Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Toward optical quantum information processing with quantum dots coupled to microstructures [Invited]

Open Access Open Access

Abstract

Major improvements have been made on semiconductor quantum dot light sources recently and now they can be seen as a serious candidate for near-future scalable photonic quantum information processing experiments. The three key parameters of these photon sources for such applications have been pushed to extreme values: almost unity single-photon purity and photon indistinguishability, and high brightness. In this paper, we review the progress achieved recently on quantum-dot-based single-photon sources. We also review some quantum information experiments where entanglement processes are achieved using semiconductor quantum dots.

© 2016 Optical Society of America

1. INTRODUCTION

Photons are good candidates to perform optical quantum information processing as they can propagate over long distances with little dissipation. Their polarization can be viewed as a two-level system. Since a quantum bit (qubit) is a linear combination of basis states in a two-level system, written α|0+β|1 following the Dirac notation with α and β complex numbers such that α2+β2=1, single-photon polarization can be used as a qubit. Moreover, only traditional polarizers and wave plates need to be used to initialize, manipulate, and project polarization-encoded single-qubit states.

The essential classical property for such single photons is the source brightness, while the essential quantum properties necessary for optically-based quantum information are single-photon-ness of the light source and the identical-ness of all the single photons. To date, most quantum optics experiments, including those with quantum information applications, use sources based on heralded spontaneous parametric downconversion (SPDC), a spontaneous process that generates two photons into two spatial modes from one higher-energy photon. Typically, one of the downconverted photons is used to herald the other one in order to obtain nearly pure single-photon emission in the low pump power regime [1,2]. The indistinguishability of the photons is usually improved either by carefully engineering the phase matching of the SPDC process or by spectrally filtering the photons with a narrow bandpass filter to generate photon states with a coherence time much longer than the pump laser coherence length [3]. SPDC sources have been essential for many advanced photon-based quantum information experiments [46].

However, the low SPDC source brightness defined here as the probability to collect a photon per excitation pulse is 106 to 104, and is becoming a strong limiting factor for further quantum information experiments. This can be increased to 102 in continuous wave excitation using a heralded process [7]. The brightness can also be increased by increasing the pump power but the quantum properties of the photons are usually reduced at high brightness. The brightness is intrinsically limited by a multiphoton component that increases with the pump power [7]. For instance, boson sampling experiments on a linear circuit have been recently performed with up to 4-folds detections (four photons in four modes) but only few five-photon events have been detected, limiting this approach for larger systems [8]. One solution being explored is to improve the brightness at constant photon purity. It involves multiplexing several heralded photons from multiple SPDC sources into a single mode [7,9].

Developing other type of sources is an alternative approach. The conditions that such sources much satisfy are demanding: single-photon emission, indistinguishable photons, and bright emission. Nonheralded single-photon solid-state sources are one class of candidates. These includes organic molecules [10], nitrogen vacancy centers in diamond [11,12], or colloidal quantum dots [13,14] that can emit single photons at room temperatures.

This review focuses on a different solid-state photon source, self-assembled InAs/GaAs semiconductor quantum dots. They are a promising alternative light source for quantum information applications since they can have very pure single-photon emission [15] and they can emit highly indistinguishable photons [16]. They have high brightness values especially when coupled to microstructures. They can also emit entangled photons pairs [17,18]. In Section 2, we present the quantum and classical properties of these quantum dot sources and techniques to improve their brightness. Section 3 describes weak light–matter coupling (Purcell effect) using microcavities where enhanced quantum dot spontaneous emission is observed. In Section 4, we present the ultrabright sources of single and indistinguishable photons that have been developed in recent years. In Section 5, we describe the polarization entangled photons that can be emitted by a quantum dot. Finally, in Section 6, we review quantum optics experiments where quantum dots are at the heart of the process.

This review focuses specifically on the photons emitted from quantum dots and thus does not cover other important topics related to these quantum dots. The spin-photon physics associated with the charge excitons is one example [1921], with a large body of on-going research that is beyond the scope of this review.

2. SEMICONDUCTOR QUANTUM DOTS FOR QUANTUM INFORMATION PROCESSING

Self-assembled InAs quantum dots are formed without lithographic patterning by strain-induced islanding, a process driven by the lattice mismatch between InAs and the host crystal, often GaAs [22]. InAs crystal growth initially occurs as a rough, planar region often called the wetting layer. As the accumulated strain increases with thickness island, growth becomes energetically favorable. Because of the lower bandgap of InAs with respect to the host crystal of GaAs, when the islands are overgrown, strong 3D quantum confinement results, forming nanoscale quantum dots.

A. Single Photons from InAs Quantum Dots

InAs-based quantum dots can emit single photons. The strong localization of the carriers inside the quantum dot enhances Coulomb interactions and the eigenenergies are strongly dependent on the carrier occupation in the quantum dot. Thus, the photon emission energies occurring during the carrier recombination processes depend on the quantum dot carrier occupation [23]. These quantum-confined states are excitonic states: a quasi-particle of localized Coulomb-bond electrons and holes. Examples include the single exciton (one electron–hole pair), biexciton (two electrons and two holes), or the singly charged exciton—the trion (one electron, one hole plus an extra electron or hole). Because of the 3D confinement and Pauli exclusion, the photons emitted from each of these excitonic states are single photons. The single-photon purity was demonstrated by Michler et al. in 2000 [15].

The single-photon purity is quantified through a normalized second-order autocorrelation measurement. It is traditionally measured using a Hanbury Brown and Twiss interferometer [24] where a beam splitter in the setup channels photons to two detectors each with single-photon resolution. Second-order correlations measurements, gHBT(2), are made on photon detections. If τ is the time difference between photon detections, particularly important are the normalized second-order statistics at τ=0, gHBT(2)[0]. A value of gHBT(2)[0]<1 is the hallmark of a quantum source. A value of zero uniquely identifies the state as a purely single-photon state. Nonzero values mean the photon state is a mixture of several photon number states, and values below 0.5 indicate a mixture of single-photon states [25]. Figure 1(a) shows an example of a gHBT(2) function obtained by exciting a quantum dot with a 12.2 ns repetition rate mode-locked laser. Here the quantum dot is excited into a discrete state either a lower-energy state of the wetting layer or an excited state of the quantum dot. We call this quasi-resonant excitation. From the area of the correlation peak at delay 0, a gHBT(2)[0]=0.01(1) is found indicating high single-photon purity of the source [27].

 figure: Fig. 1.

Fig. 1. Single-photon characterization. (a) Example of a gHBT(2)(τ) autocorrelation function under a quasi-resonant pulsed excitation. (b) Two-time second-order correlation, gHBT(2)(t1,t2), measured for an excitation in the wetting layer. Figure adapted from [26].

Download Full Size | PDF

Further information on the photon statistics can be obtained by performing a two-time correlation function gHBT(2)(t1,t2) where t1 and t2 are the delay times between the excitation pulses and the photon detection on the two detectors. This function provides information on the dynamics of the photon emission. Using this technique on quantum dot photon emission, Flagg et al. [26] observed that when the excitation laser energy is larger than the quantum dot confined states that the single-photon purity depends on the time difference between the photon emission and the excitation laser pulse. Early emitted photons have a lower purity than the average [Fig. 1(b)]. This phenomenon is explained by delayed capture processes: even though the laser pump pulse (5ps) is much shorter than the transition radiative lifetime, carriers generated in the GaAs or InAs wetting layer by the pump laser can decay through several pathways to the quantum dot. If one pathway results in early exciton emission the state can be repopulated through a slower pathway, and a second emission can occur before the next laser pulse.

Values as small as gHBT(2)(0)=3×103 have been reported recently using resonant fluorescence excitation [2831], where recapture processes are eliminated. However, resonant excitation requires additional techniques to minimize the laser scattering, such as cross-polarized excitation and detection from above the sample [28,29,32,33] or side excitations where the light field is guided in a planar cavity made of two distributed Bragg reflectors [3436].

The single-photon purity required to perform quantum information processing depends on the protocol that is being used. For example, the polarization states of two photons can be entangled (so that the state of each photon cannot be described independently) using a polarizing beam splitter. The quality of the entanglement depends on the degree of indistinguishability of the photons (see Section 2.B) but also on the single-photon purity. One can show that such entangled states can violate Bell inequalities when gHBT(2)<0.15 [7,37]. The generation of a three-photons entangled state is usually more demanding and some protocols requires gHBT(2)<0.06 (for instance, to satisfy the witness value of a Green–Horne–Zeilinger (GHZ) state, i.e., a state generated with two polarizing beam splitters [38]). Some further advance protocols require even higher purities, such as gHBT(2)<0.03 for some deterministic two-qubits gates [7].

B. Indistinguishable Photons

High-purity single photons can be used for quantum information protocols, such as the BB84 protocol [39]. However, long-distance quantum communication systems will require quantum repeaters to beat the low, but nonnegligible, photon loss over long-distance fiber propagation. Such repeaters, installed along the communication channel, will distribute the entanglement over segments of the quantum channel [40,41]. In such cases, the photons must be indistinguishable. The photon indistinguishability is also required for many quantum information protocols, for instance, where qubits interactions need to occur to perform logic operations.

When two indistinguishable single photons simultaneously enter two separate ports of a 50:50 beam splitter, because photons are bosons, theory predicts the photons will coalesce into the same mode, thus exiting the beam splitter through the same output port [42,43]. The degree of indistinguishability, C, of two single photons is usually measured using a Hong–Ou–Mandel interferometer [44], an unbalanced interferometer where one path can be adjusted to adjust the time delay of photon interference at the output beam splitter. Phase stabilization is not required. Second-order time correlations, gHOM(2), are performed between detected events at each output port. When the photons are indistinguishable and interfere at the beam splitter, no two-photon detection events occur. Thus, the value of gHOM(2)[τ=0] depends on the indistinguishability, gHOM(2)[0] goes to 0 for ideal quantum light and C goes to 1.

The indistinguishability of successively emitted quantum dot photons was shown by Santori et al. in 2002, only 2 years after the demonstration of the first quantum-dot-based single-photon source [16]. The experiment is shown in Fig. 2(a). A Hong–Ou–Mandel interferometer, as described above, and a pulsed excitation scheme are used. Several peaks appear in the gHOM(2)[τ] function because of the small, 2 ns delay and the different possible paths that can be taken by the photons. For lossless 50:50 beam splitters, the area of the peak at delay zero is proportional to gHBT(2)[0] and it fully vanishes for an ideal quantum light.

 figure: Fig. 2.

Fig. 2. Measurement of indistinguishability. (a) Schematic of a setup to measure the indistinguishability of two photons emitted from a quantum dot that is excited 2 ns apart. They are sent to an unbalanced Michelson interferometer and then to detectors. (b) Resulting second-order correlation histogram of the quantum dot light. The intensity of the peak at delay τ=0 is proportional to the photon indistinguishability. Figures adapted from [16].

Download Full Size | PDF

The required photon indistinguishability necessary for quantum information processing experiments is ideally unity but the requirement can become relaxed in some cases. For instance, in the quantum controlled-not gate case presented in Section 6.B, an indistinguishability of 0.5 is the quantum limit for the generation of entangled photon pairs [Eq. (12) and Fig. 11].

Maximum quantum interference can occur only when the photons are spatially and temporarily overlapped, and have the same polarization. In addition, the photons must be Fourier transformed limited and have the same energy. The first two parameters depend on optical alignment while the last parameters are dependent on the quantum dot source.

Fourier transformed limited photons have a coherence time, T2, limited by their radiative lifetime, T1, such that T2=2T1. Any pure dephasings, with a characteristic time T2*, will decrease the coherence time and thus coherently broaden the transition linewidth, γ:

γ=1T2=12T1+1T2*.
Pure dephasing effects occur at time scales shorter than the transition lifetime. On the contrary, slow decoherences lead to transition fluctuations and induce inhomogeneous broadening of the transition, Γin. These decoherences can result from charge fluctuations or nuclear spin flips in the quantum dot surrounding, leading to Zeeman or DC Stark shifts [45], which in turn reduces indistinguishability. A measurement integrated over several minutes will give access to a lower bound for the degree of indistinguishability of photons emitted only few nanoseconds apart [16].

