 languages while still in his teens.
His uncle, a medical doctor, took a great interest in Young’s schooling. It
was decided that he would follow in his uncle’s footsteps and so, in 1792, he
began his medical studies in London. It was there that his fascination with the
physics and physiology of vision began. At the time it was not clear how the
eye accommodates to produce a sharp image of objects that are at different
distances. After dissecting an ox eye and examining it under a microscope,
Young concluded, correctly, that it is the action of muscles that cause the
crystalline lens to change its curvature, and thus its focal length.
The paper in which he announced his findings to the Royal Society
(Young, 1793) had three remarkable consequences. First of all, it was deemed
impressive enough to merit his election as a Fellow of the Society. Second, his
result was strongly disputed by others who themselves had failed to find any
such changes in the form of the human lens. And third, the famous surgeon
John Hunter claimed that Young had stolen his own ideas of accommodation,
overhearing them at a dinner party where Young had been present. Faced
with such opposition and this unproven allegation of plagiarism, Young rec-
anted. This episode turned out to be the first of many scientific controversies
that Young would have to endure. However, he never gave up his interest in
vision and would later, based on an examination of his own eyes, be the first to
describe astigmatism (Young, 1801). This was followed, not long afterward,
by a publication in which he speculated that the human retina contains three
different kinds of receptors, one for each principle color (Young, 1802b). The
existence of these receptors, which are nowadays called cones, was confirmed
much later.
Young continued his studies in Edinburgh where he much enjoyed min-
gling with the local beau monde. Exposed to vibrant city living, he gradually
gave up the sober Quaker lifestyle of his youth, and took up activities such as
278 Greg Gbur and Taco D. Visser

fencing and dancing. After continuing his education in the German city of
G€ottingen, he obtained the title of Doctor of Physic, Surgery and
Midwifery. But in order to fulfill his plan of practicing medicine in
London, he was required to spend more time at an English university. He thus
went to Cambridge where, in reaction to his wide-ranging knowledge and
talents, his fellow students nicknamed him “phenomenon Young.” In order
to graduate, Young had to publicly subscribe to the Thirty-Nine Articles, the
foundational statement of the established Church of England. This meant a
formal separation from Young’s upbringing as a Quaker.
After completing his studies he moved to London in 1800 to set up a med-
ical practice there. However, Young turned out to be unpopular as a practical
physician. One of his biographers suggests that this may have been due to a
lack of bedside manner (Peacock, 1855a). In any case, he soon changed his
career by accepting a position as Professor of Natural Philosophy in the
Royal Institution in London. The Institution had been recently founded,
in 1799, with the aim to provide access to scientific and technical knowledge
to a lay audience and to promote its practical application. Young was tasked
with giving lectures and demonstrations on a wide range of topics.
It was at the Institution that Young began pondering the possibility of
light being a wave phenomenon. This was a rather radical thought as, nearly
100 years earlier, Newton had seemingly proven that light consists of a
stream of particles. As he had written in his Opticks (Newton, 1704):
The waves on the surface of stagnating water, passing by the sides of a broad
obstacle which stops part of them, bend afterwards and dilate themselves
gradually into the quiet water behind the obstacle. The waves, pulses or vibra-
tions of the air, wherein sounds consist, bend manifestly, though not as much
as the waves of water. For a bell or a cannon may be heard beyond a hill which
interrupts the sight of the sounding body, and sounds are propagated as read-
ily through crooked pipes as straight ones. But light is never known to follow
crooked passages nor to bend into the shadow. For the fixed stars, by the inter-
position of any of the planets, cease to be seen. And so do the parts of the Sun
by the interposition of the Moon, Mercury or Venus.

Newton’s authority and near-saintly status in England is best illustrated by


the famous epitaph that Alexander Pope wrote for him:
Nature and Nature’s laws lay hid in night: God said,“Let Newton be!” and all was
light.

Young was, of course, very much aware of the weight of Newton’s words
and knew that his results would face a highly critical reception. He hence
began his report very deferentially (Young, 1800):
Young’s interference experiment 279

Ever since the publication of Sir Isaac Newton’s incomparable writings, his doc-
trines of the emanation of particles of light from lucid substances, and of the
formal pre-existence of coloured rays in white light, have been almost univer-
sally admitted in this country, and but little opposed in others.

But he then proceeds to point out a remarkable analogy between light and
acoustic phenomena, effectively using Newton’s own observations against
him:
The phenomena of the colours of thin plates require, in the Newtonian system,
a very complicated supposition, of an ether, anticipating by its motion the
velocity of the corpuscles of light, and thus producing the fits of transmission
and reflection; and even this supposition does not much assist the explanation.
It appears, from the accurate analysis of the phenomena which Newton has
given, and which has by no means been superseded by any later observations,
that the same color recurs whenever the thickness answers to the terms of an
arithmetical progression. Now this is precisely similar to the production of the
same sound, by means of an uniform blast, from organ-pipes which are differ-
ent multiples of the same length.

It was this analogy, between what we nowadays call Newton rings and stand-
ing acoustic waves, that first made Young doubt the corpuscular theory of
light. By assuming that light is indeed a wave phenomenon, and knowing
the speed of light, as first measured in the 17th century by the Danish astron-
omer Ole Rømer, Young went on to determine the wavelength of different
colors from the location of Newton rings produced by thin plates
(Young, 1802b).
By challenging the established particle theory of light, Young was able to
make a number of important discoveries. One major contribution was his sug-
gestion that light undergoes a phase jump when it is reflected from a medium
denser than that in which it is traveling. In Young (1802a) he describes an
experiment in which the refractive index of the medium underneath a thin
plate is varied, and the resulting change in the fringe pattern clearly demon-
strates that such a jump indeed occurs. His next major step toward a convinc-
ing wave theory of light was the recognition, in 1801, of the principle of
interference. As he would recall a few years later (Young, 1804b):
It was in May 1801 that I discovered, by reflecting on the beautiful experiments
of Newton, a law which appears to me to account for a greater variety of inter-
esting phenomena than any other optical principle that has yet been made
known. I shall endeavour to explain this law by a comparison. Suppose a num-
ber of equal waves of water to move upon the surface of a stagnant lake, with a
certain constant velocity, and to enter a narrow channel leading out of the lake.
Suppose then another similar cause to have excited another equal series of
280 Greg Gbur and Taco D. Visser

waves, which arrive at the same channel, with the same velocity, and at the
same time with the first. Neither series of waves will destroy the other, but their
effects will be combined: if they enter the channel in such a manner that the
elevations of one series coincide with those of the other, they must together
produce a series of greater joint elevations; but if the elevations of one series are
so situated as to correspond to the depressions of the other, they must exactly
fill up those depressions, and the surface of the water must remain smooth; at
least I can discover no alternative, either from theory or from experiment. Now, I
maintain that similar effects take place whenever two portions of light are thus
mixed; and this I call the general law of the interference of light.

Young’s research on the nature of light culminated with his experimentum


crucis, now known as Young’s interference experiment. In November 1803
the report of his discovery was read to the Royal Society. Realizing that
he had now obtained irrefutable proof of the wave nature of light, he proudly
announced that (Young, 1804a):
In making some experiments on the fringes of colours accompanying
shadows, I have found so simple and so demonstrative a proof of the general
law of the interference of two portions of light, which I have already
endeavoured to establish, that I think it right to lay before the Royal Society,
a short statement of the facts which appear to me so decisive.

He continued his paper with a description of the experiment that is notably


different from modern textbook descriptions. Rather than using two pin-
holes or a pair of parallel slits, Young placed a thin card (1/30 in. thick)
in a narrow beam of sunlight and observed the resulting fringe pattern on
a distant wall. The fringes were of different colors, except the center one
which was white. If the light on one side of the card was intercepted, the
fringes disappeared, showing that they “were the joint effects of the portions of
light passing on each side of the slip card…” and that therefore it is interference
that produces them. Furthermore, by comparing the positions of different
fringes of the same color, he showed that they disappear whenever the dif-
ference between the path lengths of the two portions of light is a multiple of
a certain constant. This constant he found to be around 0.0000128 in.; in
modern-day language this corresponds to a very accurate estimation of
the wavelength of visible light of around 645 nm.
Young was acutely aware that his conclusions were at odds with the
Newtonian theory of light, and that they would likely stir controversy.
He therefore requested that the adherents of the corpuscular theory
…would do well to endeavour to imagine any thing like an explanation of
these experiments, derived from their own doctrines; and, if they fail in the
attempt, to refrain at least from idle declamations against a system which is
founded on the accuracy of its application to all these facts, and to a thousand
others of a similar nature.
Young’s interference experiment 281

This advice was not particularly well heeded, to say the least, and an anon-
ymous ad hominem campaign was launched in the Edinburgh Review
(Brougham, 1803). It later turned out that it was Henry Peter Brougham
(an ardent disciple of Newton, who would be made a Fellow of the
Royal Society in 1803) who wrote
As this paper contains nothing which deserves the name, either of experiment
or discovery, and as it is in fact destitute of every species of merit, we should
have allowed it to pass among the multitude of those articles which must
always find admittance into the collections of a Society which is pledged to
publish two or three volumes every year.

and
…it teaches no truth, reconciles no contradictions, arranges no anomalous
facts, suggests no new experiments and leads to new inquiries. It has not even
the pitiful merit of affording an agreeable play to the fancy.

The attacks, though in fact largely based on errors and misunderstandings,


definitely wounded Young’s pride. He responded by publishing his own
short pamphlet of rebuttals in 1804, with the title “A Reply to the
Animadversions of the Edinburgh Reviewers” (Young, 1804b). From the
very first words, it is clear that the argument between the men had gone
far beyond a simple scientific dispute:
A Man who has a proper regard for the dignity of his own character, although
his sensibility may sometimes be awakened by the unjust attacks of interested
malevolence, will esteem it in general more advisable to bear, in silence, the
temporary effects of a short lived injury, than to suffer his own pursuits to
be interrupted, in making an effort to repel the invective, and to punish the
aggressor. But it is possible that art and malice may be so insidiously combined,
as to give to the grossest misrepresentations the semblance of justice and
candor

Although Young bravely defended his scientific ideas, he seemed worn


down by the controversies that had dogged his scientific efforts from the
beginning, and announced his intention to return to medicine:
With this work, my pursuit of general science will terminate: henceforwards I
have resolved to confine my studies and my pen to medical subjects only.
For the talents which God has not given me, I am not responsible, but those
which I possess, I have hitherto cultivated and employed as diligently as my
opportunities have allowed me to do; and I shall continue to apply them with
assiduity, and in tranquility, to that profession which has constantly been the
ultimate object of all my labours.

Indeed, after the extremely hostile reception of his wave theory, Young
turned away from optical research for several years. He also gave up his
282 Greg Gbur and Taco D. Visser

position of Professor because his lectures at the Royal Institution were not
very popular. As he would later write about himself (Hilts, 1978),
…his compressed and laconic style being more adapted for the study of a man
of science than for the amusement of a lady of fashion.

This remark was also a jab at his colleague Humphry Davy, who excelled at
entertaining precisely such ladies in the Royal Institution’s audience. As
Young had earned a large inheritance from his supportive uncle, he could
afford to live without a salaried position.
Ironically, even though today the results of Young’s experiment are
often presented as a classical example of a paradigm shift, it was the doubts
of the prominent French scientist Simeon Denis Poisson that led to their
acceptance. Poisson noticed that a consequence of the wave theory, as for-
mulated in mathematical terms by Auguste Fresnel, would be the existence
of a bright spot at the center of the shadow of a circular obstacle. Since this
had never been observed, Poisson thought that this disproved the theory of
Young and Fresnel. François Arago, however, took up the experimental chal-
lenge and in 1818 succeeded in observing the predicted effect (Arago, 1819),
which is now referred to as the Poisson-Arago spot. The wave nature of light
only became widely accepted after Arago published his results confirming
Fresnel’s work.
Despite his professional setbacks, Young did not remain idle. After his
resignation, he spent the next 5 years compiling the 60 lectures he gave dur-
ing his 2-year tenure, and he oversaw their publication under the title
A Course of Lectures on Natural Philosophy and the Mechanical Arts (Young,
1845). This magnum opus, which greatly resembles an encyclopedia, covers
an enormous range of subjects. In Part I the laws of mechanics, the history
of writing techniques, the theory of perspective, the design of bridges, and
various “timekeepers,” i.e., clocks and watches, are among the many topics
that are described in great detail. In Lecture XIII, entitled On passive strength
and friction, Young introduced a measure of elasticity which is still widely
used and today is called Young’s modulus. Part II of his course is concerned
with fluids and treats capillary action, surface tension, contact angles, and
hydraulic machines. It also deals with sound and musical instruments.
Most significantly, it contains some of his aforementioned work on optics.
In this volume Young also introduced the most familiar version of his
famed experiment, using “two very small holes or slits.” Part III deals
with astronomy, the solar system, gravitation, geography, and the tides.
Young’s interference experiment 283

A common thread in all his presentations is that Young preferred qualita-


tive arguments over mathematical reasoning. As noted before, the mathe-
matical treatment of the two-slit experiment is due to Fresnel rather than
Young himself.
After the publication of his lectures Young concentrated mostly on his
medical and linguistic researches. In those days there was an enormous inter-
est in the ancient culture of Egypt, caused by Napoleon’s invasion of that
country in 1798. Along with his army, Napoleon had brought with him some
200 scientists, draftsmen, and architects. The wealth of manuscripts, objects,
and reports that they collected would later be cataloged in the celebrated
book series Description de L’Egypte. One major find of the French excavations
was the famous Rosetta stone. On it the same text, albeit paraphrased, appears
in Egyptian hieroglyphic and demotic writing, as well as in Greek writing.
Although the inscriptions were badly damaged, the Rosetta stone offered
hope to decipher the mysterious ancient Egyptian language which had frus-
trated scholars for generations. In principle, great progress could be made by
comparing the accessible Greek passages to the corresponding passages in the
dead Egyptian language. Young, who had a deep knowledge of ancient
Greek, began to try his hand at this project in 1814. He was the first to suggest
that a cartouche, an oval form that surrounds a series of hieroglyphs, indicates
a proper name. He also discovered the different ways in which plurals are
formed. As so often before, his work on the ancient language of Egypt too
would lead to a long-running dispute. This time it was an argument over
priorities with his French rival Jean–François Champollion.
Young also found other ways to broaden the knowledge of society as
a whole. From 1816 till 1823 he contributed numerous articles to the
Encyclopaedia Britannica. In them he covers an astonishing variety of topics,
including Egyptian culture, the ancient city of Herculaneum, languages,
the tides, hydrodynamics, as well as biographies of prominent scientists
such as Lagrange and Fermat. Possibly due to the hostility with which his
earlier work had been met, all these articles were published anonymously.
During the same period Young also served on several government boards.
One of these boards investigated suggestions to increase the rigidity of navy
vessels, while another one was concerned with the design of a pendulum to
standardize the definition of the second as a unit of time.
Although no longer actively working on optical problems, Young
did keep up with the many publications on light polarization that were
being written by scientists like Malus, Brewster, Biot, Arago, and Fresnel.
284 Greg Gbur and Taco D. Visser

Their discoveries were not at odds with Young’s undulatory theory, but they
could also not be readily explained by it. All this changed in 1816 when Arago
and Gay–Lussac visited Young at his home. The discussion turned to the
recent work by Fresnel by which Young was less than impressed, claiming
that he himself had already done similar work. The visitors did not agree,
and Arago offers us a rare glimpse into Young’s marriage (Arago, 1835):
…Mrs Young rose up suddenly and left the room. We immediately offered our
most urgent apologies to her husband, when Mrs Young returned, with an
enormous quarto under her arm. It was the first volume of the Natural
Philosophy [i.e., Young’s collected lectures] . She placed it on the table, opened
it up without saying a word, at page 787, and pointed with her finger to a
figure where the curved lines of the diffracted bands, on which the discussion
turned, appeared theoretically established. a

During the same visit Arago also described his own experiments in which,
using a two-slit setup, he had found the surprising result that rays of light
polarized in orthogonal planes will not interfere with each other (Arago,
1819; Arago & Fresnel, 1819). This set Young on an entirely new path
of thinking. A few months later he wrote to Arago (Peacock, 1855b) that
his result could be explained by assuming that light waves do not behave
as “the undulations of sound, consisting simply of the direct and retrograde motions
of their particles in the direction of their radius …,” but rather as “a transverse vibra-
tion, …and this is polarization.”
The suggestion that light is a transverse rather than a longitudinal wave
greatly propelled physical optics. It not only explained Arago’s observations,
but it also offered a direct way to derive Malus’ law. It also opened up the
analysis of crystalline structures in terms that present-day readers would be
familiar with, by assuming a refractive index that depends on the plane of
polarization.
Even as he grew older, Young continued to find new topics to research,
and new problems to solve. In England of the Industrial Revolution, mor-
tality rates in the countryside were significantly lower than those in the
crowded cities. Moreover, these rates were changing quickly due to
improved sanitation, the introduction of vaccination, and better access to
clean drinking water. These circumstances made it very difficult for insur-
ance companies to estimate the costs of annuities. In 1824 the Palladium
Insurance Company enlisted Young’s help and appointed him to the

a
The authors first learned of this story from the late Emil Wolf. After telling it he cried out “What a
woman!”
Young’s interference experiment 285

position of Inspector of Calculations, with the task of analyzing life expec-


tancies. This was a completely new field for him, and he approached it with
his usual thoroughness. He began by comparing different tables of mortality
rates, going all the way back to Roman times. The result of his studies was an
empirical formula for “the decrement of life” (Young, 1826), thus making
him one of the founders of actuarial studies.
Toward the end of his life Young composed an autobiographical sketch
(Hilts, 1978) in which, with disarming honesty, he explained his limited
social skills by describing himself as
…generally cheerful, and sometimes playful, though seldom very eloquent or
very entertaining in conversation. He felt some inconvenience in society from
being a little short sighted, and he used to attribute in part to this circumstance
the mistakes which he sometimes made respecting the impression produced
by what he said or did, on the feelings of others.

He also added that he lived


…neither fearing his last day, nor desiring it.

