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

Integrated InP optical unitary converter with compact half-integer multimode interferometers

Open Access Open Access

Abstract

Integrated optical unitary converters (OUCs) are vital devices for various emerging applications such as mode-multiplexed optical communication, optical neural networks, and quantum computing. In order to realize large-scale OUCs in a limited footprint, the number of elements, as well as the size of each element, is important. In this work, we present a novel type of OUC using half-integer multimode interferometers (MMIs) based on the multi-plane light conversion (MPLC) concept. A half-integer MMI enables unitary coupling among the multiple input and output ports, while requiring only half the length of a conventional uniform MMI. Although the splitting ratio is not uniform across the ports, we show both numerically and experimentally that arbitrary unitary operation can still be achieved with comparable performance. We fabricate 4×4 OUC with half-integer MMIs on the monolithic InP platform and experimentally demonstrate reconfigurable 4-mode sorting and switching with a significantly reduced footprint compared with the conventional OUCs using uniform MMIs.

© 2021 Optical Society of America under the terms of the OSA Open Access Publishing Agreement

1. Introduction

Optical unitary converters (OUCs) can convert mutually orthogonal input optical modes into arbitrary output modes. OUCs are attracting increasing interest in various application areas, including optical communication [14], neural networks [58], and quantum computing [911]. To this end, various types of integrated OUCs have been demonstrated on the Si [25,1013], silica [9,14], and InP [1517] photonic integration platforms. The conventional architecture is based on the meshed Mach-Zehnder interferometers (MZIs) [18,19], where arbitrary transformation is achieved by configuring the phase shifters. These MZI-based OUCs, however, face an issue that each MZI needs to operate with precise 50:50 splitters [6,20]. This severe fabrication tolerance may limit the scaling of these OUCs, especially on the InP platforms, where integration of compact and accurate MZIs are challenging. On the other hand, the monolithic InP platform provides a unique advantage of integrating various active and nonlinear optical components, such as lasers, semiconductor optical amplifiers (SOAs), and high-speed photodetectors (PDs), with minimal insertion losses [2123].

To solve this problem, we have recently demonstrated an integrated monolithic InP OUC with cascaded N×N mode-mixing layers and phase shifter arrays [17]. The basic mechanism relies on the concept of multi-plane light conversion (MPLC) [24], which has originally been demonstrated using free-space optics [25,26]. In our photonic integrated implementation, the device consists of N-port multimode interferometers (MMIs) and phase shifter arrays arranged in N stages, which could be fabricated with a significantly less complexity even on InP, compared with the MZI-based scheme, requiring dense integration of N(N-1)/2 MZIs.

Here, we propose and experimentally demonstrate a novel type of OUC that can further reduce the footprint and complexity of the MPLC-based InP OUC. We utilize the unique redundancy of the MPLC method that the mode-mixing layer may be implemented by other types of linear device, provided that its operation is described by a sufficiently dense unitary matrix [12,2426]. Therefore, instead of using a uniform MMI as in the previous work [4,17,27], we propose to employ a so-called half-integer MMI or an over-lapping-image MMI [28,29], which provides non-uniform splitting among the ports, but still is a unitary device. Since this half-integer MMI has half the length of the uniform MMI, the footprint of the mixing layer can be reduced by a factor of two without sacrificing the OUC performance. For proof of concept, we fabricate a 4×4 OUC with the half-integer MMIs on InP and experimentally demonstrate reconfigurable unitary conversion.

2. Principle of the MPLC-based OUC using MMIs

The MPLC-based OUC using MMIs is illustrated schematically in Fig. 1(a). It consists of cascaded phase shifter arrays and N-port MMIs. The matrix U describing the transformation of the complex amplitudes from N input ports to N output ports of the entire OUC chip is expressed as

$${\mathbf U} = {{\mathbf \Phi }_{\boldsymbol K}} \cdot {\mathbf M} \cdot {{\mathbf \Phi }_{{\boldsymbol K} - 1}} \cdots {\mathbf M} \cdot {{\mathbf \Phi }_0}$$
where M is the fixed transfer matrix of N×N MMIs, ${{\mathbf \Phi }_i}$ is the diagonal matrix representing the phase shift applied at the i-th phase shifter stage, and K is the total number of MMIs.

