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

Polarization-entangled photon-pair source obtained via type-II non-collinear SPDC process with PPKTP crystal

Open Access Open Access

Abstract

We demonstrate a polarization-entangled photon-pair source obtained via a type-II non-collinear quasi-phase-matched spontaneous parametric down-conversion process with a 10-mm periodically poled KTiOPO4 crystal, which is as stable and wavelength-tunable as the well-known Sagnac configuration scheme. A brightness of 4.2 kHz/mW is detected and a concurrence of 0.975 is estimated using quantum state tomography. Without loss of entanglement and brightness, the photon-pair wavelengths are tunable through control of the crystal temperature. This improvement is achieved using the non-collinear configuration and a stable interferometric distinguishability compensator.

© 2016 Optical Society of America

1. Introduction

Many significant experiments concerning topics related to quantum information processing, such as quantum teleportation [1], dense coding [2], quantum key distribution [3], and controlled-not gate [4], have been conducted using polarization-entangled photon pairs generated via spontaneous parametric down-conversion (SPDC) processes with β-barium borate (BBO) crystals [5–8]. However, these sources have relatively low (≤ 140 Hz/mW) brightness (BR; coincidence counts per second per pump power, without loss-correction) because of inefficient collection of the generated photons and small nonlinear coefficient. To date, considerable efforts have been made towards achieving greater brightness and higher entanglement quality in such cases. In particular, in the early 2000s, nonlinear crystals with periodically poled (PP) structures and quasi-phase-matching (QPM) conditions yielded remarkably increased brightness [9–11] without polarization entanglement. To realize polarization-entangled photon-pair sources with QPM crystals, schemes involving post-selection [12], crossed crystals [13–15], a folded Mach-Zehnder interferometer (MZI) [16, 17], a polarization Sagnac interferometer (PSI) [18–20], two crystals Sagnac interferometer [21], and linear double-pass geometry [22] have been demonstrated, which are applied under collinear conditions. In particular, two outstanding schemes have been reported by Fedrizzi et al. [20] and Steinlechner et al. [22], who detected 271- and 390-kHz/mW/nm spectral brightness (SB; BR per bandwidth) and more than 99% fidelity with one of the Bell states using type-II and type-0 SPDC processes with a PP potassium titanyl phosphate (KTiOPO4; PPKTP) crystal, respectively, with stable configurations. However, the schemes reported in those studies involve an intricate alignment or specialized optics such as a dual-wave plate, a specifically tailored wave plate, and a temperature-stabilized compensation crystal.

Here, we demonstrate a polarization-entangled photon-pair generation scheme [23] based on a type-II non-collinear PPKTP crystal and employing a continuous-wave (cw) pump laser [16] and an MZI-configured distinguishability compensator [8]. Our source is stable and relatively easy to build up because of the simple configuration and lack of specialized optics. The pump-power Pp dependence of the source is examined in order to estimate the BR and pair-production-rate (loss-corrected BR), and the crystal-temperature T dependence is monitored in order to demonstrate wavelength (λ) tunability. We also conduct quantum state tomography (QST) [24] to evaluate the entanglement quality of the generated states. The detected BR, λ tunability, and concurrence (one of proper measures of the degree of entanglement [25]) are 4.2 kHz/mW, 0.198 nm/°C, and 0.975, respectively. We expect that enhanced BR, SB and a greater degree of entanglement can be achieved if optimized focusing parameters, a longer crystal, and a shorter coincidence window are employed.

The remainder of this paper is organized as follows. In section 2, we introduce and explain the experiment setup of our polarization-entanglement photon-pair source and measurement apparatus using a schematic diagram. In section 3, we briefly describe the QPM condition of the type-II non-collinear SPDC process and the theoretical foundation behind the expected pair-production-rate through comparison with a type-II collinear PSI setup. We also provide the results of numerical calculations, which are used as a reference for comparison against the experimental results presented in section 4. From those experimental results, we estimate and discuss the BR, pair-production-rate, λ tunability, entanglement quality, and stability of our source. Finally, a summary is presented in section 5.

2. Scheme and methods

Figure 1 is a schematic diagram of our experimental setup for the polarization-entangled photon-pair source. A single-longitudinal-mode cw laser is employed as a pump, having a center wavelength λp and full-width at half-maximum (FWHM) bandwidth Δν measured at approximately 406.2 nm and 0.2 GHz, respectively; these measurements were obtained using a spectrometer (Ocean Optics, USB4000; Δλ ≃ 1 nm) and a scanning Fabry-Perot interferometer (Thorlabs, SA210; Δν = 67 MHz), respectively. The pump beam is spatial-mode-filtered by a single-mode fiber (SMF) for 405 nm and its polarization and Pp are adjusted by wave plates and a polarizing beam-splitter (PBS). The well-defined pump beam is focused by a lens (focal length f = 200 mm, ARC-A) and passes through the PPKTP crystal (Λ = 10 μm, L = 10 mm) in the Rayleigh range. Under the type-II non-collinear SPDC condition, a pump-light photon is probabilistically converted to a pair of photons with horizontal (y-axis; H) and vertical (z-axis; V) polarization and spatial or directional modes determined by the phase-matching (PM) condition. In our setup, we select two spatial modes (a and b) that are tilted at approximately 0.85° relative to the x-axis, cento-symmetric to each other, and along the xy-plane. The two spatial modes are collimated by a lens (f = 200 mm, ARC-B). Each photon (H or V polarized) of a pair can be emitted in either of the two modes (a or b), but the counterpart photon must then be emitted in the other direction (i.e., b or a, respectively), in accordance with the law of conservation of momentum for near-degenerate cases (λHλV, where λH and λV are the center wavelengths of the H and V polarized photons, respectively). Therefore, if all other degrees of freedom are indistinguishable, the polarizations of a photon pair in modes a and b are entangled as |HVab + |VHab. However, the wavelengths can be distinguishable if the difference between λH and λV in a given mode is significantly larger than the wavelength bandwidth Δλ of each photons. Furthermore, in general, the refractive indices of the H and V polarizations in the crystal differ, so photons of two polarizations in a given mode are temporally distinguishable because of the group velocity difference; this holds even for the degenerate case. Thus, the generated polarization state is entangled with other degrees of freedom as |H(λH, tH)〉a|V(λV, tV)〉b + |V(λV, tV)〉a|H(λH, tH)〉b. To compensate for this distinguishability, we adopt an additional setup [8, 16, 23] which consists of a zero-order half-wave plate (HWP) at 45°, an optical delay line (ODL), and a broadband PBS. After proper adjustment of the ODL, the polarization states of a photon-pair in modes A and B of Fig. 1 are entangled as

