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

Highly efficient visible and near-IR photon pair generation with thin-film lithium niobate

Open Access Open Access

Abstract

Efficient on-chip entangled photon pair generation at telecom wavelengths is an integral aspect of emerging quantum optical technologies, particularly for quantum communication and computing. However, moving to shorter wavelengths enables the use of more accessible silicon detector technology, and opens up applications in imaging and spectroscopy. Here, we present high brightness ((1.6 ± 0.3) × 109 pairs/s/mW/nm) visible–near-IR photon pair generation in a periodically poled lithium niobate nanophotonic waveguide. The degenerate spectrum of the photon pairs is centered at 811 nm with a bandwidth of 117 nm when pumped with a spectrally multimode laser diode. The measured on-chip source efficiency of (2.3 ± 0.5) × 1011 pairs/s/mW is on par with source efficiencies at telecom wavelengths and is also orders of magnitude higher than the efficiencies of other visible sources implemented in bulk crystal or diffused waveguide-based technologies. Further improvements in the brightness and efficiencies are possible by pumping the device with a single-frequency laser, which would also shrink the pair bandwidth. These results represent the shortest wavelength of photon pairs generated in a nanophotonic waveguide reported to date by nearly an octave.

© 2024 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Spontaneous parametric downconversion (SPDC) has been used for decades to produce quantum entanglement in various photonic degrees of freedom, serving as a workhorse in emerging quantum optical technologies. Compared with most nonlinear processes, SPDC is relatively inefficient, requiring over one million pump photons to produce one pair of entangled photons in even the highest performing crystals. However, recent advances in nanophotonics, particularly in thin-film lithium niobate (TFLN), have enabled significantly more efficient frequency conversion and quantum state generation [13] through sub-$\mathrm{\mu}$m interaction areas, high nonlinearities, and low material losses [4,5]. By exploiting this platform, many recent demonstrations of SPDC in TFLN [69] have achieved efficiencies three orders of magnitude greater than that of bulk crystal-based sources [10] and one order of magnitude greater than that of large diffused waveguide-based sources [11]. To date, most TFLN-based photon pair sources are designed for SPDC at telecom wavelengths because of the low losses in optical fibers at 1550 nm [12,13] and back-compatibility for applications such as quantum communication [14], computing [15], and a globally connected quantum network [16].

Although telecom photons are preferred for quantum information applications, visible and near-infrared photons are generally better suited for imaging and spectroscopy. Experiments at these wavelengths can take advantage of multi-pixel detectors such as electron-multiplying charge-coupled devices (EM-CCD) and single-photon avalanche detector (SPAD) arrays, enabling the measurements needed for imaging [1719] and characterization of high-dimensional entangled states [20,21]. Furthermore, the electronic transitions of molecules and atoms become accessible at near-IR wavelengths, allowing for fluorescence lifetime measurements [2224], compatibility with quantum memories [25], and fundamental studies of few-photon nonlinearities [10,26,27]. More generally, near-IR and visible photons can be detected with high quantum efficiency and low dark noise using existing mature silicon technology at room temperature, compared with near-IR–IR detectors which require cryogenic cooling [28]. Despite these advantages, all demonstrations of nanophotonic pair production have resided in the telecom region, and the best near-IR and visible photon pair sources are still large-area waveguides [2932] and bulk periodically poled crystals [10,26,29,3336]. To date, lithium niobate is the only nanophotonic platform that exhibits high transparency at near-UV pump wavelengths and supports a $\chi ^{(2)}$ nonlinearity. Thus, despite the known challenges, TFLN is uniquely posed to address this wavelength range. Potential reasons for the lack of visible TFLN devices stem from the difficulty in fabricating visible nonlinear circuits on thin-film lithium niobate due to factors such as the ultra-short poling periods required for quasi-phase matching and losses from material absorption [3739] and scattering [40]. In spite of these difficulties, high-performance visible devices in thin-film lithium niobate are becoming increasingly common for classical applications such as electro-optic modulation and second harmonic generation [4145].

Here, we extend TFLN-based SPDC sources to shorter wavelengths to produce high-brightness photon pairs in the visible and near-IR. The device produces an on-chip efficiency of $(2.3\pm \,0.5)\,\times \,10^{11}$ pairs/s/mW, which corresponds to a per-photon conversion efficiency of more than 1 photon pair converted in every 10,000 pump photons [$(1.1\,\pm \,0.2)\,\times \,10^{-4}$ pairs/s/photon]. This efficiency is nearly two orders of magnitude better than visible-light diffused waveguide SPDC sources [31] and is on par with the highest-performing TFLN sources in the telecom regime [8]. The SPDC from this device exhibits a broad spectrum centered at 811 nm with a degenerate FWHM bandwidth of 117 nm and an average brightness of $(1.6 \pm 0.3) \times 10^{9}$ pairs/s/mW/nm, a number limited by at least an order of magnitude by the pump laser linewidth (0.8 nm). Consistent with this bandwidth, we measure an ultrashort coherence time of $\sim$40 fs for the entangled photons with an indistinguishability of $100 \pm 1{\% }$. These results are the shortest wavelength entangled photons generated in TFLN by nearly an octave to date. Our results therefore show that, although pumped at wavelengths near what would usually be considered its cutoff range, TFLN can equally be a platform for visible–near-IR entangled photon applications as it is at telecom wavelengths.