At least three techniques can be used to improve the indistinguishability of quantum dot photons: reduce the radiative lifetime, use resonant fluorescence, and control charge fluctuations. In 2002, Santori et al. coupled a quantum dot to a microcavity in order to benefit from the Purcell enhancement to reduce the lifetime, T1 [16]. Indeed, Eq. (1) shows that shorter lifetimes reduce the inhomogeneous spectral broadening induced by the pure dephasing T2*. They obtained indistinguishability values up to 81%. The temporal difference between the photons was small, 2 ns, about 10 times the radiative lifetime. Further cavity quantum electrodynamics advantages will be detailed in Sections 3 and 4. Resonant pumping schemes improved the indistinguishability to 0.97 (2) in 2013 [28]. Combining cavity quantum electrodynamics and resonant pumping has resulted in near unity indistinguishability (0.985) and high brightness [46]. A better control of the charge fluctuations in the quantum dot vicinity using electrically controlled micropillars [47] further increased the photons indistinguishability to 0.9956 (45) with an extraction efficiency of 0.65 [31]. Because somewhat similar values can be obtained without electrical bias, it remains to be seen if the bias is necessary in all cases to achieve such a high level of indistinguishability.

C. Brightness

Single and indistinguishable photons are two quantum properties of light that are necessary to ensure optimum quantum information processing operations. The brightness of a light source can be defined as the probability, p1, of collecting a single photon in the first external optic every laser pump pulse. A bright source is crucial to perform quantum operations on large computational Hilbert space with many fold correlations. The probability of distributing N photons into N different modes is, at best, pN(N)=(p1)N, and thus having a large p1 is critical to any scaling. Moreover, a bright photon source is required for many advanced quantum information protocols like for the loophole-free Einstein–Podolsky–Rosen (EPR) experiment with entangled photons. The overall efficiency (source brightness, optics, and detection) needs to be above 82.8% in the EPR traditional experimental scheme [48].

The brightness, p1, and the extraction efficiency, η, are linked by

p1=ps×η,
where ps is the probability of the quantum dot to be in a target state at every laser pulse. For example, the brightness of an exciton state will be reduced if the quantum dot is occasionally in a trion state, even if η is large. The term ps will also be reduced in resonant pumping if the state spectrally wanders outside the laser linewidth since the probability of creating the target state resonantly will be reduced. There is no general method to increase low values of ps but some tricks can be used, such as applying an external electric field or a weak above band laser, to partially control the quantum dot charge state and to restore its brightness [4951].

InAs quantum dots are embedded in a GaAs medium with a refractive index of about 3.5. Thus, one can only expect an extraction efficiency of η2% (collection from the top surface) for quantum dots in bulk GaAs materials [52]. The growth of AlAs/GaAs distributed Bragg reflectors (DBRs) around the quantum dot to form a λ planar cavity (λ/n thick GaAs layer, n being the GaAs refractive index) can improve the collection efficiency to η 10%–20% when the bottom DBR has a much higher reflectivity than the top one [52] [Fig. 3(a)]. The collection efficiency can be improved, up to 10%–30%, over a wide frequency range by using solid immersion lenses [55,56].

 figure: Fig. 3.

Fig. 3. Extracting quantum dot photons. (a) A quantum dot sandwiched between two asymmetric DBR mirrors allows for up to 10%–20% collection efficiency. (b) Brightness of 75% has been seen with a quantum dot inserted in the bottom of an inverted trumpet structure. Figure adapted from [53]. (c) Ultrabright sources (p179%) of single and indistinguishability photons have been made by coupling a quantum dot to a 3μm micropillar cavity. Figure adapted from [54].

Download Full Size | PDF

When microstructures are designed around the quantum dot to further improve its brightness, the extraction efficiency, η, can be written:

η=β×(1α)withβ=ΓΓ+Γother.
The term (1α) is the fraction of the microstructure optical field that can be collected by the first lens of the collection setup. The term α includes scattering losses and the fraction of the field that cannot be collected with traditional optical setups. The β factor characterizes the fraction of the light emitted by the quantum dot that is coupled into the target microcavity mode. Γ and Γother are the spontaneous emission rates of the QD transition, respectively, into the microstructure mode and into all the other modes. Thus, increasing Γ or reducing Γother improves the β factor and hence the source brightness.

Increasing β through inhibition of Γother. The term Γother can be decreased by shaping the electromagnetic field around the quantum dot to inhibit the spontaneous emission. Inhibition effects have been observed in photonic crystal cavities [57] and in photonic crystal waveguides [58]. In addition, structures using a photonic crystal waveguide exhibit almost unity emission rates into the waveguide, up to β=0.98 [5961]. The outcoupling efficiencies to free space are not specified but quantum information processing could be performed directly using photonic crystal waveguides [62].

Structures using nanowires or inverted nanotrumpets can also inhibit the spontaneous emission of a quantum dot and increase the β factor [53,63,64] [Fig. 3(b)]. The inhibition is produced by reducing the nanowire diameter at the position of the quantum dot layer [65]. Brightness values up to p10.75 have been measured [53]. These structures are broadband in frequency and thus, they could allow for entangled photon-pairs generation using the exciton and the biexciton photons [18,29,66,67]. However, the small distance between the quantum dot and the surface is a drawback that may degrade the photons indistinguishability. Some surface passivation methods have been developed to minimize drifts of the quantum dot states [68].

Increasing β through the enhancement of Γ. The spontaneous emission rate into a particular mode, Γ, can be enhanced by coupling the transition to a microcavity to benefit from the Purcell factor, as discussed in Section 3.A. Interestingly, this improves both the brightness and the photon indistinguishability. Extraction efficiencies of 0.5 have been measured with suspended circular Bragg grating microcavities [56,69] or photonic crystal cavities [70] and up to 0.79 using micropillar cavities (Section 4 and [54]).

Inhibition and enhancement. Some structures like confined Tamm plasmon modes exhibit both spontaneous emission enhancement of Γ and inhibition of Γother: the β factor can approach unity [71]. However, the term (1α)0.72 can be a limiting factor of those structures [72].

3. ENHANCEMENT OF THE SPONTANEOUS EMISSION RATES WITH LIGHT–MATTER INTERACTIONS

Enhancement of spontaneous emission can be achieved by coupling a quantum dot to a microcavity because of the Purcell effect (Section 3.A). Several kinds of cavities have been developed and besides weak cavity-quantum dot coupling with enhanced spontaneous emission (Purcell effect), strong coupling has been observed (Section 3.B) using a number of techniques (Section 3.C).

A. Weak Coupling and the Purcell Factor

Weak light–matter interactions modify the spontaneous emission rate of an emitter compared to the free-space regime. Weak coupling occurs when the cavity losses are large enough so the coupling is only from the emitter to the cavity mode and is not reversible: 4Ω<ω/Q where Ω is the Rabi frequency, ω the emitter angular frequency, and Q the cavity quality factor [73]. If the emitter dipole is weakly coupled to a microcavity, its spontaneous emission rate can be enhanced by the Purcell effect [74] and the spontaneous emission rate Γ=1/T1 can be approximated by the Fermi Golden rule [7476]:

Γ=2π2|d⃗·E⃗^(r⃗QD)|2×ρ(ωQD),
where ρ(ωQD) is the density of photon modes at the emitter’s frequency, ωQD; the term E⃗^ is the electric field operator; and r⃗QD is the location of the quantum dot dipole, d⃗. The outer averaging is performed over the modes seen by the emitter. A microcavity can play a key role here because it locally increases the density of electromagnetic states. The maximum of spontaneous enhancement is spatially located at the maximum of the electric field, r⃗c, and at the cavity resonance frequency, ωc. The maximum of spontaneous enhancement occurs when the quantum dot transition and the cavity mode are overlapped (r⃗QD=r⃗c) and have the same frequency (ωQD=ωc). In this case, the ratio of the spontaneous emission rate in the cavity mode, Γ, over to the coupling to free space, Γ0 is proportional to the Purcell Factor Fp:
ΓΓ0Fp=3Q(λc/neff)34π2Veff,
where λc is the wavelength of the cavity and neff the effective refractive index of the cavity. Therefore, cavities with high quality factors, Q, and small effective mode volumes, Veff, can result in regimes with strong Purcell factors.

B. Microcavities

Several kinds of photonic microcavities have been coupled to semiconductor quantum dots. The most studied structures are micropillar cavities, photonic crystal cavities, and microdisk cavities. Many interesting properties and effects have been reported with these structures; in particular, radiative lifetime enhancement in the weak coupling regime [7779] and normal-mode splitting in the strong coupling regime between a single quantum dot state and the cavity [8082]. More recently, new structures have been developed. For example, fiber-based external-mirror microcavities where the quantum dot epitaxial layer sits on a bottom DBR and a curved DBR is fabricated on a single-mode fiber [83]. The advantages are that a large fraction of the light is directly coupled into a single-mode fiber and, more notably, the cavity can be spatially and spectrally coupled to a specific quantum dot by moving the fiber. Suspended circular Bragg grating microcavities have also been developed and they exhibit good properties with the observation of a Purcell enhancement and extraction efficiencies of η0.48 [56,69].

Currently, micropillar cavity structures produce the highest-quality light sources in terms of the quantum properties of single-photon purity and indistinguishability, and of brightness. A micropillar is a cylinder a few micrometers in diameter. Two DBR structures are located along the pillar axis to form a λ-cavity between them. The quantum dot layer is grown in the middle of this cavity, where the electric filed is maximum [Fig. 3(c)]. The longitudinal field confinement is created by the DBRs and it can be calculated using standard transfer-matrix calculations [84]. The transverse field is determined using similar methods as used for standard waveguides and the fundamental mode is a HE11 mode [76,85,86]. Numerical apertures are rather small, typically 0.4 for 3μm diameter micropillars [87]. Coupling rates into a single-mode fiber higher than 70% have been obtained by optimizing the mode matching [27].

C. Coupling of a Quantum Dot into a Microcavity Mode

The deterministic coupling of a single emitter to a micrometer structure requires control of the energy and the position of the emitter and of the cavity resonance (Section 3.A). Self-assembled semiconductor quantum dots, described above, are unfortunately randomly distributed on the growth layer surface. Moreover, their exciton energies are inhomogeneously distributed as they depend on the quantum dot size, shape, and composition that cannot be precisely controlled during the growth process [80,81,88]. Thus one must fabricate hundreds of cavities to hopefully get a few cavities coupled with a quantum dot. This method clearly suffers from the random spectral and spatial overlap between the quantum dots and the cavity modes.

High-resolution microscopy techniques. Scanning electron microscopy [89] and atomic force microscopy [90] techniques have been developed to precisely locate a quantum dot (30nm accuracy) and to define a suspended photonic crystal cavity centered on the quantum dots [Fig. 4(a)]. Spectral matching is achieved by enlarging the photonic crystal holes and thinning the membrane using sequential etching processes.

 figure: Fig. 4.

Fig. 4. Coupling of a quantum dot to a microcavity. (a) Atomic-force microscope topography of a photonic crystal nanocavity aligned to a quantum dot. The small hill in the middle arises from a quantum dot (63 nm below the surface). The color bar indicates the measured height. Figure adapted from [90]. (b) Schematic of an optical in situ lithography technique. Two lasers are used to find the quantum dot position and energy, and to define a cavity around it. Figure adapted from [91]. (c) Image of the photoluminescence signal of a quantum dot centered with a circular Bragg grating microcavity. The scale bar represents 5 μm. Figure adapted from [56].

Download Full Size | PDF

Optical in situ lithography technique. Optical lithography at liquid helium temperatures can be used to define a mask on a sample surface. A micropillar cavity can then be etched so the cavity is spatially and spectrally aligned to a quantum dot [91]. The method uses a photoresist deposit on the sample surface and two lasers [Fig. 4(b)]. One laser, with a wavelength above 800nm, can excite a quantum dot, but barely exposes the resist. The quantum dot photoluminescence signal is sent to a spectrometer and to a CCD camera. Looking at the quantum dot photon intensity as a function of the quantum dot’s position relative to the laser, one can determine the emitter position within a 20nm accuracy. The second laser, green at 532 nm, is overlapped on the first one and is used to expose the photoresist at the position of the quantum dot. Finally, coarse spectral matching is achieved by adjusting the photoresist exposed area since the cavity mode energy depends on the diameter of the micropillar [76,86] and is fine tuned by changing the sample temperature by a few degrees or the electric field applied to the quantum dot [82,92].

Photoluminescence imaging. More recently, another optical technique has been reported [56]. The main difference between this method with the optical in situ lithography technique is that here the optical processes are used to find a quantum dot of the desired wavelength range and its position is noted relatively to a mark on the surface. A red laser and an infrared LED illuminate the sample surface and excite the quantum dot sample so that both the quantum dot light and the markers are visualized on a camera. The cavity, a circular Bragg grating microcavity, is then fabricated using the marker position and electron-beam lithography [Fig. 4(c)]. An extraction efficiency of η=0.40 is reported.

A new technique was reported in 2015 [93]. A quantum dot is located using a low temperature cathodoluminescent technique [94]. A microlens is fabricated, in situ, using electron-beam lithography process. An indistinguishability value of 0.80 is reported and a photon extraction efficiency of η=0.23.

