You are on page 1of 11

Imaging using an Induced Coherence Setup

Nicholas Nardelli
Electrical & Computer Engineering, Boston University

April 20, 2015

Abstract
In this paper I analyze the work of Lemos et al. [11] which uses two-photon spatial entanglement and the indistinguishability of two nonlinear crystal sources to create an imaging
apparatus that can probe an object at one wavelength and collect data at a second wavelength. I first explain the fundamental physics underlying this apparatus and then analyze
the experiment in the broader context of quantum optical imaging.

in order to invent or improve technology. For example, recent technological innovation has occurred in
the areas of communication, metrology and lithography [13]. One of the most investigated applications of
quantum optics is Quantum Key Distribution (QKD),
which relies on entangled photon pairs and provides a
completely secure form of communication. Although
still in its infancy, this application is being used commercially for short-range encryption [21]. Quantum
effects are also used to reach the Standard Quantum
Limit (SQL) of measurement and even below by using
clever feedback and feed-forward setups [9]. Quantum
lithography has allowed for optical imaging of features
smaller than the wavelength of light [5].
The experiment conducted by Lemos et al. [11] employs both superposition and entanglement to realize
an imaging apparatus. I will introduce and briefly
explain these quantum effects and how they relate to
second and fourth order interference. I will then explain
how entanglement can be created using a nonlinear
medium as well as how superposition and entanglement
can be used to image an unknown object. I will then
conclude with a discussion of the shortcomings of this
approach and a few logical extensions to this research
topic.

introduction

For most of our experiences the world conforms to a


set of rules for which we have a solid innate intuition.
However, when considering the physics of the very
small, quantum rules dominate which are very counterintuitive. Since the development of quantum mechanics
in the early 20th century, researchers have relied on
mathematical models to provide insight into this reality, which often times presents paradoxes which force
a choice between fundamental axioms such as the principle of locality and Heisenbergs uncertainty principle
or causality.
In recent years, due to the accelerated development
of technology, researchers have investigated experimentally the "spookiness" of quantum mechanics, especially in the realm of optics. Superposition and interference which are generally thought of as a classical
wave phenomena, are fundamental to quantum mechanics and are the basis for many recent experiments
highlighting the quantum nature of light [12]. Furthermore, the idea of entanglement, which has been very
controversial in the past and has no part in classical
physics, plays an obvious role in quantum optics and
is readily created in the laboratory.
In some cases, these quantum effects can be exploited
1

background

One of the most interesting applications of quantum


optics is imaging, which relies on both second and
fourth order interference of electric fields as well as the
less intuitive concept of two-particle entanglement.
2.1

Superposition

The canonical example of quantum superposition is


the quantum version of Thomas Youngs double slit
experiment [20], which is set up so that a single photon
source illuminates two parallel slits in an otherwise
opaque material. The photon can either go through
one slit or the other, and after it passes through this
structure, a camera that is sensitive to single photons
records its transverse location. By placing a detector
right at the exit of one of these slits, the experimenter
can determine which slit the photon went through.
Consequently, the photon arrives at the camera in one
of two places dictated by which slit it went through.
If, on the other hand, the path remains unknown,
the photon interferes with itself and has a transverse
probability distribution at the camera proportional to
the amplitude in wave interference. This is the famous
wave-particle duality which highlights the importance
of information in quantum mechanics.
This same scenario is setup in Lemos et al., except
a Mach-Zehnder interferometer is used (see Figure 1)
instead of two slits. Thus, a photon enters one arm
or the other at a beam splitter (BS1) and recombines
with the other path at a second beam splitter (BS2).
If there is no way to measure which arm the photon
went through, then the photon goes through both
arms simultaneously and thus interferes with itself on
BS2. Mathematically, this is a superposition of two
states (|ai and |bi) each of which describes propagation
through one of the interferometer arms. The total state
is described as
1
|i = (|ai + ei |bi)
2

(1)

The phase difference between the arms is . Another

Figure 1: In a Mach-Zehnder interferometer with beam


splitters BS1 and BS2 that reflect and transmit 50%, a
single photon is in a superposition of going in path a or path
b. Depending on the difference in optical path length defined
by the optical phase delay in one arm (), the probability
that the photon emerges from port 1 or port 2 is changed.
The right plot shows the probability of emerging from port 1
or port 2 at BS2 as a function of phase delay.

way to express this superposition is that a one-photon


state (|1i) and a vacuum state (|0i) are equally probable in both arms. Equation (1) can be rewritten
as
1
|i = (|1ia |0ib + ei |0ia |1ib )
(2)
2
where subscripts a and b denote the state existing
in path a or path b. Depending on the difference in
optical path length between the two arms (), the
photon will exit from port 1 or port 2 with a probability
P1 and P2 predicted by the degree of destructive and
constructive phase interference,
1
P1,2 = (1 cos ).
2

(3)

This is illustrated by the plot in Figure 1.