|ψAB=|HHAB+eiϕ|VVAB2|λAλBAB,
where λA and λB are identical to λH and λV in modes a or b, respectively. In Eq. (1), the entire state of the wavelengths for modes A and B is separable from the polarization state, and the time information depending on the polarizations is negligible. We neglect the entanglement between the λ for modes A and B in Eq. (1), as the Δλ of each photon is sufficiently narrow for non-degenerate conditions [20, 26] and the spectral correlation can be erased using narrow interference filter (IF) or engineered by a group-velocity matching condition with a broadband pump [27–30].

 figure: Fig. 1

Fig. 1 Schematic diagram of experimental setup. The pump is a single-longitudinal-mode cw laser with a center wavelength λp of ∼406.2 nm and a full-width at half-maximum (FWHM) bandwidth of Δν = 0.2(1) GHz. SMF: Single-mode fiber for 405 or 780 nm; QWP (HWP): Zero-order quarter (half)-wave plate for 405 or 808 nm; PBS: Broadband polarizing beam-splitter; Lens: Anti-reflection (AR)-coated plano-convex lens (focal length f = 200 mm, coating designation: A or B); PPKTP: periodically poled KTiOPO4 (Λ = 10 μm; length L = 10 mm); S: temporary image screen to measure photon rings; ODL: Optical delay line consisting of right-angled prism and x-axis translation stage; ϕ: Phase shifter consisting of sequential wave plates (QWP-HWP-QWP); LP: Broadband linear polarizer; TF: Temporally used wavelength tunable filter (FWHM of Δλ = 1.82 nm); SPCM: Single-photon counting module; FPGA: Single and coincidence counting unit.

Download Full Size | PDF

A relative phase ϕ between the two states in Eq. (1) can be introduced through unexpected birefringence of the optics in the interferometric compensator; therefore, we insert a phase shifter consisting of sequential zero-order wave plates (QWP-HWP-QWP; quarter-wave plate = QWP) to adjust ϕ. The fast axes of the QWP are fixed at 45° from the horizontal direction, and we adjust the HWP angle to obtain ϕ = 0.

To measure the polarizations of the generated state, we introduce polarization analyzers consisting of a QWP and a broadband linear polarizer (LP). After passing the analyzers, the photons are coupled into SMFs for 780 nm via aspheric lenses (fe = 8 mm) and detected using single-photon counting modules (SPCM). Electric pulses from the SPCM are counted by a field-programmable gate array (FPGA) system calibrated for single and coincidence counts (coincidence window ≃ 55 ns). We also measure the photon λ values in modes A and B using a wavelength tunable filter (TF, FWHM = 1.82 nm). In Fig. 1, the temporary image screen labeled S consists of an LP, an IF (three-cavity bandpass filter at approximately 812 nm, FWHM = 3 nm), an aspheric lens (fe = 8 mm), and a multi-mode fiber (MMF), which are mounted on a yz-axis translation stage. By counting the number of photons coupled in the MMF during stage movement, we can record 1D or 2D photon ring images.

3. Theory and numerical calculations

A photon pair generated in a type-II non-collinear SPDC process with a PPKTP crystal adheres to the QPM conditions [31, 32], such as: (i) energy conservation: ωp = ωH + ωV ; and (ii) momentum conservation: k⃗p = k⃗H + k⃗V + K⃗, where ω and k⃗ are the angular frequency and wavevector of the photons, respectively, the p subscript denotes the pump, and K⃗ is the grating vector of the periodic structure of the PPKTP crystal. If Λ is the period of the poled pattern, K is defined as 2π/Λ. The coordinate system depicted in Fig. 1 corresponds to the principal axes of the PPKTP crystal, and k⃗p and K⃗ are parallel to the x-axis. In the course of the type-II SPDC process, the nonlinear coefficient d24 is utilized, which means that the pump is H polarized (in the y-direction) and the two generated photons are H and V polarized (in the y- and z-directions, respectively). See [33] for a more detailed description of the QPM condition of PPKTP.