We note that quantum dots can also be grown on predetermined sites [95,96]. The quantum properties of these quantum dots are still inferior to the self-assembled quantum dots but large improvements have been made [97]. More complicated integrated structures will likely require ordered quantum dot arrays and improvements in results from these techniques will be important.

4. BRIGHT SOURCES OF INDISTINGUISHABLE SINGLE PHOTONS USING MICROPILLAR CAVITIES

In Section 2, we have seen that semiconductor quantum dots can emit single photons (2000, [15]), indistinguishable photons (2002, [16]), and be a bright light source (2010, [63]). However, combining those three parameters remained a challenge. In this section, we summarize a method based on a micropillar structure (Section 3.B) that allows the combination of high brightness (Sections 4.A and 4.B) and high indistinguishability (Section 4.B) [54]. The quantum dot is deterministically coupled to the cavity by an optical in situ method using a technique described in Section 3.C. Further improvements of the quantum properties of the source were made by exciting the quantum dot resonantly during subsequent measurements [28,30,31,46].

A. Extracting Single Photons with a Micropillar Cavity

In the general case described in Section 2.C, the extraction efficiency η is equal to β×(1α) with β being the fraction of the photons emitted into the cavity mode and (1α) the fraction of the cavity field that can be collected. In the case of a quantum dot coupled on resonance and at the maximum electric field position of a micropillar cavity, the Purcell factor, FP=Γ/Γ0Q/Veff, where Q is the quality factor of the micropillar cavity and Veff the effective volume of the cavity [Eq. (5)]. In this case Γother=Γ0 because the emission into the other modes is barely modified by the cavity. Thus, from Eq. (3):

β=FPFP+1.
In the case that the bottom DBR reflectivity is much larger than of the top DBR, most of the light is emitted through the top surface. The cavity losses are mostly due to scattering on the micropillar sidewalls and by light escaping the cavity through the top mirror. Losses induced by the DBR layers can be small because quality factors larger than 105 have been obtained with similar structures with a wider diameter [98]. Because losses reduce a cavity quality factor and broaden its linewidth, and because the quality factor of a planar cavity, Q0—an infinitely large micropillar—is mostly given by the losses on the mirror, we can directly find the fraction of the light collected by the first lens, (1α). It is the ratio of the quality factor in the confined cavity case, Q, and in planar cavity cases, Q0 [54]:
(1α)=QQ0.
A fit to experimental measurements of Q/Q0 is shown in Fig. 5. The curve drops for small diameters because the cavity mode field is laterally less confined and scattering due to the sidewall roughness is enhanced. In addition, when the diameter decreases to 1 μm, the β factor increases because the Purcell factor inversely depends on the effective cavity mode volume. The range of diameters [2–3] μm is a good compromise between losses by scattering and low Purcell factor. Theoretical efficiencies are 80% with such micropillar cavities [54,67]. Adiabatic micropillar cavities, where the DBR thickness is gradually changed around the quantum dot layer, should exhibit even better extraction efficiencies as the field is laterally better confined [99].

 figure: Fig. 5.

Fig. 5. Brightness optimization. Fit of the experimentally measured (1α)=Q/Q0 terms (black dashed line), calculated β=FP/(FP+1) (red dotted line), and the maximum theoretical extraction efficiency β×(1α) (solid green line) as a function of micropillar diameter. Figure adapted from [54].

Download Full Size | PDF

 figure: Fig. 6.

Fig. 6. Current brightness results. (Top) Raw (open squares) and multiphoton corrected (solid squares) number of collected photons per laser pulse and the corresponding detected count rate per second. (Bottom) Values of gHBT(2)[0] values as a function of the pump power. The excitation laser is tuned to 860 nm to create carriers in the wetting layer. Figure adapted from [54].

Download Full Size | PDF

B. Bright Single-Photon Sources

Using the optical in situ lithography technique described in Section 3.C, micropillars with diameters around 3 μm with a quantum dot coupled to the cavity were fabricated to optimize the extraction efficiency [54]. For the cavity discussed here, and reported in [54], the measured radiative lifetime was T1=265 (30) ps [moderate Purcell factor, Fp=3.9 (6)], a cavity quality factor Q03000, and Q/Q0=0.95(5). During the lithography process the authors selected quantum dots to ensure ps>0.95. Thus, a brightness value of p1=0.75 (16) is expected.

The brightness is experimentally measured by exciting the sample using a pulsed laser at 860 nm (3ps at a rate of 82 MHz). The photons are collected with a microscope objective, sent to a spectrometer, and to a single-photon avalanche photodiode. The brightness of the source, p1—the probability to collect a photon per pulse at the first lens of the microscope objective—is obtained by normalizing the count rate measured on the detector by the laser repetition rate (82 MHz) and the setup detection efficiency (typically 1%). The maximum brightness of the source, p1max, is obtained in the saturation regime. In this regime, a photon is emitted from a target state whose occupation probability is maximum (Section 2.C). In lower pump regimes, the occupation probability drops and the brightness equals

p1=p1max(1eP/Psat),
where P is the laser power and Psat the saturation power. The experimental data are plotted in Fig. 6(a) (dotted line). The single-photon purity is measured simultaneously to account for multicapture processes within the same laser pulse [26]. For this they used the correction coefficient 1gHBT(2)[τ=0] [100]. Figure 6(b) shows gHBT(2)[τ=0] as a function of the pump power. Values are always smaller than 0.15 so the correction factor is small. Finally, the maximum brightness value measured from this source was 0.78 (8) [Fig. 6(a), solid line], the highest brightness reported to date. More recent works reported similar values of about 0.65 [31,46].

C. Historical Development of Bright Indistinguishable Photons Sources

Combining high brightness and high indistinguishability is not trivial. As visible in Eq. (8), to obtain maximum brightness, the pump power must be adequately strong so the system is in the saturation regime. However, this strong pump power creates carriers that may dynamically alter the local potential landscape around the quantum dot and decrease the photon indistinguishability.

In Fig. 7(a), we summarize some of the main achievements toward bright sources of indistinguishable photons. Some results have improved the photons indistinguishability [16,28,30,103,104], and others have improved the brightness [53,63,101,102], but none simultaneously. Simultaneous improvements of the two properties started to be reported in 2013 in [54] from the Senellart group where several techniques were explored to combine high brightness and high photon indistinguishability. We review some of them here.

 figure: Fig. 7.

Fig. 7. Progress toward high brightness and high indistinguishability single-photon sources. (a) Blue points: indistinguishability of successively emitted photons and extraction efficiency of some quantum-dot-based sources reported since the first single-photon demonstration in 2000 [15] (yellow point). In 2002, the indistinguishability of the photons was reported [16]. Corresponding references: 2002–2007: [101,102]; 2010–2013: [53,63]; 2013: [54]; 2013–2014: [28,30]; 2015: [31,46]. The shaded area indicates the use of resonant fluorescence excitations. The red star is for good SPDC sources with gHBT(2)0.1 [7]. (b) Indistinguishability values as a function of the brightness for different pumping conditions. Green squares, wetting layer pumping; red triangles, quasi-resonant pumping; blue stars, two-color scheme (see text). The solid black line plots the normalized laser power P/Psat as a function of the source brightness. Figure adapted from [54].

Download Full Size | PDF

Pumping into the wetting layer [105,106] was tried and the results of the indistinguishability measurements are plotted in Fig. 7(b) as a function of the pump power. Values of the indistinguishability of successively emitted photons as high as 0.86 (10) are shown at low pump powers, but the indistinguishability drops down to 0.50 when the brightness of the source is increased. One explanation is because of the high laser power required to reach the saturation regime [right scale in Fig. 7(b)]. Thus, excess carriers can be generated in the quantum dot surrounding leading to spectral diffusions and decoherences.

A solution to avoid the excessive generation of carriers is to excite the sample using a quasi-resonant pumping scheme. Several groups have reported indistinguishability values between 0.7 and 0.8 under such pumping regimes [16,103,104]. However, in [54], indistinguishability values of about 0.5 were reported at almost any excitation powers [triangles in Fig. 7(b)]. Likely, this is due to sample variation for this technique.

Inspired by the weak above-band laser technique used to restore the source brightness under resonant excitation [50,51], a two-color excitation scheme with a weak wetting layer wavelength laser (860 nm) and a quasi-resonant laser (906 nm) was used in [54]. This technique helped to improve the indistinguishability of the photons making this source the first single-photon source to be both bright and emit highly indistinguishable photons [stars in Fig. 7(b)]. High indistinguishability values up to 0.92 (11) were obtained and are almost a factor of 2 higher than without the weak 860 nm laser. The indistinguishability remains large, 0.82 (11) at higher source brightness, p1=0.65 (6) photons collected per pulse.

D. Current State of the Art: Bright Indistinguishable Sources

Almost three years later, further improvements were made by several groups and recent works report almost unity photon indistinguishability at high source brightness [31,46,107]. Resonant excitation schemes and micropillar cavities are used.

Eliminating the laser scattering into detectors is challenging but a cross-polarization scheme in the excitation and collection paths, along with narrowband filters have produced large extinction ratios (>105) between the pump laser and collected emission. Single-photon purity as low as gHBT(2)[0]=0.0028 (12) has been reported [31]. An indistinguishability value for successively emitted photons of 0.985 at an extraction efficiency of 66% (the brightness is not specified) has been recently reported [46] with a quantum dot coupled to a micropillar cavity. An even higher indistinguishability value of 0.9956 (45) at a measured brightness of 0.154 (15) (extraction efficiency of 65%) was reported in late 2015 using an electrically controlled device to further minimize the charge fluctuations [31] [see historical plot, Fig. 7(a), and also Fig. 8].

 figure: Fig. 8.

Fig. 8. Almost unity indistinguishability; resonant fluorescence. (a) Schematic of a structure that emits highly indistinguishable photon under resonant excitation and electrical control. (b),(c) Second-order correlation histograms, gHOM(2). The photon indistinguishability is determined by comparing the zero delay peak amplitude when the polarization in the two interferometer arms are parallel (b) and orthogonal (c) cases. In the parallel case, the zero delay peak should vanish for fully indistinguishable photons. Figures adapted from [31].

Download Full Size | PDF

5. ENTANGLED PHOTON SOURCES

Many photonics-based quantum information protocols require entangled states of light. For example, in long-distance quantum communication the single-photon loss can be overcome with quantum repeaters and these quantum repeaters require entangled states [40]. Several quantum computing protocols require entangled photons [108,109]; they are also required in quantum teleportation [1,110] and entanglement swapping experiments [111,112].

Entangled photon states occur naturally in the parametric downconversion process [113], and in an ideal quantum dot cascade from the biexciton state, through the exciton state and to the ground state. Decay from the biexciton state, |XX occurs radiatively through two possible channels via the two neutral exciton states, |X. In an ideal, perfectly cylindrical quantum dot these |X states are energy degenerate, but in practice are energy split by shape anisotropy (the anisotropic exchange splitting). If the frequency of this splitting is small compared to the radiative decay rate, information about the actual decay path is not available, and polarization entanglement can result [114].

The photon state generated from the quantum dot |XX|X|0 decay is a maximally entangled Bell state in the polarization basis {|H,|V}:

12(|HXXHX+|VXXVX).
If the anisotropic exchange splitting is larger than the transition linewidth the degree of entanglement is reduced. A time-dependent dephasing term can be introduced between the terms |HXXHX and |VXXVX in Eq. (9) to account for this [115].

The first experimental demonstration of this entanglement was in 2006 [66], followed quickly by a postselection experiment [18]. Muller et al. showed how to eliminate the anisotropic exchange splitting by using a far-detuned dressing laser in the AC Stark regime [35], while other researchers used strain and electric fields to remove the splitting [116].

Cavity-enhanced entangled photon states are difficult to construct because the cavity must be resonant with both the |XX and |X states. This has been accomplished with a unique coupled pillar cavity in 2010 [67]. Research continues, including resonant excitation directly into the |XX state through a two-photon absorption process [29], which is discussed in more detail in the context of time-bin entanglement.

6. TOWARD QUANTUM INFORMATION PROCESSING APPLICATIONS

The properties we reviewed here suggest that quantum light sources based on semiconductor quantum dots are possible candidates for some types of optical quantum information processing. Much work remains in the development of these sources and drawbacks still need to be solved. Still, these sources have evolved to the point where they can be used in some quantum information applications. We show some examples here.

First, we describe a time-bin entanglement protocol that leads to the generation of two entangled photons from a single quantum dot source. It uses the |XX|X|0 cascade discussed above in Section 5, but creates entanglement in time and energy rather than polarization and energy [117]. Second, we review an early realization of an entangling controlled-not gate that can entangle quantum dot single photons [27], and of a quantum dot device that acts as an optical controlled-not gate [118].

