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

Fabrication and evaluation of high-quality and low-cost quantum random number generators

Open Access Open Access

Abstract

In this paper, we have fabricated two quantum random number generators (QRNGs) based on different mechanisms. The first one is based on the photon time of arrival and produces high-quality random numbers without the need for post-processing but using expensive equipment. The second one is based on the tunneling effect in a Zener diode and produces random strings with comparable quality but using low-cost equipment. We then evaluated the random sequences from these QRNGs using a set of statistical tests and showed that they are suitable for special applications such as quantum technologies.

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

1. Introduction

Random Number Generators (RNGs) are an essential ingredient for many applications such as telecommunications [13], cryptography [4,5], simulations [6,7], games [8], and quantum technologies such as Quantum Key Distribution [911], quantum imaging [12,13], and quantum radar [14,15], etc. There are two general types of RNGs based on their mechanism: pseudo-random number generators (PRNGs) which use mathematical algorithms to generate random numbers from an initial seed [16,17], and true random number generators (TRNGs) which extracts random numbers from a physical process [18,19]. There exists a subset of TRNGs that employs quantum phenomena to generate random sequences known as Quantum random number generators (QRNGs) [20]. Although PRNGs produce high-quality random numbers, they are inherently predictable [21]. However, QRNGs produce true random sequences that are unpredictable and incomputable, thus suitable for applications that need unpredictable random numbers such as Quantum Key Distribution, quantum imaging, and quantum radar [22]. There are different types of entropy sources for QRNGs including radioactive decay [23], the quantum mechanical noise in electronic circuits known as shot noise [24], photon arrival times [25,26], quantum vacuum fluctuations [2729], laser phase fluctuations [30], optical parametric oscillators [31], and amplified spontaneous emission [32], etc.

The Iranian Center for Quantum Technologies (ICQTs) is a newly established institute which conducts researches in quantum technologies [3335]. Considering ongoing projects (including QKD and quantum imaging), high-quality and low-cost QRNGs were needed. Thus using available equipment, we built two QRNGs using different approaches. We also gathered a comprehensive set of statistical tests [22] and applied them to these QRNGs. In this work, we realized a QRNG based on photon time of arrival which produces high-quality random numbers and does not need post-processing. Like most QRNGs this one is complex and expensive, thus to satisfy our needs we also realized a simple and inexpensive QRNG based on Zener diode which also produces high-quality random numbers. We compared these two QRNGs and showed that the output from the Zener diode QRNG is very close to that of the single-photon QRNG considering its cost.

This paper is organized as follows. In Section 2. we describe the physics behind the randomness and the experimental setup of two different QRNGs realized in this work. In Section 3. we bring the results and compare the QRNGs. Finally, we give a conclusion in Section 4.

2. Materials and methods

2.1 Single photon QRNG

The wave function of a highly attenuated laser state (coherent state) can be represented as a product of all coherent states in time within the coherence time

$$\left| \phi \right\rangle = \bigotimes^{n_b}_{t=1} \left| \alpha_i \right\rangle$$
where $n_b$ is the number of time bins within the coherence time of the laser. We can rewrite the eq 1 in terms of photons generated at different time instances as
$$\left| \phi \right\rangle = \sum^{\infty}_{k=0} \sqrt{P_k} \left( \frac{\sum^{n_b}_{t=1} a^{{\dagger}}_t}{\sqrt{n_b}} \right) \left| 0 \right\rangle$$
where $P_k$ is the Poisson distribution with mean photon number $n_b \mu$ and $\mu$ is the mean photon number per time bin. If there are multiple photons from the source then each photon will be in a superposition of all time states within the coherence time of the laser. The overall wave function is the product state of individual photon states. The collapse of this wave function occurs randomly in these time states. This provides the necessary randomness in QRNG.

The experimental setup (see Fig. 1(a) for schematic and Fig. 1(b) for actual photo) consists of a DFB laser with 1310 nm wavelength and 5 mW output power operating at continuous mode, two variable attenuators, a silicon SPAD, and a time tagger with a resolution of 81 ps. One attenuator is kept fixed while the other one is changed to achieve different attenuation levels.

 figure: Fig. 1.

Fig. 1. Schematic (a) and actual photo (b) of the Single Photon QRNG experimental setup.

