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

Fast and accurate measurement of the polarization-dependent detection efficiency of superconducting nanowire single photon detectors

Open Access Open Access

Abstract

Superconducting nanowire single photon detectors (SNSPDs) have been extensively investigated due to their superior characteristics, including high system detection efficiency, low dark count rate and short recovery time. The polarization sensitivity introduced by the meandering-type superconductor nanowires is an intrinsic property of SNSPD, which is normally measured by sweeping hundreds of points on the Poincaré sphere to overcome the unknown birefringent problem of the SNSPD’s delivery fiber. In this paper, we propose an alternative method to characterize the optical absorptance of SNSPDs, without sweeping hundreds of points on the Poincaré sphere. It is shown theoretically that measurements on the system detection efficiencies (SDEs) subject to cases of four specific photon polarization states are sufficient to reveal the two eigen-absorptances of the SNSPD. We validate the proposed method by comparing the measured detection spectra with the spectra attained from sweeping points on the Poincaré sphere and the simulated absorption spectra.

© 2022 Optica Publishing Group under the terms of the Optica Open Access Publishing Agreement

1. Introduction

Superconducting nanowire single photon detectors (SNSPDs) are a new subfield of photon detectors proposed by Gol’tsman in 2001 [1] and play a vital role in various applications, including quantum computation [2], quantum key distribution technology [3], satellite communications [4] and imaging [5], due to their superior characteristics, such as high system detection efficiency (SDE) [6,7], short recovery time [8], low dark count rate and low timing jitter [9]. Among all these prominent features, SDE is the most important characteristic, which is proportional to the optical absorptance of the ultrathin superconducting film. Researchers have explored many ways to improve the SDE and have demonstrated significant achievements. The most acknowledged way is to apply meandering-type nanowire patterns when designing superconducting film chips, with an optical cavity and anti-reflection coating [10] integrated to enhance the photon confinement in the nanowire region. In addition, SNSPDs integrated with waveguides [11] and nanoantennae [12] have also been proposed for seeking better performance and applications of SNSPDs.

Due to the anisotropic nature of the periodically arranged meander nanowires, the SDE of SNSPD shows a strong dependence on the polarization states of the incident photons. The polarization feature of SNSPD is vital in quantum optical experiments and applications requiring high detection efficiency and low polarization dependency. To reduce the polarization sensitivity induced by the grating structure of the nanowire, various schemes have been proposed [1317]. While effort is almost fully focused on the device design for perfect detector performance, the measurement tactic of the optical response of SNSPD seems to be stagnant. Note that since single-mode optical fibers are in general slightly birefringent due to stresses and manufacturing imperfections, the polarization state at the detector end may differ significantly from the polarization state prepared at the input end of the device’s delivery fiber in an unknown manner. This creates a difficulty for aligning the polarization of the received photon being parallel or perpendicular to the nanowire direction of the SNSPD. To overcome this difficulty and measure the polarization-dependent SDE, one common approach is with the help of a paddle polarization controller [18], which is a manual operation and suffers from many uncertainties. Another common approach is sweeping the polarization states of the input light over the whole Poincaré sphere [6] to find the maximum SDEmax and minimum SDEmin. However, over hundreds of sweeping points are generally needed [19] for accurate SDE characterization at only one wavelength. This leads to significant consumption of measurement time, especially in spectral characterization covering a broadband wavelength range [20].

In this paper, we propose an alternative method for measuring the polarization-dependent SDEs of SNSPDs, which we term as the four-state polarization algorithm (FSPA). By employing a computational scheme inspired by the computational spectrometer developed recently [21], it is shown theoretically that by preparing incident photons with four specific polarization states and measuring the four corresponding SDEs, the maximum and minimum detection efficiencies of SNSPDs can be computed, even if the birefringent property of the delivery fiber is not known. Experimental results confirm that the SDE spectra obtained using the FSPA method agree well with the spectra obtained from the traditional method (i.e., sweeping hundreds of points on the Poincaré sphere) and the simulated absorptance spectra (with the effect of quantum efficiency being considered). The proposed method is therefore a promising candidate for fast and accurate characterization of the optical performance of SNSPDs.