A. Time-Bin Entangled Photon Pairs

Polarization entangled states can be used in quantum computing; however, polarization can degrade in some cases, for instance, in long-distance optical transmission. An alternative approach is time-bin entangled states. In particular, such states are robust against dephasings in optical fibers and can propagate over long distances [119,120]. Time-bin entangled photon sources are currently mainly based on parametric downconversion processes [121,122].

It was proposed theoretically in 2005 that coherent resonant excitation of a quantum dot state can be used to generate time-bin entangled single-photon pairs from two laser pulses [123]. The generation of time-bin entangled states requires the photons to be emitted at well-defined times, within the time windows of the |Early and |Late states. One challenging task is to coherently generate the photon pairs so that the emission time and phase remain unknown.

Here we highlight experiments by the Weihs group [117]. The photon pairs are emitted from the sequential recombination of the quantum dot |XX and |X states, and only one polarization state is used [Fig. 9(a)]. An advantage of time-bin entanglement over polarization entanglement is that it does not suffer from the anisotropic exchange splitting, s, of the |X quantum dot level since only one of the two relaxation paths is needed.

 figure: Fig. 9.

Fig. 9. (a). Energy-level scheme for resonant two-photon excitation of a quantum dot biexciton state, |XX. A pair of photons is emitted in cascade through the exciton state, |X, to the ground state, |0. The term s is the anisotropic exchange splitting of the |X level. (b) Quantum dot emission spectrum under resonant excitation of the |XX state. (c) Real part of the reconstructed density matrix. The imaginary part is plotted in [117]. (b),(c) Figures adapted from [117].

Download Full Size | PDF

Time binning is achieved by resonantly pumping the quantum dot |XX state [124]. A two-photon absorption process is used via a virtual state at half the |XX state energy [Fig. 9(a)] [125]. Two peaks corresponding to the X and XX photon energies are visible in a spectrum [Fig. 9(b)]. The quantum dot is located in the middle of a planar cavity made of two DBRs to enhance the collection efficiency. The pulsed resonant laser scattering is reduced by exciting the quantum dot laterally from the side via the in-plane planar cavity waveguide and by using a pulse shaper to adjust the laser spectral width (Fig. 9).

When the two time-bin states, |Early and |Late, are created by two coherent laser pulses generated by an unbalanced interferometer in the pump path, the output state is

Φ=12(|EarlyXX|EarlyX+eiϕp|LateXX|LateX),
where ϕp=EXXΔt/ is the phase in the pump interferometer (EXX is the biexciton photon energy and Δt the delay in the pump interferometer). This state is a maximally entangled Bell state when ϕp=0 or π.

The quality of the time-bin entangled states is characterized with a quantum-state tomography measurement to reconstruct the density matrix [126]. Two additional unbalanced interferometers are installed for each XX and X photons. The delays on the excitation and on the two analyze interferometers are set equal and the relative phases between the three interferometers have to be stabilized during the measurement. An active stabilization system can be used, but in [117] the three interferometers share the same optical elements and operate in different spatial modes. The state tomography is performed is three bases by adjusting the phase in the X and XX analyze interferometers. A fidelity of the generated state with respect to the Bell state, Φ+ [Eq. (10) with ϕp=0], of 0.69 (3) was measured. The real part of the density matrix is plotted in Fig. 9(c). The authors attribute the reduced visibility to a nonnegligible double excitation (on the “Early” and “Late” events) and dephasings. Coupling to a 3D microcavity should improve the fidelity because a higher photon extraction efficiency would allow the system to operate in a lower excitation power regime reducing the double excitation events.

B. Controlled-not Gates

Sets of single- and two-qubit quantum gates can be combined to create any other quantum gates [127,128]. Single-qubit gates are easy to implement in the polarization space since a polarizer and half- and quarter-wave plates can define any polarization states of the Poincaré sphere. Two-qubit gates are more complex to implement because they require quantum interference between single photons. Knill et al. proposed in 2001 that any optical multi-qubit gate, in particular the two-qubit gate, could be implemented using linear optic components, projective measurements, and feed-forwards [109].

A quantum controlled-not gate is the main two-qubit entangling gate. It acts on two qubits, one commonly called the target qubit and the other the controlled qubit. This gate flips the target qubit conditionally on the control qubit state. A simplified controlled-not gate scheme was proposed in 2002 [129] and realized experimentally in 2003 [4]. Quantum interferences between the control and target photons can occur for some control qubit states and lead to a conditional target state flip. Several other schemes have been proposed theoretically and realized experimentally [130135]. All of these referenced experiments were performed using parametric downconversion sources.

1. Optical Entangling Controlled-not Gate

The schematic of the gate reviewed in this section was proposed in [4] and is seen in Fig. 10. The key elements are the two calcite crystals that displace one linear polarization and transmit the other. These crystals improve the temporal stability of the interferometers and they transform the polarization encoded input qubits into path encoded qubits. The half-wave plate installed between the two crystals simulates the 1/3:2/3 beam splitter that is the main element of this gate scheme [129].

 figure: Fig. 10.

Fig. 10. Optical quantum controlled-not gate. Schematic of the gate that was used in [27]. The key element for the stability is the calcite crystal and the main element of the gate is the central half-wave plates (blue lines) [129]. Two half-wave plates (blue lines) at 45° off their optical axis in the input and output target ports act as Hadamard gates.

Download Full Size | PDF

The gate was operated using a quantum dot source of bright indistinguishable single photons, as discussed above and reported in 2013 [27,54]. Two photons successively emitted by the quantum dot are probabilistically separated using a fiber beam splitter and delayed to arrive simultaneously on the controlled-not gate input ports. The target and control qubit photons are individually prepared, in any polarization state, using polarizers and quarter- and half-wave plates. The output state is monitored using quarter- and half-wave plates to project and detect any states on polarizers and single-photon avalanche detectors. Second-order correlations are performed on the detector outputs. This gate is probabilistic and only works when detections occur at each control and target outputs. The result of the operations are read on a second-order correlation measurement at delay 0 [129]. The truth table of the gate is determined in the rectilinear polarization basis, [|H,|V], by measuring the output state for the four possible input states: {|HH,|HV,|VH,|VV} (where |ct are the control and target qubit states). Overlaps between the measured and the ideal truth table up to 0.73 were observed.

Truth tables have also been measured in one or two bases in [28,136] using quantum dot photons as well, but the entangling capabilities of the gates have not been reported.

Entangling controlled-not gate. A quantum controlled-not gate can entangle photons that were initially independent. Indeed, if the control qubit is in a diagonal state 12(|H+|V) and the target qubit in the state |H, then the ideal output state is a maximally entangled Bell state:

Φ+=12(|HH+|VV).
To show entanglement of single photons that are initially independent using this controlled-not gate, quantum dot photons were sent into the gate and the fidelity of the output state compared to the Bell state Φ+ was measured. The control photons are prepared in the state 12(|H+|V) and the target photon in the state |H. The fidelity of the generated state is determined by measuring the degree of the correlations in three bases (rectilinear |H, |V; diagonal |H±|V; and circular |H±i|V) [137,138]. Figure 11(a) plots the fidelity as a function of the postselected source brightness. The postselected brightness is defined as the product of the brightness, p1, and the fraction of postselected photons. The postselection was performed on the arrival time of the photons with the lower postselected brightness values corresponding to shorter time differences. Without postselection and at a source brightness as high as 0.5 photons collected per pulse, the fidelity referenced to the Bell state Φ+ is above the 0.5 limit for quantum correlations. With temporal postselection, the fidelity increases as the photon indistinguishability is improved [26] and at a postselected brightness of p1=0.15, the fidelity reaches 0.71.

 figure: Fig. 11.

Fig. 11. Entangling controlled-not gate. (a) Measured fidelity of the generated state compared to the Bell state ϕ+ as a function of the source postselected brightness. The dotted line indicates the quantum correlation threshold. (b) Calculated fidelity as a function of the photons indistinguishability. The square indicates a fully distinguishable single-photon source; the circle (triangle) corresponds to the two experimental points without (with) temporal postselection. Figures adapted from [27].

Download Full Size | PDF

Perspectives. The quantum dot source used for this experiment emits photons with an indistinguishability of about 0.71 with postselection and 0.5 without. Following [129] to calculate the output coincidences for all the polarization configurations and as a function of the indistinguishably, C, one can obtain a fidelity referenced to the Bell state Φ+ of [27]

FΦ+=1+C2(2C).
This function is plotted in Fig. 11(b). Using the latest sources discussed in Section 4.B where indistinguishability values are above 0.99 [31,46], the fidelity should be extremely close to one.

 figure: Fig. 12.

Fig. 12. Controlled-not gate. When a photon is sent to a coupled quantum dot cavity system, the phase of the reflected photon depends on the quantum dot state. If the quantum dot is in the ground state |g and the cooperativity, C1, no phase is added, r=1 and the incoming state in unchanged (top). However, if the quantum dot is in the excited state |, r=1 and the incident state is flipped (bottom). Figure adapted from [118].

Download Full Size | PDF

2. Controlled-not Gate Using Quantum Dot–Cavity Strong Coupling

A final example of a quantum information device utilizing quantum dots is also a controlled-not gate. However, while the last example of a controlled-not gate used quantum dot photons as the input to the controlled-not gate, in the work of Waks et al. discussed here the quantum dot–cavity system acts as the gate [118].

There are several differences between the system used by [118] and the previous quantum-cavity system. Here the microcavity is a photonic-crystal cavity (PCC) formed in a GaAs slab, less than 200 nm thick and suspended by etching a sacrificial AlGaAs layer under the GaAs. The cavity is completed in the standard PCC way by etching a periodic array of holes into the GaAs to form the in-plane confinement. Researchers search for a cavity with an InAs quantum dot in the defect region (an area without a hole) of the structure. Unlike previous devices discussed, here the quantum dot–microcavity system is in the strong coupling regime, and the success of the device is based on the cooperativity, C=2g2/κΓ, where g is the cavity–quantum dot coupling strength, κ and Γ are the cavity and the quantum dot radiative decay rates. In [118] g/2π=12.9GHz, κ/2π=31.94GHz, and Γ/2π=5.2GHz. Since g>κ/4 the system is in the strong-coupling regime, with C=2.0. A similar experiment in a micropillar cavity showed a similar cooperativity (2.5) but in the weak-coupling regime [139].

Here, the controlled-not gate relies on the change in reflectivity when one of the two neutral exciton states is on-resonance with the cavity and the other is significantly off-resonance with the cavity. With a magnetic field of 1.6 T oriented along the growth direction (often called the Faraday geometry), one of the exciton states is tuned on-resonance with the cavity. Because of the magnetic field, the exciton transitions have circular polarizations, σ+ or σ. Input states are |H and |V, rotated π/4 to cavity polarization axis, |x and |y, so that |H=12(|x+|y) and |V=12(|x+|y). The important observation is that upon reflection, the states are transformed to |H=12(r|x+|y) and |V=12(r|x+|y), where r is the reflectivity.

The logic goes as follows for light incident on the cavity (Fig. 12). If the quantum dot is in the exciton state that is detuned from the cavity, no transitions at the cavity resonance can occur since the off-resonant exciton is occupied. However, if the quantum dot is in the ground state (the empty state), strong cavity absorption will take place, modifying the reflectivity. How much the reflectivity is modified depends on C since r=(C1)/(C+1). If C1, r1. Thus, in this ideal situation under one condition the incoming state is flipped (r=1), while in the other condition the state is reflected unchanged (r=1).

Results show that when the system is in the off-resonant state, the qubit-flip probabilities are PHV=0.93 (3) and PVH=0.98 (3). This is due to the bare cavity effect since the magnetic field shifts the σ state far off the cavity resonance. Conversely, when the system in the ground state and incident light puts the system in the strong-coupling regime, the reflected state should ideally be unchanged. Here the probabilities of returning the same state are lower, PVV=0.58 (4) and PHH=0.61 (7). This makes good sense since while C is very high for a quantum dot microcavity system, C=2, it falls short of the C1 ideal condition, so that r=1/3, not 1. As with the other two demonstrations, this is an excellent first step but with improvements to the cooperatively better results can be expected in the future.

7. CONCLUSION AND PERSPECTIVES

In this review paper, we described some major progress made on quantum-dot-based light sources. These sources are still under development. For instance, there is ongoing research to improve the quality of the indistinguishability of photons emitted from the source very far apart temporally. Recently, photon indistinguishability values above 0.85 have been measured with photons emitted 160 ns apart [140] and very recently extended to over 10 μs with indistinguishability of about 0.92 [141]. Long temporal streams of more than five highly indistinguishable photons will allow new possibilities for quantum information processing. For instance, they could be sent to a multiport quantum circuit to perform scalable quantum information processing on a large Hilbert space. One could think of the boson sampling experiments that are currently limited by the source brightness [8,142,143145].