Download Full Size | PDF

For extracting the raw data from photon time of arrival, we divided the time segment $T$ into different bin numbers of 4, 8, 16, 32, 64, 128, and 256. We also used different attenuation levels to achieve photon flux of 10, 50, 100, 200, 500, and 1000 kcps and compared the randomness quality of their outputs to find the optimal photon flux. The raw bit rate depends on the number of time bins, photon flux, and extraction method. In our case, the maximum achievable raw bit rate is 8 Mbps.

In the end, we analyzed the output random sequences of length 50 Mb and 1 Gb, using 5 statistical tests including Dieharder [36], NIST SP-800-22 [37], NIST SP-800-90B [38], ENT [39], and Borel normality [40,41].

2.2 Zener diode QRNG

A Zener diode is a semiconductor device consisting of a highly doped p-n junction that allows current to flow if a certain reverse voltage -known as Zener voltage- is applied. While at higher voltages the avalanche breakdown occurs, at lower voltages the reverse conduction occurs due to electron quantum tunneling known as the Zener effect. Thus, if one applies the Zener voltage to a Zener diode, the current fluctuates randomly as the quantum tunneling is a random process [20].

In this paper, we realized a simple, compact, and low-cost QRNG based on Zener diode as an alternative for expensive single photon QRNG. The experimental setup (Fig. 2) consists of a Zener diode, a common emitter amplifier to enhance the quantum noise and an 8-bit Analog-to-Digital Converter (ADC) with a sampling rate of 25 MSa/s. The raw data is sampled by ADC and sent to the computer for post-processing. The SHA-256 algorithm is used to extract the final random data. The raw bit rate is determined by the ADC speed and raw bit extraction method. In our case, the raw bit rate is about 0.5 Mbps. This can be improved to a few Mbps by changing the extraction parameters. For investigating the quality of the generated output, the same 5 statistical tests were applied to the random sequences of length 50 Mb and 1 Gb.

 figure: Fig. 2.

Fig. 2. Schematic of the Zener diode QRNG.

Download Full Size | PDF

3. Results and discussion

3.1 Single photon QRNG

First, we investigated the effect of changing the number of time bins on the quality of output sequences at a constant photon flux of 50 kcps. As one can see in Fig. 3, the min-entropy is almost constant and independent of the number of bins. Thus, we should look for other quantities to determine the optimal bin number.

 figure: Fig. 3.

Fig. 3. Min-entropy of the Single Photon QRNG at different number of time bins.

Download Full Size | PDF

We used two quantities from the ENT test as criteria for choosing the optimal bin number, namely error in Monte Carlo value for Pi, and Serial correlation coefficient. The Monte Carlo value for Pi is a measure of performance and Serial correlation coefficient shows any correlations [22,42]. Thus the ideal value for Pi value error and correlation coefficient is zero. As it can be seen in Fig. 4, the minimum error for Pi value (Fig. 4(a)) occurs at 32 and 64 bins, but the correlation coefficient (Fig. 4(b)) of 32 bins is far from the zero line. Thus, we choose 64 as the optimal number of time bins. We repeated this for the lowest and highest photon flux of 10 kcps and 1000kcps respectively, and again we found the 64 as the optimal number of bins.

 figure: Fig. 4.

Fig. 4. Monte Carlo value for Pi error (a) and the Serial correlation coefficient (b) of the Single Photon QRNG at the different number of time bins.

Download Full Size | PDF

Next, we investigated the photon flux on the quality of output sequences. As we can see in Fig. 5 the minimum Pi value error (Fig. 5(a)) occurs at photon flux of 50 kcps and its correlation coefficient (Fig. 5(b)) is relatively low, Thus we choose 50 kcps as the optimal photon flux.

 figure: Fig. 5.

Fig. 5. Monte Carlo value for Pi error (a) and the Serial correlation coefficient (b) of the Single Photon QRNG at the different intensities (photon counts).

Download Full Size | PDF

3.2 Zener diode QRNG

As one can see in Fig. 6 when the Zener diode is off, there exists a classical noise, and when the Zener diode is on, the quantum noise is dominant with a Signal to Noise Ratio (SNR) of about 9 dB.

 figure: Fig. 6.