2. Device Design and Fabrication

The periodically poled lithium niobate waveguides [Fig. 1(a)] were simulated in Lumerical MODE to determine the quasi-phase-matching poling period. The guided modes at the design pump wavelength (406 nm) and SPDC center wavelength (812 nm) were simulated using the bulk Sellmeier coefficients of lithium niobate [46] and silicon dioxide [47] with the geometric parameters shown in Fig. 1(c) and a 60° sidewall, which is consistent with the fabrication process. To take advantage of lithium niobate’s largest nonlinear tensor element ($d_{33}$ = 28 pm/V) [48], only the fundamental quasi-TE modes of X-cut lithium niobate were considered. An etch depth of 420 nm, a top width of 1.5 $\mathrm{\mu}$m, and a total LN thin film thickness of 600 nm were targeted for ease of optical coupling, fabrication, elimination of slab-mode leakage, and near-2-$\mathrm{\mu}$m poling period while providing high performance. For these parameters, the effective refractive indices ($n_{\text {eff,pump}}$ = 2.29, $n_{\text {eff,SPDC}}$ = 2.09) result in a quasi-phase-matching poling period of $\Lambda =\lambda _{\text {pump}}/\Delta n_{\text {eff}}=2.03\,\mathrm{\mu} \text {m}$ at the target pump wavelength of 406 nm [Fig. 1(d)].

 figure: Fig. 1.

Fig. 1. (a) Schematic of the periodically poled lithium niobate nanophotonic waveguide. (b) Second harmonic microscopy image of the periodic poling. Note that the distance between the electrodes is 15 $\mathrm{\mu}$m while the waveguide top width is 1.5 $\mathrm{\mu}$m, allowing for multiple waveguides in the center of the poled region where the duty cycle is $\sim$50%. (c) Mode profiles and waveguide geometry of the fundamental quasi-TE modes at the designed pump and SPDC center wavelengths. (d) Refractive indices and corresponding poling periods for a range of SPDC wavelengths.

Download Full Size | PDF

The devices were fabricated from a 1 cm by 1 cm die of a 5% MgO-doped X-cut thin-film lithium niobate on insulator wafer (NanoLN), which consists of 600 nm of lithium niobate bonded to 2 $\mathrm{\mu}$m of silicon dioxide on a 0.4-mm silicon substrate. An MgO-doped film was chosen to lower the coercive field necessary for poling [49,50] and to reduce potential photorefractive effects from the violet laser diode [51]. MgO-doped LN has been demonstrated to lower propagation losses [38], but this doping may also result in irregular poled domains due to leakage current in the charged domain walls [50]. To quasi-phase match the SPDC sources, poling electrodes (7 mm long) were first fabricated by performing a metal lift-off through electron beam lithography with bilayer polymethyl methacrylate (PMMA) resist followed by electron beam evaporation of titanium (15 nm) and gold (55 nm). The electrodes were poled with a 490-V and 70-$\mathrm{\mu}$s square wave, and the poled domain formation was monitored with second harmonic microscopy [52] [Fig. 1(b)]. After poling, waveguides were defined through an aligned electron beam lithography step with hydrogen silesquioxane (HSQ) resist followed by argon inductive coupled plasma reactive ion etching to achieve an etch depth of 420 nm, verified through atomic force microscopy. The chip facets were manually polished to increase coupling efficiency, resulting in a final waveguide length of approximately 8 mm. A broadband oscillator was used to verify the phase-matching wavelength through second harmonic generation (Supplement 1 Section S2) and was found to match with the computationally predicted second harmonic wavelength within $\pm$3 nm. This small discrepancy was likely due to fabrication tolerances, particularly in the etch depth and film thickness.

3. Device Characterization

The spectrum, generation rate, and coherence properties of the entangled photon pairs produced from the fabricated device are characterized, as shown in Fig. 2. In these experiments, the room-temperature periodically poled waveguide is pumped with a free-running laser diode (Coherent OBIS LX 405 nm) to produce entangled pairs [Fig. 2(a)]. An antireflection-coated aspheric lens (NA = 0.58, Thorlabs C140TMD-A) couples the free-space pump beam to the fundamental TE mode of the waveguide. The photon pairs produced in the fundamental TE mode are collected off-chip and collimated using a similar aspheric lens (NA = 0.58, Thorlabs C140TMD-B).

 figure: Fig. 2.

Fig. 2. Schematic for coupling and detection of the photon pairs. I, isolator; ND, variable neutral density filter; HWP, half-wave plate; L, aspheric lens; LPF, long-pass filter; emICCD, electron-multiplying intensified charge-coupled device; BS, beamsplitter; D, single-photon avalanche detector; TT, time-tagger; M, mirror. (a) Optical setup to couple into and out of the TFLN waveguide. (b) Characterization scheme for measuring the photon pair spectra. (c) Characterization scheme for coincidence counting. (d) Optical setup for the Michelson interferometer.