The interference of photons coming from several disparate quantum-dot-based sources is a key element for long-distance quantum communications, such as optical quantum teleportation and entanglement swapping protocols [1,112]. Interferences between photons emitted from distant quantum dots have been performed since 2010 [146149]. Interference visibilities are typically in the range 25%–40%, and go up to about 80% with spectral filtering; however, this strongly reduces the source brightness of the single-photon emission [150]. The generation of cluster states is an alternative scheme of quantum computation and it might be possible to obtain large-scale cluster state by coupling several quantum dots [151153].

Several recent experiments are trying to better understand the spectral jittering effects that are a main limitation for interfering photons emitted far apart temporally (Section 2.B). To that aim, the temporal dynamics of photon energies and linewidths are being studied [45,154].

Major progress has been made on the semiconductor quantum-dot-based sources since the first demonstration of their single-photon properties in 2000. The coupling with microcavities and the use of resonant fluorescence excitation strongly improved their properties and now, these sources can deterministically emit single photons with almost unity purity and indistinguishability, and at high brightness values. Quantum-dot-based sources will be used more and more in optical quantum information processing experiments, experiments that were up to some years ago almost exclusively reserved for spontaneous parametric downconversion sources.

Funding

National Science Foundation (NSF) through PFC@JQI; Army Research Office (ARO) Multidisciplinary University Research Initiative on Hybrid quantum interactions.

Acknowledgment

We greatly appreciate the support of P. Senellart here. A portion of this review is based on work O. G. conducted in the research group of P. Senellart at the Laboratoire de Photonique et de Nanostructures, CNRS, Marcoussis, France.

REFERENCES

1. D. Bouwmeester, J.-W. Pan, K. Mattle, M. Eibl, H. Weinfurter, and A. Zeilinger, “Experimental quantum teleportation,” Nature 390, 575–579 (1997). [CrossRef]  

2. J. W. Pan, Z. B. Chen, C. Y. Lu, H. Weinfurter, A. Zeilinger, and M. Zukowski, “Multiphoton entanglement and interferometry,” Rev. Mod. Phys. 84, 777–838 (2012). [CrossRef]  

3. M. Zukowski, A. Zeilinger, and H. Weinfurter, “Entangling photons radiated by independent pulsed sources,” Ann. N.Y. Acad. Sci. 755, 91–102 (1995). [CrossRef]  

4. J. L. O’Brien, G. J. Pryde, A. G. White, T. C. Ralph, and D. Branning, “Demonstration of an all-optical quantum controlled-NOT gate,” Nature 426, 264–267 (2003). [CrossRef]  

5. B. P. Lanyon, J. D. Whitfield, G. G. Gillett, M. E. Goggin, M. P. Almeida, I. Kassal, J. D. Biamonte, M. Mohseni, B. J. Powell, M. Barbieri, A. Aspuru-Guzik, and A. G. White, “Towards quantum chemistry on a quantum computer,” Nat. Chem. 2, 106–111 (2010). [CrossRef]  

6. J. Carolan, C. Harrold, C. Sparrow, E. Martín-López, N. J. Russell, J. W. Silverstone, P. J. Shadbolt, N. Matsuda, M. Oguma, M. Itoh, G. D. Marshall, M. G. Thompson, J. C. F. Matthews, T. Hashimoto, J. L. O’Brien, and A. Laing, “Universal linear optics,” Science 349, 711–716 (2015).

7. T. Jennewein, M. Barbieri, and A. G. White, “Single-photon device requirements for operating linear optics quantum computing outside the post-selection basis,” J. Mod. Opt. 58, 276–287 (2011). [CrossRef]  

8. J. Carolan, J. D. A. Meinecke, P. J. Shadbolt, N. J. Russell, N. Ismail, K. Wörhoff, T. Rudolph, M. G. Thompson, J. L. O’Brien, J. C. F. Matthews, and A. Laing, “On the experimental verification of quantum complexity in linear optics,” Nat. Photonics 8, 621–626 (2014). [CrossRef]  

9. A. L. Migdall, D. Branning, and S. Castelletto, “Tailoring single-photon and multiphoton probabilities of a single-photon on-demand source,” Phys. Rev. A 66, 053805 (2002). [CrossRef]  

10. B. Lounis and W. E. Moerner, “Single photons on demand from a single molecule at room temperature,” Nature 407, 491–493 (2000). [CrossRef]  

11. C. Kurtsiefer, S. Mayer, P. Zarda, and H. Weinfurter, “Stable solid-state source of single photons,” Phys. Rev. Lett. 85, 290–293 (2000). [CrossRef]  

12. R. Brouri, A. Beveratos, J. P. Poizat, and P. Grangier, “Photon antibunching in the fluorescence of individual color centers in diamond,” Opt. Lett. 25, 1294–1296 (2000). [CrossRef]  

13. B. Lounis, H. A. Bechtel, and D. Gerion, “Photon antibunching in single CdSe/ZnS quantum dot fluorescence,” Chem. Phys. Lett. 329, 399–404 (2000). [CrossRef]  

14. B. Mahler, P. Spinicelli, S. Buil, X. Quelin, J.-P. Hermier, and B. Dubertret, “Towards non-blinking colloidal quantum dots,” Nat. Mater. 7, 659–664 (2008). [CrossRef]  

15. P. Michler, A. Kiraz, C. Becher, W. V. Schoenfeld, P. M. Petroff, L. Zhang, E. Hu, and A. Imamoglu, “A quantum dot single-photon turnstile device,” Science 290, 2282–2285 (2000). [CrossRef]  

16. C. Santori, D. Fattal, J. Vučković, G. S. Solomon, and Y. Yamamoto, “Indistinguishable photons from a single-photon device,” Nature 419, 594–597 (2002). [CrossRef]  

17. R. J. Young, R. M. Stevenson, P. Atkinson, K. Cooper, D. A. Ritchie, and A. J. Shields, “Improved fidelity of triggered entangled photons from single quantum dots,” New J. Phys. 8, 29 (2006). [CrossRef]  

18. N. Akopian, N. Lindner, E. Poem, Y. Berlatzky, J. Avron, D. Gershoni, B. Gerardot, and P. Petroff, “Entangled photon pairs from semiconductor quantum dots,” Phys. Rev. Lett. 96, 7–10 (2006). [CrossRef]  

19. K. De Greve, L. Yu, P. L. McMahon, J. S. Pelc, C. M. Natarajan, N. Y. Kim, E. Abe, S. Maier, C. Schneider, M. Kamp, S. Höfling, R. H. Hadfield, A. Forchel, M. M. Fejer, and Y. Yamamoto, “Quantum-dot spin-photon entanglement via frequency downconversion to telecom wavelength,” Nature 491, 421–425 (2012). [CrossRef]  

20. W. B. Gao, P. Fallahi, E. Togan, J. Miguel-Sanchez, and A. Imamoglu, “Observation of entanglement between a quantum dot spin and a single photon,” Nature 491, 426–430 (2012). [CrossRef]  

21. J. R. Schaibley, A. P. Burgers, G. A. McCracken, L. M. Duan, P. R. Berman, D. G. Steel, A. S. Bracker, D. Gammon, and L. J. Sham, “Demonstration of quantum entanglement between a single electron spin confined to an InAs quantum dot and a photon,” Phys. Rev. Lett. 110, 1–5 (2013). [CrossRef]  

22. G. W. Bryant and G. Solomon, Optics of Quantum Dots and Wires (Artech House, 2005).

23. L. Landin, M. S. Miller, M.-E. Pistol, C. E. Pryor, and L. Samuelson, “Optical studies of individual InAs quantum dots in GaAs: few-particle effects,” Science 280, 262–264 (1998). [CrossRef]  

24. R. Hanbury Brown and R. Q. Twiss, “Correlation between photons in two coherent beams of light,” Nature 177, 27–29 (1956). [CrossRef]  

25. E. A. Goldschmidt, F. Piacentini, I. R. Berchera, S. V. Polyakov, S. Peters, S. Kück, G. Brida, I. P. Degiovanni, A. Migdall, and M. Genovese, “Mode reconstruction of a light field by multiphoton statistics,” Phys. Rev. A 88, 1–5 (2013). [CrossRef]  

26. E. B. Flagg, S. V. Polyakov, T. Thomay, and G. S. Solomon, “Dynamics of nonclassical light from a single solid-state quantum emitter,” Phys. Rev. Lett. 109, 163601 (2012). [CrossRef]  

27. O. Gazzano, M. Almeida, A. Nowak, S. Portalupi, A. Lemaître, I. Sagnes, A. White, and P. Senellart, “Entangling quantum-logic gate operated with an ultrabright semiconductor single-photon source,” Phys. Rev. Lett. 110, 250501 (2013). [CrossRef]  

28. Y.-M. He, Y. He, Y.-J. Wei, D. Wu, M. Atatüre, C. Schneider, S. Höfling, M. Kamp, C.-Y. Lu, and J.-W. Pan, “On-demand semiconductor single-photon source with near-unity indistinguishability,” Nat. Nanotechnol. 8, 213–217 (2013). [CrossRef]  

29. M. Müller, S. Bounouar, K. D. Jöns, M. Glässl, and P. Michler, “On-demand generation of indistinguishable polarization-entangled photon pairs,” Nat. Photonics 8, 224–228 (2014). [CrossRef]  

30. Y.-J. Wei, Y.-M. He, M.-C. Chen, Y.-N. Hu, Y. He, D. Wu, C. Schneider, M. Kamp, S. Höfling, C.-Y. Lu, and J.-W. Pan, “Deterministic and robust generation of single photons from a single quantum dot with 99.5% indistinguishability using adiabatic rapid passage,” Nano Lett. 14, 6515–6519 (2014). [CrossRef]  

31. N. Somaschi, V. Giesz, L. De Santis, J. C. Loredo, M. P. Almeida, G. Hornecker, S. L. Portalupi, T. Grange, C. Anton, J. Demory, C. Gomez, I. Sagnes, N. D. L. Kimura, A. Lemaitre, A. Auffeves, A. G. White, L. Lanco, and P. Senellart, “Near optimal single photon sources in the solid state,” Nat. Photonics 10, 340–345 (2015). [CrossRef]  

32. X. Xu, Y. Wu, B. Sun, Q. Huang, J. Cheng, D. G. Steel, A. S. Bracker, D. Gammon, C. Emary, and L. J. Sham, “Fast spin state initialization in a singly charged InAs-GaAs quantum dot by optical cooling,” Phys. Rev. Lett. 99, 097401 (2007). [CrossRef]  

33. G. Wrigge, I. Gerhardt, J. Hwang, G. Zumofen, and V. Sandoghdar, “Efficient coupling of photons to a single molecule and the observation of its resonance fluorescence,” Nat. Phys. 4, 60–66 (2007).

34. A. Muller, E. B. Flagg, P. Bianucci, X. Y. Wang, D. G. Deppe, W. Ma, J. Zhang, G. J. Salamo, M. Xiao, and C. K. Shih, “Resonance fluorescence from a coherently driven semiconductor quantum dot in a cavity,” Phys. Rev. Lett. 99, 2–5 (2007). [CrossRef]  

35. A. Muller, W. Fang, J. Lawall, and G. S. Solomon, “Emission spectrum of a dressed exciton-biexciton complex in a semiconductor quantum dot,” Phys. Rev. Lett. 101, 027401 (2008). [CrossRef]  

36. L. Monniello, A. Reigue, R. Hostein, A. Lemaitre, A. Martinez, R. Grousson, and V. Voliotis, “Indistinguishable single photons generated by a quantum dot under resonant excitation observable without postselection,” Phys. Rev. B 90, 1–5 (2014). [CrossRef]  

37. D. E. Browne and T. Rudolph, “Resource-efficient linear optical quantum computation,” Phys. Rev. Lett. 95, 010501 (2005). [CrossRef]  

38. O. Gühne and P. Hyllus, “Investigating three qubit entanglement with local measurements,” Int. J. Theor. Phys. 42, 1001–1013 (2003). [CrossRef]  

39. C. H. Bennett and G. Brassard, “Quantum cryptography: public key distribution and coin tossing,” in Proceedings of the IEEE International Conference on Computers, Systems, and Signal Processing, Bangalore, India, 1984, p. 175.

40. H. Briegel, W. Dür, J. Cirac, and P. Zoller, “Quantum repeaters: the role of imperfect local operations in quantum communication,” Phys. Rev. Lett. 81, 5932–5935 (1998). [CrossRef]  

41. W. Dür, H.-J. Briegel, J. I. Cirac, and P. Zoller, “Quantum repeaters based on entanglement purification,” Phys. Rev. A 59, 169–181 (1999). [CrossRef]  

42. R. A. Campos, B. E. A. Saleh, and M. C. Teich, “Quantum-mechanical lossless beam splitter: SU(2) symmetry and photon statistics,” Phys. Rev. A 40, 1371–1384 (1989). [CrossRef]  

43. L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University, 1995).

44. C. K. Hong, Z. Y. Ou, and L. Mandel, “Measurement of subpicosecond time intervals between two photons by interference,” Phys. Rev. Lett. 59, 2044–2046 (1987). [CrossRef]  

45. A. V. Kuhlmann, J. Houel, A. Ludwig, L. Greuter, D. Reuter, A. D. Wieck, M. Poggio, and R. J. Warburton, “Charge noise and spin noise in a semiconductor quantum device,” Nat. Phys. 9, 570–575 (2013). [CrossRef]  

46. X. Ding, Y. He, Z.-C. Duan, N. Gregersen, M.-C. Chen, S. Unsleber, S. Maier, C. Schneider, M. Kamp, S. Höfling, C.-Y. Lu, and J.-W. Pan, “On-demand single photons with high extraction efficiency and near-unity indistinguishability from a resonantly driven quantum dot in a micropillar,” Phys. Rev. Lett. 116, 020401 (2016). [CrossRef]  

47. A. K. Nowak, S. L. Portalupi, V. Giesz, O. Gazzano, C. Dal Savio, P.-F. Braun, K. Karrai, C. Arnold, L. Lanco, I. Sagnes, A. Lemaître, and P. Senellart, “Deterministic and electrically tunable bright single-photon source,” Nat. Commun. 5, 3240 (2014). [CrossRef]  

48. P. H. Eberhard, “Background level and counter efficiencies required for a loophole-free Einstein-Podolsky-Rosen experiment,” Phys. Rev. A 47, R747–R750 (1993). [CrossRef]  

49. J. Finley, P. Fry, A. Ashmore, A. Lemaître, A. Tartakovskii, R. Oulton, D. Mowbray, M. Skolnick, M. Hopkinson, P. D. Buckle, and P. Maksym, “Observation of multicharged excitons and biexcitons in a single InGaAs quantum dot,” Phys. Rev. B 63, 2–5 (2001).

50. M. Metcalfe, S. M. Carr, A. Muller, G. S. Solomon, and J. Lawall, “Resolved sideband emission of InAs/GaAs quantum dots strained by surface acoustic waves,” Phys. Rev. Lett. 105, 037401 (2010). [CrossRef]  

51. H. S. Nguyen, G. Sallen, C. Voisin, P. Roussignol, C. Diederichs, and G. Cassabois, “Optically gated resonant emission of single quantum dots,” Phys. Rev. Lett. 108, 057401 (2012). [CrossRef]  

52. H. Benisty, H. De Neve, and C. Weisbuch, “Impact of planar microcavity effects on light extraction-Part I: basic concepts and analytical trends,” IEEE J. Quantum Electron. 34, 1612–1631 (1998). [CrossRef]  

53. M. Munsch, N. S. Malik, E. Dupuy, A. Delga, J. Bleuse, J. M. Gérard, J. Claudon, N. Gregersen, and J. Mørk, “Dielectric GaAs antenna ensuring an efficient broadband coupling between an InAs quantum dot and a Gaussian optical beam,” Phys. Rev. Lett. 110, 177402 (2013). [CrossRef]  

54. O. Gazzano, S. Michaelis de Vasconcellos, C. Arnold, A. Nowak, E. Galopin, I. Sagnes, L. Lanco, A. Lemaître, and P. Senellart, “Bright solid-state sources of indistinguishable single photons,” Nat. Commun. 4, 1425 (2013). [CrossRef]  

55. V. Zwiller and G. Björk, “Improved light extraction from emitters in high refractive index materials using solid immersion lenses,” J. Appl. Phys. 92, 660–665 (2002). [CrossRef]  

56. L. Sapienza, M. Davanço, A. Badolato, and K. Srinivasan, “Nanoscale optical positioning of single quantum dots for bright and pure single-photon emission,” Nat. Commun. 6, 7833 (2015). [CrossRef]  

57. A. Kress, F. Hofbauer, N. Reinelt, M. Kaniber, H. J. Krenner, R. Meyer, G. Böhm, and J. J. Finley, “Manipulation of the spontaneous emission dynamics of quantum dots in two-dimensional photonic crystals,” Phys. Rev. B 71, 1–4 (2005). [CrossRef]  

58. T. Lund-Hansen, S. Stobbe, B. Julsgaard, H. Thyrrestrup, T. Sünner, M. Kamp, A. Forchel, and P. Lodahl, “Experimental realization of highly efficient broadband coupling of single quantum dots to a photonic crystal waveguide,” Phys. Rev. Lett. 101, 113903 (2008). [CrossRef]  

59. A. Schwagmann, S. Kalliakos, I. Farrer, J. P. Griffiths, G. A. C. Jones, D. A. Ritchie, and A. J. Shields, “On-chip single photon emission from an integrated semiconductor quantum dot into a photonic crystal waveguide,” Appl. Phys. Lett. 99, 261108 (2011). [CrossRef]  

60. A. Laucht, S. Pütz, T. Günthner, N. Hauke, R. Saive, S. Frédérick, M. Bichler, M.-C. Amann, A. W. Holleitner, M. Kaniber, and J. J. Finley, “A waveguide-coupled on-chip single-photon source,” Phys. Rev. X 2, 011014 (2012).

61. M. Arcari, I. Söllner, A. Javadi, S. Lindskov Hansen, S. Mahmoodian, J. Liu, H. Thyrrestrup, E. H. Lee, J. D. Song, S. Stobbe, and P. Lodahl, “Near-unity coupling efficiency of a quantum emitter to a photonic crystal waveguide,” Phys. Rev. Lett. 113, 093603 (2014). [CrossRef]  

62. P. Lodahl, S. Mahmoodian, and S. Stobbe, “Interfacing single photons and single quantum dots with photonic nanostructures,” Rev. Mod. Phys. 87, 347–400 (2015). [CrossRef]  

63. J. Claudon, J. Bleuse, N. S. Malik, M. Bazin, P. Jaffrennou, N. Gregersen, C. Sauvan, P. Lalanne, and J. M. Gérard, “A highly efficient single-photon source based on a quantum dot in a photonic nanowire,” Nat. Photonics 4, 174–177 (2010).

64. M. E. Reimer, G. Bulgarini, N. Akopian, M. Hocevar, M. B. Bavinck, M. A. Verheijen, E. P. A. M. Bakkers, L. P. Kouwenhoven, and V. Zwiller, “Bright single-photon sources in bottom-up tailored nanowires,” Nat. Commun. 3, 737 (2012). [CrossRef]  

65. J. Bleuse, J. Claudon, M. Creasey, N. S. Malik, J. M. Gérard, I. Maksymov, J.-P. Hugonin, and P. Lalanne, “Inhibition, enhancement, and control of spontaneous emission in photonic nanowires,” Phys. Rev. Lett. 106, 103601 (2011). [CrossRef]  

66. R. M. Stevenson, R. J. Young, P. Atkinson, K. Cooper, D. A. Ritchie, and A. J. Shields, “A semiconductor source of triggered entangled photon pairs,” Nature 439, 179–182 (2006). [CrossRef]  

67. A. Dousse, J. Suffczyński, A. Beveratos, O. Krebs, A. Lemaître, I. Sagnes, J. Bloch, P. Voisin, and P. Senellart, “Ultrabright source of entangled photon pairs,” Nature 466, 217–220 (2010). [CrossRef]  

68. I. Yeo, N. S. Malik, M. Munsch, E. Dupuy, J. Bleuse, Y. M. Niquet, J. M. Gérard, J. Claudon, E. Wagner, S. Seidelin, A. Auffèves, J. P. Poizat, and G. Nogues, “Surface effects in a semiconductor photonic nanowire and spectral stability of an embedded single quantum dot,” Appl. Phys. Lett. 99, 233106 (2011). [CrossRef]  

69. S. Ates, L. Sapienza, M. Davanco, A. Badolato, and K. Srinivasan, “Bright single-photon emission from a quantum dot in a circular Bragg grating microcavity,” IEEE J. Sel. Top. Quantum Electron. 18, 1711–1721 (2012). [CrossRef]  

70. K. H. Madsen, S. Ates, J. Liu, A. Javadi, S. M. Albrecht, I. Yeo, S. Stobbe, and P. Lodahl, “Efficient out-coupling of high-purity single photons from a coherent quantum dot in a photonic-crystal cavity,” Phys. Rev. B 90, 155303 (2014). [CrossRef]  

71. O. Gazzano, S. M. de Vasconcellos, K. Gauthron, C. Symonds, J. Bloch, P. Voisin, J. Bellessa, A. Lemaître, and P. Senellart, “Evidence for confined Tamm plasmon modes under metallic microdisks and application to the control of spontaneous optical emission,” Phys. Rev. Lett. 107, 247402 (2011). [CrossRef]  

72. O. Gazzano, S. M. de Vasconcellos, K. Gauthron, C. Symonds, P. Voisin, J. Bellessa, A. Lemaître, and P. Senellart, “Single photon source using confined Tamm plasmon modes,” Appl. Phys. Lett. 100, 232111 (2012). [CrossRef]  

73. S. Haroche and J. Raimond, Radiative Properties of Rydberg States in Resonant Cavities, Vol. 20 of Advances in Atomic, Molecular and Optical Physics (Academic, 1985), pp. 347–411.

74. E. M. Purcell, “Spontaneous emission probabilities at radio frequencies,” Phys. Rev. 69, 681 (1946). [CrossRef]  

75. J. M. Gérard and B. Gayral, “Strong Purcell effect for InAs quantum boxes in three-dimensional solid-state microcavities,” J. Lightwave Technol. 17, 2089–2095 (1999). [CrossRef]  

76. G. Solomon, M. Pelton, and Y. Yamamoto, “Single-mode spontaneous emission from a single quantum dot in a three-dimensional microcavity,” Phys. Rev. Lett. 86, 3903–3906 (2001). [CrossRef]  

77. J. M. Gérard, B. Sermage, B. Gayral, B. Legrand, E. Costard, and V. Thierry-Mieg, “Enhanced spontaneous emission by quantum boxes in a monolithic optical microcavity,” Phys. Rev. Lett. 81, 1110–1113 (1998). [CrossRef]  

78. D. Englund, D. Fattal, E. Waks, G. S. Solomon, B. Zhang, T. Nakaoka, Y. Arakawa, Y. Yamamoto, and J. Vučković, “Controlling the spontaneous emission rate of single quantum dots in a two-dimensional photonic crystal,” Phys. Rev. Lett. 95, 013904 (2005). [CrossRef]  

79. B. Gayral, J. M. Gérard, B. Sermage, A. Lemaître, and C. Dupuis, “Time-resolved probing of the Purcell effect for InAs quantum boxes in GaAs microdisks,” Appl. Phys. Lett. 78, 2828 (2001). [CrossRef]  

80. J. Reithmaier, G. Sek, A. Löffler, C. Hofmann, S. Kuhn, S. Reitzenstein, L. Keldysh, V. Kulakovskii, T. Reinecke, and A. Forchel, “Strong coupling in a single quantum dot-semiconductor microcavity system,” Nature 432, 197–200 (2004). [CrossRef]  

81. T. Yoshie, A. Scherer, J. Hendrickson, G. Khitrova, H. M. Gibbs, G. Rupper, C. Ell, O. B. Shchekin, and D. G. Deppe, “Vacuum Rabi splitting with a single quantum dot in a photonic crystal nanocavity,” Nature 432, 200–203 (2004). [CrossRef]  

82. E. Peter, P. Senellart, D. Martrou, A. Lemaître, J. Hours, J. M. Gérard, and J. Bloch, “Exciton-photon strong-coupling regime for a single quantum dot embedded in a microcavity,” Phys. Rev. Lett. 95, 067401 (2005). [CrossRef]  

83. A. Muller, E. B. Flagg, M. Metcalfe, J. Lawall, and G. S. Solomon, “Coupling an epitaxial quantum dot to a fiber-based external-mirror microcavity,” Appl. Phys. Lett. 95, 173101 (2009). [CrossRef]  

84. P. Yeh, A. Yariv, and C. Hong, “Electromagnetic propagation in periodic stratified media. I. General theory,” J. Opt. Soc. Am. 67, 423 (1977). [CrossRef]  

85. G. Björk, H. Heitmann, and Y. Yamamoto, “Spontaneous-emission coupling factor and mode characteristics of planar dielectric microcavity lasers,” Phys. Rev. A 47, 4451–4463 (1993). [CrossRef]  