While completing his work on an Egyptian dictionary, the last day of this
most remarkable man came in 1829 when, after a brief illness, he died in
London at the age of 55. A postmortem examination showed extensive
hardening of the aorta. His friend and biographer Hudson Gurney 1831
put this down to Young’s “unwearied and incessant labor of the mind from the
earliest days of infancy.”
François Arago read a eulogy to the French Academie des Sciences, of
which Young was a foreign member. In it Arago referred to Young’s work
on interference as (Arago, 1835)
…that fine discovery which will render his name imperishable…

and then asked


…who would have imagined that we should thence come to suppose that
darkness could be engendered by adding light to light?

In the last part of the eulogy Arago laments the hostile reception of Young’s
ideas and describes how discouraging that must have been for him. Noting
that an initial rejection had been the unfortunate fate of many great scientists
who were later vindicated, he ends poetically with this comforting
suggestion:
Let us try to persuade ourselves that in the dungeons of the inquisitors a
friendly voice had caused Galileo to hear some of the delightful expressions
which posterity has kept sacred for his memory.
286 Greg Gbur and Taco D. Visser

Five years after Young’s death a white marble tablet, with a profile medal-
lion, was revealed in the chapel of St Andrew in Westminster Abbey. The
inscription reads:
Sacred to the memory of Thomas Young M.D. Fellow and Foreign Secretary
of the Royal Society Member of the National Institute of France. A man alike
eminent in almost every department of human learning. Who, equally distin-
guished in the most abstruse investigations of letters and of science, first
established the undulatory theory of light and first penetrated the obscurity
which had veiled for ages the hieroglyphicks of Egypt. Endeared to his friends
by his domestic virtues, honoured by the World for his unrivalled acquirements,
he died in the hopes of the Resurrection of the just. Born at Milverton in
Somersetshire June 13th 1773, died in Park Square London May 10th 1829, in
the 56th year of his age.

A brief sketch as presented here cannot do full justice to the complicated


historical events that we have touched upon. More biographical information
about Thomas Young can be found in Gurney (1831), Peacock (1855a),
Hilts (1978), Mollon (2002), and Robinson (2005). An authoritative over-
view of the history of electromagnetism and the concept of the ether was
presented by Whittaker (1951).

3. Young’s interference experiment


Young’s setup can take on three different forms. In its simplest version
a beam is divided into two parts by an obstacle, and the two parts are then
recombined on an observation screen. This is the design that Young himself
first reported in Young (1804a). In the two other forms the beam is passed
through an opaque screen with either two narrow, parallel slits, or two pin-
holes. This screen is positioned parallel to an observation screen on which
interference fringes are formed. A closely related form of wavefront division
and subsequent recombination is achieved in a Mach–Zehnder interferom-
eter. This approach will be discussed in Section 4.
Let us consider the pinhole configuration, as sketched in Fig. 1; the
mathematical description of all three configurations is largely similar. A plane
scalar wave of monochromatic light is normally incident on a screen A that
contains two identical pinholes. The screen is positioned in the plane z ¼ 0.
The light that emanates from the pinholes forms an interference pattern on
a second, parallel screen B that is a distance Δz away. Let the scalar field
at each pinhole be denoted by U(rj, ω), with j ¼ 1, 2. Here rj is the position
of pinhole j and ω is the angular frequency of the wave. According to
Huygens’ Principle, each pinhole will emit a spherical wave. The total field
Young’s interference experiment 287

P
R1
r1
R2
d z

r2

fringes
incident plane wave

screen A screen B
Fig. 1 The geometry of Young’s interference experiment with two pinholes. A mono-
chromatic plane wave is normally incident on screen A that contains two identical pin-
holes. An interference pattern is formed on a second, parallel screen B. For simplicity it is
assumed that the refractive index in between the two screens is unity.

at a point P on the observation screen is then the superposition of these


two contributions, i.e.,

UðP, ωÞ ¼ K 1 Uðr1 , ωÞeikR1 + K 2 Uðr2 , ωÞeikR2 , (1)


where the wavenumber k ¼ 2π/λ ¼ ω/c, with λ the wavelength and c the
speed of light, and Rj the distance from pinhole j to the point P. The two
propagators Kj are given by the expression (Born & Wolf, 1999)
idA
Kj ¼  , (2)
λRj
where dA is the area of each pinhole. The spectral density (or “intensity at
frequency ω”) on the observation screen is defined as

SðP, ωÞ ≡ jUðP, ωÞj2 : (3)


From Eq. (1) we thus find the spectral interference law

SðP,ωÞ ¼ jK1 j2 Sðr1 ,ωÞ + jK2 j2 Sðr2 , ωÞ


h ∗
i (4)
+2 Re K1 K2∗ Uðr1 , ωÞU ðr2 , ωÞeikðR1 R2 Þ :

We can simplify this expression by making two assumptions. First, when the
separation Δz between the two screens is much larger than the distance
d between the two pinholes, as is usually the case, then for a point
P close to the z axis we have to a good approximation that
288 Greg Gbur and Taco D. Visser

jK 1 j2  jK 2 j2  K 1 K *2 ¼ K 2 : (5)
Second, we assume that the spectral density of the incident field at the two
pinholes is the same. Since we are dealing with a plane wave, that implies that

Sðr1 , ωÞ ¼ Sðr2 , ωÞ ¼ Uðr2 , ωÞU * ðr2 , ωÞ ≡ SðωÞ: (6)


On making use of these two assumptions in Eq. (4) we obtain the formula

SðP, ωÞ ¼ 2K 2 SðωÞf1 + cos ½kðR1  R2 Þg: (7)


Let the positions of the two pinholes be
r1 ¼ ðd=2, 0, 0Þ, r2 ¼ ðd=2, 0, 0Þ: (8)
For a point P(x,y,Δz) in the vicinity of the z axis, we then have, approximately
R1  R2  xd=Δz: (9)
This then leads to

SðP, ωÞ ¼ 2K 2 SðωÞ½1 + cos ðkdx=ΔzÞ: (10)


The spectral density S(P, ω) is seen to attain a maximum at positions x for
which
2mπΔz
x¼ , ðm ¼ 0,  1,  2, …Þ: (11)
kd
In between these maxima the spectral density will be zero when
ð2m + 1ÞπΔz
x¼ : (12)
kd
These successive bright and dark fringes are (approximately) invariant along
the y direction. The separation distance Λ between two adjacent bright
fringes equals
2πΔz
Λ¼ : (13)
kd
In an actual experiment the distances d and Δz are both known. Therefore
the wavenumber k, and hence the wavelength λ, can be determined from
measuring the fringe spacing Λ. Equivalently, because the speed of light
is known, the angular frequency ω can be determined from the interference
pattern.
Young’s interference experiment 289

Young’s experiment therefore does not just demonstrate the wave


character of light by showing that light undergoes interference, but it
also provides a direct way to measure its wavelength and frequency.
Moreover, Arago made the observation, discussed in the previous section,
that when the two apertures are covered with orthogonally oriented polar-
izers the interference fringes disappear. This reveals that the light vibrations are
transverse rather than longitudinal. We may introduce this phenomena into
our analysis of Young’s experiment by replacing the scalar field U(r, ω)
illuminating each pinhole by an electric field vector E(r, ω) ¼ e U(r, ω),
where e is a complex unit Jones vector (Brosseau, 1998) describing the
two-dimensional state of polarization. The spectral density of the field on
the observation screen is now given as

SðP, ωÞ ≡ jEðP, ωÞj2 , (14)


and the interference law (4) takes on the form

SðP, ωÞ ¼ jK1 j2 Sðr


 1 , ωÞ∗ + jK2 j∗ Sðr2 , ωÞ
2
 (15)
+2 Re K1 K2 ðe1  e2 ÞUðr1 , ωÞU*ðr2 ,ωÞeikðR1 R2 Þ :
If we assume that the spectral density of the incident field at the two pinholes
is the same, but that the state of polarization may be different, and again
consider observation points close to the z axis, this law simplifies to
 
SðP, ωÞ ¼ 2K 2 SðωÞ 1 + Re½ðe1  e∗2 Þeikdx=Δz  : (16)
This expression clearly shows that the interference term, the last term on the
right-hand side of Eq. (16), vanishes when the fields at the two pinholes are
orthogonally polarized, i.e., when e1  e∗2 ¼ 0.
The interference pattern that is produced by nonuniformly polarized
beams, for example, beams with azimuthal or radial polarization, was inves-
tigated in Li et al. (2012).

4. Weak light
Though Young’s experiment is perhaps most famous for demonstrat-
ing the wave properties of light, it has also played a crucial role in elucidating
the quantum mechanical wave-particle duality of photons and matter.
As the 19th century drew to a close, it seemed that the wave theory of
light was on very solid footing, especially with the recognition of light as an
290 Greg Gbur and Taco D. Visser

electromagnetic wave, as first postulated by Maxwell in the 1860s. Though


there were several experimental observations that could not be readily
explained by a simple wave theory, such as blackbody radiation, these
seemed at the time to be minor puzzles that would eventually be solved
and fit into the broader wave picture.
This view changed in 1905, Einstein’s Annus Mirabilis, in which he pub-
lished four papers in Annalen der Physik. Each of these papers would lay the
foundation for a new field of physics. In one of them he formulated his
Special Theory of Relativity, and in another he suggested the equivalence
between mass and energy. A third explained Brownian motion as the result
of atomic collisions. In a fourth study, called “On a heuristic viewpoint con-
cerning the production and transformation of light” (Einstein, 1905),
Einstein sought to explain the somewhat perplexing observations of the
photoelectric effect, first reported by Heinrich Hertz in 1887, in which light
can, under the right conditions, eject electrons from the surface of a metal
plate. In particular, it was noted that electrons were only ejected from the
metal once a critical frequency threshold of light was exceeded, and that
the intensity of light only increased the number of electrons released, but
not their energy. The wave theory of light could not explain the threshold,
and it predicted that the energy of the released electrons should increase with
intensity of the illuminating light field.
Einstein solved this puzzle by arguing that light acts both as a particle and as
a wave. The energy of an individual particle of light, now known as a photon,
is proportional to the frequency of the light wave, and individual photons
knock individual electrons off of the metal. However, since electrons are
bound to the metal by electromagnetic forces, the energy or frequency of
the photon must exceed a certain threshold before electrons are ejected.
Einstein’s insight could be said to mark the true beginning of quantum physics,
and he eventually won the 1921 Nobel Prize in Physics for his work on the
photoelectric effect; his work on Relativity was not recognized by the Nobel
committee and was still controversial at that time. But in the immediate after-
math of his revelation, it was unclear how to reconcile the new particle prop-
erties of light with the long-observed wave properties. In particular, it was
unclear how a set of discrete particles could produce wave-like interference
patterns.
One early hypothesis came from the famed English physicist J.J. Thomson
in 1907 (Thomson, 1906-1908). Thomson was studying the ionization of
gases through the use of ultraviolet light, a process similar to the photo-
electric effect: instead of visible light ejecting electrons off of a metal,
Young’s interference experiment 291

ultraviolet light ejects electrons off of atoms. In trying to reconcile the


wave-particle duality problem, he suggested that the particle effects are due
to the fact that the actual energy of a wave is distributed unevenly across
the wavefront, little “bumps” of energy that are tightly packed when the light
intensity is high, and widely spaced when the intensity is low. As he wrote,
The existence of the structure in the wave front implies that the ether through
which the light is traveling has also a structure. We shall for the sake of definite-
ness make a special assumption as to the nature of this structure, and suppose
that the ether has disseminated through it discrete lines of electric force and
that these are in a state of tension and that light consists of transverse vibra-
€ntgen rays of pulses, travelling along these lines. Thus the energy trav-
tions, Ro
elling outwards with the wave is not spread uniformly over the wave front, but
is concentrated on those parts of the front where the pulses are travelling along
the lines of force; these parts correspond to the bright specks, the rest to the
dark ground. There will not necessarily be a speck at each place where the lines
of force cut the front, but only at those places where the lines of force happen
to be in vibration at the instant under consideration. The energy of the wave is
thus collected into isolated regions, these regions being the portions of the
lines of force occupied by the pulses of wave motion.

Thomson’s model was not particularly well-formed, but the natural impli-
cation of it is that the interference pattern produced in Young’s experiment
comes from the interaction of many of these little pulses of energy. But what
happens when only a very small number of them are interacting with the
two pinholes at a given time? Thomson suggested that the interference pat-
tern must look very different, because there simply are not enough bumps
around to interfere.
The first attempt to explore this question experimentally was through the
use of Young’s experiment again, and was undertaken by Thomson’s student
Taylor. In his “Interference fringes with feeble light” (Taylor, 1909) he
describes the creation of a low intensity source by having a gas flame illumi-
nating a narrow slit and then placing smoked glass screens in the path of the
emanating light. Each individual screen would absorb a significant fraction of
light passing through them, and screens which were more heavily smoked
would absorb significantly more light than those less smoked. By using mul-
tiple screens, an arbitrarily high amount of attenuation could be achieved,
even to the level where only single photons were illuminating a sewing nee-
dle, acting as the Young apparatus, at any time. A photographic image was
taken of the ensuing diffraction pattern.
For the experiment Taylor took five images of these patterns, for various
screens placed behind the slit. He calibrated his experiment first by taking
292 Greg Gbur and Taco D. Visser

photographs of the light source without any screens present, and determined
how much time was needed for the plates to darken by the same amount;
repeating this experiment with the screens present, then, one could see
how Young’s interference pattern changed when the light intensity was
reduced to the level where a only small number of energy “pulses” were
going through at once.
The experiment required quite a lot of patience. In order to get sufficient
development of the photograph of the dimmest light source, Taylor had to
run the exposure for 3 months. But his patience paid off, and he found an
unexpected result; in his own words
In no case was there any diminution in the sharpness of the pattern although
the plates did not all reach the standard blackness of the first photograph.

In other words, the interference pattern remained the same, no matter how
much the light intensity was reduced. Taylor and Thomson seem to have
initially interpreted this result as indicating that they had not used a dim
enough light source, and they estimated the maximum energy of a single
photon from their arrangement. Their estimate, based on an incorrect model
of quantum physics, ended up being much smaller than the true energy of a
photon.
What happens when we ensure that, on average, just a single photon is
present in the setup? It turns out that although the photons are detected as
individual particles that arrive at seemingly random positions on the detector
area, they travel like waves. The latter conclusion follows from two facts: if
one sums over a large number of detection events, the grainy distribution
gradually becomes smoother and eventually evolves into the classical inter-
ference pattern as first observed by Young. Also, if one of the two apertures
is closed and each photon is forced to pass through the other aperture, the
interference pattern does not emerge. In the words of Feynman, Leighton,
and Sands (1963), this is
a phenomenon which is impossible, absolutely impossible, to explain in any
classical way, and which has in it the heart of quantum mechanics.

Surprisingly, the fringe pattern ensuing in single-photon interference exper-


iments can be observed with the naked eye (Parker, 1971). Over the years
such experiments have become much easier to carry out. Using a storage
oscilloscope, the individual detection spots can be accumulated to show
the buildup of smooth interference fringes (Rueckner & Titcomb, 1996).
In later experiments CCD cameras were employed to record the arrival
Young’s interference experiment 293

of individual photons (Weis & Wynands, 2003). A single camera frame


shows an apparent random distribution of photon impact points. The inte-
gration of a series of frames reveals what might be called the transition from
particles to waves as the classical smooth fringe pattern gradually takes shape.
A very elegant and convincing demonstration of this buildup of the
fringe pattern, and other aspects of the wave-particle duality of light, was
developed by Dimitrova and Weis (2008). Fig. 2 shows their use of a tradi-
tional Young’s experiment to show the creation of fringes. To explore this
and other effects further, they used a Mach–Zehnder interferometer rather
than a double-slit configuration (see Fig. 3). The large spatial separation of
the two interfering beams permits several useful manipulations that cannot
be done with double slits, such as adjusting the path difference and the rel-
ative angle of the beams. Furthermore, it is easy to block an arm of the inter-
ferometer or tag photons by changing their polarization state.
In the setup two beams, one with a variable attenuation and the other
with a high intensity, are both split and then recombined. The superposition
of the two strong beams is detected with a photodiode, that of the weak
beams with a photomultiplier. The arrival of single photons at the photo-
multiplier can be either visualized on a oscilloscope or rendered acoustically

Fig. 2 The transition from particles to waves visualized. Detection of light diffracted by
a double slit on a photon-by-photon basis using a CCD camera. A single frame shows
an apparently random distribution of photon impact points. Their integration reveals
the classical fringe pattern. Reproduced from Dimitrova, T.L., & Weis, A. (2008). The
wave-particle duality of light: A demonstration experiment. American Journal of Physics,
76(2), 137–142, with the permission of the American Association of Physics Teachers.
294 Greg Gbur and Taco D. Visser

attenuator MZI
stack beam
blocker oscilloscope

1 M
A
2 loud-
laser pointer BS speaker

trigger
A
B
lens PM
M
screen (P1)

BS lens PD
PZT
screen (P2)
lock
PI
off
scan
ramp generator

Fig. 3 Setup for the simultaneous demonstration of the wave and particle nature of
light. The strong beam is indicated by solid lines, the weak beam by dashed lines.
BS, beam splitter; M, mirror; PM, photomultiplier; PD, photodiode; PI, feedback amplifier.
Reproduced from Dimitrova, T.L., & Weis, A. (2008). The wave-particle duality of light: A
demonstration experiment. American Journal of Physics, 76(2), 137–142, with the permis-
sion of the American Association of Physics Teachers.

by a loudspeaker. An alternating voltage across a piezo element on which a


mirror is mounted produces a harmonic variation of the path difference
between the two arms. Averaging more and more time traces of the arrival
of individual photons at the photomultiplier produces a signal that eventu-
ally becomes similar to the photodiode output caused by the two
unattenuated beams (see Fig. 4). This experiment also allows the observation
of self-interference of photons. This is achieved by keeping the moveable
mirror fixed at a position that produces a minimum photon count rate at
the photomultiplier, as shown in Fig. 5A. If now the upper arm A is
blocked, the photon count rate increases significantly, see Fig. 5B. This
demonstrates that if the photon can take either path B or path A, it has a
lower probability to be detected. By forcing it to take one specific path,
the chances of it being detected are greatly increased. This can only be
explained by the nonclassical idea that in the unblocked configuration
the photon somehow traverses both paths simultaneously and then, in
the words of Dirac, “interferes with itself.”
In 1963, Mandel and Magyar decided to put Dirac’s statement to a seri-
ous test (Magyar & Mandel, 1963). They combined the light from two inde-
pendent ruby masers and formed an image of the interference pattern
produced. In essence, they performed Young’s experiment with indepen-
dent light sources. According to the classical wave theory, these two beams
Young’s interference experiment 295

photomultiplier

single sweep

4 averages

16 averages

64 averages

128 averages

A photodiode
B

PM
pi B
ez PD
o

Fig. 4 Simultaneous demonstration of the wave and particle aspects of light. The bot-
tom trace shows the wave-like intensity distribution recorded by the photodiode. The
top trace shows the photon counts of the photomultiplier. By averaging many traces,
the signal from the photomultiplier gradually becomes smoother and approaches that
of the photodiode. Reproduced from Dimitrova, T. L., & Weis, A. (2008). The wave-particle
duality of light: A demonstration experiment. American Journal of Physics, 76(2), 137–142,
with the permission of the American Association of Physics Teachers.

will not produce a pattern on average because of random fluctuations between


them; however, over a short time period, the two fields will be in phase and
produce interference. With a 40 ns exposure, Mandel and Magyar photo-
graphed clear interference fringes. They concluded:
296 Greg Gbur and Taco D. Visser

Fig. 5 Demonstrating self-interference of the photon. The moveable mirror is set to a


minimum of the fringe pattern. (A) Photomultiplier pulses recorded when the photons
are offered the alternative of two paths; only a few photons are detected.
(B) Photomultiplier pulses when the photons are forced to take a well-defined path;
the number of registered photons increases. Reproduced from Dimitrova, T. L., &
Weis, A. (2008). The wave-particle duality of light: A demonstration experiment.
American Journal of Physics, 76(2), 137–142, with the permission of the American
Association of Physics Teachers.