Fig. 6. Quantum and classical noised in the Zener diode QRNG. The Signal to Noise Ratio (SNR) is about 9 dB.

Download Full Size | PDF

The signal voltage distribution histogram is plotted in Fig. 7 and it can be seen that it follows a symmetric normal distribution.

 figure: Fig. 7.

Fig. 7. The distribution histogram of the quantum signal obtained from the Zener diode QRNG after amplification, where a normal distribution curve is fitted.

Download Full Size | PDF

3.3 Statistical tests

Now that we have two different QRNGs, we evaluate their output random numbers using various statistical tests. The first test is Dieharder, and its results are presented in Table 1.

Tables Icon

Table 1. Results of Dieharder tests for random sequences with 1Gb length.

Dieharder is a hard test to pass and needs large amounts of random data, nonetheless both QRNGs passed it. The next famous test suite is NIST SP800-22 whose results are listed in Table 2. Here we used the significance level of $\alpha =0.01$ and for tests with several P-values, we choose the worst result. Each test is considered passed if the P-value falls in the range $0.01-0.99$.

Tables Icon

Table 2. Results of NIST SP800-22 test suite for random sequences with 50Mb length and significance level of $\alpha =0.01$.

As one can see both QRNGs passed the NIST SP800-22 tests. The outputs from both QRNGs also passed the NIST SP800-90B test. Next, we perform the ENT test for optimal configurations. Its results are presented in Table 3.

Tables Icon

Table 3. The ENT test results.

As one can see both QRNGs are comparable (or even better) than commercial QRNGs [22], and the results are very close to ideal values. At last, we perform the Borel normality test on the output of our QRNGs to show that they have the necessary condition for algorithmic randomness. The metric value from the Borel normality test and results are presented in Table 4.

Tables Icon

Table 4. The Borel normality test results.

As it can be seen from the above results the single-photon QRNG shows a great performance which is very promising for use in QKD setups. The Zener diode QRNG as an inexpensive and simple alternative also shows good performance with close results to the results of the single-photon QRNG.

4. Conclusion

In conclusion, considering our needs for random number sources, we realized two QRNGs; a high quality but complicated and expensive one based on single-photon time of arrival, and a simple and inexpensive one based on quantum tunneling in a Zener diode. By performing five different statistical tests on these two QRNGs we showed that these two QRNGs are comparable to commercial QRNGs and can be used for special applications such as quantum technologies. As the Zener diode QRNG showed a great performance (considering its cost) we decided to develop it using an FPGA to further increase the number generation rate and automate the post-processing.

Acknowledgments

We thank the Iranian Center for Quantum Technologies (ICQTs) and all colleagues who helped us in this work.

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.

References

1. H. C. van Tilborg, “Shannon’s model,” in Encyclopedia of Cryptography and Security, (Springer, 2011), pp. 1194–1195.

2. D. Torrieri, Principles of spread-spectrum communication systems, vol. 1 (Springer, 2005).

3. H.-N. Hung, P.-C. Lee, and Y.-B. Lin, “Random number generation for excess life of mobile user residence time,” IEEE Trans. Veh. Technol. 55(3), 1045–1050 (2006). [CrossRef]  

4. M. Stipcevic, “Quantum random number generators and their applications in cryptography,” in Advanced Photon Counting Techniques VI, vol. 8375 (International Society for Optics and Photonics, 2012), p. 837504.

5. A. J. Acosta, T. Addabbo, and E. Tena-Sánchez, “Embedded electronic circuits for cryptography, hardware security and true random number generation: an overview,” Int. J. Circ. Theor. Appl. 45(2), 145–169 (2017). [CrossRef]  

6. N. Metropolis and S. Ulam, “The Monte Carlo method,” J. Am. Stat. Assoc. 44(247), 335–341 (1949). PMID: 18139350. [CrossRef]  

7. R. Motwani and P. Raghavan, “Randomized algorithms,” ACM Comput. Surv. 28(1), 33–37 (1996). [CrossRef]  

8. M. Du, Q. Chen, L. Liu, and X. Ma, “A blockchain-based random number generation algorithm and the application in blockchain games,” in 2019 IEEE International Conference on Systems, Man and Cybernetics (SMC), (IEEE, 2019), pp. 3498–3503.