It is important to note that the unitary matrix M, responsible for scrambling the complex amplitudes, does not necessarily have to be a specific transformation such as Fourier transform or uniform N×N splitter as demonstrated in the previous work [4,17,27]. Instead, it is only required to be a sufficiently dense unitary transformation [12,2426]. Therefore, we expect that the uniform N-port MMI used in the previous work may be replaced with a much shorter MMI, as long as it provides sufficient coupling among all N ports.

Figure 1(b) describes a schematic of a general N × N MMI, where we define x and z coordinates as shown in the figure and W is the effective MMI width. For simplicity of explanation, we assume N to be an even number, but a similar result can be derived for the odd case as well [28,29]. The position of the input and output ports are expressed as ${x_m} \equiv W({2m - 1} )/2N$ ($m = 1,\; 2,\; \ldots ,\; N$). We define ${L_0} \equiv 4{n_{eff}}{W^2}/\lambda $ to represent the self-imaging length of the MMI, where ${n_{eff}}$ and $\lambda $ are the effective refractive index and the wavelength, respectively. At $z = {L_0}$, input optical field $E({x,0} )$ is reproduced within the range of $0 \le x \le W$, corresponding to the real MMI region [29]. More generally, at $z = {L_0}/q$, where q is a positive integer, the field is represented as [29]

$$E\left( {x,\frac{{{L_0}}}{q}} \right) = \frac{1}{C}\mathop \sum \limits_{m = 0}^{q - 1} {e^{i{\varphi _m}}}\left[ {E\left( {x - m\frac{{2W}}{q},0} \right) - E\left( {2W - x - m\frac{{2W}}{q},0} \right)} \right],$$
where C is a constant and ${\varphi _m}$ is the phase shift. The first term in the summation describes the q evenly spaced copies from the real field, whereas the second term is those from the virtual field in the extended section of the MMI.

 figure: Fig. 1.

Fig. 1. (a) Schematic of the MPLC-based OUC using cascaded MMIs. (b) Schematic of general N×N MMI.

Download Full Size | PDF

In the conventional N × N MMIs, where the input light from all N ports is split equally among the N output ports, the MMI length is set to L0/N, so that q = N. In this case, two terms inside the summation of Eq. (2) do not overlap and constitutes N equally spaced copies at the output ports. On the other hand, if we do not have to split the power uniformly, the MMI length can be reduced to half by setting q = 2N. In this case, two terms in Eq. (2) overlap at the output ports and interfere to generate non-uniform power distributions [28,29]. Nevertheless, the total power from all N outputs is constant so that we can achieve unitary coupling, which is exactly what we need in this work. We can therefore expect that the MMI length can be reduced to half without causing essential penalty to the OUC performance.

3. Numerical simulation

To confirm these characteristics, first, we numerically simulate wave propagation inside an MMI by the eigenmode expansion (EME) method [30]. Figures 2(a) and 2(b) show the top and cross-sectional designs of the 4×4 MMI on InP, considered in this work. Figures 2(c) and 2(d) show the transmittance to all output ports for various MMI length L, simulated at 1550-nm wavelength. We can see that when L = 252 µm and 504 µm, the light is split without excess loss, corresponding respectively to the non-uniform and uniform unitary splitters. Figures 2(e)-(h) show the field distribution inside the MMI for L = 504 µm [Fig. 2(e) and (f)] and 252 µm [Fig. 2(g) and (h)]. We can confirm that when L = 252 µm [Fig. 2(g) and (h)], although the input power is split non-uniformly, all power is coupled to the four output ports without loss. In this case, the MMI is a half-integer MMI and its footprint can be reduced to half compared with the uniform MMI.

 figure: Fig. 2.

Fig. 2. Numerical analyses of 4×4 MMI on InP. (a) Top and (b) cross-sectional structures. (c, d) The transmittance to all output ports (Out 1, 2, 3, 4) and their sum (Total) when the light is input from port 1 (c) and port 2 (d), calculated for various MMI length L. At L = 252 µm and 504 µm, the light is split without excess loss, corresponding respectively to the non-uniform and uniform splitters. (e-h) The wave propagation inside the MMI for L = 504 µm (e, f) and 252 µm (g, h) from the input port 1 (e, g) and port 2 (f, h).