When turning up the power (i.e., adding more photons to the system), only the wave aspects of light
are observed. The lights electric field amplitude is
functionally equivalent to the probability amplitude
of the single photon state. In other words, the same
results can be seen in a single-photon system as in a
many-photon system simply by integrating over a long
period of time [1]. The probability of detection in port
1 or port 2 is observed as intensity in the classical limit,
and has the exact form of equation (3).
2

Figure 2: One photon (p ) is split into two photons (i


and s ) by a (2) crystal, one in state |Hi and the other
in state |V i. Each output photon has half the energy of the
input photon. One photon travels in path a and the other
in path b.

2.2

Discrete variable entanglement

Entanglement relies on the fact that the state of one


particle cannot be described independently of its sister particle. Thus, entangled photons occupy a nonseparable joint state. The mathematical meaning of
this is that the two-photon state cannot be separated
into a product of states (i.e., non-separability). To illustrate this point, I will start with the case where one
photon splits into two photons: one in the horizontal
polarization state |Hi and the other in the vertical
polarization state |V i and each with half the energy
of the input photon (
hp /2). The state of the system
shown in Figure 2 is
1
|i = (|Hia |V ib + |V ia |Hib )
2

(4)

Each photon takes a different path denoted by subscript


a and b and is either horizontally polarized or vertically
polarized with equal probability. Before we measure
what polarization a photon has in one of the paths,
the two photons are linked via this non-separability
of polarization states. This correlation enables us to
instantly obtain the polarization of both photons if
we are able to determine the polarization of just one
photon.
This is the basis for the EPR paradox: we could
measure a property of one of these photons and another non-commuting property of the other photon
and have complete knowledge of the original state [15].
For instance, we could measure photon #1s position

precisely which in turn yields information about the


position of photon #2 since the pair was born in the
same place. Then, the momentum of photon #2 could
be measured precisely which gives us complete knowledge of position and momentum. This, however, would
violate the Heisenberg uncertainty principle. We can
fix this by deciding that the photon pair can communicate instantaneously over the separation distance,
which protects the uncertainty principle but violates
the principle of locality [10].
This paradox forces a choice between uncertainty
and locality. In 1964, John Stewart Bell developed
a test that showed that entangled particles can be
tied together non-locally [4]. Thus, in the polarization
example, each photon is in a superposition of horizontal and vertical states until one photon is measured.
It is only after the measurement that the two photon
states collapse into their separate anti-correlated states.
We can ascertain if a state is entangled using a Bellstate measurement, which tells us to what degree the
two photon states are correlated. If this correlation
value exceeds a certain value (2) then the state cannot be classical. This idea that non-local states exist
in quantum mechanics has withstood over 50 years
experimentation.
Note that the theory of relativity is not violated in
the "communication" between these entangled photons.
This is due to the fact that the polarization states are
correlated in a single pair but are completely random
from one pair to the next. There is no conceivable
way to send information superluminally due to this
inherent randomness. This is the non-communication
theorem [14].
2.3

Continuous variable entanglement

The previous section considered discrete variable entanglement, where photons were anti-correlated in polarization states |Hi and |V i. This section considers
continuous variable entanglement where a state is described by a continuous probability distribution and
its sister photon is either correlated or anti-correlated.
Consider again a parametric process (i.e. quantities
3

Figure 3: Demonstrating that photons originating from


one location are recombined on the object plane and that
photons with the same momentum are imaged to the same
point on the Fourier plane, a focal length f from the lens.