It seems, therefore, that, as with microwaves, these interference effects are


describable in completely classical terms.

Because the interference was between light beams from independent lasers,
the result seemed to imply that, contrary to Dirac’s view, different photons
can interfere with each other. This was not the end of the investigation,
however. In 1967, Pfleegor and Mandel repeated the experiment with single
photons (Pfleegor & Mandel, 1967). Even though only one photon was pre-
sent in the apparatus at any time with high probability, with the source of the
photon unknown, the interference pattern could still be created. They
concluded
Surprising as it might seem, the statement of Dirac quoted in the introduction
appears to be as appropriate in the context of this experiment as under the
more usual conditions of interferometry.

5. The wave-particle duality for massive particles


In a famous series of experiments by Max von Laue and coworkers
carried out around 1910, a polychromatic beam of X-rays was passed
through a crystal (Friedrich, Knipping, & von Laue, 1912). They observed
monochromatic diffraction spots (nowadays called von Laue spots) that can
be described by the scattering of short-wavelength radiation (λ  1 Å) by
Young’s interference experiment 297

regularly-spaced point scatterers. These experiments showed the wave char-


acter of the X-rays that had been discovered a few years earlier by R€ ontgen.
But the experiments also revealed the lattice structure of crystals. Swiftly
after this work, two entirely new branches of physics were created: X-ray
spectroscopy and X-ray crystallography (Eckert, 2012). The use of crystals
to diffract matter waves with a wavelength much shorter than visible light
proved to be crucial for later quantum mechanical investigations.
In 1924 Louis de Broglie hypothesized that the dual wave-particle nature
of light might also hold for material particles such as electrons (de Broglie,
2004). More specifically, he suggested the equivalence of relativistic energy,
which can expressed in terms of a corresponding mass, and the energy of
radiation quanta. As he wrote
One may imagine that, by cause of a meta law of Nature, to each portion of
energy with a proper mass m, one may associate a periodic phenomenon
of frequency ν, such that one finds

hν ¼ mc 2 : (17)

Here h denotes Planck’s constant, and c is the speed of light. The first con-
firmation of de Broglie’s conjecture was obtained by Davisson and Germer
in 1927 (Davisson & Gerner, 1927). In their experiments a beam of electrons
was scattered off a Ni crystal. It was known that a plot of the intensity of
X-rays reflected by a crystal in a fixed direction as a function of 1/λ should
show a series of equally spaced peaks. Each peak corresponds to a solution of
the Bragg condition for constructive interference. Davisson and Germer
found that the intensity of reflected electrons, as a function of the root of
their bombarding potential (i.e., their speed), shows a quite similar behavior.
This result indicated that the electrons are indeed wavelike, with a wave-
length, in agreement with Eq. (17), that is inversely proportional to their
velocity.
In the same year, the wave nature of electrons was verified independently
by G.P. Thomson and A. Reid (Thomson & Reid, 1927). G.P. Thomson
was the son of J.J. Thomson, the man who discovered the electron in 1897:
in essence, the father proved the particle properties of the electron, and the
son proved the wave properties of the electron. Passing a narrow beam of
electrons through a thin celluloid film and using a photographic plate as a
detector, Thomson and Reid observed a central spot (formed by undeflected
electrons) surrounded by a series of concentric rings. This clearly demon-
strated electron diffraction.
298 Greg Gbur and Taco D. Visser

Today the wave-particle duality remains just as mysterious as it was a


century ago. As Davisson would later write (Davisson, 1928)
It is all rather paradoxical and confusing. We must believe not only that there is
a certain sense in which rabbits are cats, but there is also a certain sense in
which cats are rabbits.

Paradoxical or not, the groundbreaking achievements of de Broglie were


soon recognized by the Nobel Prize committee, who awarded him the prize
in 1929. Davisson and Thompson would share the prize in 1937.
A later demonstration of electron diffraction by M€ ollenstedt and D€ ucker
(1956) was more akin to Taylor’s use of a thin needle to divide a stream of light
into two parts, as was described in the previous section. Such an approach is
“cleaner” than previous experiments because no assumptions about crystalline
structures and their interaction with particles are needed. M€ ollenstedt and
D€ ucker placed a gold-coated quartz filament in a beam of electrons. The gold
coating allows the thread to have a selectable positive voltage. By changing the
voltage the width and spacing of the ensuing fringe pattern can be adjusted.
Today such techniques are widely used in electron microscopy (Lichte, 2002).
In the early 1960s M€ ollenstedt suggested his student J€ onsson to use the
then-fledgling lithography technique to create thin foils with one, two, or
many parallel narrow slits to study electron diffraction (J€ onsson, 1961). The
observed fringe patterns were fully analogous to those obtained by the diffrac-
tion of light waves, but for a much smaller wavelength of 0.05 Å. Although a
great technical feat, it was later suggested (Brandt & Hirschi, 1974) that by that
time the wave-particle duality was so well established that J€ onsson’s result did
not get the attention that it perhaps deserved.
The gradual build-up of interference patterns that was discussed in the
previous section for single photons has also been observed for single elec-
trons. Merli, Missiroli, and Pozzi (1976) used an electron biprism to dem-
onstrate the statistical nature of fringe development, captured on a TV
monitor. Tonomura, Endo, Matsuda, Kawasaki, and Ezawa (1989) reported
images of interference fringes formed by 10, 100, 3000, and 20,000 individ-
ual electrons. In both these experiments the wave-particle duality is illus-
trated by the electrons being emitted as point-like particles by the source,
after which they interact as a de Broglie wave with the slits and then are
detected again as individual particles. Bach, Pope, Liou, and Batelaan
(2013) presented a double-slit setup with a moveable mask behind the slits.
With one aperture covered, a one-slit diffraction pattern was observed.
Moving the mask continuously from one slit to the other caused a gradual
transition to a two-slit pattern, followed by a return of a one-slit pattern.
Young’s interference experiment 299

A completely different form of electron interference occurs in the con-


text of geometric phases (Shapere & Wilczek, 1989). The phase of a physical
system is the sum of two contributions, a dynamic phase and a geometric
phase. Whereas the first is the result of the passage of time, the second is
accrued when the system is made to trace out a (usually closed) path.
Geometric phases can be divided into four groups. The system under con-
sideration can either be a classical system or a quantum system, and the tra-
jectory that is taken can be in real space or in some abstract parameter space.
For example, Foucault’s pendulum is a classical system that travels through
real space (von Bergmann & von Bergmann, 2007). The Pancharatnam
phase that appears in polarized light beams pertains to a classical system that
is moved across the parametric Poincare sphere (Brosseau, 1998). The Berry
phase describes quantum systems with a Hamiltonian that undergoes a cycli-
cal change of its parameters (Berry, 1984).
An example of the fourth variety is the Aharonov–Bohm effect
(Aharonov & Bohm, 1959). Conceived as an interferometric thought exper-
iment, it describes an electron beam that is split into two parts. Both parts
travel through a portion of space with a zero magnetic field but with a dif-
ferent vector potential before they are recombined to produce an interfer-
ence pattern. Since the vector potential appears in the Hamiltonian that
governs the time evolution of the quantum mechanical wavefunction, the
different vector potentials that the electrons encounter in the two arms give
rise to a phase difference. Varying the strength of the vector potential
changes this phase difference which manifests as a shift of the interference
fringes. It turns out that this phase difference can be expressed in terms of
the flux of the vector potential through the interferometer, making it a
geometric phase.
In classical physics the vector potential is seen, due to its gauge freedom,
as a mathematical tool without physical significance. As the Aharonov–
Bohm experiment shows, the situation in quantum mechanics is quite dif-
ferent. Even in the absence of an electric and a magnetic field, i.e., in the
absence of forces, the phase evolution of the electrons is governed by the
vector potential. In the thought experiment the electrons would pass on
both sides of a long solenoid. In each arm the orientation of the vector
potential is then different, and the magnetic field is approximately zero.
The first attempt to verify the effect (Chambers, 1960) was criticized for
not completely suppressing the magnetic field. In a much later study
Tonomura et al. (1986) the use of a superconducting medium around the
magnet ensured a zero magnetic field, and convincingly demonstrated
the reality of the Aharonov–Bohm effect.
300 Greg Gbur and Taco D. Visser

The Pancharatnam phase also occurs in the context of Young’s experi-


ment. Consider a time-harmonic, fully coherent, and polarized beam of
light that impinges onto a screen with two pinholes in it. If the incident
light has a uniform polarization the field’s polarization state on the observa-
tion screen is the same as at the pinholes. If, however, we allow for different
states of polarization at the openings, the situation changes considerably,
leading to a spatially varying, periodic polarization state across the second
screen. In moving the point of observation one therefore traces out a
closed path on the Poincare sphere. The resulting geometric phase was
predicted and experimentally verified by Hannonen, Partanen, Tervo,
Set€al€a, and Friberg (2019).
Ever since the discovery of the wave properties of electrons, there has
been a persistent effort to demonstrate that de Broglie’s hypothesis holds
for heavier and larger particles and aggregates of particles. The interference
properties of molecules were first demonstrated in 1930 by Estermann and
Stern in their “Diffraction of molecular beams” study (Estermann & Stern,
1930). They scattered beams of H2 and He atoms off an LiF crystal and
observed that the intensity distribution was consistent with that of waves sat-
isfying the de Broglie relation. A wave character was later confirmed for sev-
eral different kinds of particles. Single and double-slit diffraction of cold
neutrons, with a de Broglie wavelength λ ¼ 15  30 Å, was announced in
1988 by Zeilinger, G€ahler, Shull, Treimer, and Mampe (1988). The first atom
interferometer was reported in 1991 (Carnal & Mlynek, 1991), in which
Young’s double-slit experiment was used to produce interference fringes
with a beam of helium atoms with λ  1 Å. A Na2 molecular beam with
λ ¼ 0.11 Å diffracted by a grating was reported in 1995 (Chapman et al.,
1995). The first demonstration of antimatter interferometry, using single pos-
itrons, was performed with a Talbot–Lau setup in 2019 (Sala et al., 2019). The
advantage of such a configuration is that it can accept a spatially incoherent
beam from which, by the use of gratings, a coherent part is selected.
A natural question now ensues: what are the largest objects for which this
quantum-mechanical wave behavior can be observed? Clearly, this question
involves dealing with ever shorter de Broglie wavelengths. In 1994
(Sch€ollkopf & Toennies, 1994), interference of small van der Waals clusters,
like (H2)8, with a wavelength λ  1 Å, was demonstrated. In 1999, Arndt et al.
(1999) demonstrated the wave-particle duality of Buckminsterfullerene or
“Bucky balls,” i.e., C60 molecules. As the authors note, these are almost clas-
sical objects due to their many excited internal degrees of freedom. At the
temperature that was used in the experiment, each molecule emits on average
Young’s interference experiment 301

2 to 3 photons on its journey between the grating and the detector.


Nevertheless, clear interference fringes were observed corresponding to
λ  2.5 pm. As the authors would later write (Nairz, Arndt, & Zeilinger,
2002):
It is interesting to compare the de Broglie wavelength of the fullerene with its
actual size: The buckyball has a diameter of about 1 nm, which is 350 times
larger than its de Broglie wavelength. Our interference experiments clearly
show that the concept of the de Broglie wavelength is not merely academic
for objects with dimensions much larger than their wavelengths but can actu-
ally be demonstrated.

The same group later showed the wave character of C70 molecules with a
wavelength of 2 to 5 pm (Brezger et al., 2002). Exploring the wave-particle
duality for even larger molecules could in principle be done at lower tem-
peratures, thereby reducing the molecules’ speed and increasing their de
Broglie wavelength (Arnd, Aspelmeyer, & Zeilinger, 2002). It has even been
suggested that one day the wave character of bacteria and viruses might be
demonstrated (Geyer et al., 2016). In an impressive technical feat, a beam
consisting of a mixture of high-mass (>10,000 atomic mass units) molecular
compounds was studied in Fein et al. (2019). Using a mass spectrometer to
separate different kinds of molecules, the wave nature of molecules con-
sisting of 810 atoms, with a total of some 5000 protons, neutrons, and elec-
trons, was observed. The internal complexity, the number of vibrational
modes, and also the internal energy of each of these particles clearly distin-
guish them from the much simpler quantum objects that were used in earlier
matter wave experiments. In a follow-up study (Eibenberger, Gerlich,
Arndt, Mayor, & T€ uxen, 2019) molecules with a mass of 25,000 amu, con-
sisting of up to 2000 atoms, were shown to have a de Broglie wavelength of
53 fm, i.e., five orders of magnitude smaller than the diameter of the mol-
ecules themselves.
An intriguing version of the double-slit experiment on a molecular scale
was reported in Zhou, Perreault, Mukherjee, and Zare (2021). Deuterium
(D2) molecules were prepared in a quantum superposition of two orthogo-
nal orientations. When helium atoms scatter off such molecules, they expe-
rience both orientations at once. As a result, the scattered intensity is due to a
coherent sum of two processes. Whereas in the classic double-slit experi-
ment the particles pass through both slits in a superposition of trajectories,
here, in contrast, one might think of the D2 molecules as a single slit that
itself is in a superposition of positions. The observed angular scattering
counts were in good agreement with the theoretical prediction.
302 Greg Gbur and Taco D. Visser

Arguably the ultimate illustration of the weirdness of the wave-particle


duality is the famous thought experiment that was suggested by Wheeler and
Zurek (1983). Consider the Mach–Zehnder interferometer that is sketched
in Fig. 6. A single-photon pulse impinges on beam splitter BSin and proceeds
through the setup until a second beam splitter, BSout, recombines the light
from the two interfering arms. A moveable mirror in Path 1 introduces a
phase shift Φ between the arms. As this phase shift is varied, interference
appears as a modulation of the detection probabilities at D1 and D2, respec-
tively, as cos 2 ðΦ=2Þ and sin 2 ðΦ=2Þ. This is the result that one expects from
a wave. As Wheeler pointed out, “this is evidence that each arriving light quantum
has arrived by both routes.” Alternatively one might say that this path-
difference setup probes the wave character of the quantum.
Let us next consider the same configuration, but with the beam splitter
BSout now removed. In that case each detector D1 or D2 is associated with a
single path of the interferometer and, provided that true single-photon
pulses are used, either one detector clicks, or the other. This establishes that
the photon has traveled only one route. In other words, this particular
“which-way” setup probes the particle nature of the quantum. The two set-
tings, with and without BSout, support the complementarity viewpoint of
Bohr that the behavior of a quantum system is determined by the type of
measurement performed on it. In experiments where the choice between
the two settings is made long in advance, one could argue that the photon,
before entering the interferometer, has somehow received some “hidden
information” about the chosen configuration and then adjusts its behavior
accordingly. It is precisely this possibility that is ruled out in the thought

D1
Path 2
mirror
D2
BSout
Single-photon
input pulse

BSin Path 1 Φ
mirror
Fig. 6 Sketch of Wheeler’s delayed-choice experiment. The moveable mirror in Path 1
introduces a variable phase difference Φ between the two arms. The second beam split-
ter, BSout, can either be present or absent from the setup.
Young’s interference experiment 303

experiment that Wheeler suggested. He proposed a “delayed choice” in


which the decision which property will be observed is made after the photon
has passed BSin. “Thus one decides the photon shall have come by one route or by
both routes after it has already done its travel.” This then requires the insertion or
removal of BSout while the photon is already within the interferometer.
Wheeler argued that if then still interference is observed whenever BSout
is inserted, and the particle character whenever BSout has been removed,
then one in a sense forces the photon to travel back in time and change
its decision to travel along one or both paths. In other words, one can then
change the history of the quantum after that history is already past.
A quarter century after Wheeler’s suggestion, a French group led by
Aspect succeeded in carrying out this thought experiment ( Jacques et al.,
2007). The choice to remove or insert BSout is randomly decided by a quan-
tum random number generator. Its output is presented to an electro-optical
modulator (EOM) configuration that can rapidly switch (40ns) to effectively
behave like a beam splitter or as a neutral component. By having the two
path lengths to be 48 m it was ensured that the random number generator
and the EOM are relativistically separated from the entry of the photon in
the interferometer. In other words, no information about the choice can
reach the photon before it passes through BSin. The result of the delayed-
choice experiment is shown in Fig. 7. The counts of each detector as a func-
tion of the phase difference Φ vary harmonically when the second beam
splitter is present, as shown in Fig. 7A. When BSout is not present, the counts
do not depend on Φ and are evenly, and randomly, distributed over the two
detectors (see Fig. 7B). This shows that the behavior of the photons in the
interferometer indeed depends on the choice which observable is being
measured, even when that choice is made after the quantum has already
entered the configuration, and the choice is made at a position and a time
that is separated from the photon’s entry by a space-like interval. As the
authors of Jacques et al. (2007) conclude: “Once more, we find that nature
behaves in agreement with the predictions of quantum mechanics even in surprising
situations where a tension with relativity seems to appear.”
More recently, Wheeler’s delayed-choice experiment was carried out
with single ultracold metastable helium atoms instead of photons
(Manning, Khakimov, Dall, & Truscott, 2013). Laser pulses replaced the
mirrors and beam splitters that are shown in the configuration of Fig. 6.
The relatively slow-moving atoms were shown, just as photons, to confirm
the idea that attributing exclusively a wave or a particle character to quantum
objects in a given experimental configuration is indeed untenable.
304 Greg Gbur and Taco D. Visser

Fig. 7 Results of the delayed-choice experiment. (A) when the beam splitter BSout is pre-
sent, the photon counts of D1 (blue) and D2 (red) vary harmonically with the phase dif-
ference Φ. When the beamsplitter BSout is not present, no interference is observed and
an equal probability of detection at the two detectors is measured. Reproduced from
Jacques, V., et al., (2007). Experimental realization of Wheeler’s delayed-choice Gedanken
experiment. Science, 315, 966–968.