Download Full Size | PDF

We then simulate the performance of the entire 4×4 OUC as shown in Fig. 1(a) with the uniform and half-integer MMIs, designed in Fig. 2. We set a random 4×4 complex unitary matrix as the target matrix ${\mathbf U^{\prime}}$ that we want to realize. Then, all phase shifters in the OUC are optimized through the simulated annealing algorithm, so that ${\mathbf U}$ in Eq. (1) approaches ${\mathbf U^{\prime}}$. More specifically, we minimize the mean-squared error (MSE), defined as

$$\textrm{MSE} = \frac{1}{{{N^2}}}\mathop \sum \nolimits_{i = 1,\; \; j = 1}^{N,\; \; N} {|{{U_{ij}} - U_{ij}^{\prime}} |^2},$$
where ${U_{ij}}$ and $U_{ij}^{\prime}$ are the matrix components of ${\mathbf U}$ and ${\mathbf U}{^{\prime}}$, respectively. Figure 3 shows the MSEs after optimizing all phase shifters for different number of stages K. The bars represent the distributed results for 100 different Haar random 4×4 unitary matrices [31], tested as the target matrix ${\mathbf U^{\prime}}$. We can confirm that the performance of the OUC is comparable for both types of MMIs. By using the half-integer MMIs, therefore, we can reduce the footprint of the mixing layer by a factor of two without sacrificing the performance.

 figure: Fig. 3.

Fig. 3. Simulated performance of 4 × 4 OUC with uniform (blue) and half-integer (red) MMIs. Average mean-square error (MSE) of the obtained unitary operation after optimizing the OUCs with different number of stages K. The bars represent the distributed results for 100 random unitary matrixes, tested as the target unitary operations.

Download Full Size | PDF

4. Experimental results

To investigate the validity of the proposed concept experimentally, we fabricated a proof-of-concept 4 × 4 OUC on InP and tested its performance in several mode-sorting scenarios as the example cases of unitary operation.

4.1 Device fabrication

A 4×4 OUC with K = 4 and the half-integer MMI designed in Fig. 2 was fabricated on InP for experimental demonstration. We employed 800-µm-long p-i-n InP/InGaAsP phase shifters, operated under carrier injection to induce efficient phase shift through the band-filling and free-carrier plasma effects in InGaAsP. The designed OUC was fabricated by electron-beam lithography and inductively-coupled-plasma reactive-ion etching followed by SiO2 deposition. The electrodes of the phase shifters were patterned by evaporating Ti/Au and lift-off process. Figure 4 shows the micrograph and scanning electron microscope (SEM) images of the fabricated OUC. The entire circuit (excluding the electrode pads) fits inside a 2×13 mm2 footprint.

 figure: Fig. 4.

Fig. 4. (a) Fabricated OUC using half-integer MMI on InP platform. (b-d) The SEM image of (b) cross-sectional image at passive waveguide, (c) top-view of the 4×4 MMI, and (d) cross-sectional image of the phase shifter.

Download Full Size | PDF

4.2 Measurement

The experimental setup and procedures were similar to those reported in [4]. In addition, we employed silica-based pitch converters to reduce the array pitch from 127 μm at the fiber array to 30 μm at the InP chip for efficient coupling [32]. A microcontroller was employed to sequentially select one of the four input ports and to collect the signals from four photodetectors to measure the transmittance to all output ports. After convergence, all phase shifters were set to the optimized condition. A continuous-wave light at 1550 nm wavelength and transverse-electric (TE) polarization mode was employed. The total fiber-to-fiber loss was measured to be around 35 dB, which consists of the coupling loss of 9.5 dB/facet, the propagation loss of 5.8 dB/cm, excess loss at the MMI of 1.5 dB/interferometer, and additional loss of around 2 dB including the phase shifter insertion loss.

Figure 5(a) shows the measured transmittance from all input ports to all output ports before tuning the phase shifters. Due to the strong coupling inside the chip, severe crosstalk is observed at the output ports. In contrast, Fig. 5(b) shows the result when the phase shifters are set to switch {In1, In2, In3, In4} to {Out 1, Out 2, Out3, Out4}. Similarly, Fig. 5(c)–5(f) show the results under different configurations. The crosstalk and MSE are suppressed below -5.9 dB and -14.6 dB in all cases, respectively. The discrepancy from the numerical simulation is attributed to the large insertion loss and the insufficient resolution of the driver circuit, which can be improved in future. We should note that the obtained performance in this work is comparable to the previous work using the uniform MMIs [17].

 figure: Fig. 5.

Fig. 5. Measured transmittance at the wavelength 1550 nm from each input and output port (a) before optimization, after optimization from {In1, In2, In3, In4} to (b) {Out1, Out2, Out3, Out4}, (c) {Out1, Out3, Out2, Out4}, (d) {Out1, Out4, Out2, Out3}, (e) {Out2, Out1, Out3, Out4}, and (f) {Ou41, Out3, Out2, Out1}. The crosstalk and MSE after optimization in the respective cases are (b) (-6.1 dB, -15.8 dB), (c) (-6.1 dB, -14.9 dB), (d) (-6.1 dB, -14.8 dB), (e) (-6.2 dB, -14.7 dB), and (f) (-5.9 dB, -14.6 dB).

Download Full Size | PDF

5. Conclusion

In conclusion, we have numerically and experimentally demonstrated a novel reconfigurable OUC using half-integer MMIs and phase shifter arrays on InP. By optimizing the phase shifters, we experimentally realized 4-mode sorting as well as mode switching. Since the MPLC method only requires a dense unitary matrix, the OUC with half-integer MMIs allows us to shrink the MMI size by half without inducing a fundamental penalty to the OUC performance.

Funding

Japan Society for the Promotion of Science (JP20J221861, JP21K1816).

Acknowledgments

Portions of the proposed were presented at 24th OptoElectronics and Communications Conference (OECC) and the International Conference on Photonics in Switching and Computing (PSC) in 2019, “Compact reconfigurable optical unitary converter based on non-uniform multimode interference coupler” (paper TuF3-1). We would like to thank Eisaku Kato and Shun Takahashi for the technical support.

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. D. A. B. Miller, “Self-configuring universal linear optical component,” Photonics Res. 1(1), 1–15 (2013). [CrossRef]  

2. N. K. Fontaine, C. R. Doerr, M. A. Mestre, R. R. Ryf, P. J. Winzer, L. L. Buhl, Y. Sun, X. Jiang, and R. Lingle Jr, “Space-division multiplexing and all-optical MIMO demultiplexing using a photonic integrated circuit,” in Opt. Fiber Commun. Conf. (OFC) (Optical Society of America, 2012), paper PDP5B.1.

3. A. Annoni, E. Guglielmi, M. Cartminati, G. Ferrari, M. Sampietro, D. A. B. Miller, A. Melloni, and F. Morichetti, “Unscrambling light - automatically undoing strong mixing between modes,” Light Sci Appl 6(12), e17110 (2017). [CrossRef]  

4. R. Tang, T. Tanemura, and Y. Nakano, “Reconfigurable all-optical on-chip MIMO 3-mode demultiplexing based on multi-plane light conversion,” Opt. Lett. 43(8), 1798–1801 (2018). [CrossRef]  

5. Y. Shen, N. C. Harris, S. Skirlo, M. Prabhu, T. B. Jones, M. Hochberg, X. Sun, S. Zhao, H. Larochelle, D. Englund, and M. Soljačić, “Deep learning with coherent nanophotonic circuits,” Nat. Photonics 11(7), 441–446 (2017). [CrossRef]  

6. M. Y. S. Fang, S. Manipatruni, C. Wierzynski, A. Khosrowshahi, and M. R. DeWeese, “Design of optical neural networks with component imprecisions,” Opt. Express 27(10), 14009–14029 (2019). [CrossRef]  

7. T. W. Hughes, M. Minkov, Y. Shi, and S. Fan, “Training of photonic neural networks through in situ backpropagation,” Optica 5(7), 864–871 (2018). [CrossRef]  

8. S. Pai, I. A. D. Williamson, T. W. Hughes, M. Minkov, O. Solgaard, S. Fan, and D. A. B. Miller, “Parallel programming of an arbitrary feedforward photonic network,” IEEE J. Sel. Top. Quantum Electron. 26(5), 6100813 (2020). [CrossRef]  

9. J. Carolan, C. Harrold, C. Sparrow, E. M. López, N. J. Russell, J. W. Silverstone, P. J. Shadbolt, N. Matsuda, M. Oguma, M. Itoh, G. D. Marshall, M. G. Thompson, J. C. F. Matthews, T. Hashimoto, J. L. O’Brien, and A. Laing, “Universal linear optics,” Science 349(6249), 711–716 (2015). [CrossRef]  

10. N. C. Harris, G. R. Steinbrecher, M. Prabhu, Y. Lahini, J. Mower, D. Bunandar, C. Chen, F. N. C. Wong, T. B. Jones, M. Hochberg, S. Lloyd, and D. Englund, “Quantum transport simulations in a programmable nanophotonic processor,” Nat. Photonics 11(7), 447–452 (2017). [CrossRef]  

11. J. M. Arrazola, V. Bergholm, K. Brádler, T. R. Bromley, M. J. Collins, I. Dhand, A. Fumagalli, T. Gerrits, A. Goussev, L. G. Helt, J. Hundal, T. Isacsson, R. B. Israel, J. Izaac, S. Jahangiri, R. Janik, N. Killoran, S. P. Kumar, J. Lavoie, A. E. Lita, D. H. Mahler, M. Menotti, B. Morrison, S. W. Nam, L. Neuhaus, H. Y. Qi, N. Quesada, A. Repingon, K. K. Sabapathy, M. Schuld, D. Su, J. Swinarton, A. Száva, K. Tan, P. Tan, V. D. Vaidya, Z. Vernon, Z. Zabaneh, and Y. Zhang, “Quantum circuits with many photons on a programmable nanophotonic chip,” Nature 591(7848), 54–60 (2021). [CrossRef]  

12. R. Tanomura, R. Tang, S. Ghosh, T. Tanemura, and Y. Nakano, “Robust integrated optical unitary converter using multiport directional couplers,” J. Lightwave Technol. 38(1), 60–66 (2020). [CrossRef]  

13. R. Tang, R. Tanomura, T. Tanemura, and Y. Nakano, “Ten-port unitary optical processor on a silicon photonics chip,” ACS Photonics 8(7), 2074–2080 (2021). [CrossRef]  

14. P. L. Mennea, W. R. Clements, D. H. Smith, J. C. Gates, B. J. Metcalf, R. H. Bannerman, R. Burgwal, J. J. Renema, W. S. Kolthammer, I. A. Walmsley, and P. G. R. Smith, “Modular linear optical circuits,” Optica 5(9), 1087–1090 (2018). [CrossRef]  

15. D. Melati, A. Alippi, and A. Melloni, “Reconfigurable photonic integrated mode (de)multiplexer for SDM fiber transmission,” Opt. Express 24(12), 12625–12634 (2016). [CrossRef]  

16. D. Melati, A. Alippi, A. Annoni, N. Peserico, and A. Melloni, “Integrated all-optical MIMO demultiplexer for mode- and wavelength-division-multiplexed transmission,” Opt. Lett. 42(2), 342–345 (2017). [CrossRef]  

17. R. Tanomura, R. Tang, T. Suganuma, K. Okawa, E. Kato, T. Tanemura, and Y. Nakano, “Monolithic InP optical unitary converter based on multi-plane light conversion,” Opt. Express 28(17), 25392–25399 (2020). [CrossRef]  

18. M. Reck, A. Zeilinger, H. J. Bernstein, and P. Bertani, “Experimental realization of any discrete unitary operator,” Phys. Rev. Lett. 73(1), 58–61 (1994). [CrossRef]  

19. W. R. Clements, P. C. Humphreys, B. J. Metcalf, W. S. Kolthammer, and I. A. Walmsley, “Optimal design for universal multi-port interferometers,” Optica 3(12), 1460–1465 (2016). [CrossRef]  

20. R. Tanomura, R. Tang, T. Umezaki, G. Soma, T. Tanemura, and Y. Nakano, “Scalable and robust photonic integrated unitary converter,” arXiv:2103.14782 (2021).

21. S. Arafin and L. A. Coldren, “Advanced InP photonic integrated circuits for communication and sensing,” IEEE J. Sel. Top. Quantum Electron. 24(1), 1–12 (2018). [CrossRef]  

22. I. M. Soganci, T. Tanemura, and Y. Nakano, “Integrated phased-array switches for large-scale photonic routing on chip,” Laser & Photon. