Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

www.sciencemag.

org/content/351/6278/1176/suppl/DC1

Supplementary Materials for


Generation of multiphoton entangled quantum states by means of
integrated frequency combs
Christian Reimer,* Michael Kues,*† Piotr Roztocki, Benjamin Wetzel, Fabio Grazioso,
Brent E. Little, Sai T. Chu, Tudor Johnston, Yaron Bromberg, Lucia Caspani, David J.
Moss,‡ Roberto Morandotti†
*These authors contributed equally to this work.
†Corresponding author. E-mail: michael.kues@emt.inrs.ca (M.K.); morandotti@emt.inrs.ca (R.M.)

Published 11 March 2016, Science 351, 1176 (2016)


DOI: 10.1126/science.aad8532

This PDF file includes:

Materials and Methods


Supplementary Text
Figs. S1 to S4
Full Reference List
Materials and methods
Device fabrication and characterization
The microring resonator was fabricated using UV photolithography and reactive ion
etching in a CMOS-compatible, high refractive index silica glass (Hydex) deposited by
chemical vapour deposition without the need for high temperature annealing. Hydex is
featured by very low linear (<0.06 dB/cm) and negligible nonlinear optical losses (no
nonlinear losses measured up to 25 GW/cm2) with low and anomalous waveguide
dispersion at 1556 nm, with the zero-dispersion wavelength at 1560 nm. The microring
resonators are vertically coupled to two bus waveguides, forming a 4-port configuration.
The input and output bus waveguides are featured with mode converters and are
connected to polarization maintaining fibers, resulting in coupling losses of <1.6 dB per
facet. More details on the material platform Hydex can be found in (14).
In this specific case, the resonator was characterized using an optical spectrum
analyzer with 5 MHz (40 fm) resolution and 3 pm wavelength accuracy, see Fig. S1. A
free spectral range of 200.39 GHz and an average resonator bandwidth of 803 MHz was
measured (leading to an average resonator finesse of 250) with a variation in bandwidth
of less than 7% over the full C-Band, which is within the measurement accuracy of the
used optical spectrum analyzer. At the excitation frequency of 192.65 THz, the measured
spectral resonator bandwidth corresponds to a Q-factor of 240,000.

Single photon detection and system losses


For the single photon spectrum measurement (Fig. 1), we used an idQuantique
ID210 single photon detector at 5% quantum efficiency (corresponding to 13 dB
detection loss), 10 µs dead time, 500 ps timing jitter, and 1.3 kHz dark count rates.
For the entanglement measurement, we used superconducting nanowire single
photon detectors from Quantum Opus with 75% detection efficiency (corresponding to
1.25 dB detection loss), 50 ns dead time, 50 ps timing jitter, and 800 Hz dark count rates.
The detector signals were acquired with a Time to Digital Converter (TDC) PicoQuant
HydraHarp 400 with 1 ps timing resolution.
The losses for the photons in the entanglement measurement add up to 14.75 dB,
which break down as follows: 1.6 dB coupling off the chip, 1.5 dB per polarization
controller, 5.4 dB per interferometer, 5 dB for the frequency filters (to remove the
excitation field and separate the photons), and 1.25 dB detection loss (Quantum Opus
detectors).

Single photon spectrum measurements


We characterized the emitted spectral shape of the quantum frequency comb using a
tunable grating-based monochromator with an operating range from 1470 to 1620 nm in
combination with a single photon detector. The raw measured single photon counts are
measured with the monochromator as a function of wavelength. In addition to the
broadband measurement, we also investigated the single photon spectrum of the 6
resonances around the pump frequency using a digital tunable wavelength filter, limited
to the C-Band, with a higher resolution (see inset in Fig. 1). From the Single Photon
Count Rate (SPCR), it is possible to extract the Pair Production Rate (PPR) per pulse:
𝑃𝑃𝑅 = 𝑆𝑃𝐶𝑅 /( 𝜂! 𝜂! 𝑓! ), where 𝜂! and 𝜂! are the signal losses and detection
efficiency, respectively, and 𝑓! is the repetition rate of the pulsed pump laser (24), here

2
16.8 MHz. Another way to measure the pair production rate is to use photon coincidence
measurements between channels symmetric to the pump frequency. We collected such
data at the 6 channel pairs using two identical single photon detectors (ID210 detectors).
From the Coincidence Counts (CC) it is possible to extract the
𝑃𝑃𝑅 = 𝐶𝐶 /( 𝜂! 𝜂! 𝜂!! 𝜂!! 𝑓! ). With the monochromator, we measured signal and idler
losses of 7.3 dB (1.6 dB coupling, 1.5 dB polarization control, 4.2 dB monochromator),
corresponding to 𝜂! = 𝜂! = 0.18621. On the other hand, using the digital wavelength
filter, we estimate similar losses of 7.3 dB (1.6 dB coupling, 5.7 dB digital filter). For the
first resonance pair symmetric to the pump frequency, we measured a single photon count
rate of 2,759 Hz and coincidence counts of 26 Hz, corresponding to pair production rates
of 0.0176 and 0.0179 pairs per pulse, respectively, showing very good agreement
between the two measurements. The PPR values extracted from the single photon count
rate (not coincidence measurements) are displayed in Fig. 1 on the right axis.

Two-photon time-bin entanglement measurement


Time-bin entanglement is a discrete form of energy-time entanglement (26),
equivalent in formalism, e.g., to polarization entanglement. Using a double pulse
configuration, a photon pair is generated in a coherent superposition of two time-bins
referred to as short S and long L . The generated quantum state can be written as
!
Ψ = ! S, S + 𝑒 !" L, L , where S, S = S !"#!"# S !"#$% refers to the signal and idler
photons both being in the first time-bin, and L, L in the second time-bin. The phase 𝜃 is
related to the phase difference between the pump pulses depending on the nonlinear
process used, as discussed in the Methods section “Phase-dependency of the nonlinear
process”. In order to characterize this form of entanglement, the generated photons can be
passed through unbalanced interferometers with a delay equal to the time-bin temporal
separation of the entangled state, leading to quantum interference in the photon
coincidence counts in the time-bins projected on the central time slot after the second
interferometer. In order to achieve the required synchronization with the source, a trigger
from the pulsed laser was connected to the TDC. The entanglement measurements shown
in Fig. 3 are averaged over 7 individual measurements with the standard deviation shown
with error bars. The background was extracted from the coincidence analysis of non-
entangled photons generated by different pump pulses. For this purpose, coincidences
between photons generated up to 20 pump pulses apart were collected and averaged.

Four-photon time-bin entanglement measurement


Starting with the entangled state
!
Ψ!!!!!"!# = ! S!! , S!! , S!! , 𝑆!! + 𝑒 !!! 𝑆!! , S!! , L!! , L!! + 𝑒 !!! L!! , L!! , S!! , S!! +
𝑒 !!! 𝐿!! , L!! , L!! , L!! , with 𝜑 being the phase of the pump interferometer, the photons
are passed through additional interferometers, adding a phase for each photon. After the
second set of interferometers, the photons can be projected into 81 different combinations
of time-slots, first, central (C) and last, where four-photon quantum interference is only
seen on the projection in which all photons are in the central slot. This projection is given
by Ψ!!!! = (𝑒 !(!!!!!!!) + 𝑒 !(!!!!!!) + 𝑒 !(!!!!!!) + 𝑒 !!! ) C!! , C!! , C!! , C!! , with
𝛼, 𝛽, 𝛾, 𝛿 being the phases added on the four individual photons. This leads to an expected

3
four-photon interference proportional to 2 + 𝑐𝑜𝑠 𝛼 + 𝛽 + 𝛾 + 𝛿 − 4𝜑 + 𝑐𝑜𝑠 𝛼 + 𝛽 −
𝛾 − 𝛿 + 2 cos 𝛼 + 𝛽 − 2𝜑 + 2cos (𝛾 + 𝛿 − 2𝜑).

Quantum state tomography


A very detailed and instructive summary of how to perform quantum state
tomography is presented by (30), while a detailed description for the particular case of
two-photon time-bin entangled qubits is given in the work by Takesue and Noguchi (32).
Quantum state tomography is an experimental method that allows the extraction of the
density matrix of a quantum state. We can represent the time-bin entangled two-photon
!
state as Ψ = ! S, S + L, L with its density matrix 𝜌.
1
𝜌= Ψ Ψ = S, S + L, L 𝑆, 𝑆 + 𝐿, 𝐿
2
Using the basis 𝐵 = S, S , S, L , L, S , L, L this results in the ideal theoretical two-
photon density matrix:
1 1
2 0 0 2
0 0 0 0
𝜌!! = .
0 0 0 0
1 1
2 0 0 2
It is possible to measure this density matrix through projection measurements of the
signal and idler qubits, resulting in a minimum of 16 required values for the full
description of the two-photon state. To reconstruct the density matrix of a four-photon
entangled state, consisting of 16x16 entries, 256 measurements are needed. An immediate
advantage in using time-bin entangled photons is that these required values can be
extracted for the two (or four) photon case through 4 (or 16) separate projection
measurements, respectively, as each phase setting of the interferometer immediately
results in 9 (or 81) projection measurements, see Ref. (32). Each of these projection
measurements can be described through a wave vector Ψ! . The photon counts collected
in this projection are then given by:
𝑛! = 𝐶 Ψ! 𝜌 Ψ! ,
where C is a measurement constant dependent on the integration time.
The measured density matrix can then be reconstructed using the relations:
𝜌 = 𝐶 !! ! 𝑀! 𝑛! ,
𝑀! = ! Γ! 𝐵!! !,! ,
𝐵!,! = Ψ! Γ! Ψ! , and
𝐶= 𝑛! , 𝑓𝑜𝑟 𝑇𝑟 𝑀! = 1 .
!
In addition to the measured values, a set of linearly-independent and normalized matrices
Γ is required. Any set of such matrices can be used as long as they fulfill the requirement
𝑇𝑟 Γ! Γ! = 𝛿!,! , see Ref. (30).
A fast way to construct this set of matrices is to use the generators of the SU(2) (two
photon) or SU(4) (four photon) special unitary groups plus the identity matrix, and
construct the Γ matrices through the Kronecker product of all combinations. For the two-
photon case, the SU(2) generators are simply the three Pauli matrices, which together
with the identity matrix can be used to generate the 16 (4x4) matrices, which can also be

4
seen in the Appendix A of Ref. (30). For the four-photon case, the 15 SU(4) generator
matrices (which can be found in Ref. (33)) plus the identity matrix can be used to create
the required 256 (16x16) matrices.
We first calibrated the interferometers by keeping the pump interferometer phase
constant and adjusting the phases of the signal and idler interferometers in such a way
that we reach a minimum in the quantum interference, and there define all phase values
as zero. For the two-photon case, we then performed four separate measurements, where
the signal and idler interferometers are set to [0, 0], [0, π/2], [π/2, 0] and [π/2, π/2],
respectively. Post-processing of the time-tags collected from these four measurements
enables the extraction of the required 16 projection values needed to reconstruct the
density matrix. Instead of using four independent interferometers, the four-photon
tomography can be performed with two interferometers by exploiting a filter, routing
different combinations of photons into different interferometers. By doing so, the 16
different phase settings (all combinations of phase 0 and phase π/2 for each photon) can
be generated, enabling the reconstruction of the required 256 measurement values.
A density matrix associated to a physical system has to be Hermitian and positive.
However, the matrix extracted from measurements usually does not comply with this
ideal requirement. To retrieve a physically-meaningful density matrix, we performed a
maximum-likelihood estimation, which is a method used to find the physically-realistic
density matrix closest to the measured one (30). Once the density matrix 𝜌!"# is found, it
is possible to extract the fidelity of the measured state. The fidelity describes the overlap
between the ideal theoretical and measured density matrix, and is given by 𝐹 =
𝑇𝑟(𝜌!"# 𝜌!! ). A fidelity of one corresponds to a perfect overlap with the ideal entangled
state.

Supplementary Text
Single photon purity and mode measurement
The amount of modes (spatial, spectral, temporal) is an important measure for the
complete description of a quantum state. In free-space generation, such as with the use of
SPDC, spatial mode-matching is very important in order to achieve good coupling into
optical fibers or sufficient mode-overlap at beam splitters, a requirement to perform
quantum interference experiments. In our integrated platform, the photons are directly
generated within the single-mode waveguide, thus eliminating the issue of spatial mode-
matching.
The spectral mode of the photon pair is often measured through the joint spectral
density, where a pure separable two-photon state is characterized by a Schmidt mode
decomposition equal to 1, see Ref. (13, 34). As the resonator modes are too narrow (803
MHz) to achieve single photon measurements with sufficient spectral resolution, the
temporal equivalent of this measurement can be performed through a single photon
autocorrelation measurement (35, 36). We performed the autocorrelation measurement
both for the signal and idler photons (see Fig. S2 A and B), where the peak is given by
1+1/(Number of Modes). Here, we extracted 1.08 ± 0.03 and 1.03 ± 0.02 modes for
signal and idler photons respectively, without any correction for background noise. We
repeated the measurement combining multiple resonances, and obtained a clear linear
increase of the number of measured modes, as a function of the number of selected

5
resonances (i.e. 1 mode for 1 resonance, 2 modes for 2 resonances, 3 modes for 3
resonances, etc.). The corresponding slopes of 1.029 ± 0.029 and 1.039 ± 0.046, confirm
once more the single mode nature of the signal and idler photons, respectively. This
measurement shows that a single photon pair can be described by a single Schmidt mode
and a single photon pair forms a pure separable state. Furthermore, each additional
resonator frequency pair increases the Schmidt number by one, as theoretically predicted
(37).

Four-photon entanglement at different mode combinations


The presented four-photon states are constructed as the product of two-photon Bell
states. As a Bell state is generated at each resonance couple symmetric to the pump
frequency, any combination of two mode sets can be chosen to construct a four-photon
entangled state. We confirmed this by measuring the four-photon quantum interference at
3 different frequency mode combinations, see Fig. S3.

Entanglement after 40km of fiber propagation


We performed quantum state tomography of the qubit state before and after 40 km
fiber propagation, showing that the state is well-preserved and suitable for quantum
communications, see Fig. S4.

6
Resonance bandwidth characterization
Measured Linear fit
1000

800
Bandwidth (MHz)

600
1.0
Measured
Lorentzian fit 0.8

Intensity
400 0.6

0.4
200 0.2

192.636 192.638 192.640


0
190 191 192 193 194 195 196
Frequency (THz)

Fig. S1
Microring resonance bandwidth characterization. Measured spectral bandwidth of all
resonances within the C telecommunications band. The inset shows the measured
transmission spectrum of the resonance used for the excitation, as well as the
corresponding Lorentzian fit. The error bars correspond to the measured accuracy of the
optical spectrum analyzer, which is mainly limited by drifts of the device reference
frequency during the measurement integration-time.

7
Mode characterization
Autocorrelation – Signal 1 Autocorrelation – Idler 1 Signal Idler
2 2
A B C
Normalized counts

Normalized counts
6

Measured number
Mode number Mode number
1.5 1.08±0.03 1.5 1.03±0.02 5

of modes
4
1 1 3
2
0.5 0.5
1
0 0 0
-3 -2 -1 0 1 2 3 -3 -2 -1 0 1 2 3 0 1 2 3 4 5 6
Pulse number Pulse number Number of selected
resonances

Fig. S2
Signal and idler frequency mode characterization. (A) and (B) autocorrelation
characterization for the signal and idler photon, revealing a measured number of modes
of 1.08 ± 0.03 and 1.03 ± 0.02, respectively. (C) Mode measurement repeated combining
multiple resonances, showing a linear scaling between the number of modes and number
of resonances selected, with a slope of 1.029 ± 0.029 and 1.039 ± 0.046, for signal (red)
and idler (blue), respectively, thus confirming that each resonance represents a single
frequency mode.