In the collinear case, QPM condition (ii) is reduced to kp = kH + kV + K, and the refractive indices of the H and V photons are functions of the wavelength and T of the crystal. KTP is a biaxial crystal and its refractive indices for principal axes depending on wavelength and T have been well studied [34–38]. Here, we adopted the T-dependent Sellmeier formulas and thermal expansion coefficients for KTP from [16, 37, 38]. Using numerical calculations with given parameters, such as λp = 406.2 nm and Λ = 10 μm, we obtained the T value of the degenerate collinear case as Tdc = 98.98°C. We should note that numerically calculated values of Tdc differ depending on the references used for the Sellmeier formulas, as these expressions are empirical relationships obtained from individual finite experimental datasets. Therefore, the following numerical results for the non-collinear conditions are not ideal references for comparison against the experimental results. Instead, they provide a rough outline only.

Under general non-collinear SPDC in biaxial crystals, accurate prediction of PM conditions in the entire space is sophisticated [39]. However, it is much easier to calculate the PM condition only on the two orthogonal principal planes (xy- and xz-planes), as the limiting cases. Furthermore, for a near-collinear case where the angle between the wave-vectors of the pump and the SPDC is small (3.0° or smaller as in our experiment), one can neglect the Poynting vector walk-off from the wave-vector (0.3° or less), significantly simplifying the calculation process.

Under near-collinear conditions, the ring images of the H and V photons are overlapped and indistinguishable in the degenerate case, and the effective nonlinear coefficient is also almost identical to that of the collinear case (deff = 2d24/π in the ideal case). At this point, we note that photon pairs are accumulated in two spatial modes using the non-collinear setup, so that twice the number of photon pairs can be collected compared with the collinear setup. This means that our scheme, which is based on non-collinear conditions, can yield the twice brightness of the collinear configuration in an ideal experimental setup, i.e., with the same L and optimized spatial mode parameters for pump and generated photons.

Figure 2 shows the approximated (assuming near-collinearity) numerical results for: (a) The emission angle (θ) values outside the crystal exit plane of the degenerate (λH = λV) SPDC cone as a function of T for the photon pairs emitted on the xy- and xz-planes; (b) ring (or cone) images on the yz-plane at several fixed T values for the degenerate case; and (c) the θ values on the xy-plane according to the wavelengths of the H and V photons at a fixed T = 95°C. As shown in Fig. 2(a), the θ values of the degenerate photon pairs on the xy- and xz-planes for fixed T differ slightly, because of the difference between the principal indices ny (∼ nx) and nz [40]. Therefore, the ring images of the photon pairs in Fig. 2(b) are slightly elliptic (e ≃ 0.22) [16]. The specific value of T = 95°C for Fig. 2(c) was chosen to satisfy the fact that the θ value for the degenerate case is approximately 0.85°, which is the real experimental parameter of our setup. At the spatial modes (θ ≃ 0.85, the degenerate condition for T = 95°), the derivative of the emission angle against the wavelength dθ/dλ is approximately 0.45 °/nm. This is a characteristic of the non-collinear setup that contrasts with the collinear configuration, as dθ/dλ for the degenerate collinear condition (θ = 0) is approximately infinite.

 figure: Fig. 2

Fig. 2 Numerically calculated emission angles θ for photons emitted from PPKTP crystal exit plane for: (a) Degenerate case on xy- and xz-planes as a function of T; (b) degenerate case for several fixed T values on yz-plane; and (c) fixed T = 95°C on xy-plane as a function of photon wavelength.

Download Full Size | PDF

Figure 3 shows the numerical results for the (a) λH and λV and (b) spectrum FWHM Δλ of the H and V photons as a function of T, where the spatial modes were fixed to the degenerate condition in Fig. 2(c). In fact, for non-degenerate cases (λHλV), the θ values of the H and V photons differ slightly, in accordance with the law of conservation of momentum. However, this difference is negligible, as ΔθHVλHV ≃ 18 μrad/nm, where ΔθHV and ΔλHV are the differences in the emission angles and wavelengths between the H and V photons of a given pair, respectively. λ tunability via control of T is shown by the slopes of the trendlines in Fig. 3(a), which were calculated as 0.23 nm/°C. The numerical results in Fig. 3(b) demonstrate that the photon Δλ decrease slightly as T increases, and the degenerate case (λH = λV = 812.4 nm) is given by 0.553 nm, which is almost the same result as that obtained for the collinear case [20].

 figure: Fig. 3

Fig. 3 T dependence of (a) λH and λV and (b) FWHM Δλ of H and V photons of pair generated under non-collinear condition, with spatial modes determined by degenerate condition at T = 95°C.

Download Full Size | PDF

4. Experimental results and discussions

4.1. Photon ring images

Firstly, to mark the center of the ring images or to determine the position of the collinear condition, we measured the photon ring images at T = 68.6°C, as shown in Fig. 4, using the S shown in Fig. 1. To show the ring images clearly, we decreased T to a value significantly lower than the calculated Tdc = 98.98°C, as the IF Δλ (FWHM = 3 nm) was broader than the numerically calculated photon Δλ (< 0.6 nm) in 10-mm PPKTP. Figures 4(a) and 4(b) show ring images for H and V polarized photons, respectively. The measured ring image in Fig. 4(a) is slightly larger than that of Fig. 4(b), because the IF center wavelength λIF may not have been identical to the degenerate wavelength (2λp). By means of data fitting, the ring-image ellipticities were estimated to be 0.25(1).

 figure: Fig. 4

Fig. 4 Measured photon ring images for (a) H and (b) V polarized photons with IF (IF center wavelength λIF ≃ 812 nm, FWHM = 3 nm).