86. J. M. Gérard, D. Barrier, J. Y. Marzin, R. Kuszelewicz, L. Manin, E. Costard, V. Thierry-Mieg, and T. Rivera, “Quantum boxes as active probes for photonic microstructures: the pillar microcavity case,” Appl. Phys. Lett. 69, 449 (1996). [CrossRef]  

87. H. Rigneault, J. Broudic, B. Gayral, and J. M. Gérard, “Far-field radiation from quantum boxes located in pillar microcavities,” Opt. Lett. 26, 1595–1597 (2001). [CrossRef]  

88. C. Santori, M. Pelton, G. Solomon, Y. Dale, and Y. Yamamoto, “Triggered single photons from a quantum dot,” Phys. Rev. Lett. 86, 1502–1505 (2001). [CrossRef]  

89. A. Badolato, K. Hennessy, M. Atatüre, J. Dreiser, E. Hu, P. M. Petroff, and A. Imamoglu, “Deterministic coupling of single quantum dots to single nanocavity modes,” Science 308, 1158–1161 (2005). [CrossRef]  

90. K. Hennessy, A. Badolato, M. Winger, D. Gerace, M. Atatüre, S. Gulde, S. Fält, E. L. Hu, and A. Imamoglu, “Quantum nature of a strongly coupled single quantum dot-cavity system,” Nature 445, 896–899 (2007). [CrossRef]  

91. A. Dousse, L. Lanco, J. Suffczynski, E. Semenova, A. Miard, A. Lemaître, I. Sagnes, C. Roblin, J. Bloch, and P. Senellart, “Controlled light-matter coupling for a single quantum dot embedded in a pillar microcavity using far-field optical lithography,” Phys. Rev. Lett. 101, 267404 (2008). [CrossRef]  

92. E. Grilli, M. Guzzi, R. Zamboni, and L. Pavesi, “High-precision determination of the temperature dependence of the fundamental energy gap in gallium arsenide,” Phys. Rev. B 45, 1638–1644 (1992). [CrossRef]  

93. M. Gschrey, A. Thoma, P. Schnauber, M. Seifried, R. Schmidt, B. Wohlfeil, L. Krüger, J.-H. Schulze, T. Heindel, S. Burger, F. Schmidt, A. Strittmatter, S. Rodt, and S. Reitzenstein, “Highly indistinguishable photons from deterministic quantum-dot microlenses utilizing three-dimensional in situ electron-beam lithography,” Nat. Commun. 6, 7662 (2015). [CrossRef]  

94. M. Gschrey, F. Gericke, A. Schußler, R. Schmidt, J.-H. Schulze, T. Heindel, S. Rodt, A. Strittmatter, and S. Reitzenstein, “In situ electron-beam lithography of deterministic single-quantum-dot mesa-structures using low-temperature cathodoluminescence spectroscopy,” Appl. Phys. Lett. 102, 251113 (2013). [CrossRef]  

95. P. Gallo, M. Felici, B. Dwir, K. A. Atlasov, K. F. Karlsson, A. Rudra, A. Mohan, G. Biasiol, L. Sorba, and E. Kapon, “Integration of site-controlled pyramidal quantum dots and photonic crystal membrane cavities,” Appl. Phys. Lett. 92, 263101 (2008). [CrossRef]  

96. D. Dalacu, K. Mnaymneh, V. Sazonova, P. J. Poole, G. C. Aers, J. Lapointe, R. Cheriton, A. J. SpringThorpe, and R. Williams, “Deterministic emitter-cavity coupling using a single-site controlled quantum dot,” Phys. Rev. B 82, 033301 (2010). [CrossRef]  

97. K. D. Jöns, P. Atkinson, M. Müller, M. Heldmaier, S. M. Ulrich, O. G. Schmidt, and P. Michler, “Triggered indistinguishable single photons with narrow line widths from site-controlled quantum dots,” Nano Lett. 13, 126–130 (2013). [CrossRef]  

98. C. Arnold, V. Loo, A. Lemaître, I. Sagnes, O. Krebs, P. Voisin, P. Senellart, and L. Lanco, “Optical bistability in a quantum dots/micropillar device with a quality factor exceeding 200 000,” Appl. Phys. Lett. 100, 111111 (2012). [CrossRef]  

99. M. Lermer, N. Gregersen, F. Dunzer, S. Reitzenstein, S. Höfling, J. Mørk, L. Worschech, M. Kamp, and A. Forchel, “Bloch-wave engineering of quantum dot micropillars for cavity quantum electrodynamics experiments,” Phys. Rev. Lett. 108, 057402 (2012). [CrossRef]  

100. M. Pelton, C. Santori, J. Vučković, B. Zhang, G. S. Solomon, J. Plant, and Y. Yamamoto, “Efficient source of single photons: a single quantum dot in a micropost microcavity,” Phys. Rev. Lett. 89, 233602 (2002). [CrossRef]  

101. E. Moreau, I. Robert, L. Manin, V. Thierry-Mieg, J. M. Gérard, and I. Abram, “A single-mode solid-state source of single photons based on isolated quantum dots in a micropillar,” Physica E 13, 418–422 (2002). [CrossRef]  

102. S. Strauf, N. G. Stoltz, M. T. Rakher, L. A. Coldren, P. M. Petroff, and D. Bouwmeester, “High-frequency single-photon source with polarization control,” Nat. Photonics 1, 704–708 (2007). [CrossRef]  

103. S. Laurent, S. Varoutsis, L. Le Gratiet, A. Lemaître, I. Sagnes, F. Raineri, A. Levenson, I. Robert-Philip, and I. Abram, “Indistinguishable single photons from a single-quantum dot in a two-dimensional photonic crystal cavity,” Appl. Phys. Lett. 87, 163107 (2005). [CrossRef]  

104. S. Ates, S. M. Ulrich, S. Reitzenstein, A. Löffler, A. Forchel, and P. Michler, “Post-selected indistinguishable photons from the resonance fluorescence of a single quantum dot in a microcavity,” Phys. Rev. Lett. 103, 167402 (2009). [CrossRef]  

105. W. Wu, J. R. Tucker, G. S. Solomon, and J. S. Harris, “Atom-resolved scanning tunneling microscopy of vertically ordered InAs quantum dots,” Appl. Phys. Lett. 71, 1083 (1997). [CrossRef]  

106. G. S. Solomon, J. A. Trezza, and J. S. Harris, “Substrate temperature and monolayer coverage effects on epitaxial ordering of InAs and InGaAs islands on GaAs,” Appl. Phys. Lett. 66, 991 (1995). [CrossRef]  

107. S. Unsleber, Y.-M. He, S. Maier, S. Gerhardt, C.-Y. Lu, J.-W. Pan, M. Kamp, C. Schneider, and S. Höfling, “Highly indistinguishable on-demand resonance fluorescence photons from a deterministic quantum dot micropillar device with 74% extraction efficiency,” Opt. Express 24, 8539–8546 (2015).

108. D. Bouwmeester, A. K. Ekert, and A. T. Zeilinger, The Physics of Quantum Information (Springer, 2000).

109. E. Knill, R. Laflamme, and G. Milburn, “A scheme for efficient quantum computation with linear optics,” Nature 409, 46–52 (2001). [CrossRef]  

110. H. de Riedmatten, I. Marcikic, W. Tittel, H. Zbinden, D. Collins, and N. Gisin, “Long distance quantum teleportation in a quantum relay configuration,” Phys. Rev. Lett. 92, 047904 (2004). [CrossRef]  

111. J.-W. Pan, D. Bouwmeester, H. Weinfurter, and A. Zeilinger, “Experimental entanglement swapping: entangling photons that never interacted,” Phys. Rev. Lett. 80, 3891–3894 (1998). [CrossRef]  

112. H. De Riedmatten, I. Marcikic, J. A. W. Van Houwelingen, W. Tittel, H. Zbinden, and N. Gisin, “Long-distance entanglement swapping with photons from separated sources,” Phys. Rev. A 71, 2–5 (2005). [CrossRef]  

113. P. G. Kwiat, K. Mattle, H. Weinfurter, A. Zeilinger, A. Sergienko, and Y. Shih, “New high-intensity source of polarization-entangled photon pairs,” Phys. Rev. Lett. 75, 4337–4341 (1995). [CrossRef]  

114. O. Benson, C. Santori, M. Pelton, and Y. Yamamoto, “Regulated and entangled photons from a single quantum dot,” Phys. Rev. Lett. 84, 2513–2516 (2000). [CrossRef]  

115. R. Stevenson, A. Hudson, A. Bennett, R. Young, C. Nicoll, D. Ritchie, and A. Shields, “Evolution of entanglement between distinguishable light states,” Phys. Rev. Lett. 101, 170501 (2008). [CrossRef]  

116. R. Trotta, J. S. Wildmann, E. Zallo, O. G. Schmidt, and A. Rastelli, “Highly entangled photons from hybrid piezoelectric-semiconductor quantum dot devices,” Nano Lett. 14, 3439–3444 (2014). [CrossRef]  

117. H. Jayakumar, A. Predojević, T. Kauten, T. Huber, G. S. Solomon, and G. Weihs, “Time-bin entangled photons from a quantum dot,” Nat. Commun. 5, 4251 (2014). [CrossRef]  

118. H. Kim, R. Bose, T. C. Shen, G. S. Solomon, and E. Waks, “A quantum logic gate between a solid-state quantum bit and a photon,” Nat. Photonics 7, 373–377 (2013). [CrossRef]  

119. T. Honjo, H. Takesue, H. Kamada, Y. Nishida, O. Tadanaga, M. Asobe, and K. Inoue, “Long-distance distribution of time-bin entangled photon pairs over 100 km using frequency up-conversion detectors,” Opt. Express 15, 13957–13964 (2007). [CrossRef]  

120. J. F. Dynes, H. Takesue, Z. L. Yuan, A. W. Sharpe, K. Harada, T. Honjo, H. Kamada, O. Tadanaga, Y. Nishida, M. Asobe, and A. J. Shields, “Efficient entanglement distribution over 200 kilometers,” Opt. Express 17, 11440–11449 (2009). [CrossRef]  

121. J. Brendel, N. Gisin, W. Tittel, and H. Zbinden, “Pulsed energy-time entangled twin-photon source for quantum communication,” Phys. Rev. Lett. 82, 2594–2597 (1999). [CrossRef]  

122. I. Marcikic, H. de Riedmatten, W. Tittel, V. Scarani, H. Zbinden, and N. Gisin, “Time-bin entangled qubits for quantum communication created by femtosecond pulses,” Phys. Rev. A 66, 062308 (2002). [CrossRef]  

123. C. Simon and J. P. Poizat, “Creating single time-bin-entangled photon pairs,” Phys. Rev. Lett. 94, 030502 (2005). [CrossRef]  

124. H. Jayakumar, A. Predojević, T. Huber, T. Kauten, G. S. Solomon, and G. Weihs, “Deterministic photon pairs and coherent optical control of a single quantum dot,” Phys. Rev. Lett. 110, 135505 (2013). [CrossRef]  

125. S. Stufler, P. Machnikowski, P. Ester, M. Bichler, V. M. Axt, T. Kuhn, and A. Zrenner, “Two-photon Rabi oscillations in a single InGaAs/GaAs quantum dot,” Phys. Rev. B 73, 125304 (2006). [CrossRef]  

126. H. Takesue and Y. Noguchi, “Implementation of quantum state tomography for time-bin entangled photon pairs,” Opt. Express 17, 10976–10989 (2009). [CrossRef]  

127. M. A. Nielsen and I. L. Chuang, Quantum Computation and Quantum Information (Cambridge University, 2000).

128. D. P. DiVincenzo and D. Loss, “Quantum information is physical,” Superlattices Microstruct. 23, 419–432 (1998). [CrossRef]  

129. T. C. Ralph, N. K. Langford, T. Bell, and A. G. White, “Linear optical controlled-NOT gate in the coincidence basis,” Phys. Rev. A 65, 062324 (2002). [CrossRef]  

130. J. Franson, M. Donegan, M. Fitch, B. Jacobs, and T. Pittman, “High-fidelity quantum logic operations using linear optical elements,” Phys. Rev. Lett. 89, 137901 (2002). [CrossRef]  

131. T. Pittman, M. Fitch, B. Jacobs, and J. Franson, “Experimental controlled-NOT logic gate for single photons in the coincidence basis,” Phys. Rev. A 68, 032316 (2003). [CrossRef]  

132. M. Fiorentino and F. Wong, “Deterministic controlled-NOT gate for single-photon two-qubit quantum logic,” Phys. Rev. Lett. 93, 070502 (2004). [CrossRef]  