such as energy and momentum of the light must be


conserved and nothing is lost to the material [6]) that
converts a single photon into two lower-energy photons.
Since entangled photons are born together, they will
originate from the same location and have transverse
momenta that add up to the original transverse momentum. For example, in the case of zero original
transverse momentum the photon pair exhibits equal
and opposite transverse momenta. Using a lens, these
two photons can be projected onto the Fourier plane
where their momenta will be anti-correlated, or onto
the image plane where their positions will be correlated
[16]. This concept is illustrated by blue and red rays
in Figure 3.
The two-photon spatially-entangled state depicted
in Figure 3 at the image plane is expressed as
ZZ
|i =
dr1 dr2 (r1 , r2 )|1r1 i|1r2 i(r1 r2 ) (5)
where |1r1 i and |1r2 i describe the individual photon
spatial states where r1 and r2 are the transverse position vectors of each separated photon and the delta
function indicates that these states are always coincident. The probability amplitude (r1 , r2 ) describes
the probability of detecting a photon at a transverse
location and is dependent on the second-order nonlinear coefficient (2) and the degree of phase-mismatch
between pump, signal and idler waves [10]. This prob-

ability must be unity RR


when integrated over all possible
states: P = h|i =
dr1 dr2 |(r1 , r2 )|2 = 1.
If, as the joint probability amplitude suggests, two
entangled particles travel together in both paths, then
any object encountered by one photon will be encountered by both. Namely, if a transmitting, refractive or
dispersive element is located in one path, both photons
will experience the intensity or phase change imparted
by this object even though one of these photons is
detected in an entirely different path.
This phenomenon is measurable: if the entangled
parameter for both of these two states is detected
(transverse position or momentum in this case) and
the result of each measurement is combined in a coincidence circuit, fourth-order electric field interference
is observed. This is the main physical principle underlying ghost imaging which will be briefly introduced
in the next section [3].

2.3.1

Imaging via coincidence detection

Entangled photons exhibit fourth-order coherence:


when separated, these photons can be detected by
different detectors in a coincidence circuit, which measures if two photons arrive at the two detectors within a
given time window. Photons arriving at the same time
are correlated meaning that they share some properties
such as polarization, position or momentum. However, two entangled photons can be viewed as a single
bi-photon traveling in a superposition of two paths.
This property allows the experimenter to put an
amplitude or phase altering object in one path with
a single-pixel (bucket) detector behind it, while only
placing a multi-pixel detector at the other paths end.
The bucket detector serves as a trigger while the multipixel detector senses the transverse position of the
photon. Since this bi-photon has in principle interacted
with the object to be imaged, the alteration of the two
entangled photons is detectable in the fourth-order
interference uncovered by the coincidence circuit. This
technique is ghost imaging.
4

Figure 5: Spontaneous down conversion relies on the conservation of energy and momentum. The pump (subscript p)
yields an idler (subscript i) and signal (subscript s) photon,
the combination of which conserves energy and momentum.
Figure 4: The apparatus used in the experiment conducted
by Lemos et al. [11] Pumped by the green laser, spatial and
momentum entanglement is produced in NL1 and NL2. D1,
D2 and D3 are all dichroic mirrors which only reflect at the
red frequency and transmit at the other frequencies used. O
is the object to be imaged. The yellow beams are combined
on BS2 where the second order interference is detected. The
red beam is discarded.

Methods

This section discusses the methods used to realize this


experiment by employing the quantum phenomena
introduced in the previous sections.
3.1

Generating entanglement

One of the most widely used ways to create entanglement is through spontaneous parametric down conversion (SPDC). This is usually created using a nonlinear
material whose properties allow one higher energy photon to split up into two lower energy photons. The
reverse process can also occur in such a material, so
to enhance one process and suppress others requires
matching the phase velocities of the combining electric fields such that only the phase-matched fields are
allowed to build up within the crystal.
SPDC (in this case, collinear and non-degenerate)
relies on energy matching,
Ep = Ei + Es p = i + s

(6)

and phase matching,


~kp = ~ki + ~ks np = ni + ns
p
i
s

(7)

where subscripts p, i and s stand for pump, idler and


signal respectively and nj is the refractive index at
j . These two constraints are depicted in Figure 5.
The process of downconversion is spontaneous, so the
input pump field starts to couple with a combination
of vacuum fluctuations whose energy and momentum
matches that of the pump wave. Over the length of
the crystal, the amplitude of the signal and idler fields
grow, or in terms of single photons, the probability
that a photon pair is born grows.
There are several methods used to phase-match fields
including pivoting a birefringent crystal or, as I will
cover in the next section, using a periodically-poled
material to realize quasi-phase-matching.
3.1.1