2. Concept and methods

2.1 Measurement setup and theory model

We first briefly introduce the optical measurement setup before elaborating on the FSPA method. As shown in Fig. 1, the whole measurement configuration consists of two main compositions, the optical part in the green dashed box and the electrical section in the blue dashed box. The red line denotes the optical fiber, while the black line represents the coaxial cable. A continuous-wave (CW) laser (Keysight N7778C) spans a wavelength range from 1490 nm to 1640 nm, and the generated light is directed into a polarization synthesizer (Keysight N7786B). The polarization synthesizer enables us to prepare the polarization of the input photon with arbitrary states. An attenuator A1 functions as a first-stage attenuator and is followed by a fiber-coupled beam splitter with a splitting ratio of 50:50. One branch of the splitter connects to a power meter PM1 (Thorlabs PM320E) and monitors the real-time optical power incident onto the SNSPD, while the other branch is directed toward the second attenuator A2. With proper settings, the light is attenuated to the single-photon level and sent to the fiber-connected SNSPD. Note that before we take the measurement, the splitting ratio between PM2 and PM1 was calibrated over the whole spectra. Thus, the dispersion effect of the measurement system is eliminated.

 figure: Fig. 1.

Fig. 1. Schematics of the polarization-dependent SDE measurement setup.

Download Full Size | PDF

Inspired by the computational spectrometer developed recently [21], here we attempt to measure the two eigen-absorptances of SNSPDs using a polarization related computational scheme. Consider the case where the polarization synthesizer outputs a light field with two orthogonal components Ex and Eyexp(), where Ex and Ey denote the amplitude of the two polarization components and φ denotes the phase difference between them. Dictated by the square law of intensity detection, the absorptance of the SNSPD can be in general expressed as an interferometric quadrature form:

$$\textrm{A} = a \cdot \textrm{E}_\textrm{x}^\textrm{2} + b \cdot \textrm{E}_\textrm{y}^\textrm{2} + c \cdot {\textrm{E}_\textrm{x}}{\textrm{E}_\textrm{y}}\cos ({\varphi _b} - {\varphi _\Delta })$$
where a, b, c and φb are constants that are related to the birefringent property of the delivery fiber and the eigen-aborptance of the SNSPD. Since Ex and Ey satisfy the normalization condition Ex2 + Ey2 = 1, the extremums of A can be found by searching the parameter space spanned by Ex/Ey and φ. Using some mathematical manipulations, one can prove that the maximum and minimum absorptance reads as:
$${\textrm{A}_{\max ,\min }} = \frac{{a + b}}{2} \pm \sqrt {{{\left( {\frac{{a - b}}{2}} \right)}^2} + {{\left( {\frac{c}{2}} \right)}^2}}$$

It follows that Amax and Amin are computable, if a, b and c can be determined using some sort of measurements.

Note that in Eq. (1), Ex, Ey and φ can be adjusted at will by programming the polarization synthesizer. Taking advantage of this fact, one can simply choose: Ex = 1, Ey = 0 (horizontally polarization), Ex = 0, Ey = 1 (vertically polarization), Ex =${{\sqrt 2 } / 2}$, Ey =${{\sqrt 2 } / 2}$, φ△ = 0 (45° oblique linearly polarization) and Ex =${{\sqrt 2 } / 2}$, Ey =${{\sqrt 2 } / 2}$, φ = π/2 (circularly polarization). Substituting the above parameters into Eq. (1), one ends up with:

$$\left\{ {\begin{array}{{c}} {{\textrm{A}_ \to } = a}\\ {{\textrm{A}_ \uparrow } = b}\\ {\textrm{ }{\textrm{A}_\mathrm{\ \nearrow }} = \frac{{a + b}}{2} + \frac{c}{2}\textrm{cos}{\varphi_\textrm{b}}\textrm{ }}\\ {\textrm{ }{\textrm{A}_\mathrm{\bigcirc }} = \frac{{a + b}}{2} + \frac{c}{2}\textrm{sin}{\varphi_\textrm{b}}\textrm{ }} \end{array}} \right.$$
where ${\textrm{A}_ \to }$, ${\textrm{A}_ \uparrow }$, ${\textrm{A}_\mathrm{\ \nearrow }}$ and ${\textrm{A}_\mathrm{\bigcirc }}$ denote the measured absorptance for cases of horizontally, vertically, 45° oblique linearly and circularly polarized light respectively. It follows from Eq. (3) that $a = {\textrm{A}_ \to }$, $b = {\textrm{A}_ \uparrow }$ and $c = \sqrt {{{({\textrm{2}{\textrm{A}_\mathrm{\ \nearrow }} - {\textrm{A}_ \to } - {\textrm{A}_ \uparrow }} )}^2} + {{({\textrm{2}{\textrm{A}_\mathrm{\bigcirc }} - {\textrm{A}_ \to } - {\textrm{A}_ \uparrow }} )}^2}}$. Therefore Eq. (2) becomes:
$${\textrm{A}_{\max ,\min }} = \frac{{{\textrm{A}_ \to } + {\textrm{A}_ \uparrow }}}{2} \pm \sqrt {{{\left( {\frac{{{\textrm{A}_ \to } - {\textrm{A}_ \uparrow }}}{2}} \right)}^2} + {{\left( {{\textrm{A}_\mathrm{\ \nearrow }} - \frac{{{\textrm{A}_ \to } + {\textrm{A}_ \uparrow }}}{2}} \right)}^2} + {{\left( {{\textrm{A}_\mathrm{\bigcirc }} - \frac{{{\textrm{A}_ \to } + {\textrm{A}_ \uparrow }}}{2}} \right)}^2}}$$