133. N. Kiesel, C. Schmid, U. Weber, R. Ursin, and H. Weinfurter, “Linear optics controlled-phase gate made simple,” Phys. Rev. Lett. 95, 210505 (2005). [CrossRef]  

134. R. Okamoto, H. F. Hofmann, S. Takeuchi, and K. Sasaki, “Demonstration of an optical quantum controlled-NOT gate without path interference,” Phys. Rev. Lett. 95, 210506 (2005). [CrossRef]  

135. N. K. Langford, T. J. Weinhold, R. Prevedel, K. J. Resch, A. Gilchrist, J. L. O’Brien, G. J. Pryde, and A. G. White, “Demonstration of a simple entangling optical gate and its use in Bell-state analysis,” Phys. Rev. Lett. 95, 210504 (2005). [CrossRef]  

136. M. A. Pooley, D. J. P. Ellis, R. B. Patel, A. J. Bennett, K. H. A. Chan, I. Farrer, D. A. Ritchie, and A. J. Shields, “Controlled-NOT gate operating with single photons,” Appl. Phys. Lett. 100, 211103 (2012). [CrossRef]  

137. D. F. V. James, P. G. Kwiat, W. J. Munro, and A. G. White, “Measurement of qubits,” Phys. Rev. A 64, 052312 (2001). [CrossRef]  

138. A. G. White, A. Gilchrist, G. J. Pryde, J. L. O’Brien, M. J. Bremner, and N. K. Langford, “Measuring two-qubit gates,” J. Opt. Soc. Am. B 24, 172–183 (2007). [CrossRef]  

139. C. Arnold, J. Demory, V. Loo, A. Lemaître, I. Sagnes, M. Glazov, O. Krebs, P. Voisin, P. Senellart, and L. Lanco, “Macroscopic rotation of photon polarization induced by a single spin,” Nat. Commun. 6, 6236 (2015). [CrossRef]  

140. J. C. Loredo, N. A. Zakaria, N. Somaschi, C. Anton, L. De Santis, V. Giesz, T. Grange, M. A. Broome, O. Gazzano, G. Coppola, I. Sagnes, A. Lemaitre, A. Auffeves, P. Senellart, M. P. Almeida, and A. G. White, “Scalable performance in solid-state single-photon sources,” Optica 3, 433–440 (2016). [CrossRef]  

141. H. Wang, Z. C. Duan, Y. H. Li, S. Chen, J. P. Li, Y. M. He, M. C. Chen, Y. He, X. Ding, C.-Z. Peng, C. Schneider, M. Kamp, S. Höfling, C.-Y. Lu, and J.-W. Pan, “Near transform-limited single photons from an efficient solid-state quantum emitter,” Phys. Rev. Lett. 116, 213601 (2016). [CrossRef]  

142. J. B. Spring, B. J. Metcalf, P. C. Humphreys, W. S. Kolthammer, X.-M. Jin, M. Barbieri, A. Datta, N. Thomas-Peter, N. K. Langford, D. Kundys, J. C. Gates, B. J. Smith, P. G. R. Smith, and I. A. Walmsley, “Boson sampling on a photonic chip,” Science 339, 798–801 (2013). [CrossRef]  

143. M. A. Broome, A. Fedrizzi, S. Rahimi-Keshari, J. Dove, S. Aaronson, T. C. Ralph, and A. G. White, “Photonic boson sampling in a tunable circuit,” Science 339, 794–798 (2013). [CrossRef]  

144. A. M. Childs, D. Gosset, and Z. Webb, “Universal computation by multiparticle quantum walk,” Science 339, 791–794 (2013). [CrossRef]  

145. N. Spagnolo, C. Vitelli, M. Bentivegna, D. J. Brod, A. Crespi, F. Flamini, S. Giacomini, G. Milani, R. Ramponi, P. Mataloni, R. Osellame, E. F. Galvão, and F. Sciarrino, “Experimental validation of photonic boson sampling,” Nat. Photonics 8, 615–620 (2014). [CrossRef]  

146. E. Flagg, A. Muller, S. Polyakov, A. Ling, A. Migdall, and G. Solomon, “Interference of single photons from two separate semiconductor quantum dots,” Phys. Rev. Lett. 104, 1–4 (2010). [CrossRef]  

147. R. B. Patel, A. J. Bennett, I. Farrer, C. A. Nicoll, D. A. Ritchie, and A. J. Shields, “Two-photon interference of the emission from electrically tunable remote quantum dots,” Nat. Photonics 4, 632–635 (2010). [CrossRef]  

148. P. Gold, A. Thoma, S. Maier, S. Reitzenstein, C. Schneider, S. Höfling, and M. Kamp, “Two-photon interference from remote quantum dots with inhomogeneously broadened linewidths,” Phys. Rev. B 89, 035313 (2014). [CrossRef]  

149. V. Giesz, S. L. Portalupi, T. Grange, C. Anton, L. De Santis, J. Demory, N. Somaschi, I. Sagnes, A. Lemaître, L. Lanco, A. Auffèves, and P. Senellart, “Cavity-enhanced two-photon interference using remote quantum dot sources,” Phys. Rev. B 92, 1–5 (2015). [CrossRef]  

150. W. Gao, P. Fallahi, E. Togan, A. Delteil, Y. Chin, J. Miguel-Sanchez, and A. Imamoğlu, “Quantum teleportation from a propagating photon to a solid-state spin qubit,” Nat. Commun. 4, 2744 (2013).

151. R. Raussendorf and H. J. Briegel, “A one-way quantum computer,” Phys. Rev. Lett. 86, 5188 (2001). [CrossRef]  

152. N. H. Lindner and T. Rudolph, “Proposal for pulsed on-demand sources of photonic cluster state strings,” Phys. Rev. Lett. 103, 1–4 (2009). [CrossRef]  

153. S. E. Economou, N. Lindner, and T. Rudolph, “Optically generated 2-dimensional photonic cluster state from coupled quantum dots,” Phys. Rev. Lett. 105, 1–4 (2010). [CrossRef]  

154. A. V. Kuhlmann, J. H. Prechtel, J. Houel, A. Ludwig, D. Reuter, A. D. Wieck, and R. J. Warburton, “Transform-limited single photons from a single quantum dot,” Nat. Commun. 6, 8204 (2015). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (12)

Fig. 1.
Fig. 1. Single-photon characterization. (a) Example of a g HBT ( 2 ) ( τ ) autocorrelation function under a quasi-resonant pulsed excitation. (b) Two-time second-order correlation, g HBT ( 2 ) ( t 1 , t 2 ) , measured for an excitation in the wetting layer. Figure adapted from [26].
Fig. 2.
Fig. 2. Measurement of indistinguishability. (a) Schematic of a setup to measure the indistinguishability of two photons emitted from a quantum dot that is excited 2 ns apart. They are sent to an unbalanced Michelson interferometer and then to detectors. (b) Resulting second-order correlation histogram of the quantum dot light. The intensity of the peak at delay τ = 0 is proportional to the photon indistinguishability. Figures adapted from [16].
Fig. 3.
Fig. 3. Extracting quantum dot photons. (a) A quantum dot sandwiched between two asymmetric DBR mirrors allows for up to 10%–20% collection efficiency. (b) Brightness of 75% has been seen with a quantum dot inserted in the bottom of an inverted trumpet structure. Figure adapted from [53]. (c) Ultrabright sources ( p 1 79%) of single and indistinguishability photons have been made by coupling a quantum dot to a 3 μm micropillar cavity. Figure adapted from [54].
Fig. 4.
Fig. 4. Coupling of a quantum dot to a microcavity. (a) Atomic-force microscope topography of a photonic crystal nanocavity aligned to a quantum dot. The small hill in the middle arises from a quantum dot (63 nm below the surface). The color bar indicates the measured height. Figure adapted from [90]. (b) Schematic of an optical in situ lithography technique. Two lasers are used to find the quantum dot position and energy, and to define a cavity around it. Figure adapted from [91]. (c) Image of the photoluminescence signal of a quantum dot centered with a circular Bragg grating microcavity. The scale bar represents 5 μm. Figure adapted from [56].
Fig. 5.
Fig. 5. Brightness optimization. Fit of the experimentally measured ( 1 α ) = Q / Q 0 terms (black dashed line), calculated β = F P / ( F P + 1 ) (red dotted line), and the maximum theoretical extraction efficiency β × ( 1 α ) (solid green line) as a function of micropillar diameter. Figure adapted from [54].
Fig. 6.
Fig. 6. Current brightness results. (Top) Raw (open squares) and multiphoton corrected (solid squares) number of collected photons per laser pulse and the corresponding detected count rate per second. (Bottom) Values of g HBT ( 2 ) [ 0 ] values as a function of the pump power. The excitation laser is tuned to 860 nm to create carriers in the wetting layer. Figure adapted from [54].
Fig. 7.
Fig. 7. Progress toward high brightness and high indistinguishability single-photon sources. (a) Blue points: indistinguishability of successively emitted photons and extraction efficiency of some quantum-dot-based sources reported since the first single-photon demonstration in 2000 [15] (yellow point). In 2002, the indistinguishability of the photons was reported [16]. Corresponding references: 2002–2007: [101,102]; 2010–2013: [53,63]; 2013: [54]; 2013–2014: [28,30]; 2015: [31,46]. The shaded area indicates the use of resonant fluorescence excitations. The red star is for good SPDC sources with g HBT ( 2 ) 0.1 [7]. (b) Indistinguishability values as a function of the brightness for different pumping conditions. Green squares, wetting layer pumping; red triangles, quasi-resonant pumping; blue stars, two-color scheme (see text). The solid black line plots the normalized laser power P / P sat as a function of the source brightness. Figure adapted from [54].
Fig. 8.
Fig. 8. Almost unity indistinguishability; resonant fluorescence. (a) Schematic of a structure that emits highly indistinguishable photon under resonant excitation and electrical control. (b),(c) Second-order correlation histograms, g HOM ( 2 ) . The photon indistinguishability is determined by comparing the zero delay peak amplitude when the polarization in the two interferometer arms are parallel (b) and orthogonal (c) cases. In the parallel case, the zero delay peak should vanish for fully indistinguishable photons. Figures adapted from [31].
Fig. 9.
Fig. 9. (a). Energy-level scheme for resonant two-photon excitation of a quantum dot biexciton state, | X X . A pair of photons is emitted in cascade through the exciton state, | X , to the ground state, | 0 . The term s is the anisotropic exchange splitting of the | X level. (b) Quantum dot emission spectrum under resonant excitation of the | X X state. (c) Real part of the reconstructed density matrix. The imaginary part is plotted in [117]. (b),(c) Figures adapted from [117].
Fig. 10.
Fig. 10. Optical quantum controlled-not gate. Schematic of the gate that was used in [27]. The key element for the stability is the calcite crystal and the main element of the gate is the central half-wave plates (blue lines) [129]. Two half-wave plates (blue lines) at 45° off their optical axis in the input and output target ports act as Hadamard gates.
Fig. 11.
Fig. 11. Entangling controlled-not gate. (a) Measured fidelity of the generated state compared to the Bell state ϕ + as a function of the source postselected brightness. The dotted line indicates the quantum correlation threshold. (b) Calculated fidelity as a function of the photons indistinguishability. The square indicates a fully distinguishable single-photon source; the circle (triangle) corresponds to the two experimental points without (with) temporal postselection. Figures adapted from [27].
Fig. 12.
Fig. 12. Controlled-not gate. When a photon is sent to a coupled quantum dot cavity system, the phase of the reflected photon depends on the quantum dot state. If the quantum dot is in the ground state | g and the cooperativity, C 1 , no phase is added, r = 1 and the incoming state in unchanged (top). However, if the quantum dot is in the excited state | , r = 1 and the incident state is flipped (bottom). Figure adapted from [118].

Equations (12)

Equations on this page are rendered with MathJax. Learn more.

γ = 1 T 2 = 1 2 T 1 + 1 T 2 * .
p 1 = p s × η ,
η = β × ( 1 α ) with β = Γ Γ + Γ other .
Γ = 2 π 2 | d⃗ · E⃗ ^ ( r⃗ QD ) | 2 × ρ ( ω QD ) ,
Γ Γ 0 F p = 3 Q ( λ c / n eff ) 3 4 π 2 V eff ,
β = F P F P + 1 .
( 1 α ) = Q Q 0 .
p 1 = p 1 max ( 1 e P / P sat ) ,
1 2 ( | H XX H X + | V XX V X ) .
Φ = 1 2 ( | Early XX | Early X + e i ϕ p | Late XX | Late X ) ,
Φ + = 1 2 ( | H H + | V V ) .
F Φ + = 1 + C 2 ( 2 C ) .
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.