9. N. Gisin, G. Ribordy, W. Tittel, and H. Zbinden, “Quantum cryptography,” Rev. Mod. Phys. 74(1), 145–195 (2002). [CrossRef]  

10. N. Walenta, M. Soucarros, D. Stucki, D. Caselunghe, M. Domergue, M. Hagerman, R. Hart, D. Hayford, R. Houlmann, and M. Legré, “Practical aspects of security certification for commercial quantum technologies,” Proc. SPIE 9648, 199–209 (2015). [CrossRef]  

11. A. Martin, B. Sanguinetti, C. C. W. Lim, R. Houlmann, and H. Zbinden, “Quantum random number generation for 1.25-ghz quantum key distribution systems,” J. Lightwave Technol. 33(13), 2855–2859 (2015). [CrossRef]  

12. G. Brida, M. Genovese, and I. R. Berchera, “Experimental realization of sub-shot-noise quantum imaging,” Nat. Photonics 4(4), 227–230 (2010). [CrossRef]  

13. O. S. Magaña-Loaiza and R. W. Boyd, “Quantum imaging and information,” Rep. Prog. Phys. 82(12), 124401 (2019). [CrossRef]  

14. M. Lanzagorta, “Quantum radar,” Synth. Lect. on Quantum Comput. 3(1), 1–139 (2011). [CrossRef]  

15. D. Luong, B. Balaji, C. W. S. Chang, V. M. A. Rao, and C. Wilson, “Microwave quantum radar: An experimental validation,” in 2018 International Carnahan Conference on Security Technology (ICCST), (IEEE, 2018), pp. 1–5.

16. L. Blum, M. Blum, and M. Shub, “A simple unpredictable pseudo-random number generator,” SIAM J. Comput. 15(2), 364–383 (1986). [CrossRef]  

17. F. James, “A review of pseudorandom number generators,” Comput. Phys. Commun. 60(3), 329–344 (1990). [CrossRef]  

18. B. Barak, R. Shaltiel, and E. Tromer, “True random number generators secure in a changing environment,” in International Workshop on Cryptographic Hardware and Embedded Systems, (Springer, 2003), pp. 166–180.

19. M. Stipčević and Ç. K. Koç, “True random number generators,” in Open Problems in Mathematics and Computational Science, (Springer, 2014), pp. 275–315.

20. M. Herrero-Collantes and J. C. Garcia-Escartin, “Quantum random number generators,” Rev. Mod. Phys. 89(1), 015004 (2017). [CrossRef]  

21. J. T. Kavulich, B. P. Van Deren, and M. Schlosshauer, “Searching for evidence of algorithmic randomness and incomputability in the output of quantum random number generators,” Phys. Lett. A 388, 127032 (2021). [CrossRef]  

22. S. Hajibaba, A. Dadahkhani, and S. A. Madani, “Verifying quantum random number generators using a comprehensive set of statistical tests,” Manuscript submitted for publication.

23. H. Schmidt, “Quantum-mechanical random-number generator,” J. Appl. Phys. 41(2), 462–468 (1970). [CrossRef]  

24. Y. Shen, L. Tian, and H. Zou, “Practical quantum random number generator based on measuring the shot noise of vacuum states,” Phys. Rev. A 81(6), 063814 (2010). [CrossRef]  

25. M. A. Wayne, E. R. Jeffrey, G. M. Akselrod, and P. G. Kwiat, “Photon arrival time quantum random number generation,” J. Mod. Opt. 56(4), 516–522 (2009). [CrossRef]  

26. M. Stipčević and B. M. Rogina, “Quantum random number generator based on photonic emission in semiconductors,” Rev. Sci. Instrum. 78(4), 045104 (2007). [CrossRef]  

27. C. Gabriel, C. Wittmann, D. Sych, R. Dong, W. Mauerer, U. L. Andersen, C. Marquardt, and G. Leuchs, “A generator for unique quantum random numbers based on vacuum states,” Nat. Photonics 4(10), 711–715 (2010). [CrossRef]  

28. T. Symul, S. Assad, and P. K. Lam, “Real time demonstration of high bitrate quantum random number generation with coherent laser light,” Appl. Phys. Lett. 98(23), 231103 (2011). [CrossRef]  