Periodic polling

Using a periodically-poled material to phase-match


fields takes advantage of the fact that, without phasematching, input and output fields will exchange energy
as a system of coupled oscillators over the length of
the nonlinear crystal. Figure 6(c) shows one of these
fields oscillating in energy. Since the phase-matching
condition is sensitive to the polarity of the nonlinear
crystal itself, the polarity of the crystal can be abruptly
5

3.2

Figure 6: Illustration of amplitude build-up over some


propagation length in (a) a perfectly phase-matched crystal,
(b) a quasi-phase-matched crystal and (c) a crystal with a
large phase mismatch. (b) is the equivalent of (c) when the
crystal polarity is periodically flipped. [6]

reversed when the field that needs to build up is starting to deplete. This is called quasi-phase-matching.
In this way, the selected output field can be built up
according to the quasi-phase-matching condition
k = kp ks ki
where
= 2Lcoh =

(8)

2
kp ks ki

where is the polling period. Equation (8) illustrates


that the periodic structure has its own momentum that
can negate the momentum mismatch of the photons.
Since there is a mismatch between wavevectors (k)
in the absence of polling, the total energy will oscillate between pump and signal and idler fields, which
prohibits the building up of the desired output waves.
If however, at the very peak of this oscillation, when
the signal and idler fields have maximum energy, the
polarity of the crystal is flipped, the phase of the input
and output fields are artificially advanced by . This
allows the signal and idler fields to continually increase
despite the wavevector mismatch and is illustrated in
Figure 6(b).
To produce different emergent frequencies, the nonlinear crystal can be heated so that different phasematching conditions hold.

Indistinguishability of the source

This experiment by Lemos et al. borrows heavily


from the work of Mandel [22] [18], which demonstrates second-order interference by coupling two MachZehnder interferometers created by overlapping the
idlers produced by two nonlinear crystals. This setup
exhibits induced coherence which is described in more
detail below.
Referring to Figure 4, a single green photon travels
in a superposition of paths a and b. If the photon
is in path a it will be down-converted to a red and
yellow photon in NL1 and similarly in path b the downconversion will happen in NL2. At this point, spatial
entanglement between down-converted photons is realized by splitting the red photon with a dichroic mirror
(D1 or D3). By placing a detector in paths c or d,
we would know in which crystal the down-conversion
event occurred.
If instead of discerning which crystal down-converted
the pump photon we allow the two systems to couple
so that the idler of NL1 becomes collinear with the
idler from NL2, we have "erased" the information of
which crystal produced the signal and idler photons.
If we measure the red photon after NL2, we have
no way of knowing which nonlinear crystal the red
photon originated from. Similarly, by combining the
yellow photons from NL1 and NL2 on a beam splitter,
we cannot discern which crystal produced the signal
photon either.
If the idler is blocked between NL1 and NL2, the origin of the signal photon can be determined (assuming
100% detector efficiency) by detecting a coincidence
between idler and signal if the origin was NL2, or no coincidence if the origin was NL1. In this case, we would
not detect any interference at beamsplitter BS2. It
does not matter whether or not the observer chooses to
notice this which-path information, the mere possibility
of obtaining this information destroys the second-order
interference [22].
If path d is blocked, the state of the system is in a

superposition of down-conversion in NL1 and NL2:


|i = |cis |dii + |eis |f ii

(9)

Now if we unblock path d and let the transmittance of


the object in path d be T and its phase shift be and
align the two idler beams perfectly, then the two idler
modes can be expressed as a single mode:
|dii T ei |f ii

(10)

and the overall state after NL2 becomes


|i = (T ei |cis + |eis )|f ii

(11)

Ive omitted the normalization constants in the above


equations to preserve simplicity. Once the idler has
been discarded (via dichroic mirror D3) the state looks
like a single-photon superposition from which we would
expect to see second-order interference.
Assuming a spatial dependence T (x, y) and (x, y)
of the object O, an intensity image can be created by
blocking certain parts of the image plane and breaking
the indistinguishability of photons at each "pixel".
Note that since a photon pair can be spontaneously
created anywhere along the length of the crystal, a
typical SPDC source is spatially and temporally incoherent with respect to itself [17]. However, since the
experiment has been setup in such a way that the observer can have no way of identifying in which crystal
the down-conversion occurred, coherence is induced
[22] [18].
3.3

Imaging experiments

To better illustrate the imaging system, I have created


a simplified lens picture in Figure 8. A typical 4f system is used to image NL1 (plane 1) onto NL2 (plane
3) so that a photon originating from one crystal is
indistinguishable from a photon from the other crystal,
while collecting information from object O in the middle. Planes 1, 2 and 3 are depicted in Figure 7 as grey
dotted lines. The object (plane 2) then goes through
its own 4f lens system so that it is imaged onto the
EMCCD camera. In the other two paths starting on

Figure 7: The experimental apparatus from Lemos et al.