Download Full Size | PDF

Figure 5 shows experimentally observed cross-section images of photon rings along the y-axis for T varied from 48.6 to 98.6°C. Because of the broad Δλ of the IF and the discrepancy that 2λpλIF, we could not accurately determine the experimental value of Tdc by means of the cross-section results shown in Fig. 5. However, we roughly estimated that its value was approximately 89(5)°C. It is possible to obtain an accurate value of T for the degenerate conditions of the collinear and non-collinear cases by measuring the spectra of H and V photons using the TF. The spectral results for a non-collinear case are discussed below. In Fig. 5, the maximum count rates of the H photons are larger than those of the V photons for T = 98.6 and 93.6°C, which are above the experimental Tdc value of 89(5)°C, and the angles of the peak counts for the H photons are larger than those of the V photons at values lower than the Tdc. From these results, we inferred that λIF was slightly longer than the degenerate wavelength, i.e., λIF − 2λp ≃ 0.07 nm. Further, we noted that the peak counts of the H and V photons are reduced when T decreases, as the near-collinear approximation may not work for lower T, and deff is decreased. Therefore, we chose the SPDC cone angle approximately 0.85° with respect to the pump beam in order to maintain a similar counting rate in each spatial mode and to collect approximately twice the number of photon pairs compared to the collinear case.

 figure: Fig. 5

Fig. 5 Measured cross-section (1D) images of photon rings along y-axis for various T (48.6 – 98.6 °C) with IF (FWHM = 3 nm).

Download Full Size | PDF

4.2. Temperature dependence of photon-spectrum

Figure 6 shows (a) an example of measured spectral distributions for photons in modes A and B, which were obtained using the TF at T = 42°C, and (b) λH,V and Δλ (represented by error bars) estimated from measured photon spectra at T = 42–122°C. The error bars in Fig. 6(b) represent the FWHMs of the Gaussian fitted curves of the data shown in Fig. 6(a). From the linear fitted lines of the estimated center wavelengths in Fig. 6(b), the experimental wavelength tunability via control of T was estimated as 0.198(1) nm/°C, and Tdc was obtained as 81.7°. The average value of the estimated Δλ (FWHM) of the photons in modes A and B was 2.11(3) nm. This is significantly larger than the numerically calculated values (< 0.6 nm), because of the broad TF bandwidth (Δλ = 1.82 nm). For the reason, we could not measure the photon Δλ variation in response to the crystal T directly, as shown in Fig. 3(b).

 figure: Fig. 6

Fig. 6 (a) Measured spectral distributions for photons in modes A and B obtained via TF at T = 42°C. (b) Center wavelength T dependence for photons in modes A and B. FWHM Δλ are represented by error bars. o: measured or estimated data. −: Gaussian or linear fitted lines.

Download Full Size | PDF

4.3. Pump power dependence of single and coincidence counts

We measured the polarization-dependent single and coincidence count rates for modes A and B in Fig. 1 with Pp of 0.5–4 mW at T = 82°C around the degenerate condition. As shown in Fig. 7, the single and coincidence count rates were linearly proportional to Pp. The estimated single count rates for each mode were approximately 31–35 kHz/mW (except the result of 42.7 kHz/mW for VB), whereas the coincidence count rates for modes HAHB and VAVB were 2.0 and 2.2 kHz/mW, respectively. The coincidence count rates for the unexpected modes HAVB and VAHB were approximately 36 Hz/mW, because of the finite extinction ratio of the PBS. The detected BR of 4.2 kHz/mW mainly from modes HAHB and VAVB include several losses due to the finite transmittance (≃ 0.8) of the LP (Thorlabs LPVIS050), the detection efficiency (≃ 0.4 at around 810 nm) of the SPCM (PerkinElmer SPCM-AQ4C), and mode coupling inefficiency to the SMF, for example. Therefore, the total pair-production-rate of our source under ideal conditions is approximately 41 kHz/mW at minimum. Using the numerical result of the photon bandwidth ΔλH,V = 0.553 nm, the estimated spectral-pair-production-rate (pair-production-rate per bandwidth) is approximately 74 kHz/mW/nm. If we use a longer crystal with L = 25 mm and scaling factor LL [20], the expected spectral-pair-production-rate is approximately 293 kHz/mW/nm. We note and expect that greater count rates can be achieved if spatial-mode optimization for the pump and photons [20] is implemented.

 figure: Fig. 7

Fig. 7 Pump-power Pp dependence of count rates for (a) single counts of modes HA, VA, HB and VB, and (b–e) coincidence counts of modes HAHB, HAVB, VAHB and VAVB. Raw: Raw data of coincidence counts. Net: Accidental counts removed data. Fit: Linear fitted line of Net.

Download Full Size | PDF

Very recently, Jeong et al. reported a bright source of polarization-entangled photons via SPDC process of type-II non-collinear PPKTP crystal and broadband multi-mode cw laser [41]. The configurations of ours and [41] are the same except longitudinal modes of the pump. Jeong et al. showed that detected BR using MMF collection are approximately 13 times larger than that of SMF at the cost of a little amount of concurrence. Thus, we expect that approximately one order of magnitude improvement in count rates can be achieved by MMF collection, without longer crystal and optimization of focusing parameters.