Equation (4) shows that the maximum absorptance Amax and minimum absorptance Amin can be computed from measurements that involve only four special points on the Poincaré sphere. We note here that for the case of a regular SNSPD, which contains only one absorptive layer, i.e. the nanowire meander, we have Amax = ATE and Amin = ATM. A discussion on a more general case that involving multiple absorptive layers can be found in Supplement 1. Also, we note that the four special polarization states proposed here are not the only choice for measuring Amax and Amin. In other word, one can also choose a different set of four-polarization states to complete the absorptance measurement, as long as a, b and c can be uniquely determined from measurements of Eq. (1).

2.2 Device fabrication

SNSPD was fabricated on the substrate of a distributed Bragg reflector (DBR), as shown in Fig. 2(a). The DBR consists of 13.5 pairs of alternative SiO2 and Ta2O5, on top of which the NbN thin film was deposited by sputtering. The gold conduct was then fabricated by lift-off processes. The meandering nanowire pattern was written by electron beam lithography. After development, the nanowire pattern is transferred to the NbN film by reactive ion etching process. Finally, a deep silicon etch step was implemented to obtain the key-hole for self-aligning SNSPD [22]. Figure 2(b) presents the typical scanning electron microscopy (SEM) image of the meandering nanowire, with a nanowire width of 65 ± 2 nm and a pitch of ∼160 nm. The active area of the nanowire is a circle with a diameter of 16 µm. The zoomed SEM images in the right panel of Fig. 2(b) indicate that the fabricated nanowires feature a good uniformity.

 figure: Fig. 2.

Fig. 2. (a) Schematic of the NbN nanowire fabricated on a DBR substrate. Light is directly coupled to the nanowire region by self-aligning coupling approach. (b) SEM image of the fabricated SNSPD. The nanowire shows a uniform width of 65 ± 2 nm and a pitch of ∼160 nm. (c) Measured SDE (scatters) and sigmoid fitting curve (lines) for TE (red) and TM (blue) polarized wave as a function of normalized bias current at a wavelength of 1550 nm. DCR is indicated by the black dots referring to the right y-axis. (d) Timing jitter of SNSPD at three different bias currents of the measured SNSPD.

Download Full Size | PDF

Figure 2(c) shows the measured SDE of SNSPD as a function of the normalized bias current at a wavelength of 1550 nm. The red and blue scatters indicate the response for TE- and TM-polarized light while the red and blue lines show the sigmoid-curve fittings of the corresponding photon-responses. The dark count rate (DCR) is presented by the black dots, referring to the right y-axis of Fig. 2(c). The DCR is less than 100 when Ib is less than 0.85 Ic, where Ic indicates the critical current. The SDE of SNSPD for TE- and TM- polarized photon at Ib = 0.85 Ic is 60% and 13.5%, respectively. Figure 2(d) shows the typical timing jitter of the fabricated SNSPD, with full width at half maximums (FWHMs) of 170 ps, 146 ps, 124 ps for the histogram, at a bias current of Ib = 0.76 Ic, 0.85 Ic, and 0.93 Ic, respectively. The inset of Fig. 2(d) shows the photon-response pulse waveforms recorded with a high-speed oscilloscope at a working current of 5.5 µA. The pulse peak is ∼100 mV in total, and fits well with the exponential decay function, with a pulse-decay time of ∼40 ns.