8
4 Photon quantum interference
Measured Theory 4 Photon
Frequency mode pair 1 and 3 Frequency mode pair 1 and 2
1.5 A 1.5 B

Normalized counts
Normalized counts

1 1

0.5 0.5

0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Phase (π) Phase (π)
Frequency mode pair 2 and 3
1.5 C
Normalized counts

0.5

0
0 0.2 0.4 0.6 0.8 1
Phase (π)

Fig. S3
Four-photon entanglement measurement for different combinations of frequency
mode pairs. Quantum interference measurement is performed for frequency mode pairs 1
and 3 (A), 1 and 2 (B) and 2 and 3 (C), showing that four-photon quantum states are
generated for any combination of frequency mode pairs.

9
A Ideal

0.5
0.5

Re (ρ^ )

Im (ρ^ )
0.25 0

0 -0.5
|11> |11>
|12> <22| |12> <22|
|21> <21| |21> <21|
<12| <12|
|22> <11| |22> <11|

B Measured

0.5
0.5
Re (ρ^ )

Im (ρ^ )
0.25 0

0 -0.5
|11> |11>
|12> <22| |12> <22|
|21> <21| |21> <21|
<12| <12|
|22> <11| |22> <11|

C Measured after 40 km propagation

0.5
0.5
Re (ρ^ )

Im (ρ^ )

0.25 0

0 -0.5
|11> |11>
|12> <22| |12> <22|
|21> <21| |21> <21|
<12| <12|
|22> <11| |22> <11|

Fig. S4
Quantum state tomography measurement after 40 km of fiber propagation. Real
(left) and imaginary (right) part of the ideal (A), directly measured (B) and measured
after 40 km of fiber propagation (C) density matrix of the time-bin entangled qubit. The
measured fidelity is only slightly reduced from 96% in (B) to 87% in (C) after
propagation in the fiber.