6. Optical coherence
Though optical science often applies simple models of monochro-
matic planar light waves to explain physical phenomena, all real sources
of light possess inherent random fluctuations. This randomness has observ-
able effects on the behavior of emitted light, affecting the spectrum, polar-
ization, directionality, and interference-causing capabilities of the light
wave. Optical coherence theory was developed to characterize the random-
ness of light and its effects on the measurable behavior of light, and also use
that randomness to improve existing optical applications and develop new
ones (Korotkova & Gbur, 2020). Young’s experiment has played a pivotal
role in understanding and measuring optical coherence properties.
Young’s interference experiment 305

The first person to ponder the subject of optical coherence was the
 mile Verdet. As he notes in his textbook (Verdet, 1869)
French scientist E
…interfering rays always have a common source: this is a necessary condition
because with two distinct radiating sources the interference fringes disappear
completely.

Although Verdet did not use the term, one would nowadays say that is due
to the fact that the two sources are uncorrelated. He then goes on to explain
that the superposition, at a particular frequency, of two completely unrelated
point sources (“molecules”) radiating with the same amplitude will vary rap-
idly between constructive and destructive interference. On average the
superposition will have the value between that of a bright fringe and a dark
fringe. Furthermore, he argues that there exists a duration of time and a spa-
tial area over which the light emitted by the two sources may be considered
to vibrate in unison. It is the concordance of these vibrations that is necessary
to produce interference fringes. All these considerations make Verdet the
founding father of optical coherence theory. The random changes of phase
and amplitude that occur in optical wavefields play out at such short time-
scales that they cannot be observed directly. Even an ultra-fast detector pro-
vides only a time-averaged value of these quantities. One of the attractive
points of coherence theory, as we shall see, is that it is formulated in terms
of observable quantities that can be determined with the help of Young’s
experiment.
In the modern approach, as formulated by Wolf, the statistical properties
of a stochastic wavefield are characterized by correlation functions (Wolf,
2007b). If the field is stationary, at least in the wide sense, then its temporal
correlations at two moments t1 and t2 depend only on the time difference
τ ¼ t2  t1. For such a field, represented by an analytic signal V (r, t)
(Mandel & Wolf, 1995), its mutual coherence function is defined as

Γðr1 , r2 , τÞ ¼ hV * ðr1 , tÞV ðr2 , t + τÞi (18)


Z ∞
1
¼ lim V * ðr1 , tÞV ðr2 , t + τÞ dt: (19)
T !∞ 2T ∞

Here the angular brackets indicate a time average. The average intensity at a
position r is defined as the mutual coherence function evaluated at τ ¼ 0, i.e.,

IðrÞ ¼ Γðr, r, 0Þ: (20)

A normalized version of Γ(r1, r2, τ), called the complex degree of coherence,
is obtained by defining
306 Greg Gbur and Taco D. Visser

Γðr1 , r2 , τÞ
γðr1 , r2 , τÞ ¼ : (21)
½Iðr1 ÞIðr2 Þ1=2
It can be shown that this function is bounded, namely

0  jγðr1 , r2 , τÞj  1: (22)


The upper bound corresponds to complete coherence, whereas the lower
bound indicates the absence of any coherence. For all intermediate values
the field is said to be partially coherent.
In the space-frequency domain two analogous correlation functions
exist. The first is the so-called cross-spectral density that is defined as the
temporal Fourier transform of the mutual coherence function
Z
1 ∞
W ðr1 , r2 , ωÞ ¼ Γðr , r , τÞeiωτ dτ: (23)
2π ∞ 1 2

It can be shown (Mandel & Wolf, 1995) that W(r1, r2, ω) can also be rep-
resented as a correlation function. More specifically,

W ðr1 , r2 , ωÞ ¼ hU * ðr1 , ωÞUðr2 , ωÞi: (24)


In this expression the average is not over time, but over an ensemble of mono-
chromatic field realizations. The spectral density (the average “intensity at fre-
quency ω”) at a position r is given by the cross-spectral density evaluated at
two coincident points, i.e.,

Sðr, ωÞ ¼ W ðr, r, ωÞ: (25)


The normalized version of W(r1, r2, ω), called the spectral degree of coher-
ence, is defined as

W ðr1 , r2 , ωÞ
μðr1 , r2 , ωÞ ¼ : (26)
½Sðr1 , ωÞSðr2 , ωÞ1=2
Just like the complex degree of coherence, the magnitude of this quantity is
bounded. One finds that

0  jμðr1 , r2 , ωÞj  1: (27)


The bounds have the same physical meaning as before. Zero indicates the
complete absence of spatial coherence (at frequency ω), whereas the value
1 corresponds to a spatially fully coherent field.
Young’s interference experiment 307

The notion of coherence was given an operational meaning (Wolf,


2007a) by the Nobel laureate Frits Zernike. In his landmark study
(Zernike, 1938), he connected the interference fringes that are formed in
Young’s experiment with the complex degree of coherence of the incident
field. To describe this, we return to the configuration sketched in Fig. 1, but
let us now assume that the incident field is statistically stationary and partially
coherent, rather than monochromatic and fully coherent. The analysis is
greatly simplified by the additional assumption that the light field is quasi-
monochromatic and has a mean wavelength denoted λ.  In the space-time
domain the field at a point P consists of the sum of the fields emanating from
the two pinholes, but at different times. Hence
V ðP, tÞ ¼ K 1 V ðr1 , t  t 1 Þ +K 2 V ðr2 , t  t 2 Þ, (28)
with the purely imaginary propagators K1 and K2, given by Eq. (2), evalu-
 Furthermore, tj denotes the time it takes the light to travel from
ated at λ.
pinhole j to the point P, i.e.,
Rj
tj ¼ , ð j ¼ 1, 2Þ, (29)
c
with c denoting the speed of light. The intensity at P is then given by the
expression

IðPÞ ¼ jK 1 j2 jV ðr1 , t  t 1 Þj2 + jK 2 j2 jV ðr2 , t  t 2 Þj2


  (30)
+ K 1 K *2 V ðr1 , t  t 1 ÞV * ðr2 , t  t 2 Þ +c:c: ,
where c.c. denotes the complex conjugate of the neighboring term. Because
the field is assumed to be stationary, the origin of time may be shifted in each
of these four terms. The result is that

IðPÞ ¼ jK1 j2 Iðr1 Þ + jK2 j2 Iðr2 Þ + 2jK1 K2 jRe½Γðr1 , r2 , τÞ: (31)


The first term in Eq. (31) is the intensity that would be observed at P if only
the pinhole at r1 is open. A similar interpretation holds for the second term.
In the last term of the expression we encounter the mutual cross-spectral
density function Γ, which was defined in Eq. (19), evaluated at time differ-
ence τ ¼ t1  t2. Eq. (31) can be further simplified by assuming that the point
P is close to the z axis, and that the distance between the two screens is much
larger than the separation of the two pinholes. In that case

jK 1 j2  jK 2 j2  jK 1 K 2 j ¼ K 2 : (32)
308 Greg Gbur and Taco D. Visser

If furthermore the time-averaged intensities at the two pinholes are equal, as


is usually the case, i.e.,
Iðr1 Þ ¼ Iðr1 Þ ¼ I, (33)
then Eq. (31) reduces to

IðPÞ ¼ 2jKj2 I f1 + Re½γðr1 , r2 , τÞg, (34)


where we have used Eq. (21). This result shows that the time-averaged
intensity that is observed at a point P is governed by (the real part of ) the
complex degree of coherence of the field at the pinholes at time differ-
ence τ. In turn, Re½γðr1 , r2 , τÞ can be determined from taking measure-
ments of I(P) and I. This is an illustration of coherence theory expressed
in terms of “observable quantities” as pioneered by Wolf (1954). Let us
next write

γðr1 , r2 , τÞ ¼ jγðr1 , r2 , τÞjei½a12 ðτÞωτ , (35)
 , and with the mean frequency
where a12 ðτÞ ¼ arg ½γðr1 , r2 , τÞ + ωτ
 On using this notation in Eq. (34) we get the result
 ¼ 2πc=λ.
ω

 g:
IðPÞ ¼ 2jKj2 I f1 + jγðr1 , r2 , τÞj cos ½a12 ðτÞ  ωτ (36)
It follows from general considerations of the analytic signal representation
of narrow-band signals (Mandel & Wolf, 1995) that in Eq. (36) both

jγðr1 , r2 , τÞj and a12(τ) vary much slower with τ than the term cos ðωτÞ.
From Eq. (9) we have that
R1  R 2 xd
τ¼ ¼ : (37)
c cΔz
Hence, moving the observation point P along the x direction (i.e., the ver-
tical direction in Fig. 1) results in a change of the time delay τ that is pro-
portional to the displacement of P. This means that the intensity near the
observation point consists of a constant background (the first term within
the curly brackets of Eq. (36)) with a cosine-like term superposed. Thus
the maximum and minimum intensity in the vicinity of P are given by

I max ¼ 2jKj2 I f1 + jγðr1 , r2 , τÞjg, (38)


2
I min ¼ 2jKj I f1  jγðr1 , r2 , τÞjg, (39)
Young’s interference experiment 309

respectively. The visibility V of the fringe pattern is defined as

I max  I min
V≡ ¼ jγðr1 , r2 , τÞj: (40)
I max + I min

Eq. (40) states that, unlike the optical field itself that oscillates too rapidly to
be observed, the modulus of the complex degree of coherence can be
directly established from intensity measurements in Young’s experiment.
From the bounds of jγðr1 , r2 , τÞj , given by Eq. (22), it follows that the
visibility can only attain its maximum value of unity when the field at the
two pinholes is fully coherent. This is illustrated in Fig. 8 where the normal-
ized intensity is plotted as a function of position x of the point of observation
for three different values of the modulus of the complex degree of coher-
ence. Complete destructive and constructive interference, i.e., V ¼ 1, occurs
when jγðr1 ,r2 ,τÞj ¼ 1(green curve). The other extreme, V ¼ 0, correspond-
ing to uncorrelated light with jγðr1 , r2 , τÞj ¼ 0 , gives rise to a constant
intensity distribution (red curve). An example of partially coherent light,
with jγðr1 , r2 , τÞj ¼ 0:5, is given by the dashed blue curve.
The phase of the complex degree of coherence can be determined by
comparing the fringe pattern with the fringes that are produced when the
two pinholes are illuminated by fully coherent and cophasal light, i.e., with
γ(r1, r2, τ) ¼ 1. This procedure is outlined in Section 10.4.1 of Born and
Wolf (1999).

I(x)

x [a.u]
Fig. 8 The normalized intensity pattern that is formed in Young’s experiment for three
selected values of the modulus of the complex degree of coherence. Green curve:
jγðr1 , r2 , τÞj ¼ 1, dashed blue curve: jγðr1 , r2 , τÞj ¼ 0:5, and red curve: jγðr1 , r2 , τÞj ¼ 0.
310 Greg Gbur and Taco D. Visser

P
F1 R1
r1
R2
d z

r2
F2

fringes
incident partially
coherent wave
screen A screen B
Fig. 9 Young’s interference experiment with the pinholes covered by two identical
narrow-band frequency filters F1 and F2.

We next turn our attention to the space-frequency domain and consider


the situation that is sketched in Fig. 9, in which the two apertures are both
covered by a narrow-band filter with center frequency ω. The field at a point
P on the observation screen is then given by the superposition

UðP, ωÞ ¼ K 1 Uðr1 , ωÞeikR1 + K 2 Uðr2 , ωÞeikR2 : (41)


The spectral density at P equals the ensemble-averaged value of jU(P, ω)j2,
and hence

SðP, ωÞ ¼ hjUðP, ωÞj2 i


(42)
¼ jK1 j2 Sðr1 , ωÞ + jK2 j2 Sðr2 ,ωÞ
h i
+ 2 Re K 1 K *2 hUðr1 , ωÞU * ðr2 ,ωÞieikðR1 R2 Þ : (43)

In the often-encountered situation that the spectral densities at the two pin-
holes are the same, i.e.,
Sðr1 , ωÞ ¼ Sðr2 , ωÞ ¼ SðωÞ, (44)
and again making the assumption described by Eq. (32), we find that
 n o
SðP, ωÞ ¼ 2K 2 SðωÞ 1 + Re μðr1 , r2 , ωÞei½kðR1 R2 Þ (45)
¼ 2K 2 SðωÞf1 + jμðr1 , r2 , ωÞj cos ½kðR1  R2 Þ +δg, (46)
where
δ ¼ arg½μðr1 , r2 , ωÞ: (47)
Young’s interference experiment 311

This expression is of the same form as its counterpart in the space-time


domain, Eq. (36), and states that the spectral density on the observation
screen is given by the sum of a constant term and a cosine. The maximum
and minimum values of the spectral density in the vicinity of P are,
respectively,
Smax ¼ 2K 2 SðωÞ½1 + jμðr1 , r2 , ωÞj, (48)
Smin ¼ 2K SðωÞ½1  jμðr1 , r2 , ωÞj:
2
(49)
It is therefore natural to introduce the spectral visibility as
Smax  Smin
Vω ≡ ¼ jμðr1 , r2 , ωÞj: (50)
Smax + Smin
This equation shows that the modulus of the spectral degree of coherence
can be deduced from the spectral visibility of the interference fringes in
Young’s experiment as illustrated in Fig. 9. The formula (46), first derived
by Mandel and Wolf (1976), is known as the spectral interference law. It
expresses the fact that in general, the spectral density of the field due to
the superposition of two partially coherent fields differs, at any fixed point,
from the spectral density of the incident field. Examples of such coherence-
induced spectral changes, to be distinguished from the effects of diffraction,
were studied (James & Wolf, 1991a, 1991b). Especially for the case of broad-
band light these spectral changes were predicted to be quite significant. An
experimental study by Santarsiero and Gori (1992) confirmed these
predictions. It was later shown that in the vicinity of dark lines (of certain
spectral components) the spectrum can either be red-shifted, blue-shifted,
or split into two parts (Pu, Cai, & Nemoto, 2004).
Eq. (45) and its space-time counterpart (36) show that complete coher-
ence is required in Young’s experiment in order to see complete destructive
interference, and this example may give the impression that complete coher-
ence is a general requirement for complete destructive interference.
However, in 2004, Gbur et al. showed theoretically that in fact this is not
the case (Gbur, Visser, & Wolf, 2004a). Using a “three-pinhole interfero-
meter,” they demonstrated that there are circumstances in which the light
emerging from the three pinholes is mutually partially coherent, but the
interference pattern contains points of complete darkness. If the degree of
coherence between all pairs of pinholes is the same and equal to μ0, the con-
dition for complete destructive interference is that μ0 satisfies the cubic
equation,
1  3μ20 + 2μ30 ¼ 0: (51)
312 Greg Gbur and Taco D. Visser

Two of the roots of this equation are μ0 ¼ 1; the third root is μ0 ¼ 1/2.
This theoretical prediction was soon verified using both acoustics
(Basano & Ottonello, 2005) and visible light (Ambrosini, Gori, &
Paoletti, 2005). More recent work has studied the problem of coherence
and destructive interference more generally, looking at multiple pinholes
(Gan & Gbur, 2007) as well as the case when the degree of coherence is dif-
ferent for different pinhole pairs (Rosenbury, Gu, & Gbur, 2012). For the
case of N pinholes, with the same degree of coherence μ0 between the pin-
holes, the nontrivial degree of coherence that produces destructive interfer-
ence is μ0 ¼ 1/(N  1).
Even with two pinholes, Young’s experiment has revealed unexpected
behavior. In 2003, Schouten et al. demonstrated that there are always pairs of
points in the superposition region of Young’s experiment where the field is
fully coherent, regardless of the state of coherence of the light at the pinholes
(Schouten, Visser, & Wolf, 2003).
When nontrivial polarization is added to Young’s experiment, even
richer coherence properties are exhibited. Traditionally, the statistical prop-
erties of the state of polarization have been treated separately from the sta-
tistical properties of the complex wavefield. The polarization was studied at
one point in space, often described by the so-called coherency matrix (Wolf,
1959),
" #
hE*x ðr, tÞE x ðr, tÞi hE *x ðr, tÞE y ðr, tÞi
J≡ , (52)
hE*y ðr, tÞE x ðr, tÞi hE *y ðr, tÞE y ðr, tÞi

whereas the coherence properties are characterized by two points in space


and time, through Γ(r1, r2, τ).
In 2003, Wolf introduced a “unified theory of coherence and polarization”
Wolf (2003), joining together the statistical properties of the complex field and
state of polarization into a cross-spectral density matrix,