[11] redrawn with lenses to highlight the fact that transverse
imaging is the main objective. Lenses image plane 1 onto
plane 3 so that photons from NL1 and NL2 are indistinguishable. Plane 2 is imaged onto the plane of the EMCCD
camera.

Figure 8: The pump laser is focused onto NL1 via L1


which is projected onto the object via L3 and focused back
onto NL2 via L3. The signal beam is focused by L5 onto
the camera so that the object O is imaged onto its plane.
L3 and L3 have focal length f and L5 has focal length 2f .

plane 1 and ending on plane 3, the pump and signal


idler take similar paths through their own lenses so
that at the beam splitter, beams from both each crystal
has an identical beam waist and divergence [11].
For the two experiments in the next two sections,
the object O is replaced with an object changing the
beams phase and amplitude.
3.3.1

Intensity imaging

The first experiment performed by Lemos et al. is to


create an image by blocking some pixels in the object
plane so that photons at the blocked pixels become
distinguishable and thus quantum interference ceases
7

Figure 9: The left image is the image from one of the


BS2 ports and the image on the right is from the other port.
The interfering part appears light in one port and dark in
the other while the non-interfering region remains constant
due to which-path information. This image was taken from
Lemos et al. [11]

to exist in these transverse regions. Unblocked parts of


the object plane still produce interference. At the output ports of the last beam splitter, the unblocked region
will appear lighter or darker depending on the relative phase between interferometer arms. The blocked
portion will remain a constant intensity, not effected
by the phase difference. The imaging results of this
scenario are displayed in Figure 9.

3.3.2

Phase imaging

In the second experiment Lemos et al. use an object


which is transparent to all frequencies in the setup.
This objects main purpose is to act as an optical delay
so that different parts of the beam have a relative phase
shift compared to other parts. A silicon figure of a cat
is used which creates a phase delay between different
parts of the probing beam of = which creates a
stark contrast in the interference of different parts of
the beam at the image plane.
The phase delay between different parts of the signal
beam (yellow) is = 2 so that the object used is
effectively invisible to this beam. To prove this, the
image is placed in the signal photons path on the
dotted line labeled "2" in Figure 7 and no image is
observed on the EMCCD camera.

Figure 10: (a) Right and left are the images from the two
ports of the beamsplitter. There is no which-path information about any part of the beam, but part of the beam is
radians out of phase with respect to the rest of the beam.
This shows up as a difference in intensity at the output. In
(b) a diagram of the cat-shaped transparent glass plate is
shown. This image was taken from Lemos et al. [11]

This experiment is repeated with a new set of downconverted frequencies to show the versatility of such a
setup. This is achieved by heating the nonlinear crystals so that a different wave vector mismatch exists that
can only be nullified by a different set of frequencies.

Discussion

This study, though flashy and seemly very impressive,


is not very novel. This study has effectually been
done in [22] and [18] where the induced coherence
between SPDC photons was examined. Two nonlinear
crystals were used to create an indistinguishable pair
of down-converted photons that had second and fourth
order interference depending on if the idler beam were
blocked or not. The only change in [11] was to make
this setup into an imaging experiment where a 2D
8

object was used to block or delay the idler beam.


4.1

Quantum vs classical operation

Although this experiment provides an interesting insight into quantum entanglement and coherence, it is
hard to see that it will have any immediate practical
applications due to the required long integration time.
If the sample to be imaged is scattering, then due
to the fact that SPDC events happen at a very low
rate, and if the area to be scanned is large at all, the
integration time will be very long [2].
In fact, because photons from NL1 can stimulate
emission of down-converted photons in NL2 [19], increasing the pump power will induce coherence between
the two crystals in the classical regime. Thus, integrating over a long time in the quantum regime will
produce an almost identical output intensity distribution as in the classical regime. This comes from the fact
that a quantum probability amplitude is analogous to
an electric field amplitude, both of which can interfere
with themselves.
This concept is displayed graphically in Figure 11.
The first-order coherence function g (1) (1, 2) varies with
the transmittivity T of the object in the idler path
differently in the quantum and classical case. This is
illustrated in the following equation:
r
1+n