10
REFERENCES AND NOTES
1. H. J. Kimble, The quantum internet. Nature 453, 1023–1030 (2008). Medline
doi:10.1038/nature07127
2. D. Deutsch, Quantum theory, the Church-Turing principle and the universal quantum
computer. Proc. R. Soc. A Math. Phys. Eng. Sci. 400, 97 (1985).
3. P. Walther, K. J. Resch, T. Rudolph, E. Schenck, H. Weinfurter, V. Vedral, M. Aspelmeyer, A.
Zeilinger, Experimental one-way quantum computing. Nature 434, 169–176 (2005).
Medline doi:10.1038/nature03347
4. P. C. Humphreys, B. J. Metcalf, J. B. Spring, M. Moore, X. M. Jin, M. Barbieri, W. S.
Kolthammer, I. A. Walmsley, Linear optical quantum computing in a single spatial mode.
Phys. Rev. Lett. 111, 150501 (2013). Medline doi:10.1103/PhysRevLett.111.150501
5. M. Kolobov, The spatial behavior of nonclassical light. Rev. Mod. Phys. 71, 1539–1589
(1999). doi:10.1103/RevModPhys.71.1539
6. M. Chen, N. C. Menicucci, O. Pfister, Experimental realization of multipartite entanglement of
60 modes of a quantum optical frequency comb. Phys. Rev. Lett. 112, 120505 (2014).
Medline doi:10.1103/PhysRevLett.112.120505
7. M. Pysher, Y. Miwa, R. Shahrokhshahi, R. Bloomer, O. Pfister, Parallel generation of
quadripartite cluster entanglement in the optical frequency comb. Phys. Rev. Lett. 107,
030505 (2011). Medline doi:10.1103/PhysRevLett.107.030505
8. J. Roslund, R. M. de Araújo, S. Jiang, C. Fabre, N. Treps, Wavelength-multiplexed quantum
networks with ultrafast frequency combs. Nat. Photonics 8, 109–112 (2013).
doi:10.1038/nphoton.2013.340
9. N. C. Menicucci, Fault-tolerant measurement-based quantum computing with continuous-
variable cluster states. Phys. Rev. Lett. 112, 120504 (2014).
10. M. D. Eisaman, J. Fan, A. Migdall, S. V. Polyakov, Invited review article: Single-photon
sources and detectors. Rev. Sci. Instrum. 82, 071101 (2011). Medline
doi:10.1063/1.3610677
11. W. Wieczorek, N. Kiesel, C. Schmid, W. Laskowski, M. Zukowski, H. Weinfurter,
Multiphoton interference as a tool to observe families of multiphoton entangled states.
IEEE J. Sel. Top. Quantum Electron. 15, 1704–1712 (2009).
doi:10.1109/JSTQE.2009.2025697
12. T. J. Kippenberg, R. Holzwarth, S. A. Diddams, Microresonator-based optical frequency
combs. Science 332, 555–559 (2011). Medline doi:10.1126/science.1193968
13. D. Bonneau, J. W. Silverstone, M. G. Thompson, in Silicon Photonics III, L. Pavesi, D. J.
Lockwood, Eds. (Springer, ed. 3, 2016), pp. 41–82.
14. D. J. Moss, R. Morandotti, A. L. Gaeta, M. Lipson, New CMOS-compatible platforms based
on silicon nitride and Hydex for nonlinear optics. Nat. Photonics 7, 597–607 (2013).
doi:10.1038/nphoton.2013.183
15. S. Azzini, D. Grassani, M. J. Strain, M. Sorel, L. G. Helt, J. E. Sipe, M. Liscidini, M. Galli,
D. Bajoni, Ultra-low power generation of twin photons in a compact silicon ring
resonator. Opt. Express 20, 23100–23107 (2012). Medline doi:10.1364/OE.20.023100
16. C. Reimer, M. Kues, L. Caspani, B. Wetzel, P. Roztocki, M. Clerici, Y. Jestin, M. Ferrera, M.
Peccianti, A. Pasquazi, B. E. Little, S. T. Chu, D. J. Moss, R. Morandotti, Cross-polarized
photon-pair generation and bi-chromatically pumped optical parametric oscillation on a
chip. Nat. Commun. 6, 8236 (2015). Medline doi:10.1038/ncomms9236
17. N. C. Harris, D. Grassani, A. Simbula, M. Pant, M. Galli, T. Baehr-Jones, M. Hochberg, D.
Englund, D. Bajoni, C. Galland, Integrated source of spectrally filtered correlated photons
for large-scale quantum photonic systems. Phys. Rev. X 4, 041047 (2014).
18. D. Grassani, S. Azzini, M. Liscidini, M. Galli, M. J. Strain, M. Sorel, J. E. Sipe, D. Bajoni,