4.4. Entanglement quality

We estimated the entanglement quality of the generated states using several indicators: (i) The visibility of the Hong-Ou-Mandel (HOM) interferences; (ii) the correlation function visibilities; (iii) the fidelity with the ideal state (|Φ+〉) in Eq. (1) for ϕ = 0; and (iv) the concurrence. We first investigated the Pp dependence of the entanglement quality. To ensure fairness, we attempted to equalize the total number of resources by adjusting the counting duration as follows: 1, 2, 4, and 8 s for 4(3), 2, 1, and 0.5 mW, respectively. The measured quantities of the entanglement indicators for various Pp are summarized in Table 1, apart from the correlation function visibilities, which were approximately 0.92(1)–0.99(1). As an example, Fig. 8 shows measured results for: (a) the HOM dip and peak interferences; (b) the polarization correlation functions; and (c, d) the reconstructed density matrix obtained via QST at Pp = 0.5 mW.

Tables Icon

Table 1. Estimated entanglement qualities for various Pp

 figure: Fig. 8

Fig. 8 Experimental results at Pp = 0.5 mW for: (a) HOM dip (blue) and peak (red) interferences; (b) correlation functions of polarizations for four cases; and (c) real and (d) imaginary parts of reconstructed density matrix obtained via QST. For the correlation function measurements in (b), the LP from mode A (LPA) was fixed to 0° (red); 45° (green); 90° (blue); and 135° (black). Dotted lines: Sine-fitted curves.

Download Full Size | PDF

For the HOM interference and polarization correlation measurements, the QWPs for modes A and B in Fig. 1 were fixed to 0° and 90°, respectively, so as to compensate for the phases introduced by the QWPs themselves. The HOM peak (dip) interference was measured using the coincidence counts between modes A and B by sweeping the position of the right-angled prism (ODL) in Fig. 1. The LPs for modes A and B (LPA and LPB, respectively) were set to 45° and (−)45°, respectively. We also measured the correlation functions of the polarizations in modes A and B for four LPA values of 0°, 45°, 90°, and 135°, the results of which were fit using raised sine functions. The visibilities obtained from the fitting results were 0.994, 0.982, 0.975 and 0.930, respectively. To calculate the fidelities with the ideal state and concurrences, we performed 2-qubit QST and reconstructed the density matrices. As shown in Table 1, when Pp increases the entanglement qualities decrease, as the multi-photon effects are increased and the coincidence window is relatively large. We expect that higher visibility and concurrence (> 0.97) can be measured with a shorter coincidence window at lower Pp.

To investigate the entanglement qualities for non-degenerate cases, we also measured the quality indicators for varying T. Figure 9 shows that HOM dip (in blue) and peak (in red) interferences at T = 62–102°C, Pp = 0.5 mW, and t = 8 s. The plotted points are net coincidence counts for modes A and B. All the visibilities of the interference patterns in Fig. 9 are larger than 0.97. The fidelities and concurrences estimated from the QST results are summarized in Table 2. The entanglement qualities for non-degenerate cases (|λAλB| < 16 nm) are almost identical in our setup, as we used zero-order wave plates for 808 nm and broadband PBS and LP to reduce wavelength-dependent systematic errors.

 figure: Fig. 9

Fig. 9 Measured HOM dip and peak interferences for T = 62–102°C at Pp = 0.5 mW and t = 8 s. Peak (dip) data in red (blue) are measured based on coincidence counts of modes A and B with LPA and LPB angles set to 45° and (−)45°, respectively. To show the beat patterns clearly, lines are drawn between neighboring points.

Download Full Size | PDF

Tables Icon

Table 2. Entanglement qualities of generated states for various T

4.5. Phase modulation and stability

The ϕ between |HHAB and |VVAB in Eq. (1) can be modulated using the phase shifter shown in Fig. 1. To check the phase modulability, we measured the HOM dip interferences for the cases of ϕ = /2, where n = 0, 1, 2 and 3. Figure 10 shows the normalized HOM dip (LPA = 45° and LPB = −45°) beating patterns at Pp = 0.5 mW, t = 8 s, and T = 77°C. The datapoints colored blue, green, red, and black correspond to n = 0, 1, 2, and 3, respectively. The total duration of the data measurement for the results shown in Fig. 10 was more than 1 h.

 figure: Fig. 10

Fig. 10 HOM dip interferences of generated state, |λAλBAB(|HHAB+eiϕ|VVAB)/2, where ϕ = /2 and λAλB. The datapoints colored in blue, green, red, and black correspond to n = 0, 1, 2, and 3, respectively.

Download Full Size | PDF

The experimental configuration, especially the compensation section, employed in the approach presented in this study is based on MZI. Thus, the phase stability is fundamentally limited, because of the vibrations of the optic mounts. However, the Sagnac configuration is robust against this effect. If we denote the two optical path lengths from the PPKTP crystal to the PBS in modes a and b in Fig. 1 as la and lb, respectively, the ϕ difference between the superposed two states in Eq. (1) induced by the difference in optical path length is (lalb)(kAkB), where kA(B) = 2π/λA(B). Therefore, the total phase fluctuation Δϕ due to the optics vibration Δl in the MZI can be expressed as m|δk|Δl, where δk = kAkB and m is the number of independent optics in the MZI. The order of Δl is typically sub-micrometer in the case on a vibration-isolated optical table. For example, for parameters of δλ = λBλA ≃ 50 nm, λp = 406.2 nm, m = 5, and Δl = 0.1 μm, the expected value of Δϕ is approximately 0.017 × 2π (< 2%). Thus, the phase fluctuations are negligible (< 0.4%) in our experimental results (δλ < 10 nm). Moreover, for the degenerate case (λA = λB), Δϕ can be fully neglected (as in the Sagnac configuration), because δk = 0. This may be comprehensive in terms of quantum mechanics. The two superposed states |HVab and |VHab of a photon pair immediately after generation via SPDC are simultaneously influenced by the phase fluctuations of the optics in the MZI; thus, this phase can be factored out as an overall phase.

5. Summary

We have demonstrated a stable and bright polarization-entangled photon-pair source obtained via a type-II non-collinear SPDC process with a 10-mm-long PPKTP crystal. The detected brightness was 4.2 kHz/mW, and the spectral-pair-production-rate was estimated as being 74 kHz/mW/nm. Further, we expect that a spectral-pair-production-rate of more than 293 kHz/mW/nm can be achieved via optimization of the focusing conditions and through use of a 25-mm crystal, or photon collection using MMF as previous studies [20,41]. This expectation is based on the fact that the non-collinear setup can collect twice the number of photon pairs as a collinear setup. Our setup is as stable as the Sagnac configuration for the degenerate case. Although this stability decreases with increased wavelength non-degeneracy, the instability is acceptable if the non-degeneracy is not excessive compared with the MZI compensator instability. For various photon-pair wavelengths (λA,B: 808–816 nm) obtained by controlling the crystal temperature (62–102°), the fidelities with one of the Bell states (|Φ+〉) and the concurrences of the generated states were estimated as being approximately 0.97–0.98, despite the large coincidence window (55 ns). Therefore, our source is bright, wavelength-tunable, and relatively stable (comparable to the Sagnac configuration setup) with no requirement for specialized optical components. Further, this source has high entanglement quality; thus, we expect that it will be utilized in various experiments concerning quantum information processing.

Acknowledgments

This work was supported by the National Research Foundation of Korea (NRF) grant funded by the Korea government (MSIP) (No. 2015R1A2A1A05001819, No. 2014R1A1A2054719, and No. 2014R1A1A2055488), and by the Measurement Research Center (MRC) Program for Korea Research Institute of Standards and Science.

References and links

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. K. Mattle, H. Weinfurter, P. G. Kwiat, and A. Zeilinger, “Dense coding in experimental quantum communication,” Phys. Rev. Lett. 76, 4656–4659 (1996). [CrossRef]   [PubMed]  

3. A. Poppe, A. Fedrizzi, R. Ursin, H. Böhm, T. Lörunser, O. Maurhardt, M. Peev, M. Suda, C. Kurtsiefer, H. Weinfurter, T. Jennewein, and A. Zeilinger, “Practical quantum key distribution with polarization entangled photons,” Opt. Express 12, 3865–3871 (2004). [CrossRef]   [PubMed]  

4. S. Gasparoni, J.-W. Pan, P. Walther, T. Rudolph, and A. Zeilinger, “Realization of a photonic controlled-NOT gate sufficient for quantum computation,” Phys. Rev. Lett. 93, 020504 (2004). [CrossRef]   [PubMed]  

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

6. P. G. Kwiat, E. Waks, A. G. White, I. Appelbaum, and P. H. Eberhard, “Ultrabright source of polarization-entangled photons,” Phys. Rev. A 60, R773–R776 (1999). [CrossRef]  

7. G. Giorgi, G. D. Nepi, P. Mataloni, and F. D. Martini, “A high brightness parametric source of entangled photon states,” Laser Phys. 13, 350–354 (2003).

8. Y.-H. Kim, S. P. Kulik, M. V. Chekhova, W. P. Grice, and Y. Shih, “Experimental entanglement concentration and universal bell-state synthesizer,” Phys. Rev. A 67, 010301 (2003). [CrossRef]  

9. S. Tanzilli, H. de Riedmatten, H. Tittel, H. Zbinden, P. Baldi, M. De Micheli, D. Ostrowsky, and N. Gisin, “Highly efficient photon-pair source using periodically poled lithium niobate waveguide,” Electron. Lett. 37, 26–28 (2001). [CrossRef]  

10. K. Banaszek, A. B. U’Ren, and I. A. Walmsley, “Generation of correlated photons in controlled spatial modes by downconversion in nonlinear waveguides,” Opt. Lett. 26, 1367–1369 (2001). [CrossRef]  

11. K. Sanaka, K. Kawahara, and T. Kuga, “New high-efficiency source of photon pairs for engineering quantum entanglement,” Phys. Rev. Lett. 86, 5620–5623 (2001). [CrossRef]   [PubMed]  

12. C. E. Kuklewicz, M. Fiorentino, G. Messin, F. N. C. Wong, and J. H. Shapiro, “High-flux source of polarization-entangled photons from a periodically poled KTiOPO4 parametric down-converter,” Phys. Rev. A 69, 013807 (2004). [CrossRef]  

13. M. Pelton, P. Marsden, D. Ljunggren, M. Tengner, A. Karlsson, A. Fragemann, C. Canalias, and F. Laurell, “Bright, single-spatial-mode source of frequency non-degenerate, polarization-entangled photon pairs using periodically poled KTP,” Opt. Express 12, 3573–3580 (2004). [CrossRef]   [PubMed]  

14. H. Hübel, M. R. Vanner, T. Lederer, B. Blauensteiner, T. Lorünser, A. Poppe, and A. Zeilinger, “High-fidelity transmission of polarization encoded qubits from an entangled source over 100 km of fiber,” Opt. Express 15, 7853–7862 (2007). [CrossRef]   [PubMed]  