Wðr1 , r2 , ωÞ ¼ W ij ðr1 , r2 , ωÞ ¼ hE *i ðr1 , ωÞE j ðr2 , ωÞi, (53)

where i, j ¼ {x, y}. This matrix can be used to determine the coherence and
polarization properties of the field and interrelations between them.
A degree of coherence for this general case can readily be determined by
returning to Young’s experiment, but now evaluated with fluctuating elec-
tric fields. Again working with a frequency-filtered version of Young’s
experiment, we have
Young’s interference experiment 313

EðP, ωÞ ¼ K 1 Eðr1 , ωÞeikR1 + K 2 Eðr2 , ωÞeikR2 : (54)


We define the spectral density on the observation screen as the ensemble
averaged version of Eq. (14), i.e.,

SðP, ωÞ ≡ hjEðP, ωÞj2 i ¼ TrWðP, P, ωÞ, (55)


where Tr represents the trace of the matrix. On substitution from Eq. (54)
into Eq. (55), we may find a generalization of Eq. (42),

SðP, ωÞ ¼ jK 1 j2 Sðr1 , ωÞ +jK 2 j2 Sðr2 , ωÞ


  (56)
+ 2 Re K 1 K *2 ηðr1 , r2 , ωÞeikðR1 R2 Þ ,
where η(r1, r2, ω) is the spectral degree of coherence for general electromag-
netic fields, defined as
TrWðr1 , r2 , ωÞ
ηðr1 , r2 , ωÞ ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (57)
Sðr1 , ωÞSðr2 , ωÞ
This definition of the degree of coherence has an absolute value bounded
between zero and one, with zero representing incoherence and unity rep-
resenting full coherence. In this electromagnetic case, however, the visibility
depends not only on the spatial correlations between the pinholes but also on
the polarization properties of the field—in accordance with the Fresnel–
Arago laws, it is possible for a field to be fully coherent but produce no inter-
ference fringes and thus be “incoherent.” It is to be noted that Tervo et al.
introduced an alternative definition of a degree of coherence that quantifies
the statistical similarity of the electromagnetic fields at two points (Tervo,
Set€al€a, & Friberg, 2003); this degree of coherence can be measured with
a sequence of Young-type experiments.
The unified theory of coherence and polarization allowed a systematic
study of the interplay between these two optical phenomena, and Young’s
experiment remained a natural approach for exploring the implications of
the theory. In 2004, Mujat et al. used Young’s experiment to introduce gen-
eralizations of the classic Fresnel-Arago laws (Mujat, Dogariu, & Wolf, 2004).
In 2006, Gori et al. examined the effect of spatial coherence on the degree of
polarization using Young’s experiment (Gori, Santarsiero, Borghi, & Wolf,
2006), showing that the degree of polarization can have a complex structure
in the image plane, even if the light is unpolarized at the pinholes. Agarwal
et al. demonstrated that there are pairs of points in Young’s experiment where
the field is always fully coherent, regardless of the source coherence and degree
314 Greg Gbur and Taco D. Visser

of polarization (Agarwal, Dogariu, Visser, & Wolf, 2005); this work general-
ized the aforementioned work by Schouten et al. for scalar fields (Schouten,
Visser, & Wolf, 2003). Gori et al. interpreted the output of Young’s exper-
iment using a vector mode theory (Gori, Santarsiero, & Borghi, 2006). Also in
2006, Set€al€a, Tervo, and Friberg (2006 a, 2006 b) derived spectral interference
laws for the Stokes parameters in Young’s experiment.
In concluding this section on coherence, we note that Young’s exper-
iment can not only be used to analyze the statistical properties of light,
but also synthesize light with tailored coherence properties. In 2021, Lv
et al. (2021) showed that Young’s experiment can be used to generate non-
uniformly correlated fields, i.e., fields that possess a spatially varying, non-
homogeneous, degree of coherence.

7. Surface plasmons
In 1957, a novel type of electromagnetic guided mode was discovered
by Ritchie in experiments studying electron transmission through thick
metallic foils (Ritchie, 1957). These modes are known as surface plasma
polaritons, often simply called surface plasmons (SPs) (Raether, 1988). At
optical frequencies, they propagate along the interface between a dielectric
and a metal. As is characteristic for surface waves, their maximum intensity is
at the interface, and the fields decay exponentially away from it. The prop-
agation of SPs includes a longitudinal density fluctuation of surface electrons.
Owing to the strong localization of intensity, SPs are routinely employed
in surface-enhanced Raman scattering (SERS) and second harmonic gener-
ation (SHG). Whereas optical communication speeds are limited by the
conversion of light waves into electronic signals for purposes of routing, sig-
nal processing, and amplification, the use of SPs in so-called all-optical net-
works might one day make this electronic bottleneck obsolete. For our
present discussion, we are interested in the role that SPs play in optical trans-
mission through small apertures in metal films. As we will discuss, the total
transmission, the state of coherence, and the state of polarization are all
strongly affected by SPs, and these effects can be demonstrated using
Young’s experiment. A simple intuitive model can explain the mechanism
behind these phenomena.
The diffraction of light by an aperture in an opaque screen can often be
adequately described with classical scalar theory. The main formulas that are
commonly used are one by Kirchhoff, and two expressions due to Rayleigh
and Sommerfeld (Goodman, 1996). One feature these approaches have in
Young’s interference experiment 315

common is the assumption that the field in the aperture is identical to the
incident field, i.e., the field in the absence of the screen, which is often taken
to be a normally incident plane wave. In reality, the actual field will differ
from the incident field because of scattering caused by the screen. In spite of
this approximation, when the aperture is much larger than the wavelength,
the predictions of the three formulas that were just mentioned are very sim-
ilar and quite accurate.
A theoretical model of the behavior of light transmission through a sub-
wavelength hole in an infinitely thin, perfectly conducting screen was intro-
duced by Bethe in 1944 (Bethe, 1944). Bethe predicted the transmission
efficiency to scale as (r/λ)4, where r is the radius of the hole. For holes sig-
nificantly smaller than the wavelength, it was therefore typically assumed
that almost no light would be transmitted.
When dealing with subwavelength apertures in metal films with a finite
thickness and conductivity, a scalar approach no longer suffices and, follow-
ing Bethe, Maxwell’s equations must be used. An example is presented in
Fig. 10 where the time-averaged Poynting vector near a narrow slit in thin
metal film is plotted. The result is seen to be quite intricate and contains
structures like vortices and saddle points. This field structure is not only very
sensitive to the wavelength and polarization of the incident field, but also to
the slit width, the thickness of the film, and its conductivity.

a b

c d

100 nm 0 1

Fig. 10 Time-averaged Poynting vector near a 200-nm wide slit in a 100-nm thick silver
plate. Plane wave light, with a wavelength λ ¼ 500 nm, is incident below. Left-handed (A
and D) and right-handed (B and C) optical vortices are visible, just as two saddle points (E
and F). The color coding indicates the modulus of the normalized Poynting vector.
Reproduced from Schouten, H. F., Visser, T. D., Lenstra, D., & Blok, H. (2003). Light transmis-
sion through a sub-wavelength slit: Waveguiding and optical vortices. Physical Review E, 67,
036608.
316 Greg Gbur and Taco D. Visser

In 1998, Ebbesen et al. reported observations of a phenomenon that they


referred to as Extraordinary Optical Transmission (EOT) (Ebbesen, Lezec,
Ghaemi, Thio, & Wolff, 1998). In their experiment, they illuminated peri-
odic arrays of submicrometer cylindrical holes in metallic films. For several
wavelengths, comparable with the array period, the array of holes transmits
more light than a large macroscopic hole with the same area as all the small
holes combined. Clearly, this implies that the array itself is an active element,
and not just a passive geometrical object in the path of the incident beam.
The authors suggested that SPs might play a role in the mechanism under-
lying EOT. One clue being that an array of holes in a Ge film, which does
not support SPs, does not exhibit enhanced transmission.
The original observation of EOT and the role of SPs were initially met
with great skepticism, possibly due to the complex interactions near sub-
wavelength apertures, as shown in Fig. 10. For example, in Cao and
Lalanne (2002) it was argued that surface plasmons do not help but rather hin-
der light transmission through gratings, and that EOT is due to waveguide
mode resonances. For many subwavelength transmission geometries only
numerical results are available, and in general it is quite hard to disentangle
the roles of SPs, waveguide modes, surface corrugations, and singularities like
vortices and saddle points in the transmission process (Garcia-Vidal,
Martin-Moreno, Ebbesen, & Kuipers, 2010; Genet & Ebbesen, 2007). It is
only through a simple and “clean” setup like Young’s double-slit experiment
that the crucial role of SPs can be determined.
In 2005, a combined experimental and theoretical study by Schouten
et al. explored the role of SPs in transmission through Young’s
double-slit experiment (Schouten et al., 2005). It was found not only that
SPs do in fact modulate the transmission of light through subwavelength
holes, but that SPs can either increase or decrease the optical transmission.
The total intensity of the far-field double-slit pattern was shown to be
reduced or enhanced as a function of the wavelength of the incident light
beam. This modulation was attributed to an interference phenomenon
occurring at each of the slits, rather than at the detector. SPs that are excited
at one slit can travel to the other slit where they are converted back into a
freely propagating field; this means that interference can occur between the
incident field that is directly transmitted and the field contribution due to
backconverted SPs.
This process can be explained using a simple, phenomenological model
(Schouten et al., 2005) which captures the essence of this SP-modulated
Young’s interference experiment 317

U (out) (r1 ) U (out) (r2 )


dielectric

metal
z
SP SP dielectric

x d
U (in) (r1 ) U (in) (r2 )
Fig. 11 A metal film with two identical subwavelength parallel slits is embedded in a
dielectric. The film is illuminated from below (blue arrows). At each side of the metal film
and at each slit SPs are launched both to the left and to the right (red arrows). Cylindrical
waves (blue) emanate from the slits. The distance between the two slits is d.

transmission process; and is illustrated in Fig. 11. Moreover, this model also
adequately describes changes in coherence and polarization.
Let U(in)(ri), with i ¼ 1 or 2, denote the monochromatic incident field at
slit i at frequency ω, and U(out)(ri) the transmitted field. When the incident
field is TE polarized (along the y direction) part of it will be reflected, and a
fraction, α say, will be directly transmitted, i.e.,

U ðoutÞ ðr1 Þ ¼ αU ðinÞ ðr1 Þ, (58)


ðoutÞ ðinÞ
U ðr2 Þ ¼ αU ðr2 Þ, ðTE polarization:Þ (59)
When the incident field is TM polarized (along the x direction), part of it
will be reflected, and a fraction β will be directly transmitted. In addition
to these effects, surface plasmons will be generated near the slits at both
the top and bottom side of the film. These SPs travel to the left and to
the right. The SPs that travel a distance d to the adjacent slit can be coupled
back into an optical wave that propagates in the positive z direction. This
phenomenological model was later partly confirmed in Kuzmin et al.
(2007), in which it was shown that Young’s interference fringes appear even
when only one slit is illuminated with TM light—SPs from the illuminated
slit are converted into propagating light at the unilluminated slit, thus pro-
ducing a two-slit diffraction pattern.
If the efficiency of this light-to-SP-back-to-light scattering process is
called γ, then

U ðoutÞ ðr1 Þ ¼ βU ðinÞ ðr1 Þ +γU ðinÞ ðr2 Þeiksp d , (60)


ðoutÞ ðinÞ ðinÞ
U ðr2 Þ ¼ βU ðr2 Þ +γU ðr1 Þe iksp d
, ðTM polarization:Þ (61)
318 Greg Gbur and Taco D. Visser

In these expressions the SP wavenumber is given by


rffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Em Ed
ksp ¼ k0 , (62)
Em + Ed

where Em and Ed are the relative complex dielectric constants of the metal and
the dielectric, respectively, and k0 ¼ 2π/λ0 is the free-space wavenumber.
The SP wavelength is related to the real part of ksp by λsp ¼ 2π/Re(ksp).
The amplitude decay length is given by 1/Im(ksp). It is seen from
Eqs. (60) and (61) that the SP-mediated amplitude interferes with the ampli-
tude of the light that is directly transmitted by each slit. When the slit widths
are subwavelength these two contributions are of comparable magnitude.
From now on we concentrate on TM polarization and consider the case
of a normally incident, monochromatic plane wave, and hence

U ðinÞ ðr1 Þ ¼ U ðinÞ ðr2 Þ ¼ U ðinÞ , (63)


and define S(in) ¼ jU(in)j2. Then the spectral density of the transmitted field at
each slit equals

SðoutÞ ¼ jU ðoutÞ j2 (64)


h  i
¼ SðinÞ jβj2 + jγj2 + 2Re γβ* eiksp d : (65)

This result implies that, depending on the slit separation d, the SPs can either
help or hinder the transmission process. Notice that the intensity modulation
varies with the SP wavenumber ksp, rather than the free-space wavenumber
k0. Even though this intuitive model contains three unknown complex con-
stants, namely the transmission coefficients α and β and the scattering coef-
ficient γ, it provides a clear physical insight into the mechanism that
underlies EOT.
A rigorous numerical analysis presented in Schouten et al. (2005), based
on a Green’s tensor scattering formalism, showed the SPs form a standing
wave pattern between the two slits, as seen in Fig. 12. When the slit sepa-
ration is such that the antinodes coincide with the slit positions the transmis-
sion reaches a maximum, when the nodes of the standing-wave pattern are
at the slits, the transmission is suppressed. In the same study, an experimental
verification using a tunable laser was reported. As shown in Fig. 13, the
predicted modulation of the detected transmitted field (integrated over
many diffraction orders) with frequency, and hence with ksp, was only
observed for TM illumination. This is evidence that SPs indeed play a
pivotal role in the transmission process of light through nanoapertures.
Young’s interference experiment 319

Fig. 12 Color-coded intensity distribution in the vicinity of a double-slit system


with TM-polarized illumination when the transmission reaches a maximum and the
slit separation d ¼ 5λsp/2 (top frame), and when the transmission is at a minimum,
and d ¼ 4λsp/2. All lengths are in nm. Reproduced from Schouten, H. F., Kuzmin, N.,
Dubois, G., Visser, T. D., Gbur, G., Alkemade, P. F. A., …Eliel, E. R. (2005). Plasmon-assisted
two-slit transmission: Young’s experiment revisited. Physical Review Letters, 94, 053901.

We next turn our attention to the case that the field that is incident on the
slits is not fully coherent, as was assumed so far, but is spatially partially
coherent. Surface plasmons traveling between the slits are then found to
modify the state of coherence as well. Since many properties of a light
field—such as its spectrum, polarization, and directionality—may change
on propagation and are dependent on the spatial coherence of the (second-
ary) source, this suggest that the use of surface plasmons provides a new way
to alter or even tailor the statistical properties of a light field. We recall
(cf. Eq. (26) of Section 6) that the spectral degree of coherence of the
incident field is defined as

ðinÞ hU ðinÞ* ðr1 ÞU ðinÞ ðr2 Þi


μ12 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ffi , (66)
SðinÞ ðr1 ÞSðinÞ ðr2 Þ

and, in strict analogy, the spectral degree of the transmitted field is given by
the expression
320 Greg Gbur and Taco D. Visser

Fig. 13 The top frame shows Young’s interference pattern as observed with a CCD. The
other frames show diffraction-order-integrated transmission spectra for TM illumina-
tion, for selected values of the slit separation d. In the bottom frame results for TE illu-
mination are also included (open squares). The vertical axis on the right corresponds to
this choice of polarization. Reproduced from Schouten, H. F., Kuzmin, N., Dubois, G.,
Visser, T. D., Gbur, G., Alkemade, P. F. A., …Eliel, E. R. (2005). Plasmon-assisted two-slit trans-
mission: Young’s experiment revisited. Physical Review Letters, 94, 053901.