Download Full Size | PDF

The spectra of the entangled photon pairs are measured to assess the phase-matching properties and tunability of the device (Fig. 3). Pairs collected from the waveguide are transmitted to a grating spectrometer and measured using an electron-multiplying intensified camera [Fig. 2(b)]. To tune the SPDC emission, the center wavelength of the pump wavelength is varied from 405 to 406.4 nm by changing the drive current of the laser diode. Variable neutral-density filters are used to keep the pump power for all the collected spectra at a consistent 10 $\mathrm{\mu}$W. Each spectrum was divided by the spectrometer wavelength response and then normalized to the maximum count rate to infer the relative intensity. Three distinct phase-matching regions are explored [Fig. 3(a)]: (1) at long pump wavelengths, the phase-matching condition is not satisfied and SPDC emission is not observed; (2) from 405.6 to 405.9 nm, the degenerate wavelengths are phase matched; and (3) at short pump wavelengths, the spectrum splits and nondegenerate emission extending to the cutoff wavelength of the filter is observed. The dip in intensity in Fig. 3(a) around 755 nm is likely due to on-chip loss, potentially from a mode-crossing [53], and is investigated in further detail with a transmission measurement (Supplement 1 Section S2). Due to the linewidth of the laser used in the experiment (0.8-nm FWHM), the spectra are considerably broadened compared with the spectra expected from a single-frequency pump laser (Supplement 1 Figure S4). Nevertheless, experiment and theory reach qualitative agreement by accounting for the pump linewidth [Fig. 3(b)], described further in Supplement 1 Section S6. Further spectral broadening can also arise from propagation loss at the pump or SPDC wavelengths [54] or index variations [5557]. For all subsequent experiments, a laser center wavelength of 405.7 nm is used for degenerate phase matching. The resulting spectrum [Fig. 3(c)] is centered at 811.4 $\pm$ 0.7 nm with a FWHM bandwidth of 117 nm (53 THz) that accounts for 85% of the overall flux.

 figure: Fig. 3.

Fig. 3. (a) Measured and (b) theoretical SPDC spectra as a function of pump wavelength. Note that the arbitrary units for intensity could also be interpreted as counts per max counts due to the normalization of the spectra. (c) Lineouts of the measured and theoretical SPDC spectra at a pump wavelength of 405.7 nm, corresponding to the dashed horizontal lines in panels (a) and (b). Note that the frequency bandwidth per nm and the downconversion efficiency are wavelength dependent (Supplement 1 Eq. (S1)), resulting in higher intensities for shorter wavelengths. The theoretical spectra in panels (b) and (c) are calculated with the laser linewidth taken into account, described further in Supplement 1 Section S6.

Download Full Size | PDF

The pair generation efficiency of the device is measured through coincidence counting between two SPADs [Fig. 2(c)]. In this measurement, the pairs from the chip were split at a 50:50 broadband beamsplitter (Thorlabs BS014) and coupled to multimode optical fibers (Thorlabs M122L01) connected to the detectors. Coincidence detection events between the SPADs (Laser Components Count) were recorded with a time-tagger (Picoquant PicoHarp 300). Figure 4(a) shows a representative raw coincidence histogram recorded at 4.3 nW of on-chip pump power. The temporal correlation in this graph (3.4 ns FWHM) is given by the response time of the SPADs and not the entangled photon correlations (see Fig. 5 later in the text). The coincidence counts are corrected by background subtraction of the number of counts in a 9.5-ns window at the histogram peak from the number of counts in another 9.5-ns window in a background region far from the peak. Sweeping the laser power with a neutral density filter yields the curves in Figs. 4(b) and 4(c), which are linearly fit to determine the pair generation efficiency. Additional experimental details and the raw data from these measurements are given in Section S4 and Table S1. To account for the wavelength dependence of the SPAD quantum efficiency (Supplement 1 Section S3), all wavelengths in the spectrum are integrated over to calculate the average detection efficiency for single photons ($\eta _{1}=0.52$) as well as the average joint pair detection probability ($\eta _{12}=0.27$). Including a factor of 2 due to the probability of splitting pairs at the beamsplitter yields Eq. (1) for the measured efficiency of the source:

$$E = \frac{m_1 m_2}{m_c} \frac{\eta_{12}}{2 \eta_1^2}.$$
Here, $E$ is the pair generation efficiency, $m_1$ and $m_2$ are the singles rates at the two detectors, and $m_c$ is the rate of coincidences, all in units of counts/s/mW. A derivation of Eq. (1) can be found in Supplement 1 Section S5. Accounting for the 10.2-dB transmission loss of the pump laser into the waveguide, a pair generation efficiency of $(2.3\pm 0.5) \times 10^{11}$ pairs/s/mW is measured, which is equivalent to a per-pump-photon efficiency of $(1.1\pm 0.2) \times 10^{-4}$ pairs/s/photon. Over the 117-nm FWHM bandwidth of the spectrum, this efficiency translates to an average brightness of $(1.6 \pm 0.3) \times 10^{9}$ pairs/s/nm/mW. The uncertainties here and throughout this work are reported as one standard deviation, derived from the standard error in the fits of Figs. 4(b) and 4(c) and the uncertainty in the detector quantum efficiency. The ratio of singles counts to coincidence counts suggests that the transmission of the SPDC from the waveguide to each of the two detectors is 8.4 dB and 8.0 dB, respectively, which includes losses out of the waveguide and of the free-space optics. Our theoretical efficiency (Supplement 1 Section S6), including the FWHM of the pump laser, is $2.66 \times 10^{11}$ pairs/s/mW, in close agreement with our experimental results.

 figure: Fig. 4.

Fig. 4. (a) Raw coincidence histogram, including accidentals, at an input on-chip pump power of 4.3 nW. Measured (b) coincidence counts with accidentals subtracted and (c) singles counts while sweeping the input power. The vertical axes are measured in coincidence or singles counts per second (cps). The fitted slopes (lines) produce the pair generation efficiency of our device. Note that error bars are included in panels (b) and (c), but are smaller than the data markers. Additional experimental details and the raw data can be found in Supplement 1 Section S4 and Table S1.