(1)
g (1, 2) = T
(12)
1 + T 2n

where n
is the mean number of photons in the setup.
The degree of coherence of the signal photons increases
linearly with the transmission coefficient only for extremely low average photon numbers, i.e., the quantum
case. As the system becomes more classical, i.e., more
photons are in the apparatus at once, this relation
between g (1) (1, 2) and T becomes less and less linear.
Coherence is induced despite a high degree of distinguishability between photon sources. Physically, the
coherence is seen as the fringe visibility at the beam
splitter.
For classical measurements, this type of device would
be useful in imaging. It is easy to imagine a situation

Figure 11: Figure from [19]: the theoretical relation between coherence and transmittivity coefficient T . The curve
is plotted for various average photon numbers: n
is 102 ,
4
1, 10, 100 and 10 . As more photons are allowed into the
system, the curve becomes less and less linear, indicating a
deviation from the quantum reality as down-conversion is
stimulated in NL2.

where probing an object works best at one wavelength


but transportation and imaging may work best at
another wavelength, such as 1550 nm.
This characterizes a larger problem in the field of applied quantum optics, namely, quantum measurements
can sometimes provide a slightly better signal-to-noise
ratio than classical measurements [17], yet often times
these quantum measurements require longer acquisition times, higher-quality detectors and a much more
stable setup.
4.2

Imaging limits

As exhibited, this apparatus can image transverse features of an object by sensing the amplitude and phase
differences in the beam. This allows for the imaging
of a transparent or absorptive material, or by a reflection scheme, a reflecting material. This setup cannot,
however, distinguish between different planes in the
longitudinal direction. This fact excludes the imaging
of samples that are longer than a certain length.
Also, a real object will probably have absorptive,
9

refractive and reflective features and this setup precludes measurement of all three features. From the
two output ports of the beam splitter, it is impossible to distinguish between a pixel that is blocked by
the object and a pixel whose phase shift is 2 relative to the signal photon. A variable delay line would
need to be introduced in the signal arm to break the
indistinguishability of these two cases.
An interesting addition to this experiment would
be to use the concept of coherent diffraction imaging to also see longitudinal variations in the sample.
Depending on the wavelength sent into the object, a
different diffraction (phase) pattern would be created
and seen via second-order interference on the beam
splitter. By varying the wavelength used to probe
the object and various post-processing techniques, a
three-dimensional image of the object could be constructed [8]. This is completely feasible since Lemos
et al. pointed out that the frequency of the photon
pair can be changed by changing the phase-matching
condition via temperature control in the two nonlinear
crystals.
4.3

Quantum feed-forward measurement

After the two idlers are overlapped, the signal photons


have the ability to interfere. There is no need to detect
the idler photons in coincidence since indistinguishability of the photon pair source allows for second-order
interference. The idler photon, however, is in a superposition of passing through the object and not passing
through the object, and so it must carry the same
phase information as the signal photon. Since this
idler photon is being thrown away, it can be used to
couple this setup to another system.
Such an extraction of information about a quantum
state without absorbing the photon energy itself can
be fed-forward in order to influence another quantum
optical setup. These types of measurements have been
investigated for quantum computational algorithms,
for instance, to provide a feed-forward mechanism in
order to control a quantum state [7].

Conclusion

In this paper I have studied the work of Lemos et al.


who used two-photon spatial entanglement and the
indistinguishability of two nonlinear crystal sources to
create an imaging apparatus that can probe an object
at one wavelength and collect data about that object
at a second wavelength. Although this study does not
uncover any novel quantum behavior, it highlights the
potential of such a setup for quantum imaging. This
would be useful in the case where the probing light
needs to be a different frequency than the signal light,
i.e., 1550 nm to be sent in fiber. Unfortunately, this
quantum system, especially second-order interference,
is extremely sensitive to path-length and other variations that cause uncertainty in the collected data. A
classical system is conceivable that gives almost identical results via different physical principles.
This experiment, however, is not doomed as there are
several potential methods for improvement. I have discussed the idea of transforming this setup into a threedimensional imaging setup by using post-processing
and the wavelength versatility of periodically-polled
structures. Ive also mentioned the fact that information is encoded into the photon pair by the object and
extracted by detection of the signal photon, leaving
the idler photon available for possible quantum feedforward algorithms. Further work and clever insight
into these problems will most certainly lead to the
development of these quantum optical phenomena into
real world applications.