ðoutÞ hU ðoutÞ* ðr1 ÞU ðoutÞ ðr2 Þi


μ12 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi : (67)
ðoutÞ ðoutÞ
S ðr1 ÞS ðr2 Þ
Young’s interference experiment 321

Using the definition (64) for the spectral density of the transmitted field now
gives
h  i
ðinÞ
SðoutÞ ðr1 Þ ¼ SðinÞ jβj2 + jγj2 + 2Re γβ* μ12 eiksp d , (68)
h  i
ðinÞ*
SðoutÞ ðr2 Þ ¼ SðinÞ jβj2 + jγj2 + 2Re γβ* μ12 eiksp d , (69)

which leads to
ðinÞ ðinÞ*

ðoutÞ jβj2 μ12 + jγj2 μ12 + 2Re β* γeiksp d


μ12 ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi :
ðinÞ ðinÞ*
½jβj2 + jγj2 + 2Reðγβ* μ12 eiksp d Þ½jβj2 + jγj2 + 2Reðγβ* μ12 eiksp d Þ
(70)
This formula shows that the spectral degree of coherence of the field that is
radiated by the two apertures is not equal to the spectral degree of coherence
of the incident field. Because of the presence of oscillating terms in this expres-
sion, the modulus of the former can either be larger or smaller than that of the
latter. In other words, varying the distance that separates the two slits will
modulate the visibility of the observed interference fringes. That this effect
is solely due to the action of surface plasmons is easily verified by setting γ,
their relative contribution, to zero. In that case Eq. (70) reduces to
ðoutÞ ðinÞ
μ12 ¼ μ12 , (71)
i.e., the spectral degree of coherence of the transmitted field is then equal to
that of the incident field. In Gan, Gbur, and Visser (2007) this SP-modulated
coherence in a two-slit configuration was studied using a rigorous Green’s
tensor method. It was verified that, depending on the slit separation, SPs can
indeed increase or decrease the spectral degree of coherence. Some illustrative
results are shown in Fig. 14. Notice that it is even possible for SPs to convert
completely incoherent light (i.e., a field with μ(in)
12 ¼ 0) into light that is sig-
nificantly coherent. Experimental confirmation of the modulating influence
of SPs on the state of spatial coherence in Young’s experiment was presented
in Divitt, Frimmer, Visser, and Novotny (2016) and Li and Pacifici (2017).
Generalizations of this experiment, using more holes, have been performed
with the ultimate goal of introducing coherence-converting plasmonic
devices. In Gan and Gbur (2008), exact numerical simulations of a three-slit
interferometer demonstrated that the intermediate hole would not obstruct
the plasmons significantly or reduce the changes of coherence. A more general
configuration, consisting of an array of subwavelength circular holes in a metal
322 Greg Gbur and Taco D. Visser

|μ12(ω)|

μ
|μ12(ω)|

Fig. 14 The absolute value of the spectral degree of coherence of the transmitted field
as a function of the slit separation d, for selected values of the spectral degree of coher-
ence of the incident field. In (A) the slit width w ¼ 100 nm, in (b) w ¼ 100 nm.
Reproduced from Gan, C. H., Gbur, G., & Visser, T. D. (2007). Surface plasmons modulate
the spatial coherence of light. Physical Review Letters, 98, 043908.

film, was analyzed numerically in Gan, Gu, Visser, and Gbur (2012). In such a
setup there is no direct relation between the visibility of the interference pat-
tern produced by the transmitted field and an overall degree of coherence.
Nevertheless, the coherence of neighboring pairs of apertures can be studied.
It was found that this depended strongly on the separation of the holes and
their light-to-SP scattering strength. A further study into the effects of array
Young’s interference experiment 323

geometry, lattice constant, and hole size on coherence conversion was pres-
ented in Gbur and Smith (2021). The same authors had previously discovered
the existence of so-called coherence bandgaps in plasmonic hole arrays
(Smith & Gbur, 2019). In Saastamoinen and Lajunen (2013) a subwavelength
metallic grating with an incident Gaussian Schell-model beam was investi-
gated numerically. The authors introduced a global degree of coherence, as
an intensity-weighted average of the spectral degree of coherence of the
diffracted field. They observed that, for certain grating parameters, this quan-
tity can be enhanced significantly.
A third form of modulation of the diffracted field caused by SPs is that of
the state of polarization (SOP). This effect was examined by Leinonen et al.
(2021). The electric field at the exit plane z ¼ h of the metallic film can be
propagated to an arbitrary plane z > h using the angular spectrum technique
(Mandel & Wolf, 1995). In such an approach each electric field component
is written as an integral of plane waves, i.e.,
Z ∞
E j ðx, zÞ ¼ Aj ðkx Þ exp ½iðkx x + kz ΔzÞ dkx , ðj ¼ x, yÞ, (72)
∞

where Δz ¼ z  h,
Z
1 ∞
Aj ðkx Þ ¼ E ðx, tÞ exp ðikx xÞ dx, ðj ¼ x, yÞ, (73)
2π ∞ j

and
8 qffiffiffiffiffiffiffiffiffiffiffiffiffiffi
>
< k20  k2x when jkx j < k0 ,
kz ¼ qffiffiffiffiffiffiffiffiffiffiffiffiffiffi (74)
>
: i k2  k2
x 0 when jkx j < k0 ,

and k0 denotes the free-space wavenumber. At distances Δz ≫ λ0 ¼ 2π/


k0 the contributions of the evanescent waves with imaginary values of kz
can be neglected. Under such circumstances the spatial integration in
Eq. (73) may be limited to the close proximity of the slits. If we take the slits
to be line sources located at x ¼ d/2 and x ¼ d/2, then, apart from an
inconsequential overall prefactor,

E x ðx, hÞ ¼ ½α + γ exp ðiksp dÞ½δðx + d=2Þ + δðx  d=2Þ, (75)


Ey ðx, hÞ ¼ β½δðx + d=2Þ + δðx  d=2Þ: (76)
324 Greg Gbur and Taco D. Visser

This gives
Ax ðkx Þ ¼ ½α + γ exp ðiksp dÞ cos ðdkx =2Þ, (77)
Ay ðkx Þ ¼ β cos ðdkx =2Þ: (78)
With this formalism the SP-modulated State of Polarization (SOP) of the
transmitted field can be calculated. In Leinonen et al. (2021), the phenom-
enological model that we just described was validated. With the help of a
rigorous Fourier Modal Method it was shown that the values of the three
parameters α, β, and γ remain fairly constant over a range of configuration
parameters such as the slit separation distance. It was shown that not just the
far-zone SOP can be controlled by varying the slit separation, but also the
degree of polarization can be manipulated. In particular, it was shown
numerically that a linearly partially polarized field can, under suitable con-
ditions, be rendered completely unpolarized by the action of surface
plasmons.
The interference of SPs, i.e., without the usual back conversion into light,
has been studied in which was called an SP analog to Young’s double-slit
experiment (Zia & Brongersma, 2007). Two narrow metallic stripes that sup-
port plasmons play the same role as the traditional slits. At the end of the slits
the SPs enter a broad Au film where their interference can be observed with a
photon scanning tunneling microscope.

8. Entanglement and duality


Entanglement is often considered to be a phenomenon that is purely
confined to the quantum domain. However, recent researches have pointed
out that classical systems can also be entangled. The key notion here is that
the different degrees of freedom of a classical system, for example, the three
Cartesian directions and their respective field amplitudes, can not always be
factorized. This is similar to the nonfactorizability of the wave function of
entangled quantum systems. An overview of classical entanglement was
presented in Forbes, Aiello, and Ndagano (2019); an early discussion of
the concept is given in Spreeuw (1998). In this section we review results that
apply both to the quantum domain and to classical wave fields. Obviously
then, the concepts that are discussed can have various interpretations.
In recent years there has been a renewed interest in the question of how
to define a quantitative measure of the wave-particle duality. In Young’s
double-slit setup, the intensity at a point P on the observation screen is
Young’s interference experiment 325

due to contributions of both apertures, A and B. If aperture A is blocked, the


intensity at P is solely due to the emission from the unblocked aperture B,
which we denote as IB. The intensity that reaches P when B is blocked is
called IA. If we assume the quanta behave as particles, then it stands to reason
to introduce the probability that the quantum passed through either aperture
A or B as
IA
PA ¼ , (79)
IA + IB
IB
PB ¼ , (80)
IA + IB
respectively. The so-called which-path distinguishability D, i.e., the degree to
which the signal at P can be attributed to one of the two slits, is then intro-
duced as the magnitude of the difference between these probabilities, i.e.,

jI A  I B j
D ¼ jP A  P B j ¼ : (81)
IA + IB
This quantity is also called “predictability” by Jaeger, Horne, and Shimoney
(1993). It is considered as a marker of the amount of “particleness.” Similarly,
a measure of the amount of “waveness” is provided by the visibility, or fringe
contrast V, which was introduced in Section 6. The usual view is that com-
plete which-way information, i.e., D ¼ 1, rules out the formation of inter-
ference fringes, implying that V ¼ 0. Which-way information can be
obtained, for example, by placing orthogonal polarizers behind each of the
two slits and thus tagging the photons. Remarkably, this information is not
absolute, in the sense that it can be erased to restore the formation of inter-
ference fringes (Scully, Englert, & Walther, 1991; Walborn, Cunha,
Pádua, & Monken, 2002). Conversely, perfectly sharp fringes with V ¼ 1
can only be formed when the intensity from both apertures is equal, i.e., when
D ¼ 0 (Brezger et al., 2002).
In Wootters and Zurek (1979) it was pointed out that duality need not
represent an absolute incompatibility. In fact, in Young-type scenarios an
amount of particleness can be given up for an increase in waveness, and vice
versa. This implies a sort of duality balance between wave and particle
interpretations. More precisely, one can show that (for a list of the many dif-
ferent derivations of this result, see Ref. 6 in Eberly, Qian, and Vamivakas
(2017))

V 2 + D2  1: (82)
326 Greg Gbur and Taco D. Visser

In Qian, Vamivakas, and Eberly (2018) it was argued that this picture cannot
tell the whole story. For example, if the contributions from A and B are
equally strong, then the distinguishability D ¼ 0. If these two fields are also
orthogonally polarized, then no fringes will be produced and the visibility
V ¼ 0. A nonzero signal must be characterized by something more than
V ¼ D ¼ 0, a clear indication that something is absent from this picture.
This missing information can be identified by a recently derived
“polarization coherence theorem” (Eberly et al., 2017), which connects
the degree of polarization P, the ratio of the intensity of the polarized por-
tion of the field, and its total intensity, with the visibility and the distinguish-
ability through the relation

P 2 ¼ V 2 + D2 : (83)

Notice that since the degree of polarization is bounded by zero and unity,
there is no conflict between Eqs. (82) and (83). An experimental verification
of this theorem was reported in Kanseri and Sethuraj (2019). A further ingre-
dient to consider is the degree of entanglement, of which one of the possible
measures is the concurrence C. The concurrence and the degree of polar-
ization are, within the context of Young’s experiment, perfect opposites. By
considering the aforementioned analogy of factorization of classical polari-
zation states and entanglement, it can be derived that (Qian, Malhotra,
Vamivakas, & Eberly, 2016)

P2 ¼ 1  C2: (84)

This expression shows that gaining some of P2 comes at the expense of losing
an equal amount of C2. This result is quite surprising, since P and C have
such different traditional origins and interpretations, entanglement being a
bipartite property, whereas polarization is a single-party property. On mak-
ing use of Eq. (84) in Eq. (83) it immediately follows that

V 2 + D2 + C 2 ¼ 1: (85)

This formula expresses a profound connection between wave-particle dual-


ity, as expressed by V and D, and the entanglement C. In fact, one might say
that entanglement limits duality and vice versa. The expression (85) calls to
mind a unit sphere in a space with axes that are the independent degrees of
freedom V, D, and C. This relationship between three observables in
Young’s experiment was validated in Qian et al. (2018). An illustration of
Young’s interference experiment 327

C
13 1

12
11
9 10

8
5
0 7
6

1
D
4
1
V 1 3
2

Fig. 15 Thirteen measurements of (V, D, C), represented by the red dots, are seen to fit
on the surface of one octant of a unit sphere. Reproduced from Qian, X.-F.,
Vamivakas, A. N., & Eberly, J. H. (2018). Entanglement limits duality and vice versa.
Optica, 5, 942–947.

the experimental verification, with all observables exploring their maximum


range, is shown in Fig. 15. As the authors say
It is fascinating that the two elements of quantum theory most often said to be
conceptually mysterious or weird, duality and entanglement, are intimately
related, and in control of each other.

9. Multi-order interference and exotic paths


A multislit configuration can be used to test one of the basic tenets of
Quantum Mechanics, the so-called Born Rule. According to Born’s interpre-
tation of the wave function Ψ(r, t), the probability density to find a particle at
position r and at time t is given by

Pðr, tÞ ¼ jΨðr, tÞj2 : (86)


As shown out by Sinha, Couteau, Jennewein, Laflamme, and Weihs (2010),
the double-slit diffraction experiment can be used to test Born’s premise,
because the interference that is observed in the experiment is a direct con-
sequence of this rule. The probability to detect a particle at r after passing
through a screen with two slits, A and B, is

P AB ðrÞ ¼ jΨA ðrÞ +ΨB ðrÞj2 (87)


328 Greg Gbur and Taco D. Visser

¼ jΨA j2 + jΨB j2 + Ψ*A ΨB + Ψ*B ΨA (88)


¼ P A + P B + I AB , (89)
where the position argument has been omitted for brevity, and with Pi
defined as the detection probability when only slit i (i ¼ A, B) is open.
The corresponding second-order interference can be defined in terms of
these three probabilities as

IAB ≡ PAB  ðPA + PB Þ: (90)


Adding more slits does not add higher-order complexity. For example, for
three slits, A, B, and C, one finds that

PABC ¼ PA + PB + PC + IAB + IAC + IBC : (91)


In words, interference always occurs in pairs of possibilities and is defined as
the deviation from the classical additivity of the possibility of mutually exclu-
sive events (i.e., the passage through one particular aperture). As a conse-
quence, a possible third-order interference term IABC for a three-slit setup
can be written as the difference between PABC and the sum of the individual
(mutually exclusive) probabilities and all second-order interference terms,
namely as

IABC ≡ PABC  ðPA + PB + PC + IAB + IAC + IBC Þ (92)


¼ PABC  PAB  PBC  PAC + PA + PB + PC : (93)
The term IABC, named the Sorkin parameter (Sorkin, 1994), equals zero in
all wave theories with a square-law relation between the field energy and
field amplitude, as is the case for quantum mechanics. In Sinha et al.
(2010) an arrangement, as shown in Fig. 16, of one, two and three slits
was used to measure the seven individual terms on the right-hand side of
Eq. (93). For their null test the authors considered a scaled version of the
Sorkin parameter, namely
IABC
κ≡ : (94)
jIAB j + jIBC j + jIAC j
Measurements were taken using true single photons from a nonclassical
source, as well as with weakened laser pulses with an average occupation
of less than one photon. In both cases it was found that κ < 0.01. This then
Young’s interference experiment 329

Blocking Mask
Position
60µm
0
600µm
C
1000µm

(Inset) Slit Mask BC


30µm

AC
300µm

AB

100µm
ABC

Fig. 16 The seven different slit arrangements that were used to test Born’s Rule.
Reproduced from Sinha, U., Couteau, C., Jennewein, T., Laflamme, R., & Weihs, G. (2010).
Ruling out multi-order interference in quantum mechanics. Science, 329, 418–421.

confirms Born’s Rule: quantum interference between many different path-


ways is simply the sum of the effects from all pairs of pathways.
A further observation to make is that multislit interference is typically
simplified in textbooks. For example, in a double-slit experiment carried
out with, e.g., photons or electrons, the wave function at the detector when
only slit A is open is denoted ΨA. The wave function when only slit B is
open is called ΨB. What is the wave function when both slits are open? It
is usually assumed that then ΨAB ¼ ΨA + ΨB. However, these three cases
correspond to different boundary conditions and therefore a direct applica-
tion of the superposition principle can at best be approximate. Numerical
simulations of Maxwell’s equations for the case of two slits milled in a thin
metal film demonstrated this in the classical domain (Raedt, Michielsen, &
Hess, 2012). The conclusion is that a three-slit configuration can only be
analyzed approximately in terms of two-slit configurations as is tacitly
assumed in Eq. (93). Furthermore, are passages through different slits really
mutually exclusive events? This was explicitly assumed in the preceding dis-
cussion. But in the quantum domain one has to take into account so-called
nonclassical paths which, according to Feynman’s path integral formalism,
are also possible. As illustrated in Fig. 17, apart from the classical path
(green curve), a particle may also first travel through slit A, then loop back
330 Greg Gbur and Taco D. Visser

Fig. 17 A three-slit configuration showing both a classical path (green) and a non-
classical path (blue). Reproduced from Sawant, R., Samuel, J., Sinha, A., Sinha, S., &
Sinha, U. (2014). Nonclassical paths in quantum interference experiments. Physical
Review Letters, 113, 120406.

via slit B, and emerge again from slit C (blue curve). For a double-slit
arrangement, this means that Eq. (87) would have be altered to a more gen-
eral form, namely (Sawant, Samuel, Sinha, Sinha, & Sinha, 2014)

PAB ðrÞ ¼ jΨA ðrÞ + ΨB ðrÞ + ΨL ðrÞj2 , (95)

where ΨL is the contribution due to the nonclassical, looped paths. In the


path integral formalism any path can be divided into several subpaths and
the resulting propagator is the product of the individual propagators.
Under typical circumstances, using the Fresnel approximation to describe
the propagator results in a very small but nonzero estimate of the scaled
Sorkin parameter introduced above, namely (Sawant et al., 2014)

κ  105 : (96)
To overcome the technical difficulties involved in observing such a very
small nonzero value of κ, Magana-Loaiza et al. (2016) designed a three-slit
setup illuminated with a polarization that allows for the excitation of SPs
traveling between the closely spaced slits. This enhances the near field
and thus increases the contribution of nonclassical paths. In Fig. 18 the
Poynting vector is shown for the case of only the leftmost slit A being illumi-
nated with an x-polarized field. It can be seen that the Poynting vector clearly
exhibits a loop-like behavior. This results in a strongly increased visibility of
the far-field interference pattern as compared to y-polarized illumination that
Young’s interference experiment 331

d A B C

|P| z

0 x
E

Fig. 18 Simulation of the normalized Poynting vector P in the vicinity of three slits. The
illumination assumes an x-polarized illumination that onto slit A. The Poynting vector is
seen to follow a looped trajectory. Reproduced from Magana-Loaiza, O. S., Leon, I. D.,
Mirhosseini, M., Fickler, R., Safari, A., Mick, U., …Boyd, R. W. (2016). Exotic looped trajectories
of photons in three-slit interference. Nature Communications, 13987, 159–162.

does not generate SPs. Using single heralded photons the individual terms of
Eq. (93) were measured. The observed value of κ  0.3, in agreement with
simulations and well above the level of uncertainty, demonstrates that under
certain conditions nonclassical paths do indeed occur.