3. Results and discussion

Using the FSPA method summarized in Eq. (4), the maximum (TE) and minimum (TM) SDEs of the fabricated SNSPD are measured. The results are presented in Fig. 3(a) as a function of wavelength for cases of three different bias currents. The blue, orange and magenta circles represent the SDE spectra of the TE-polarized photons, while the green, violet and cyan squares represent the SDE spectra of the TM-polarized photons. The difference in the SDE spectra at different biases is caused by the change of quantum efficiency, since the detector is not operated at the saturation regime. To validate the proposed FSPA method, Fig. 3(b) compares the spectra obtained by the FSPA with the spectra obtained by the Poincaré sphere scan approach. The orange and violet symbols denote the SDE for TE and TM polarized light measured by the FSPA approach at a bias current of 0.85 Ic, and the light blue triangles denote the SDE measured by sweeping 100 points on the Poincaré sphere. The 100 points are chosen to be evenly distributed on the Poincaré sphere. Two observations can be made from Fig. 3(b). First of all, the unfilled portion of the blue area indicates that 100 sampling points on the Poincaré sphere are insufficient to precisely measure the SDE. Secondly and most importantly, ATE and ATM obtained by the FSPA method agree well with the maximum and minimum bounds obtained by sweeping points on the Poincaré sphere, suggesting a validation of the FSPA approach.

 figure: Fig. 3.

Fig. 3. (a) Measured SDE as a function of wavelength for cases of three different bias currents by means of FSPA. (b) Comparison of the SDE spectra obtained by FSPA and sweeping 100 points on the Poincaré sphere at Ib = 0.85 Ic. The measurement uncertainty of SDE by means of FSPA is determined as σSDE,max = ± 3.13% and σSDE,min = ± 3.19% (see Supplement 1).

Download Full Size | PDF

We next provide a further validation of the FSPA by comparing the absorption spectra obtained from the FSPA with the absorption spectra obtained from numerical simulations. For this purpose, the quantum efficiency (QE) needs to be determined and excluded from the measured SDE. We determine the QEs by measuring SDEs as a function of the bias current and normalizing them by the saturated value of their current-related sigmoid-curve fittings [23]. Since literature studies reveal [24,25] that the quantum efficiency depends on the polarization state of the incident photons, QEs for both TE and TM cases are measured (deduced). The results are shown in Fig. 4(a) and 4(b) as a function of the bias current for cases of several different photon wavelength, where the dotted lines correspond to the measured (deduced) data and the solid lines correspond to their sigmoid-curve fittings.

 figure: Fig. 4.

Fig. 4. QE as a function of bias current at different wavelengths for TE (a) and TM (b) polarized photons. (c) The sigmoid-curve extracted (scatters) and power-law fitted (lines) QEs. (d) Deduced absorptance as a function of wavelength at different bias currents.

Download Full Size | PDF

Based on the above results, QEs at three different Ib (0.93 Ic, 0.85 Ic and 0.76 Ic) are plotted as a function of wavelength in Fig. 4(c), where the scatters denote the measured (deduced) QEs and lines denote their wavelength-related power law fittings [26]. The absorptance of the fabricated SNSPD is then obtained by dividing the measured SDE spectra with respect to the QE spectra, i.e. Figure 4(c). The results are presented in Fig. 4(d), where the blue, orange and magenta scatters represent the measured (deduced) absorption spectra of the TE-polarized photons, and the green, violet and cyan dots represent the measured (deduced) absorption spectra of the TM-polarized photons. We note that with the effect of quantum efficiency being excluded, the three absorption spectra at different bias currents coincide well with each other, suggesting that the removal of QE is successful.

The simulated absorption spectrum is also shown in Fig. 4(d) using the red lines for TE polarization and blue lines for TM polarization. In our numerical simulation, the DBR substrate is set to be 13.5 pairs of alternative SiO2 and Ta2O5, with each layer have a quarter-wavelength (at 1550 nm) thickness. The dielectric constants of SiO2 and Ta2O5 are assumed to be 2.08 and 4.33, respectively. The thickness of the ultrathin NbN film is 5.5 nm, and the linewidth of the nanowire is 65 nm with a pitch of 160 nm. The dielectric constant of NbN is -5 + 45i (i.e. nNbN = 4.49 + 5.01i) and is assumed to be constant in the wavelength range from 1490 nm to 1640 nm [27]. Note that to better account for the real situation, we also assume that there exists an airgap between the detector and the fiber end facet (shown in Fig. 2(a)) [18]. In our simulation, the airgap of the SNSPD is chosen as 7.68 µm. Excellent agreement between the measured and simulated absorptance spectra can be identified, suggesting a further validation of the FSPA.

4. Conclusion

In conclusion, we have proposed a four-state polarization algorithm for measuring polarization-dependent SDEs, which requires only four polarization state-related measurements at a given wavelength. The experimental results of the SDE spectra obtained using the FSPA method agree well with the spectra obtained using the traditional method (i.e., sweeping hundreds of points on the Poincaré sphere). The simulated absorptance spectra also coincide well with the measured spectra when the effect of quantum efficiency has been considered. The proposed measurement method is fast and accurate, and is expected to be a useful tool for characterizing the optical performance of SNSPDs.

Funding

the Innovation Program for Quantum Science and Technology (2021ZD0301701); National Key Research and Development Program of China (2017YFA0304002); National Natural Science Foundation of China (11227904, 12033002, 61521001, 61801206, 62071214, 62071218, 62101240); Natural Science Foundation of Jiangsu Province (BK20210177); Key-Area Research and Development Program of Guangdong Province (2020B0303020001); Fundamental Research Funds for the Central Universities; Priority Academic Program Development of Jiangsu Higher Education Institutions; Recruitment Program for Young Professionals; Qinglan Project of Jiangsu Province of China Key Laboratory of Advanced Manipulating Technique of Electromagnetic Waves.

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. G. N. Gol’tsman, O. Okunev, G. Chulkova, A. Lipatov, A. Semenov, K. Smirnov, B. Voronov, A. Dzardanov, C. Williams, and R. Sobolewski, “Picosecond superconducting single-photon optical detector,” Appl. Phys. Lett. 79(6), 705–707 (2001). [CrossRef]  

2. T. Zhong, X. Hu, F. N. C. Wong, K. K. Berggren, T. D. Roberts, and P. Battle, “High-quality fiber-optic polarization entanglement distribution at 1.3 µm telecom wavelength,” Opt. Lett. 35(9), 1392–1394 (2010). [CrossRef]  

3. N. T. Islam, C. C. W. Lim, C. Cahall, J. Kim, and D. J. Gauthier, “Provably secure and high-rate quantum key distribution with time-bin qudits,” Sci. Adv. 3(11), e1701491 (2017). [CrossRef]  

4. H. Li, S. Chen, L. You, W. Meng, Z. Wu, Z. Zhang, K. Tang, L. Zhang, W. Zhang, X. Yang, X. Liu, Z. Wang, and X. Xie, “Superconducting nanowire single photon detector at 532 nm and demonstration in satellite laser ranging,” Opt. Express 24(4), 3535–3542 (2016). [CrossRef]  

5. S. Chen, D. Liu, W. Zhang, L. You, Y. He, W. Zhang, X. Yang, G. Wu, M. Ren, H. Zeng, Z. Wang, X. Xie, and M. Jiang, “Time-of-flight laser ranging and imaging at 1550 nm using low-jitter superconducting nanowire single-photon detection system,” Appl. Opt. 52(14), 3241–3245 (2013). [CrossRef]  

6. F. Marsili, V. B. Verma, J. A. Stern, S. Harrington, A. E. Lita, T. Gerrits, I. Vayshenker, B. Baek, M. D. Shaw, R. P. Mirin, and S. W. Nam, “Detecting single infrared photons with 93% system efficiency,” Nat. Photonics 7(3), 210–214 (2013). [CrossRef]  

7. D. V. Reddy, R. R. Nerem, S. W. Nam, R. P. Mirin, and V. B. Verma, “Superconducting nanowire single-photon detectors with 98% system detection efficiency at 1550 nm,” Optica 7(12), 1649–1653 (2020). [CrossRef]  

8. D. Rosenberg, A. J. Kerman, R. J. Molnar, and E. A. Dauler, “High-speed and high-efficiency superconducting nanowire single photon detector array,” Opt. Express 21(2), 1440–1447 (2013). [CrossRef]  