References

[1] A F Abouraddy, B E Saleh, A V Sergienko, and M C


Teich. Role of entanglement in two-photon imaging.
Physical review letters, 87(12):123602, 2001.
[2] Ayman F. Abouraddy, Magued B. Nasr, Bahaa E. a.
Saleh, Alexander V. Sergienko, and Malvin C. Teich.
Quantum optical coherence tomography with dispersion cancellation. Physical Review A, 65, 2002.

10

[3] Ayman F. Abouraddy, Patrick R. Stone, Alexander V.


Sergienko, Bahaa E a Saleh, and Malvin C. Teich.
Entangled-photon imaging of a pure phase object.
Physical Review Letters, 93(21):1922, 2004.
[4] J. S. Bell. On the Einstein-Podolsky-Rosen Paradox,
1964.
[5] A.N. Boto, D.S. Abrams, C.P. Williams, and J.P. Dowling. Quantum interferometric lithography: exploiting
entanglement to beat the diffraction limit. Physical
Review Letters, 85(13):27332736, 2000.
[6] Robert Boyd. Nonlinear Optics. 2003.
[7] Ben Buchler, Ping Lam, Hans-A. Bachor, Ulrik Andersen, and Timothy Ralph. Squeezing more from a
quantum nondemolition measurement. Physical Review
A, 65(1):25, 2001.
[8] James R Fienup, Richard G Paxman, Michael F Reiley,
and Brian J Thelen. 3D imaging correlography and
coherent image reconstruction. Proc. SPIE Vol. 3815,
3815(July):60, 1999.
[9] Vittorio Giovannetti, Seth Lloyd, and Lorenzo Maccone. Quantum-enhanced measurements: beating the
standard quantum limit. Science, 306(5700):13301336,
2004.
[10] John C. Howell, Ryan S. Bennink, Sean J. Bentley,
and R. W. Boyd. Realization of the einstein-podolskyrosen paradox using momentumand position-entangled
photons from spontaneous parametric down conversion.
Physical Review Letters, 92(21):14, 2004.

[13] Jeremy L. OBrien, Akira Furusawa, and Jelena


Vukovi. Photonic quantum technologies. Nature
Photonics, 3:687695, 2009.
[14] Anirban Pathak. Elements of Quantum Computation
and Quantum Communication.pdf. Taylor & Francis,
2013.
[15] T. B. Pittman, Y. H. Shih, D. V. Strekalov, and A. V.
Sergienko. Optical imaging by means of two-photon
quantum entanglement. Physical Letters A, 52(5),
1995.
[16] D. S. Tasca, S. P. Walborn, P. H. Souto Ribeiro,
F. Toscano, and P. Pellat-Finet. Propagation of transverse intensity correlations of a two-photon state. Physical Review A - Atomic, Molecular, and Optical Physics,
79(3):19, 2009.
[17] S. P. Walborn, C. H. Monken, S. Pdua, and P. H.
Souto Ribeiro. Spatial correlations in parametric downconversion. Physics Reports, 495(4-5):87139, 2010.
[18] L. J. Wang, X. Y. Zou, and Mandel L. Induced coherence without induced emission. Physical Review A,
44(7):46144622, 1991.
[19] H. M. Wiseman. Using feedback to eliminate backaction in quantum measurements. Physical Review A,
51(3):24592468, 1995.
[20] Thomas Young. A course of lectures on natural philosophy and the mechanical arts, 1807.

[11] Gabriela B. Lemos, Victoria Borish, Garrett D. Cole,


Sven Ramelow, Radek Lapkiewicz, and Anton Zeilinger.
Quantum Imaging with Undetected Photons. Nature,
512(7515):409412, 2014.

[21] Guohong Zhao, Wanrong Yu, Baokang Zhao, and Chunqing Wu. Towards a Key Consuming Detection in
QKD-VoIP Systems Security VoIP Application in QKD
Network. In Availability, Reliability, and Security in
Information Systems, pages 281285. Springer International Publishing, 2014.

[12] L Mandel. Quantum effects in one-photon and twophoton interference. Reviews of Modern Physics,
71(2):274282, 1999.

[22] X. Y. Zou, L. J. Wang, and L. Mandel. Induced coherence and indistinguishability in optical interference.
Physical Review Letters, 67(3):168171, 1991.

11

You might also like