15. F. Steinlechner, P. Trojek, M. Jofre, H. Weier, D. Perez, T. Jennewein, R. Ursin, J. Rarity, M. W. Mitchell, J. P. Torres, H. Weinfurter, and V. Pruneri, “A high-brightness source of polarization-entangled photons optimized for applications in free space,” Opt. Express 20, 9640–9649 (2012). [CrossRef]   [PubMed]  

16. M. Fiorentino, C. E. Kuklewicz, and F. N. C. Wong, “Source of polarization entanglement in a single periodically poled KTiOPO4 crystal with overlapping emission cones,” Opt. Express 13, 127–135 (2005). [CrossRef]   [PubMed]  

17. F. König, E. J. Mason, F. N. C. Wong, and M. A. Albota, “Efficient and spectrally bright source of polarization-entangled photons,” Phys. Rev. A 71, 033805 (2005). [CrossRef]  

18. T. Kim, M. Fiorentino, and F. N. C. Wong, “Phase-stable source of polarization-entangled photons using a polarization Sagnac interferometer,” Phys. Rev. A 73, 012316 (2006). [CrossRef]  

19. F. Wong, J. Shapiro, and T. Kim, “Efficient generation of polarization-entangled photons in a nonlinear crystal,” Laser Phys. 16, 1517–1524 (2006). [CrossRef]  

20. A. Fedrizzi, T. Herbst, A. Poppe, T. Jennewein, and A. Zeilinger, “A wavelength-tunable fiber-coupled source of narrowband entangled photons,” Opt. Express 15, 15377–15386 (2007). [CrossRef]   [PubMed]  

21. T. E. Stuart, J. A. Slater, F. Bussières, and W. Tittel, “Flexible source of nondegenerate entangled photons based on a two-crystal Sagnac interferometer,” Phys. Rev. A 88, 012301 (2013). [CrossRef]  

22. F. Steinlechner, S. Ramelow, M. Jofre, M. Gilaberte, T. Jennewein, J. P. Torres, M. W. Mitchell, and V. Pruneri, “Phase-stable source of polarization-entangled photons in a linear double-pass configuration,” Opt. Express 21, 11943–11951 (2013). [CrossRef]   [PubMed]  

23. Y.-C. Jeong and Y.-H. Kim, “Ultra-bright bell state synthesizer,” in Optical Society of Korea Winter Annual Meeting (OSA, 2014), pp. 19–21.

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

25. W. K. Wootters, “Entanglement of formation of an arbitrary state of two qubits,” Phys. Rev. Lett. 80, 2245–2248 (1998). [CrossRef]  

26. S. Ramelow, L. Ratschbacher, A. Fedrizzi, N. K. Langford, and A. Zeilinger, “Discrete tunable color entanglement,” Phys. Rev. Lett. 103, 253601 (2009). [CrossRef]  

27. W. P. Grice and I. A. Walmsley, “Spectral information and distinguishability in type-II down-conversion with a broadband pump,” Phys. Rev. A 56, 1627–1634 (1997). [CrossRef]  

28. T. E. Keller and M. H. Rubin, “Theory of two-photon entanglement for spontaneous parametric down-conversion driven by a narrow pump pulse,” Phys. Rev. A 56, 1534–1541 (1997). [CrossRef]  

29. P. J. Mosley, J. S. Lundeen, B. J. Smith, P. Wasylczyk, A. B. U’Ren, C. Silberhorn, and I. A. Walmsley, “Heralded generation of ultrafast single photons in pure quantum states,” Phys. Rev. Lett. 100, 133601 (2008). [CrossRef]   [PubMed]  

30. R.-B. Jin, R. Shimizu, K. Wakui, M. Fujiwara, T. Yamashita, S. Miki, H. Terai, Z. Wang, and M. Sasaki, “Pulsed Sagnac polarization-entangled photon source with a PPKTP crystal at telecom wavelength,” Opt. Express 22, 11498–11507 (2014). [CrossRef]   [PubMed]  

31. J. A. Armstrong, N. Bloembergen, J. Ducuing, and P. S. Pershan, “Interactions between light waves in a nonlinear dielectric,” Phys. Rev. 127, 1918–1939 (1962). [CrossRef]  

32. M. Fejer, G. Magel, D. H. Jundt, and R. Byer, “Quasi-phase-matched second harmonic generation: tuning and tolerances,” IEEE J. Quantum Electron. 28, 2631–2654 (1992). [CrossRef]  

33. H. Lee, H. Kim, M. Cha, and H. Moon, “Simultaneous type-0 and type-II spontaneous parametric down-conversions in a single periodically poled KTiOPO4 crystal,” Appl. Phys. B 108, 585–589 (2012). [CrossRef]  

34. T. Y. Fan, C. E. Huang, B. Q. Hu, R. C. Eckardt, Y. X. Fan, R. L. Byer, and R. S. Feigelson, “Second harmonic generation and accurate index of refraction measurements in flux-grown KTiOPO4,” Appl. Opt. 26, 2390–2394 (1987). [CrossRef]   [PubMed]  

35. J. D. Bierlein and H. Vanherzeele, “Potassium titanyl phosphate: properties and new applications,” J. Opt. Soc. Am. B 6, 622–633 (1989). [CrossRef]  