29. J.-Y. Haw, S. Assad, A. Lance, N. Ng, V. Sharma, P. K. Lam, and T. Symul, “Maximization of extractable randomness in a quantum random-number generator,” Phys. Rev. Appl. 3(5), 054004 (2015). [CrossRef]  

30. H. Guo, W. Tang, Y. Liu, and W. Wei, “Truly random number generation based on measurement of phase noise of a laser,” Phys. Rev. E 81(5), 051137 (2010). [CrossRef]  

31. A. Marandi, N. C. Leindecker, K. L. Vodopyanov, and R. L. Byer, “All-optical quantum random bit generation from intrinsically binary phase of parametric oscillators,” Opt. Express 20(17), 19322–19330 (2012). [CrossRef]  

32. C. R. Williams, J. C. Salevan, X. Li, R. Roy, and T. E. Murphy, “Fast physical random number generator using amplified spontaneous emission,” Opt. Express 18(23), 23584–23597 (2010). [CrossRef]  

33. A. Motazedifard, S. A. Madani, and N. Vayaghan, “Measurement of entropy and quantum coherence properties of two type-I entangled photonic qubits,” Opt. Quantum Electron. 53(1), 1–26 (2021). [CrossRef]  

34. A. Motazedifard, S. A. Madani, J. J. Dashkasan, and N. S. Vayaghan, “Nonlocal realism tests and quantum state tomography in Sagnac-based type-II polarization-entanglement SPDC-source,” Heliyon 7(6), e07384 (2021). [CrossRef]  

35. A. Motazedifard and S. A. Madani, “High-precision quantum transmittometry of DNA and methylene-blue using a frequency-entangled twin-photon beam in type-I SPDC,” OSA Continuum 4(3), 1049–1069 (2021). [CrossRef]  

36. https://webhome.phy.duke.edu/rgb/General/dieharder.php.

37. A. Rukhin, J. Soto, J. Nechvatal, M. Smid, and E. Barker, “A statistical test suite for random and pseudorandom number generators for cryptographic applications,” Tech. rep., Booz-Allen Hamilton, Inc., McLean, VA, USA (2001).

38. https://doi.org/10.6028/NIST.SP.800-90B.

39. https://www.fourmilab.ch/random/.

40. C. S. Calude, Information and randomness: an algorithmic perspective (Springer Science & Business Media, 2002).

41. M. Émile Borel, “Les probabilités dénombrables et leurs applications arithmétiques,” Rend. Circ. Matem. Palermo 27(1), 247–271 (1909). [CrossRef]  

42. D. E. Knuth, Art of computer programming, volume 2: Seminumerical algorithms (Addison-Wesley Professional, 2014).

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 (7)

Fig. 1.
Fig. 1. Schematic (a) and actual photo (b) of the Single Photon QRNG experimental setup.
Fig. 2.
Fig. 2. Schematic of the Zener diode QRNG.
Fig. 3.
Fig. 3. Min-entropy of the Single Photon QRNG at different number of time bins.
Fig. 4.
Fig. 4. Monte Carlo value for Pi error (a) and the Serial correlation coefficient (b) of the Single Photon QRNG at the different number of time bins.
Fig. 5.
Fig. 5. Monte Carlo value for Pi error (a) and the Serial correlation coefficient (b) of the Single Photon QRNG at the different intensities (photon counts).
Fig. 6.
Fig. 6. Quantum and classical noised in the Zener diode QRNG. The Signal to Noise Ratio (SNR) is about 9 dB.
Fig. 7.
Fig. 7. The distribution histogram of the quantum signal obtained from the Zener diode QRNG after amplification, where a normal distribution curve is fitted.

Tables (4)

Tables Icon

Table 1. Results of Dieharder tests for random sequences with 1Gb length.

Tables Icon

Table 2. Results of NIST SP800-22 test suite for random sequences with 50Mb length and significance level of α = 0.01 .

Tables Icon

Table 3. The ENT test results.

Tables Icon

Table 4. The Borel normality test results.

Equations (2)

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

| ϕ = t = 1 n b | α i
| ϕ = k = 0 P k ( t = 1 n b a t n b ) | 0
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.