Download Full Size | PDF

 figure: Fig. 5.

Fig. 5. (a) Measured and (b) simulated two-photon Michelson interferogram for the device. (c) Corresponding singles counts out of the interferometer, demonstrating single-photon interference within the coherence length of the source. (d) Coincidence counts and (e) singles counts far from the coherence length of the source. All measured coincidence counts [panels (a) and (d)] do not have accidentals subtracted. Note that error bars are included in panels (d) and (e), but are smaller than the data markers. Details of the simulated two-photon Michelson interferogram in panel (b) and the sinusoidal fit in panel (d) can be found in Supplement 1 Section S7.

Download Full Size | PDF

Finally, the two-photon interference is measured [Fig. 2(d)] to demonstrate the non-classical behavior of the produced photon pairs. Figure 5 shows the measured two-photon interferogram [Fig. 5(a)] obtained from the device without subtracting accidentals, as well as the one-photon interferogram [Fig. 5(c)] for comparison. Due to the aforementioned temporal resolution of the SPADs used here, the four unique paths through the interferometer are indistinguishable and combine to yield the interference pattern. The important features of the interferogram are as follows. (1) Near the zero time delay, photons are delayed within the coherence length of the source and exhibit both one- and two-photon interference. A visibility of $100 \pm 1$% is measured within the coherence length, with an uncertainty derived assuming Poissonian statistics. This near-perfect visibility indicates good mode overlap in the interferometer and indistinguishability of the photons within the pair. (2) Far from the coherence length of the source (delays greater than $\pm 20$ fs), interference between two photons taking different paths disappears, which explains why the single-photon interference [Fig. 5(c)] disappears in this region, shown in more detail in Fig. 5(e). Notably, interference between pairs of photons that travel together through the interferometer persists with a fringe period at half the pair wavelength [Fig. 5(d)]. This feature suggests quantum interference due to the energy-time entanglement of the pairs, and would not be observed if the light was generated from a coherent or a thermal source with a similar spectrum [58]. A fringe visibility of $43\pm 3{\% }$ is observed in this region far from the coherence length, which is close to the theoretical maximum of 50% for this experiment due to the temporal resolution of the detectors. The uncertainty in visibility is given from the standard error in the fit of Fig. 5(d), the details of which are outlined in Supplement 1 Section S7. The qualitative agreement between the measured and theoretical two-photon interferogram [Figs. 5(a) and 5(b)] suggests that SPDC and genuine energy-time entanglement are being produced.

4. Discussion

Compared with the state-of-the-art for visible photon pair sources, the device presented in this work exhibits substantially improved brightness and efficiency due to the small effective area of the nanophotonic waveguide. The device’s performance against reported literature devices spanning from the visible to IR is plotted in Fig. 6 using efficiency, brightness, and wavelength as the figures of merit. Our visible–near-IR device demonstrates improved efficiency and brightness even over other TFLN sources at the better-explored telecom band [6,7,9] and has comparable performance to the most efficient telecom TFLN source (Ref. [8]) to date. At short wavelengths, the phase-matching bandwidth decreases due to group velocity dispersion in lithium niobate, but the efficiency remains high because the downconversion spectral power density scales with the inverse fifth power of the SPDC wavelength ($\lambda _s^{-5}$) [29]. Thus, in addition to the benefits of higher-energy photon pairs, decreasing the SPDC wavelength to access higher efficiencies could enable single-photon nonlinearities when integrated with a resonator [59]. The high efficiency of this device has significant implications for practical uses of entangled photons, including allowing the use of low-power laser diodes for pair generation, reducing integration times, allowing high signal-to-noise coincidence measurements at low (nW) laser powers even with losses present, and reducing fluorescence and stray light.

 figure: Fig. 6.

Fig. 6. Comparison of relevant literature SPDC sources to this work with respect to (a) efficiency and (b) brightness against the SPDC wavelength. Horizontal error bars represent the reported bandwidth of the sources. Data for the efficiency, brightness, center wavelength, and bandwidth are taken from Refs. [10,26,29,33,35,36,60,61] for bulk crystal sources, Refs. [11,2932,6267] for large waveguide sources (including micromachined and diffused waveguides), and Refs. [69] for TFLN sources. Note that Refs. [11,64] are not included in panel (b) since the brightness and bandwidth were not reported. The shaded region represents the typical telecommunication wavelength window.

Download Full Size | PDF

Although the efficiency and brightness of the device compares well with literature, these values can be further improved by narrowing the bandwidth of the pump laser, as discussed further in Supplement 1 Section S6. A single-frequency pump is estimated to shrink the phase-matching bandwidth from 117 nm to 35 nm, increasing the brightness by a comparable factor. The SPDC process is also most efficient near degeneracy because the group velocity of the signal and idler are equal to first order, so the efficiency could increase by a factor of 5. Furthermore, compared with SPDC produced with a monochromatic pump, the radiation produced here with a multimode pump will exhibit reduced time-energy entanglement (and greater separability) due to the uncertainty in the energy of the pump photons. The coherence properties of the pump laser are also transferred to the SPDC state, which can reduce the N00N state interference visibility at delays longer than the coherence length of the pump laser. Note that this negative effect is not observed here since the path length differences in Fig. 5 cover more than the full coherence length of the SPDC source but less than the coherence length of the pump laser [68,69], indicating that the pump laser coherence does not impair the visibilities measured in this work. Conversely, one benefit of using a multifrequency pump is that the device sensitivity to the laser wavelength is reduced, yielding a higher stability average response with greater bandwidth. Another benefit of the multifrequency pump is in time-bin entanglement, which can be generated from revivals in the biphoton interference at integer multiples of the laser cavity length [70]. For these benefits, as well as for the greater accessibility of multimode lasers, a multimode pump laser was used in this work to represent an alternate and more applied use of photon pair sources. Furthermore, despite the aforementioned drawbacks of a multimode pump, the device was still able to achieve high efficiencies on par with the current state-of-the-art telecom TFLN pair sources.