36. W. Wiechmann, T. Fukui, H. Masuda, and S. Kubota, “Refractive-index temperature derivatives of potassium titanyl phosphate,” Opt. Lett. 18, 1208–1210 (1993). [CrossRef]   [PubMed]  

37. K. Kato and E. Takaoka, “Sellmeier and thermo-optic dispersion formulas for KTP,” Appl. Opt. 41, 5040–5044 (2002). [CrossRef]   [PubMed]  

38. S. Emanueli and A. Arie, “Temperature-dependent dispersion equations for KTiOPO4 and KTiOAsO4,” Appl. Opt. 42, 6661–6665 (2003). [CrossRef]   [PubMed]  

39. H. Ito, H. Naito, and H. Inaba, “Generalized study on angular dependence of induced second-order nonlinear optical polarizations and phase matching in biaxial crystals,” J. Appl. Phys. 46, 3992–3998 (1975). [CrossRef]  

40. The authors of [16] have claimed that the slightly elliptic ring image obtained under the noncollinear degenerate condition stems from the biaxiality of the KTP crystal. However, this is ambiguous, as elliptic ring images can be generated from uniaxial crystals (nx = ny ≠ nz). In addition, KTP is approximately uniaxial (nx ≃ ny < nz).

41. Y.-C. Jeong, K.-H. Hong, and Y.-H. Kim, “Bright source of polarization-entangled photons using a PPKTP pumped by a broadband multi-mode diode laser,” Opt. Express 24, 1165–1174 (2016). [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 (10)

Fig. 1
Fig. 1 Schematic diagram of experimental setup. The pump is a single-longitudinal-mode cw laser with a center wavelength λp of ∼406.2 nm and a full-width at half-maximum (FWHM) bandwidth of Δν = 0.2(1) GHz. SMF: Single-mode fiber for 405 or 780 nm; QWP (HWP): Zero-order quarter (half)-wave plate for 405 or 808 nm; PBS: Broadband polarizing beam-splitter; Lens: Anti-reflection (AR)-coated plano-convex lens (focal length f = 200 mm, coating designation: A or B); PPKTP: periodically poled KTiOPO4 (Λ = 10 μm; length L = 10 mm); S: temporary image screen to measure photon rings; ODL: Optical delay line consisting of right-angled prism and x-axis translation stage; ϕ: Phase shifter consisting of sequential wave plates (QWP-HWP-QWP); LP: Broadband linear polarizer; TF: Temporally used wavelength tunable filter (FWHM of Δλ = 1.82 nm); SPCM: Single-photon counting module; FPGA: Single and coincidence counting unit.
Fig. 2
Fig. 2 Numerically calculated emission angles θ for photons emitted from PPKTP crystal exit plane for: (a) Degenerate case on xy- and xz-planes as a function of T; (b) degenerate case for several fixed T values on yz-plane; and (c) fixed T = 95°C on xy-plane as a function of photon wavelength.
Fig. 3
Fig. 3 T dependence of (a) λH and λV and (b) FWHM Δλ of H and V photons of pair generated under non-collinear condition, with spatial modes determined by degenerate condition at T = 95°C.
Fig. 4
Fig. 4 Measured photon ring images for (a) H and (b) V polarized photons with IF (IF center wavelength λIF ≃ 812 nm, FWHM = 3 nm).
Fig. 5
Fig. 5 Measured cross-section (1D) images of photon rings along y-axis for various T (48.6 – 98.6 °C) with IF (FWHM = 3 nm).
Fig. 6
Fig. 6 (a) Measured spectral distributions for photons in modes A and B obtained via TF at T = 42°C. (b) Center wavelength T dependence for photons in modes A and B. FWHM Δλ are represented by error bars. o: measured or estimated data. −: Gaussian or linear fitted lines.
Fig. 7
Fig. 7 Pump-power Pp dependence of count rates for (a) single counts of modes HA, VA, HB and VB, and (b–e) coincidence counts of modes HAHB, HAVB, VAHB and VAVB. Raw: Raw data of coincidence counts. Net: Accidental counts removed data. Fit: Linear fitted line of Net.
Fig. 8
Fig. 8 Experimental results at Pp = 0.5 mW for: (a) HOM dip (blue) and peak (red) interferences; (b) correlation functions of polarizations for four cases; and (c) real and (d) imaginary parts of reconstructed density matrix obtained via QST. For the correlation function measurements in (b), the LP from mode A (LPA) was fixed to 0° (red); 45° (green); 90° (blue); and 135° (black). Dotted lines: Sine-fitted curves.
Fig. 9
Fig. 9 Measured HOM dip and peak interferences for T = 62–102°C at Pp = 0.5 mW and t = 8 s. Peak (dip) data in red (blue) are measured based on coincidence counts of modes A and B with LPA and LPB angles set to 45° and (−)45°, respectively. To show the beat patterns clearly, lines are drawn between neighboring points.
Fig. 10
Fig. 10 HOM dip interferences of generated state, | λ A λ B A B ( | H H A B + e i ϕ | V V A B ) / 2, where ϕ = /2 and λAλB. The datapoints colored in blue, green, red, and black correspond to n = 0, 1, 2, and 3, respectively.

Tables (2)

Tables Icon

Table 1 Estimated entanglement qualities for various Pp

Tables Icon

Table 2 Entanglement qualities of generated states for various T

Equations (1)

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

| ψ A B = | H H A B + e i ϕ | V V A B 2 | λ A λ B A B ,
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.