10. Singular optics


The subfield of optics known as singular optics is concerned with the
phenomenon that at some points in space certain parameters that character-
ize a wavefield may be undefined, or singular, and that the singular structures
that arise in the vicinity of such points have a predictable and stable pattern.
Extensive overviews of singular optics can be found in Soskin and Vasnetsov
(2001), Angelsky (2007), Dennis, O’Holleran, and Padgett (2009), and Gbur
(2017). The best known type of singularity is the phase singularity which
occurs, for example, at the center of a Laguerre–Gauss beam. On axis the
intensity is zero, meaning that the phase of the complex amplitude is
undefined there. Characteristically, the phase increases by an amount of 
m2π along any closed contour that contains the beam axis, and where the inte-
ger m denotes the angular order of the beam. The index m also indicates the
so-called topological charge of the phase singularity. These singularities are
stable under small perturbations of the wavefield. In polychromatic fields,
phase singularities can be ‘hidden’ in the sense that at a certain position they
only occur for one particular frequency (Gbur, Visser, & Wolf, 2004b).
332 Greg Gbur and Taco D. Visser

A second type of singularity pertains to polarization ellipses of electro-


magnetic beams. Such an ellipse is characterized by three parameters, namely
the orientation of its major axis, its eccentricity, and its handedness. In regions
where the polarization is circular, the orientation of the ellipse is undefined.
Such C singularities occurs generically along lines in three-dimensional
space, and intersect the transverse plane at points; they are typically called
“C-points.” The generic types of C-points are shown in Fig. 19. In regions
where the polarization is linear, the handedness is undefined. For paraxial
fields, these L singularities typically occur on surfaces and intersect the plane
in lines, and are thus called “L-lines.” For nonparaxial fields, L singularities
occur on lines and manifest as L-points in a transverse plane (Berry &
Dennis, 2001).
The diffraction of vector vortex beams by a two-slit arrangement was
studied by Khan, Joshi, and Senthilkumaran (2021). Such beams are char-
acterized by the Poincare–Hopf index η, which is defined as
I
1
η≡ rγ  dl, (97)

here γ is the azimuth of the polarization ellipse, and the integral is along a
closed circuit enclosing the beam axis. The far-zone polarization morphol-
ogy was found to be lattice like, with the precise pattern depending on the
index η of the incident beam.
The occurrence of polarization singularities in an N-pinhole inter-
ferometer has been detailed in Schoonover and Visser (2009). For the case
of N ¼ 2 and with the fields at the two apertures assumed to be cophasal,

(a) (b) (c)


Fig. 19 Illustration of the three typical, or generic, types of polarization singularities:
(A) lemon, (B) star, and (C) monstar, showing the local orientation of the polarization
ellipses. These singularities are characterized by the number of separatrices (bold lines)
and the sense in which the major axis rotates in a closed path around the singularity.
Adapted from Gbur, G. J. (2017). Singular optics. Boca Raton: CRC Press.
Young’s interference experiment 333

with equal amplitude and orthogonal polarization, obviously no interfer-


ence fringes are formed. However, L-lines in the shape of semi-hyperbolas
are present in such a setup, whereas no points of circular polarization exist.
When N ¼ 3 and with the pinholes located at the vertices of an equilateral
triangle, a richer polarization behavior is observed. When the polarization at
the pinholes is taken to be radial, L-lines as well as C-points in the form of
stars and lemons are created. It was shown numerically that even when the
pinholes are not located symmetrically, the singularities remain present.
A third type of singularity is found in partially coherent wavefields. In
scalar wavefields the spectral degree of coherence (defined in Eq. (26)
μ(r1, r2, ω) can be zero. The pair of points r1 and r2 then form a coherence
singularity with the phase of μ being undefined (Schouten, Gbur, & Wolf,
2003). Likewise, in vector beams the electromagnetic degree of coherence
(see Eq. (57)) can be singular when η(r1, r2, ω) ¼ 0. To distinguish these
coherence singularities from their scalar counterparts they are sometimes
referred to as η singularities. Their occurrence in partially coherent beams
was described in Raghunathan, Schouten, and Visser (2012).
All classes of singularities discussed above can be annihilated or created in
topological reactions in which certain conservation laws must be satisfied;
this conservation for coherence singularities was discussed, for example,
in Raghunathan, Schouten, and Visser (2013). The three types of singular-
ities that we have mentioned occur generically in optical wavefields. That
means that no special preparation or symmetry in an experiment is needed
for them to exist.
It has gradually become clear that these different types of singularities are
not unrelated and that they can in fact transform into one another. For
example, in Flossmann, Schwarz, Maier, and Dennis (2005) the unfolding
of a phase singularity through a birefringent crystal was studied both theo-
retically and experimentally. The evolution of a coherence vortex into a
phase singularity was described in Gbur et al. (2004b). Young’s experiment
in two different forms has been crucial in fully elucidating these interrela-
tions, as we will now describe.
In Visser and Schoonover (2008) a two-pinhole setup was examined.
When the incident fields at the two apertures share the same linear polari-
zation and are spatially partially coherent, there will always be a multitude of
line pairs on the observation screen that are completely uncorrelated. These
lines therefore are coherence singularities. When the magnitude of the spec-
tral degree of coherence of the incident field is gradually increased, lines of
different pairs move closer together, and eventually annihilate to form phase
334 Greg Gbur and Taco D. Visser

singularities, at least within the approximations leading up to Eq. (10). If


next the direction of polarization of the fields at the two apertures is changed
from being parallel to being under a small angle, each phase singularity
unfolds into a triplet of lines. These triplets consist of an L-line flanked
by two C-lines of opposite handedness. However, this cascade of two-
dimensional singular field patterns is not generic. For example, no isolated
star C-points with their characteristic 120° symmetry can be formed in such
a geometry. A more general setup, consisting of a triangular arrangement of
pinholes as shown in Fig. 20, was studied in Pang, Gbur, and Visser (2015).
Both fields at the lower two pinholes are horizontally polarized, whereas the
field at the top pinhole is polarized under an angle θ. Pinholes 1 and 3 are
fully correlated. Pinholes 2 is partially correlated with the other two pin-
holes, as quantified by the parameter μ. By varying the orientation angle
θ and the correlation μ the system can be taken through a continuous
cycle in which three types of generic singularities can be produced on the
observation screen. The cycle begins with an array of phase singularities,
from which polarization singularities, correlation singularities, and general
electromagnetic correlation singularities are created.
Initially, the polarization angle θ ¼ 0 and the three fields are fully coher-
ent (μ ¼ 1). The ensuing spectral density, the modulus squared of Ex(x, y), is

x
(a) O
(b)
1

z
2 3

exp (i2 /3) exp (i4 /3)

Fig. 20 (A) Three pinholes in an opaque screen producing a hexagonal interference pat-
tern. (B) The fields at the bottom two pinholes are horizontally polarized, whereas the
field at the top pinhole is polarized under an angle θ. Pinholes 1 and 3 are fully corre-
lated. Pinhole 2 is partially correlated with the other two pinholes, quantified by the
parameter μ. Reproduced from Pang, X., Gbur, G., & Visser, T. D. (2015). Cycle of phase,
coherence and polarization singularities in Young’s three-pinhole experiment. Optics
Express, 23, 34093–34108.
Young’s interference experiment 335

Fig. 21 Color-coded plot of the spectral density (left), and the phase of Ex(x, y) when
θ ¼ 0 and μ ¼ 1. Reproduced from Pang, X., Gbur, G., & Visser, T. D. (2015). Cycle of phase,
coherence and polarization singularities in Young’s three-pinhole experiment. Optics
Express, 23, 34093–34108.

shown in the left panel of Fig. 21. The phase of Ex(x, y) is plotted on the
right. Several phase singularities, points where the different phase contours
intersect, can be seen. If now the angle θ is increased to a nonzero value, the
highly symmetric field distribution is perturbed, and the central phase sin-
gularity at x ¼ y ¼ 0 decays into a pair of two polarization singularities,
as illustrated in Fig. 22. The orientation of the polarization axis shows the
characteristic pattern of a lemon (top) and a star singularity (bottom). In
the right-hand panel contours of the corresponding normalized Stokes
parameters are plotted. The two C-points are separated by an L-line
(orange). Furthermore, it is seen that the two C-points have opposite hand-
edness, with s3 ¼ +1 and  1, respectively. By further changing the param-
eters θ and μ electromagnetic coherence singularities, zeros of η(r1, r2, ω),
can be created and annihilated in this configuration.
Beams of light can carry both spin and orbital angular momentum. The
spin angular momentum arises from the vectorial, or polarization, nature of
light, while the orbital angular momentum originates from the spatial distri-
bution of the wavefront. Laguerre-Gauss (LG) beams of light carry a well-
defined amount of orbital angular momentum (OAM) that is proportional
to the variation of the phase structure of the beam. An overview of different
methods to detect OAM is presented in Emile, Emile, and Brousseau (2019).
In Young’s double-slit experiment, the presence of OAM in the incident field
profoundly alters the observed interference pattern. This was demonstrated
336 Greg Gbur and Taco D. Visser

(a) (b)
0.01
0.01

y [mm]
y [mm]

0 0

-0.01
-0.01
-0.01 0 0.01 -0.01 0 0.01
x [mm] x [mm]
Fig. 22 (A) The local orientation of the major axis of the polarization ellipse after the
angle of polarization at pinhole 1 has been increased from zero to θ ¼ 0.03. The phase
singularity of Ex at (0, 0) in Fig. 21 has decayed into two polarization singularities: a
lemon (top) and a star (bottom). (B) Selected contours of the normalized Stokes param-
eters for the same region as in the left-hand panel: s1 ¼ 0 (blue), s2 ¼ 0 (green), s3 ¼ 0.998
(red), s3 ¼ 0.998 (black). Points with horizontal polarization (s3 ¼ 0) are represented by
the orange curve. Reproduced from Pang, X., Gbur, G., & Visser, T. D. (2015). Cycle of phase,
coherence and polarization singularities in Young’s three-pinhole experiment. Optics
Express, 23, 34093–34108.

experimentally in Sztul and Alfano (2006) in which the phase singularity on


the beam axis of an LG beam was positioned between the slits. The variation
of the phase difference along the direction of the slits produces a twist of the
fringes, with the direction of the twist being opposite for beams with a topo-
logical charge of l ¼ +1 and l ¼ 1, respectively. This twist then can be used
to detect the presence of OAM. Simulations of fringes produced by beams
with higher topological charge were reported in Zhou, Yan, Dong, and
Zhang (2014). Another approach to detect OAM in LG beams was described
in Berkhout and Beijersbergen (2008). There a multipinhole configuration,
with the apertures divided uniformly over a circle, was shown to produce
intricate interference patterns that strongly depend on both the number of
pinholes as well as the azimuthal mode index l. The authors suggest that
the scalability of this detection method makes it potentially useful for situa-
tions where the optical vortices are expected to be large, such as in astrophys-
ics. In a variation of this approach (Emile, Emile, de Lesegno, Pruvost, &
Brousseau, 2015), one uses a series of different screens, each with a different
number of pinholes. In that case the OAM can be deduced from just the
intensity on the central axis that each screen produces. An inherent restriction
Young’s interference experiment 337

of this “hole wheel” method is that the number of point apertures limits the
upper bound of the topological charge that can be distinguished. This can be
partly remedied by distributing the apertures nonuniformly over a circle, as
was demonstrated numerically in Shi, Tian, and Chen (2012).

11. Conclusions
After conducting an informal poll among physicists, asking them to
name the experiment they thought was the most beautiful, Crease
(Crease, 2003) described the 10 most popular choices in his engaging book
The Prism and the Pendulum. Remarkably, Young’s experiment is listed
twice. The top spot was taken by the double-slit experiment with electrons
carried out by J€onsson (1961), and the original optical experiment, per-
formed by Young himself, ended in fifth place. It goes without saying that
the present authors feel that these accolades are well deserved. Indeed, it is
more than impressive that more than two centuries after its inception
Young’s experiment still plays a fundamental role in the development of
quantum mechanics, coherence theory, singular optics, and plasmonics.
The much maligned Thomas Young has left us a legacy that stands out in
the history of physics.

Acknowledgments
We wish to thank Dr. Hugo Schouten, Prof. Joe Eberly, Prof. Dorinda Outram, Prof. Ari
Friberg, Prof. Pang Xiaoyan, and the late Prof. Emil Wolf for the many discussions that have
helped shape our understanding.

References
Agarwal, G. S., Dogariu, A., Visser, T. D., & Wolf, E. (2005). Generation of complete coher-
ence in Young’s interference experiment with random mutually uncorrelated electro-
magnetic beams. Optics Letters, 30, 120–122.
Aharonov, Y., & Bohm, D. (1959). Significance of electromagnetic potentials in the quan-
tum theory. Physical Review, 115, 485–491.
Ambrosini, D., Gori, F., & Paoletti, D. (2005). Destructive interference from three partially
coherent point sources. Optics Communications, 254(1), 30–39.
Angelsky, O. V. (2007). Optical correlation: Techniques and applications. Bellingham,
Washington: SPIE.
Arago, F. (1819). Rapport fait par M. Arago. Annales de Chimie et de Physique, 11, 5–30.
Arago, F. (1835). Biographical memoir of Dr Thomas Young. The Edinburgh New
Philosophical Journal, 20, 213–240.
Arago, F., & Fresnel, A. J. (1819). L’action que les rayons de lumière polarisee exercent les uns
sur les autres. Annales de Chimie et de Physique, 10, 288–300.
Arnd, M., Aspelmeyer, M., & Zeilinger, A. (2002). How to extend quantum experiments.
Fortschritte der Physik, 57, 1153–1162.
338 Greg Gbur and Taco D. Visser

Arndt, M., Nairz, O., Vos-Andreae, J., Keller, C., van der Zouw, G., & Zeilinger, A. (1999).
Wave–particle duality of C60 molecules. Nature, 401, 680–682.
Bach, R., Pope, D., Liou, S. H., & Batelaan, H. (2013). Controlled double-slit electron dif-
fraction. New Journal of Physics, 15, 033018.
Basano, L., & Ottonello, P. (2005). Complete destructive interference of partially coherent
sources of acoustic waves. Physical Review Letters, 94(17), 173901.
Berkhout, G. C. G., & Beijersbergen, M. W. (2008). Method for probing the orbital angular
momentum of optical vortices in electromagnetic waves from astronomical objects.
Physical Review Letters, 101, 100801.
Berry, M. V. (1984). Quantal phase factors accompanying adiabatic changes. Proceedings of the
Royal Society of London. Series A, Mathematical and Physical Sciences, 392, 45–57.
Berry, M. V., & Dennis, M. R. (2001). Polarization singularities in isotropic random vector
waves. Proceedings of The Royal Society A Mathematical Physical and Engineering Sciences, 457,
141–155.
Bethe, H. A. (1944). Theory of diffraction by small holes. The Physical Review, 66, 163–182.
Born, M., & Wolf, E. (1999). Principles of Optics. Cambridge: Cambridge University Press.
Brandt, D., & Hirschi, S. (1974). Electron diffraction at multiple slits. American Journal of
Physics, 42, 4–11.
Brezger, B., Hackerm€ uller, L., Uttenthaler, S., Petschinka, J., Arndt, M., &
Zeilinger, A. (2002). Matter-wave interferometer for large molecules. Physical Review
Letters, 88, 100404.
Brosseau, C. (1998). Fundamentals of polarized light. New York: John Wiley and Sons.
Brougham, H. P. (1803). Art. XVI. The Edinburgh Review, 1, 450–460.
Cao, Q., & Lalanne, P. (2002). Negative role of surface plasmons in the transmission of metal-
lic gratings with very narrow slits. Physical Review Letters, 88, 057403.
Carnal, O., & Mlynek, J. (1991). Young’s double-slit experiment with atoms: A simple atom
interferometer. Physical Review Letters, 66, 2689–2692.
Chambers, R. G. (1960). Shift of an electron interference pattern by enclosed magnetic flux.
Physical Review Letters, 5, 3–5.
Chapman, M. S., Ekstrom, C. R., Hammond, T. D., Rubenstein, R. A., Schmiedmayer, J.,
Wehinger, S., & Pritchard, D. E. (1995). Optics and interferometry with Na2 molecules.
Physical Review Letters, 74, 4783–4786.
Crease, R. P. (2003). The Prism and the Pendulum. New York: Random House.
Davisson, C. J. (1928). Are electrons waves? Journal of the Franklin Institute, 206, 597–623.
Davisson, C. J., & Gerner, L. H. (1927). The scattering of electrons by a single crystal of
Nickel. Nature, 119, 558–560.
de Broglie, L.-V. (2004). On the theory of quanta. Privately published translation by A.F.
Kracklauer.
Dennis, M. R., O’Holleran, K., & Padgett, M. J. (2009). Singular optics: Optical vortices and
polarization singularities. In E. Wolf (Ed.), Progress in optics: Vol. 53 (pp. 293–363).
Amsterdam: Elsevier.
Dimitrova, T. L., & Weis, A. (2008). The wave-particle duality of light: A demonstration
experiment. American Journal of Physics, 76, 137–142.
Divitt, S., Frimmer, M., Visser, T. D., & Novotny, L. (2016). Modulation of optical spatial
coherence by surface plasmon polaritons. Optics Letters, 41, 3094–3097.
Ebbesen, T. W., Lezec, H. J., Ghaemi, H. F., Thio, T., & Wolff, P. A. (1998). Extraordinary
optical transmission through sub-wavelength hole arrays. Nature, 391, 667–669.
Eberly, J. H., Qian, X.-F., & Vamivakas, A. N. (2017). Polarization coherence theorem.
Optica, 4, 1113–1114.
Eckert, M. (2012). Max von Laue and the discovery of X-ray diffraction in 1912. Annalen der
Physik, 524, A83–A85.
Young’s interference experiment 339

Eibenberger, S., Gerlich, S., Arndt, M., Mayor, M., & T€ uxen, J. (2019). Matter-wave inter-
ference of particles selected from a molecular library with masses exceeding 10, 000 amu.
Physical Chemistry Chemical Physics, 15, 14696.