9. B. Korzh, Q.-Y. Zhao, J. P. Allmaras, S. Frasca, T. M. Autry, E. A. Bersin, A. D. Beyer, R. M. Briggs, B. Bumble, M. Colangelo, G. M. Crouch, A. E. Dane, T. Gerrits, A. E. Lita, F. Marsili, G. Moody, C. Peña, E. Ramirez, J. D. Rezac, N. Sinclair, M. J. Stevens, A. E. Velasco, V. B. Verma, E. E. Wollman, S. Xie, D. Zhu, P. D. Hale, M. Spiropulu, K. L. Silverman, R. P. Mirin, S. W. Nam, A. G. Kozorezov, M. D. Shaw, and K. K. Berggren, “Demonstration of sub-3 ps temporal resolution with a superconducting nanowire single-photon detector,” Nat. Photonics 14(4), 250–255 (2020). [CrossRef]  

10. K. M. Rosfjord, J. K. W. Yang, E. A. Dauler, A. J. Kerman, V. Anant, B. M. Voronov, G. N. Gol’tsman, and K. K. Berggren, “Nanowire single-photon detector with an integrated optical cavity and anti-reflection coating,” Opt. Express 14(2), 527–534 (2006). [CrossRef]  

11. M. Csete, Á Sipos, A. Szalai, F. Najafi, G. Szabó, and K. K. Berggren, “Improvement of infrared single-photon detectors absorptance by integrated plasmonic structures,” Sci. Rep. 3(1), 2406 (2013). [CrossRef]  

12. X. Hu, E. A. Dauler, R. J. Molnar, and K. K. Berggren, “Superconducting nanowire single-photon detectors integrated with optical nano-antennae,” Opt. Express 19(1), 17–31 (2011). [CrossRef]  

13. J. Huang, W. J. Zhang, L. X. You, X. Y. Liu, Q. Guo, Y. Wang, L. Zhang, X. Y. Yang, H. Li, Z. Wang, and X. M. Xie, “Spiral superconducting nanowire single-photon detector with efficiency over 50% at 1550 nm wavelength,” Supercond. Sci. Technol. 30(7), 074004 (2017). [CrossRef]  

14. L. Redaelli, V. Zwiller, E. Monroy, and J. M. Gérard, “Design of polarization-insensitive superconducting single photon detectors with high-index dielectrics,” Supercond. Sci. Technol. 30(3), 035005 (2017). [CrossRef]  

15. R. Xu, F. Zheng, D. Qin, X. Yan, G. Zhu, L. Kang, L. Zhang, X. Jia, X. Tu, B. Jin, W. Xu, J. Chen, and P. Wu, “Demonstration of Polarization-Insensitive Superconducting Nanowire Single-Photon Detector With Si Compensation Layer,” J. Lightwave Technol. 35(21), 4707–4713 (2017). [CrossRef]  

16. Y. Meng, K. Zou, N. Hu, L. Xu, X. Lan, S. Steinhauer, S. Gyger, V. Zwiller, and X. Hu, “Fractal Superconducting Nanowires Detect Infrared Single Photons with 84% System Detection Efficiency, 1.02 Polarization Sensitivity, and 20.8 ps Timing Resolution,” ACS Photonics 9(5), 1547–1553 (2022). [CrossRef]  

17. D. V. Reddy, N. Otrooshi, S. W. Nam, R. P. Mirin, and V. B. Verma, “Broadband polarization insensitivity and high detection efficiency in high-fill-factor superconducting microwire single-photon detectors,” APL Photonics 7(5), 051302 (2022). [CrossRef]  

18. J. Chang, J. W. N. Los, J. O. Tenorio-Pearl, N. Noordzij, R. Gourgues, A. Guardiani, J. R. Zichi, S. F. Pereira, H. P. Urbach, V. Zwiller, S. N. Dorenbos, and I. Esmaeil Zadeh, “Detecting telecom single photons with 99.5−2.07 + 0.5% system detection efficiency and high time resolution,” APL Photonics 6(3), 036114 (2021). [CrossRef]  

19. V. Anant, “Engineering the optical properties of subwavelength devices and materials,” Doctor of Philosophy in Electrical Engineering, Massachusetts Institute of Technology (2007).