Micrometer-scale integrated silicon source of time-energy entangled photons. Optica 2,
88 (2015). doi:10.1364/OPTICA.2.000088
19. A. Politi, M. J. Cryan, J. G. Rarity, S. Yu, J. L. O’Brien, Silica-on-silicon waveguide
quantum circuits. Science 320, 646–649 (2008). Medline
20. J. C. F. Matthews, A. Politi, A. Stefanov, J. L. O’Brien, Manipulation of multiphoton
entanglement in waveguide quantum circuits. Nat. Photonics 3, 346–350 (2009).
doi:10.1038/nphoton.2009.93
21. Materials and methods are available as supplementary materials on Science Online.
22. Z. Ou, Y. Lu, Cavity enhanced spontaneous parametric down-conversion for the prolongation
of correlation time between conjugate photons. Phys. Rev. Lett. 83, 2556–2559 (1999).
doi:10.1103/PhysRevLett.83.2556
23. C. Reimer, L. Caspani, M. Clerici, M. Ferrera, M. Kues, M. Peccianti, A. Pasquazi, L.
Razzari, B. E. Little, S. T. Chu, D. J. Moss, R. Morandotti, Integrated frequency comb
source of heralded single photons. Opt. Express 22, 6535–6546 (2014).
doi:10.1364/OE.22.006535
24. A. S. Clark, M. J. Collins, A. C. Judge, E. C. Mägi, C. Xiong, B. J. Eggleton, Raman
scattering effects on correlated photon-pair generation in chalcogenide. Opt. Express 20,
16807 (2012). doi:10.1364/OE.20.016807
25. T. Pittman, It’s a good time for time-bin qubits. Physics 6, 110 (2013).
doi:10.1103/Physics.6.110
26. J. Brendel, N. Gisen, W. Tittel, H. Zbinden, Pulsed energy-time entangled twin-photon source
for quantum communication. Phys. Rev. Lett. 82, 2594–2597 (1999).
27. J. F. Clauser, M. A. Horne, A. Shimony, R. A. Holt, Proposed experiment to test local
hidden-variable theories. Phys. Rev. Lett. 23, 880–884 (1969).
doi:10.1103/PhysRevLett.23.880
28. C. C. Gerry, P. L. Knight, Introductory Quantum Optics (Cambridge Univ. Press, 2005).
29. Z. Y. Ou, S. F. Pereira, H. J. Kimble, K. C. Peng, Realization of the Einstein-Podolsky-Rosen
paradox for continuous variables. Phys. Rev. Lett. 68, 3663–3666 (1992). Medline
doi:10.1103/PhysRevLett.68.3663
30. D. F. V. James, P. G. Kwiat, W. J. Munro, A. G. White, Measurement of qubits. Phys. Rev. A
64, 052312 (2001). doi:10.1103/PhysRevA.64.052312
31. G. Vallone, E. Pomarico, P. Mataloni, F. De Martini, V. Berardi, Realization and
characterization of a two-photon four-qubit linear cluster state. Phys. Rev. Lett. 98,
180502 (2007). doi:10.1103/PhysRevLett.98.180502
32. H. Takesue, Y. Noguchi, Implementation of quantum state tomography for time-bin
entangled photon pairs. Opt. Express 17, 10976–10989 (2009). Medline
doi:10.1364/OE.17.010976
33. M. A. A. Sbaih, M. K. Srour, M. S. Hamada, H. M. Fayad, Lie algebra and representation of
SU(4). Electron. J. Theor. Phys. 10, 9–26 (2013).
34. A. Ekert, P. L. Knight, Entangled quantum systems and the Schmidt decomposition. Am. J.
Phys. 63, 415 (1995). doi:10.1119/1.17904
35. J. B. Spring, P. S. Salter, B. J. Metcalf, P. C. Humphreys, M. Moore, N. Thomas-Peter, M.
Barbieri, X. M. Jin, N. K. Langford, W. S. Kolthammer, M. J. Booth, I. A. Walmsley, On-
chip low loss heralded source of pure single photons. Opt. Express 21, 13522–13532
(2013). Medline doi:10.1364/OE.21.013522
36. M. Förtsch, J. U. Fürst, C. Wittmann, D. Strekalov, A. Aiello, M. V. Chekhova, C.
Silberhorn, G. Leuchs, C. Marquardt, A versatile source of single photons for quantum
information processing. Nat. Commun. 4, 1818 (2013). Medline
doi:10.1038/ncomms2838
37. J. Chen, Z. H. Levine, J. Fan, A. L. Migdall, Frequency-bin entangled comb of photon pairs
from a Silicon-on-Insulator micro-resonator. Opt. Express 19, 1470–1483 (2011).
Medline doi:10.1364/OE.19.001470

You might also like