Einstein, A. (1905). Uber einen die Erzeugung und Verwandlung des Lichtes betreffenden
heuristischen Gesichtspunkt. Annalen der Physik, 17, 132–148.
Emile, O., Emile, J., & Brousseau, C. (2019). Detection of the orbital angular momentum in optics.
U. Rennes.
Emile, O., Emile, J., de Lesegno, B. V., Pruvost, L., & Brousseau, C. (2015). Analysis of the
topological charge of vortex beams using a hole wheel. EPL, 111, 34001.
Estermann, I., & Stern, O. (1930). Beugung von Molekularstrahlen. Zeitschrift f€ ur Physik, 61,
95–125.
Fein, Y. Y., Geyer, P., Zwick, P., Kialka, F., Pedalino, S., Mayor, M., … Arndt, M. (2019).
Quantum superposition of molecules beyond 25 kDa. Nature Physics, 15, 1242–1245.
Feynman, R., Leighton, R., & Sands, M. (1963). Vol. I. The feynman lectures on physics.
Reading Mass: Addison-Wesley.
Flossmann, F., Schwarz, U. T., Maier, M., & Dennis, M. R. (2005). Polarization singularities
from unfolding an optical vortex through a birefringent crystal. Physical Review Letters,
95, 253901.
Forbes, A., Aiello, A., & Ndagano, B. (2019). Classically entangled light. In T. D. Visser
(Ed.), Vol. 64. Progress in optics (pp. 99–153).
Friedrich, W., Knipping, P., & von Laue, M. (1912). Diffraction. Sber. Bayer. Akad. Wiss.,
524, 303–304.
Gan, C. H., & Gbur, G. (2007). Phase and coherence singularities generated by the interfer-
ence of partially coherent fields. Optics Communications, 280(2), 249–255.
Gan, C. H., & Gbur, G. (2008). Spatial coherence conversion with surface plasmons using a
three-slit interferometer. Plasmonics, 3, 111.
Gan, C. H., Gbur, G., & Visser, T. D. (2007). Surface plasmons modulate the spatial coher-
ence of light. Physical Review Letters, 98, 043908.
Gan, C. H., Gu, Y., Visser, T. D., & Gbur, G. (2012). Coherence converting plasmonic hole
arrays. Plasmonics, 7, 313–322.
Garcia-Vidal, F. J., Martin-Moreno, L., Ebbesen, T. W., & Kuipers, L. (2010). Light passing
through subwavelength apertures. Reviews of Modern Physics, 82, 729–787.
Gbur, G., & Smith, M. (2021). Controlled coherence plasmonic light sources. Photonics,
8, 268.
Gbur, G., Visser, T. D., & Wolf, E. (2004a). Complete destructive interference of partially
coherent fields. Optics Communications, 239(1), 15–23.
Gbur, G., Visser, T. D., & Wolf, E. (2004b). Hidden singularities in partially coherent
wavefields. Journal of Optics A Pure and Applied Optics, 6, S239–S242.
Gbur, G. J. (2017). Singular Optics. Boca Raton: CRC Press.
Genet, C., & Ebbesen, T. W. (2007). Light in tiny holes. Nature, 445, 39–46.
Geyer, P., Sezer, U., Rodewald, J., Mairhofer, L., D€ orre, N., Haslinger, P., …
Arndt, M. (2016). Perspectives for quantum interference with biomolecules and biomo-
lecular clusters. Physica Scripta, 91, 063007.
Goodman, J. W. (1996). Introduction to Fourier Optics. New York: McGraw-Hill.
Gori, F., Santarsiero, M., & Borghi, R. (2006). Vector mode analysis of a Young interfer-
ometer. Optics Letters, 31(7), 858–860.
Gori, F., Santarsiero, M., Borghi, R., & Wolf, E. (2006). Effects of coherence on the degree
of polarization in a Young interference pattern. Optics Letters, 31(6), 688–690.
Gurney, H. (1831). Memories of the life of Thomas Young. London: John and Arthur.
Hannonen, A., Partanen, H., Tervo, J., Set€al€a, T., & Friberg, A. T. (2019).
Pancharatnam-Berry phase in electromagnetic double-pinhole interference. Physical
Review A, 99, 053826.
340 Greg Gbur and Taco D. Visser

Hilts, V. L. (1978). Thomas Young’s autobiographical sketch. Proceedings of the American


Philosophical Society, 122, 248–260.
Jacques, V., Wu, E., Grosshans, F., Treussart, F., Grangier, P., Aspect, A., & Roch, J. F.
(2007). Experimental realization of Wheeler’s delayed-choice Gedanken experiment.
Science, 315, 966–968.
Jaeger, G., Horne, M. A., & Shimoney, A. (1993). Complementarity of one-particle and
two-particle interference. Physical Review A, 48, 1023–1027.
James, D. F. V., & Wolf, E. (1991a). Some new aspects of Young’s interference experiment.
Physics Letters A, 157, 6–10.
James, D. F. V., & Wolf, E. (1991b). Spectral changes produced in Young’s interference
experiment. Optics Communications, 81, 150–154.
J€
onsson, C. (1961). Elektroneninterferenzen an mehreren k€ unstlich hergestellten
Feinspalten. Zeitschrift f€ur Physik, 161, 454–474.
Kanseri, B., & Sethuraj, K. R. (2019). Experimental observation of the polarization coher-
ence theorem. Optics Letters, 44, 159–162.
Khan, S. N., Joshi, S., & Senthilkumaran, P. (2021). Young’s double-slit experiment with
vector vortex beams. Optics Letters, 46, 4136–4139.
Korotkova, O., & Gbur, G. (2020). Applications of optical coherence theory. In T. D. Visser
(Ed.), Progress in optics: Vol. 65 (pp. 43–104). Elsevier.
Kuzmin, N., ’t Hooft, G. W., Eliel, E. R., Gbur, G., Schouten, H. F., & Visser, T. D. (2007).
Enhancement of spatial coherence by surface plasmons. Optics Letters, 32, 445–447.
Leinonen, A., Saastamoinen, K., Pesonen, H., Wu, G. F., Visser, T. D., Turunen, J., &
Friberg, A. T. (2021). Polarization modulation by surface plasmons in Young’s
double-slit setup. Physical Review A, 104, 043503.
Li, D., & Pacifici, D. (2017). Strong amplitude and phase modulation of optical spatial coher-
ence with surface plasmon polaritons. Science Advances, 3, e1700133.
Li, Y., Wang, X.-L., Zhao, H., Kong, L.-J., Lou, K., Gu, B., … Wang, H.-T. (2012).
Young’s two-slit interference of vector light fields. Optics Letters, 37, 1790–1792.
Lichte, H. (2002). Electron interference: Mystery and reality. Philosophical Transactions. Series
A, Mathematical, Physical, and Engineering Sciences, 360, 897–920.
Lv, H., Zhou, T., Wang, F., Chen, Y., Cai, Y., & Korotkova, O. (2021). Young’s interfer-
ence experiment for generating light with non-uniform coherence states. Optics Letters,
46(3), 693–696.
Magana-Loaiza, O. S., Leon, I. D., Mirhosseini, M., Fickler, R., Safari, A., Mick, U., …
Boyd, R. W. (2016). Exotic looped trajectories of photons in three-slit interference.
Nature Communications, 13987, 159–162.
Magyar, G., & Mandel, L. (1963). Interference fringes produced by superposition of two
independent maser light beams. Nature, 198, 255–256.
Mandel, L., & Wolf, E. (1976). Spectral coherence and the concept of cross-spectral purity.
Journal of the Optical Society of America, 66, 529–535.
Mandel, L., & Wolf, E. (1995). Optical coherence and quantum optics. Cambridge University
Press.
Manning, A. G., Khakimov, R. I., Dall, R. G., & Truscott, A. G. (2013). Wheeler’s
delayed-choice gedanken experiment with a single atom. Nature Physics, 11, 539–542.
Merli, P. G., Missiroli, G. F., & Pozzi, G. (1976). On the statistical aspect of electron inter-
ference phenomena. American Journal of Physics, 44, 306–307.
M€ ollenstedt, G., & D€ ucker, H. (1956). Beobachtungen und Messungen an
Biprisma-lnterferenzen mit Elektronenwellen. Zeitschrift fr Physik, 145, 377–397.
Mollon, J. D. (2002). The origins of the concept of interference. Philosophical Transactions.
Series A, Mathematical, Physical, and Engineering Sciences, 360, 807–819.
Mujat, M., Dogariu, A., & Wolf, E. (2004). A law of interference of electromagnetic beams
of any state of coherence and polarization and the Fresnel–Arago interference laws.
Journal of the Optical Society of America. A, 21(12), 2414–2417.
Young’s interference experiment 341

Nairz, J., Arndt, M., & Zeilinger, A. (2002). Quantum interference experiments with large
molecules. American Journal of Physics, 71, 319–325.
Newton, I. (1704). Opticks. London: Smith and Walford.
Pang, X., Gbur, G., & Visser, T. D. (2015). Cycle of phase, coherence and polari-
zation singularities in Young’s three-pinhole experiment. Optics Express, 23,
34093–34108.
Parker, S. (1971). A single-photon double-slit interference experiment. American Journal of
Physics, 39(4), 420–424.
Peacock, G. (1855). Life of Thomas Young. London: J. Muray.
Peacock, G. (Ed.). (1855). Miscellaneous works of the late Thomas Young. London: J. Murray.
Pfleegor, R. L., & Mandel, L. (1967). Interference of independent photon beams. The
Physical Review, 159, 1084–1088.
Pu, J., Cai, C., & Nemoto, S. (2004). Spectral anomalies in Young’s double-slit interference
experiment. Optics Express, 12, 5131–5139.
Qian, X.-F., Malhotra, T., Vamivakas, A. N., & Eberly, J. H. (2016). Coherence constraints
and the last hidden optical coherence. Physical Review Letters, 117, 153901.
Qian, X.-F., Vamivakas, A. N., & Eberly, J. H. (2018). Entanglement limits duality and vice
versa. Optica, 5, 942–947.
Raedt, H. D., Michielsen, K., & Hess, K. (2012). Analysis of multipath interference in
three-slit experiments. Physical Review A, 85, 012101.
Raether, H. (1988). Surface plasmons on smooth and rough surfaces and on gratings. Berlin:
Springer.
Raghunathan, S. B., Schouten, H. F., & Visser, T. D. (2012). Correlation singularities in
partially coherent electromagnetic beams. Optics Letters, 37, 4179–4181.
Raghunathan, S. B., Schouten, H. F., & Visser, T. D. (2013). Topological reactions of
correlation functions in partially coherent electromagnetic beams. Journal of the Optical
Society of America, 30, 582–588.
Ritchie, R. H. (1957). Plasma losses by fast electrons in thin films. The Physical Review, 106,
874–881.
Robinson, A. (2005). The last man who knew everything. New York: Pi Press.
Rosenbury, C., Gu, Y., & Gbur, G. (2012). Phase singularities, correlation singularities, and
conditions for complete destructive interference. Journal of the Optical Society of America A,
29(4), 410–416.
Rueckner, W., & Titcomb, P. (1996). A lecture demonstration of single photon interference.
American Journal of Physics, 64(2), 184–188.
Saastamoinen, T., & Lajunen, H. (2013). Increase of spatial coherence by subwavelength
metallic gratings. Optics Letters, 38, 5000–5003.
Sala, S., Ariga, A., Ereditato, A., Ferragut, R., Giammarchi, M., Leone, M., …
Scampoli, P. (2019). First demonstration of antimatter wave interferometry. Science
Advances, 5, eaav7610.
Santarsiero, M., & Gori, F. (1992). Spectral changes in a Young interference pattern. Physics
Letters A, 167, 123–128.
Sawant, R., Samuel, J., Sinha, A., Sinha, S., & Sinha, U. (2014). Nonclassical paths in quan-
tum interference experiments. Physical Review Letters, 113, 120406.
Sch€ollkopf, W., & Toennies, J. P. (1994). Non-destructive mass selection of small van der
Waals clusters. Science, 266, 1345–1348.
Schoonover, R. W., & Visser, T. D. (2009). Creating polarization singularities with an
N-pinhole interferometer. Physical Review A, 79, 043809.
Schouten, H. F., Gbur, G., & Wolf, T. D. V. E. (2003). Phase singularities of the coherence
functions in Young’s interference pattern. Optics Letters, 28, 968–970.
Schouten, H. F., Kuzmin, N., Dubois, G., Visser, T. D., Gbur, G., Alkemade, P. F. A., …
Eliel, E. R. (2005). Plasmon-assisted two-slit transmission: Young’s experiment
revisited. Physical Review Letters, 94, 053901.
342 Greg Gbur and Taco D. Visser

Schouten, H. F., Visser, T. D., & Wolf, E. (2003). New effects in Young’s interference
experiment with partially coherent light. Optics Letters, 28(14), 1182–1184.
Scully, M. O., Englert, B.-G., & Walther, H. (1991). Quantum optical tests of complemen-
tarity. Nature, 351, 111–116.
Set€al€a, T., Tervo, J., & Friberg, A. T. (2006a). Contrasts of Stokes parameters in Young’s
interference experiment and electromagnetic degree of coherence. Optics Letterrs,
31(18), 2669–2671.
Set€al€a, T., Tervo, J., & Friberg, A. T. (2006b). Stokes parameters and polarization contrasts in
Young’s interference experiment. Optics Letters, 31(14), 2208–2210.
Shapere, A., & Wilczek, F. (1989). Geometric phases in physics. Singapore: World Scientific.
Shi, L., Tian, L., & Chen, X. (2012). Characterizing topological charge of optical vortex
using non-uniformly distributed multi-pinhole plate. Chinese Optics Letters, 10, 120501.
Sinha, U., Couteau, C., Jennewein, T., Laflamme, R., & Weihs, G. (2010). Ruling out
multi-order interference in quantum mechanics. Science, 329, 418–421.
Smith, M., & Gbur, G. (2019). Coherence resonances and band gaps in plasmonic hole arrays.
Physical Review A, 99, 023812.
Sorkin, R. D. (1994). Quantum mechanics as quantum measure theory. Modern Physics Letters
A, 33, 3119–3127.
Soskin, M. S., & Vasnetsov, M. V. (2001). Singular optics. In E. Wolf (Ed.), Progress in optics:
Vol. 42 (p. p. 219). Amsterdam: Elsevier.
Spreeuw, R. J. C. (1998). A classical analogy of entanglement. Foundations of Physics, 28,
361–374.
Sztul, H. I., & Alfano, R. R. (2006). Double-slit interference with Laguerre-Gaussian beams.
Optics Letters, 31, 999–1001.
Taylor, G. I. (1909). Interference fringes with weak light. Proceedings—Cambridge Philosophical
Society, 15, 114–115.
Tervo, J., Set€al€a, T., & Friberg, A. T. (2003). Degree of coherence for electromagnetic fields.
Optics Express, 11(10), 1137–1143.
Thomson, G. P., & Reid, A. (1927). Diffraction of cathode rays by a thin film. Nature,
119, 890.
Thomson, J. J. (1906-1908). On the ionization of gases by ultra-violet light and on the evi-
dence as to the structure of light afforded by its electrical effects. Proceedings of the
Cambridge Philosophical Society, 14, 417–424.
Tonomura, A., Endo, J., Matsuda, T., Kawasaki, T., & Ezawa, H. (1989). Demonstration of
single-electron buildup of an interference pattern. American Journal of Physics, 57,
117–120.
Tonomura, A., Osakabe, N., Matsuda, T., Kawasaki, T., Endo, J., Yano, S., &
Yamada, H. (1986). Evidence for Aharonov-Bohm effect with magnetic field
completely shielded from electron wave. Physical Review Letters, 56, 792–801.
Verdet, E. (1869). Leçons d’Optique Physique. Paris: G. Masson.
Visser, T. D., & Schoonover, R. W. (2008). A cascade of singular field patterns in Young’s
interference experiment. Optics Communications, 281, 1–6.
von Bergmann, J., & von Bergmann, H. (2007). Foucault pendulum through basic geometry.
American Journal of Physics, 75, 888–892.
Walborn, S. P., Cunha, M. O. T., Pádua, S., & Monken, C. H. (2002). Double-slit quantum
eraser. Physical Review A, 65, 033818.
Weis, A., & Wynands, R. (2003). Three demonstration experiments on the wave and particle
nature of light. Physik und Didaktik in Schule und Hochschule, I,II, 67–73.
Wheeler, J. A., & Zurek, W. H. (Eds.). (1983). Quantum theory and measurement. Princeton:
Princeton University Press.
Whittaker, E. T. (1951). A history of the theories of aether and electricity, 2 Vols. London: Nelson.
Wolf, E. (1954). Optics in terms of observable quantities. Nuovo Cimento, 12, 884–888.
Young’s interference experiment 343

Wolf, E. (1959). Coherence properties of partially polarized electromagnetic radiation.


Nuovo Cimento, 13, 1165–1181.
Wolf, E. (2003). Unified theory of coherence and polarization of random electromagnetic
beams. Physics Letters A, 312, 263–267.
Wolf, E. (2007a). The influence of Young’s interference experiment on the development of
statistical optics. In E. Wolf (Ed.), Vol. 50. Progress in optics (pp. 251–273). Amsterdam:
Elsevier.
Wolf, E. (2007b). Introduction to the theory of coherence and polarization of light. Cambridge
University Press.
Wootters, W. K., & Zurek, W. H. (1979). Complementarity in the double-slit experiment:
Quantum nonseparability and a quantitative statement of Bohr’s principle. Physical
Review D, 19, 473–484.
Young, T. (1793). Observations on vision. Philosophical Transactions of the Royal Society of
London, 83, 169–181.
Young, T. (1800). Outlines of experiments and inquiries respecting sound and light.
Philosophical Transactions of the Royal Society of London, 90, 106–150.
Young, T. (1801). On the mechanism of the eye. Philosophical Transactions of the Royal Society
of London, 91, 23–88.
Young, T. (1802). An account of some cases of the production of colours, not hitherto
described. Philosophical Transactions of the Royal Society of London, 92, 387–397.
Young, T. (1802). On the theory of light and colours. Philosophical Transactions of the Royal
Society of London, 92, 12–48.
Young, T. (1804). Experiments and calculations relative to physical optics. Philosophical
Transactions of the Royal Society of London, 94, 1–16.
Young, T. (1804). A reply to the Animadversions of the Edinburgh reviewers. London: Longman,
Hurst, Ress, and Orme.
Young, T. (1826). A formula for expressing the decrement of human life. Philosophical
Transactions of the Royal Society of London, 116, 281–303.
Young, T. (1845). A course of lectures on natural philosophy and the mechanical arts. London:
Taylor and Walton.
Zeilinger, A., G€ahler, R., Shull, C. G., Treimer, W., & Mampe, W. (1988). Single- and
double-slit diffraction of neutrons. Reviews of Modern Physics, 60, 1067–1073.
Zernike, F. (1938). The concept of degree of coherence and its application to optical prob-
lems. Physica, 5, 785–795.
Zhou, H., Perreault, W. E., Mukherjee, N., & Zare, R. N. (2021). Quantum mechanical
double slit for molecular scattering. Science, 374, 960–964.
Zhou, H., Yan, S., Dong, J., & Zhang, X. (2014). Double metal subwavelength slit arrays
interference to measure the orbital angular momentum and the polarization of light.
Optics Letters, 39, 3173–3176.
Zia, R., & Brongersma, M. (2007). Surface plasmon polariton analogue to Young’s
double-slit experiment. Nature Nanotechnology, 2, 426–429.

View publication stats

You might also like