20. H. Li, H. Wang, L. You, P. Hu, W. Shen, W. Zhang, X. Yang, L. Zhang, H. Zhou, Z. Wang, and X. Xie, “Multispectral superconducting nanowire single photon detector,” Opt. Express 27(4), 4727–4733 (2019). [CrossRef]  

21. Z. Yang, T. Albrow-Owen, H. Cui, J. Alexander-Webber, F. Gu, X. Wang, T.-C. Wu, M. Zhuge, C. Williams, P. Wang, A. V. Zayats, W. Cai, L. Dai, S. Hofmann, M. Overend, L. Tong, Q. Yang, Z. Sun, and T. Hasan, “Single-nanowire spectrometers,” Science 365(6457), 1017–1020 (2019). [CrossRef]  

22. A. J. Miller, A. E. Lita, B. Calkins, I. Vayshenker, S. M. Gruber, and S. W. Nam, “Compact cryogenic self-aligning fiber-to-detector coupling with losses below one percent,” Opt. Express 19(10), 9102–9110 (2011). [CrossRef]  

23. J. W. N. Chang, R. Los, S. Gourgues, S. N. Steinhauer, S. F. Dorenbos, H. P. Pereira, V. Urbach, I. Zwiller, and Esmaeil Zadeh, “Efficient mid-infrared single-photon detection using superconducting NbTiN nanowires with high time resolution in a Gifford-McMahon cryocooler,” Photon. Res. 10(4), 1063–1070 (2022). [CrossRef]  

24. V. Anant, A. J. Kerman, E. A. Dauler, K. M. Rosfjord, and K. K. Berggren, “Optical properties of superconducting nanowire single-photon detectors,” Opt. Express 16(14), 10750–10761 (2008). [CrossRef]  

25. J. J. Renema, Q. Wang, R. Gaudio, I. Komen, K. o. ‘. Hoog, D. Sahin, A. Schilling, M. P. van Exter, A. Fiore, A. Engel, and M. J. A. de Dood, “Position-Dependent Local Detection Efficiency in a Nanowire Superconducting Single-Photon Detector,” Nano Lett. 15(7), 4541–4545 (2015). [CrossRef]  

26. R. Lusche, A. Semenov, K. Ilin, M. Siegel, Y. Korneeva, A. Trifonov, A. Korneev, G. Goltsman, D. Vodolazov, and H.-W. Hübers, “Effect of the wire width on the intrinsic detection efficiency of superconducting-nanowire single-photon detectors,” J. Appl. Phys. 116(4), 043906 (2014). [CrossRef]  

27. A. Semenov, B. Günther, U. Böttger, H.-W. Hübers, H. Bartolf, A. Engel, A. Schilling, K. Ilin, M. Siegel, R. Schneider, D. Gerthsen, and N. A. Gippius, “Optical and transport properties of ultrathin NbN films and nanostructures,” Phys. Rev. B 80(5), 054510 (2009). [CrossRef]  

Supplementary Material (1)

NameDescription
Supplement 1       Supplemental Document for theory model and uncertainty analysis

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 (4)

Fig. 1.
Fig. 1. Schematics of the polarization-dependent SDE measurement setup.
Fig. 2.
Fig. 2. (a) Schematic of the NbN nanowire fabricated on a DBR substrate. Light is directly coupled to the nanowire region by self-aligning coupling approach. (b) SEM image of the fabricated SNSPD. The nanowire shows a uniform width of 65 ± 2 nm and a pitch of ∼160 nm. (c) Measured SDE (scatters) and sigmoid fitting curve (lines) for TE (red) and TM (blue) polarized wave as a function of normalized bias current at a wavelength of 1550 nm. DCR is indicated by the black dots referring to the right y-axis. (d) Timing jitter of SNSPD at three different bias currents of the measured SNSPD.
Fig. 3.
Fig. 3. (a) Measured SDE as a function of wavelength for cases of three different bias currents by means of FSPA. (b) Comparison of the SDE spectra obtained by FSPA and sweeping 100 points on the Poincaré sphere at Ib = 0.85 Ic. The measurement uncertainty of SDE by means of FSPA is determined as σSDE,max = ± 3.13% and σSDE,min = ± 3.19% (see Supplement 1).
Fig. 4.
Fig. 4. QE as a function of bias current at different wavelengths for TE (a) and TM (b) polarized photons. (c) The sigmoid-curve extracted (scatters) and power-law fitted (lines) QEs. (d) Deduced absorptance as a function of wavelength at different bias currents.

Equations (4)

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

A = a E x 2 + b E y 2 + c E x E y cos ( φ b φ Δ )
A max , min = a + b 2 ± ( a b 2 ) 2 + ( c 2 ) 2
{ A = a A = b   A   = a + b 2 + c 2 cos φ b     A = a + b 2 + c 2 sin φ b  
A max , min = A + A 2 ± ( A A 2 ) 2 + ( A   A + A 2 ) 2 + ( A A + A 2 ) 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.