5. Conclusion

Efficient photon pair generation has been demonstrated with an integrated thin-film lithium niobate waveguide at visible and NIR wavelengths (720–900 nm). An on-chip SPDC efficiency of $(2.3\pm 0.5) \times 10^{11}$ pairs/s/mW, which is on par with reported TFLN literature at the better-explored telecom wavelengths even with a multimode pump laser, has been produced near the usually associated cut-off wavelengths for the pump (406 nm) in TFLN. The photon pair spectra has an average brightness of $(1.6 \pm 0.3) \times 10^{9}$ pairs/s/mW/nm, centered at 811 nm with a 117-nm bandwidth. To date, these results are the shortest wavelength photon pairs generated in a thin-film platform by nearly an octave. The work opens up opportunities to exploit the quantum advantage of integrated entangled photon circuits beyond telecom to imaging and spectroscopy applications in the visible and NIR.

Funding

Air Force Office of Scientific Research (FA9550-20-1-0040); Army Research Office (W911NF-23-1-0048); U.S. Department of Energy (DE-SC0020151); National Science Foundation (DGE-1745301, EECS 1846273).

Acknowledgments

The authors gratefully acknowledge the critical support and infrastructure provided for this work by The Kavli Nanoscience Institute (KNI) and the Beckman Biological Imaging Facility at Caltech. This work was additionally supported by the KNI-Wheatley Scholar in Nanoscience and the Rothenberg Innovation Initiative. N.A.H. was supported by the Department of Defense (DoD) through the National Defense Science and Engineering Graduate (NDSEG) Fellowship Program. E.Y.H. was supported by the National Science Foundation Graduate Research Fellowship Program under Grant no. DGE-1745301. Any opinion, findings, and conclusions or recommendations expressed in this material are those of the authors(s) and do not necessarily reflect the views of the National Science Foundation.

Disclosures

The authors declare no conflicts of interest.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

Supplemental document

See Supplement 1 for supporting content.

References

1. M. Jankowski, C. Langrock, B. Desiatov, et al., “Ultrabroadband nonlinear optics in nanophotonic periodically poled lithium niobate waveguides,” Optica 7, 40 (2020). [CrossRef]  

2. C. Wang, C. Langrock, A. Marandi, et al., “Ultrahigh-efficiency wavelength conversion in nanophotonic periodically poled lithium niobate waveguides,” Optica 5, 1438 (2018). [CrossRef]  

3. R. Nehra, R. Sekine, L. Ledezma, et al., “Few-cycle vacuum squeezing in nanophotonics,” Science 377, 1333–1337 (2022). [CrossRef]  

4. A. Boes, L. Chang, C. Langrock, et al., “Lithium niobate photonics: unlocking the electromagnetic spectrum,” Science 379, eabj4396 (2023). [CrossRef]  

5. D. Zhu, L. Shao, M. Yu, et al., “Integrated photonics on thin-film lithium niobate,” Adv. Opt. Photonics 13, 242 (2021). [CrossRef]  

6. J. Zhao, C. Ma, M. Rüsing, et al., “High quality entangled photon pair generation in periodically poled thin-film lithium niobate waveguides,” Phys. Rev. Lett. 124, 163603 (2020). [CrossRef]  

7. U. A. Javid, J. Ling, J. Staffa, et al., “Ultrabroadband entangled photons on a nanophotonic chip,” Phys. Rev. Lett. 127, 183601 (2021). [CrossRef]  

8. G.-T. Xue, Y.-F. Niu, X. Liu, et al., “Ultrabright multiplexed energy-time-entangled photon generation from lithium niobate on insulator Chip,” Phys. Rev. Appl. 15, 064059 (2021). [CrossRef]  

9. Y. Zhang, H. Li, T. Ding, et al., “Scalable, fiber-compatible lithium niobate-on-insulator microwaveguides for efficient nonlinear photonics,” Optica 10, 688 (2023). [CrossRef]  

10. B. Dayan, A. Pe’er, A. A. Friesem, et al., “Nonlinear Interactions with an ultrahigh flux of broadband entangled photons,” Phys. Rev. Lett. 94, 043602 (2005). [CrossRef]  

11. M. Bock, A. Lenhard, C. Chunnilall, et al., “Highly efficient heralded single-photon source for telecom wavelengths based on a PPLN waveguide,” Opt. Express 24, 23992 (2016). [CrossRef]  

12. T. Miya, Y. Terunuma, T. Hosaka, et al., “Ultimate low-loss single-mode fibre at 1.55 μm,” Electron. Lett. 15, 106 (1979). [CrossRef]  

13. K. Nagayama, M. Kakui, M. Matsui, et al., “Ultra-low-loss (0.1484 dB/km) pure silica core fibre and extension of transmission distance,” Electron. Lett. 38, 1168 (2002). [CrossRef]  

14. N. Gisin and R. Thew, “Quantum communication,” Nat. Photonics 1, 165–171 (2007). [CrossRef]  

15. J. L. O’Brien, “Optical quantum computing,” Science 318, 1567–1570 (2007). [CrossRef]  

16. H. J. Kimble, “The quantum internet,” Nature 453, 1023–1030 (2008). [CrossRef]  

17. E. Toninelli, P.-A. Moreau, T. Gregory, et al., “Resolution-enhanced quantum imaging by centroid estimation of biphotons,” Optica 6, 347 (2019). [CrossRef]  

18. H. Defienne, P. Cameron, B. Ndagano, et al., “Pixel super-resolution with spatially entangled photons,” Nat. Commun. 13, 3566 (2022). [CrossRef]  

19. P.-A. Moreau, E. Toninelli, T. Gregory, et al., “Imaging with quantum states of light,” Nat. Rev. Phys. 1, 367–380 (2019). [CrossRef]  

20. D. Zia, N. Dehghan, A. D’Errico, et al., “Interferometric imaging of amplitude and phase of spatial biphoton states,” Nat. Photonics 17, 1009–1016 (2023). [CrossRef]  

21. Y. Zhang, D. England, A. Nomerotski, et al., “High speed imaging of spectral-temporal correlations in Hong-Ou-Mandel interference,” Opt. Express 29, 28217 (2021). [CrossRef]  

22. N. Harper, B. P. Hickam, M. He, et al., “Entangled photon correlations allow a continuous-wave laser diode to measure single-photon, time-resolved fluorescence,” J. Phys. Chem. Lett. 14, 5805–5811 (2023). [CrossRef]  

23. A. Eshun, X. Yi, A. Wilson, et al., “Fluorescence lifetime measurements using photon pair correlations generated via spontaneous parametric down conversion (SPDC),” Opt. Express 31, 26935 (2023). [CrossRef]  

24. G. Scarcelli and S. H. Yun, “Entangled-photon coincidence fluorescence imaging,” Opt. Express 16, 16189 (2008). [CrossRef]  

25. H. Zhang, X.-M. Jin, J. Yang, et al., “Preparation and storage of frequency-uncorrelated entangled photons from cavity-enhanced spontaneous parametric downconversion,” Nat. Photonics 5, 628–632 (2011). [CrossRef]  

26. D. Tabakaev, A. Djorovic, L. La Volpe, et al., “Spatial properties of entangled two-photon absorption,” Phys. Rev. Lett. 129, 183601 (2022). [CrossRef]  

27. A. Pe’er, B. Dayan, A. A. Friesem, et al., “Temporal shaping of entangled photons,” Phys. Rev. Lett. 94, 073601 (2005). [CrossRef]  

28. F. Ceccarelli, G. Acconcia, A. Gulinatti, et al., “Recent advances and future perspectives of single-photon avalanche diodes for quantum photonics applications,” Adv. Quantum Technol. 4, 2000102 (2021). [CrossRef]  

29. M. Fiorentino, S. M. Spillane, R. G. Beausoleil, et al., “Spontaneous parametric down-conversion in periodically poled KTP waveguides and bulk crystals,” Opt. Express 15, 7479 (2007). [CrossRef]  

30. S. Tanzilli, H. De Riedmatten, W. Tittel, et al., “Highly efficient photon-pair source using periodically poled lithium niobate waveguide,” Electron. Lett. 37, 26 (2001). [CrossRef]  

31. T. B. Gäbler, P. Hendra, N. Jain, et al., “Photon Pair Source based on PPLN-Waveguides for Entangled Two-Photon Absorption,” arXiv, arXiv:2303.16584 (2023). [CrossRef]  

32. A. Jechow, A. Heuer, and R. Menzel, “High brightness, tunable biphoton source at 976 nm for quantum spectroscopy,” Opt. Express 16, 13439 (2008). [CrossRef]  

33. M. B. Nasr, S. Carrasco, B. E. A. Saleh, et al., “Ultrabroadband biphotons generated via chirped quasi-phase-matched optical parametric down-conversion,” Phys. Rev. Lett. 100, 183601 (2008). [CrossRef]  

34. A. Tanaka, R. Okamoto, H. H. Lim, et al., “Noncollinear parametric fluorescence by chirped quasi-phase matching for monocycle temporal entanglement,” Opt. Express 20, 25228 (2012). [CrossRef]  

35. M. Okano, H. H. Lim, R. Okamoto, et al., “0.54 μm resolution two-photon interference with dispersion cancellation for quantum optical coherence tomography,” Sci. Rep. 5, 18042 (2015). [CrossRef]  

36. S. Sensarn, G. Y. Yin, and S. E. Harris, “Generation and compression of chirped biphotons,” Phys. Rev. Lett. 104, 253602 (2010). [CrossRef]  

37. R. Bhatt, I. Bhaumik, S. Ganesamoorthy, et al., “Urbach tail and bandgap analysis in near stoichiometric LiNbO3 crystals,” Phys. Status Solidi A 209, 176–180 (2012). [CrossRef]  

38. J. R. Schwesyg, M. C. C. Kajiyama, M. Falk, et al., “Light absorption in undoped congruent and magnesium-doped lithium niobate crystals in the visible wavelength range,” Appl. Phys. B 100, 109–115 (2010). [CrossRef]  

39. M. V. Ciampolillo, A. Zaltron, M. Bazzan, et al., “Quantification of iron (Fe) in lithium niobate by optical absorption,” Appl. Spectrosc. 65, 216–220 (2011). [CrossRef]  

40. R. Maurer, “Glass fibers for optical communications,” Proc. IEEE 61, 452–462 (1973). [CrossRef]  

41. S. Xue, Z. Shi, J. Ling, et al., “Full-spectrum visible electro-optic modulator,” Optica 10, 125–126 (2023). [CrossRef]  

42. D. Renaud, D. R. Assumpcao, G. Joe, et al., “Sub-1 Volt and high-bandwidth visible to near-infrared electro-optic modulators,” Nat. Commun. 14, 1496 (2023). [CrossRef]  

43. T. Park, H. S. Stokowski, V. Ansari, et al., “High-efficiency second harmonic generation of blue light on thin-film lithium niobate,” Opt. Lett. 47, 2706–2709 (2022). [CrossRef]  

44. B. Desiatov, A. Shams-Ansari, M. Zhang, et al., “Ultra-low-loss integrated visible photonics using thin-film lithium niobate,” Optica 6, 380–384 (2019). [CrossRef]  

45. A. A. Sayem, Y. Wang, J. Lu, et al., “Efficient and tunable blue light generation using lithium niobate nonlinear photonics,” Appl. Phys. Lett. 119, 231104 (2021). [CrossRef]  

46. O. Gayer, Z. Sacks, E. Galun, et al., “Temperature and wavelength dependent refractive index equations for MgO-doped congruent and stoichiometric LiNbO3,” Appl. Phys. B 91, 343–348 (2008). [CrossRef]  

47. T. Toyoda and M. Yabe, “The temperature dependence of the refractive indices of fused silica and crystal quartz,” J. Phys. D: Appl. Phys. 16, L97–L100 (1983). [CrossRef]  

48. I. Shoji, T. Kondo, A. Kitamoto, et al., “Absolute scale of second-order nonlinear-optical coefficients,” J. Opt. Soc. Am. B 14, 2268 (1997). [CrossRef]  

49. K. Mizuuchi, A. Morikawa, T. Sugita, et al., “Electric-field poling in Mg-doped LiNbO3,” J. Appl. Phys. 96, 6585–6590 (2004). [CrossRef]  

50. J. T. Nagy and R. M. Reano, “Reducing leakage current during periodic poling of ion-sliced x-cut MgO doped lithium niobate thin films,” Opt. Mater. Express 9, 3146 (2019). [CrossRef]  

51. Y. Furukawa, K. Kitamura, S. Takekawa, et al., “Photorefraction in LiNbO3 as a function of [Li]/[Nb] and MgO concentrations,” Appl. Phys. Lett. 77, 2494–2496 (2000). [CrossRef]  

52. M. Rúsing, J. Zhao, and S. Mookherjea, “Second harmonic microscopy of poled x-cut thin film lithium niobate: understanding the contrast mechanism,” J. Appl. Phys. 126, 114105 (2019). [CrossRef]  

53. Y. Hu, M. Yu, B. Buscaino, et al., “High-efficiency and broadband on-chip electro-optic frequency comb generators,” Nat. Photonics 16, 679–685 (2022). [CrossRef]  

54. L. G. Helt, M. J. Steel, and J. E. Sipe, “Spontaneous parametric downconversion in waveguides: what’s loss got to do with it?” New J. Phys. 17, 013055 (2015). [CrossRef]  

55. J. Zhao, X. Li, T.-C. Hu, et al., “Unveiling the origins of quasi-phase matching spectral imperfections in thin-film lithium niobate frequency doublers,” APL Photonics 8, 126106 (2023). [CrossRef]  

56. P. S. Kuo, “Noncritical phasematching behavior in thin-film lithium niobate frequency converters,” Opt. Lett. 47, 54 (2022). [CrossRef]  

57. P.-K. Chen, I. Briggs, C. Cui, et al., “Adapted poling to break the nonlinear efficiency limit in nanophotonic lithium niobate waveguides,” Nat. Nanotechnol. 19, 44–50 (2024). [CrossRef]  

58. Z. Y. Ou, X. Y. Zou, L. J. Wang, et al., “Observation of nonlocal interference in separated photon channels,” Phys. Rev. Lett. 65, 321–324 (1990). [CrossRef]  

59. M. Jankowski, J. Mishra, and M. M. Fejer, “Dispersion-engineered nanophotonics: a flexible tool for nonclassical light,” J. Phys. Photonics 3, 042005 (2021). [CrossRef]  

60. M. Okano, R. Okamoto, A. Tanaka, et al., “Generation of broadband spontaneous parametric fluorescence using multiple bulk nonlinear crystals,” Opt. Express 20, 13977 (2012). [CrossRef]  

61. J. Lavoie, T. Landes, A. Tamimi, et al., “Phase-modulated interferometry, spectroscopy, and refractometry using entangled photon pairs,” Adv. Quantum Technol. 3, 1900114 (2020). [CrossRef]  

62. D. Kang, A. Anirban, and A. S. Helmy, “Monolithic semiconductor chips as a source for broadband wavelength-multiplexed polarization entangled photons,” Opt. Express 24, 15160 (2016). [CrossRef]  

63. H. Jin, F. Liu, P. Xu, et al., “On-chip generation and manipulation of entangled photons based on reconfigurable lithium-niobate waveguide circuits,” Phys. Rev. Lett. 113, 103601 (2014). [CrossRef]  

64. E. Meyer-Scott, D. McCloskey, K. Golos, et al., “Certifying the presence of a photonic qubit by splitting it in two,” Phys. Rev. Lett. 116, 070501 (2016). [CrossRef]  

65. D.-S. Ding, W. Zhang, S. Shi, et al., “Hybrid-cascaded generation of tripartite telecom photons using an atomic ensemble and a nonlinear waveguide,” Optica 2, 642 (2015). [CrossRef]  

66. P. Sarrafi, E. Y. Zhu, B. M. Holmes, et al., “High-visibility two-photon interference of frequency-time entangled photons generated in a quasi-phase-matched AlGaAs waveguide,” Opt. Lett. 39, 5188 (2014). [CrossRef]  

67. G. Fujii, N. Namekata, M. Motoya, et al., “Bright narrowband source of photon pairs at optical telecommunication wavelengths using a type-II periodically poled lithium niobate waveguide,” Opt. Express 15, 12769 (2007). [CrossRef]  

68. A. V. Burlakov, M. V. Chekhova, O. A. Karabutova, et al., “Biphoton interference with a multimode pump,” Phys. Rev. A 63, 053801 (2001). [CrossRef]  

69. O. Kwon, Y.-S. Ra, and Y.-H. Kim, “Coherence properties of spontaneous parametric down-conversion pumped by a multi-mode cw diode laser,” Opt. Express 17, 13059–13069 (2009). [CrossRef]  

70. O. Kwon, K.-K. Park, Y.-S. Ra, et al., “Time-bin entangled photon pairs from spontaneous parametric down-conversion pumped by a cw multi-mode diode laser,” Opt. Express 21, 25492–25500 (2013). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental document with additional details on the SPDC background theory, coincidence raw data and detection probabilities, multimode pump theory, and theoretical Michelson interferogram.

Data availability

Data underlying the results presented in this paper are not publicly available at this time but may be obtained from the authors upon reasonable request.

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 (6)

Fig. 1.
Fig. 1. (a) Schematic of the periodically poled lithium niobate nanophotonic waveguide. (b) Second harmonic microscopy image of the periodic poling. Note that the distance between the electrodes is 15 $\mathrm{\mu}$m while the waveguide top width is 1.5 $\mathrm{\mu}$m, allowing for multiple waveguides in the center of the poled region where the duty cycle is $\sim$50%. (c) Mode profiles and waveguide geometry of the fundamental quasi-TE modes at the designed pump and SPDC center wavelengths. (d) Refractive indices and corresponding poling periods for a range of SPDC wavelengths.
Fig. 2.
Fig. 2. Schematic for coupling and detection of the photon pairs. I, isolator; ND, variable neutral density filter; HWP, half-wave plate; L, aspheric lens; LPF, long-pass filter; emICCD, electron-multiplying intensified charge-coupled device; BS, beamsplitter; D, single-photon avalanche detector; TT, time-tagger; M, mirror. (a) Optical setup to couple into and out of the TFLN waveguide. (b) Characterization scheme for measuring the photon pair spectra. (c) Characterization scheme for coincidence counting. (d) Optical setup for the Michelson interferometer.
Fig. 3.
Fig. 3. (a) Measured and (b) theoretical SPDC spectra as a function of pump wavelength. Note that the arbitrary units for intensity could also be interpreted as counts per max counts due to the normalization of the spectra. (c) Lineouts of the measured and theoretical SPDC spectra at a pump wavelength of 405.7 nm, corresponding to the dashed horizontal lines in panels (a) and (b). Note that the frequency bandwidth per nm and the downconversion efficiency are wavelength dependent (Supplement 1 Eq. (S1)), resulting in higher intensities for shorter wavelengths. The theoretical spectra in panels (b) and (c) are calculated with the laser linewidth taken into account, described further in Supplement 1 Section S6.
Fig. 4.
Fig. 4. (a) Raw coincidence histogram, including accidentals, at an input on-chip pump power of 4.3 nW. Measured (b) coincidence counts with accidentals subtracted and (c) singles counts while sweeping the input power. The vertical axes are measured in coincidence or singles counts per second (cps). The fitted slopes (lines) produce the pair generation efficiency of our device. Note that error bars are included in panels (b) and (c), but are smaller than the data markers. Additional experimental details and the raw data can be found in Supplement 1 Section S4 and Table S1.
Fig. 5.
Fig. 5. (a) Measured and (b) simulated two-photon Michelson interferogram for the device. (c) Corresponding singles counts out of the interferometer, demonstrating single-photon interference within the coherence length of the source. (d) Coincidence counts and (e) singles counts far from the coherence length of the source. All measured coincidence counts [panels (a) and (d)] do not have accidentals subtracted. Note that error bars are included in panels (d) and (e), but are smaller than the data markers. Details of the simulated two-photon Michelson interferogram in panel (b) and the sinusoidal fit in panel (d) can be found in Supplement 1 Section S7.
Fig. 6.
Fig. 6. Comparison of relevant literature SPDC sources to this work with respect to (a) efficiency and (b) brightness against the SPDC wavelength. Horizontal error bars represent the reported bandwidth of the sources. Data for the efficiency, brightness, center wavelength, and bandwidth are taken from Refs. [10,26,29,33,35,36,60,61] for bulk crystal sources, Refs. [11,2932,6267] for large waveguide sources (including micromachined and diffused waveguides), and Refs. [69] for TFLN sources. Note that Refs. [11,64] are not included in panel (b) since the brightness and bandwidth were not reported. The shaded region represents the typical telecommunication wavelength window.

Equations (1)

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

E = m 1 m 2 m c η 12 2 η 1 2 .
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.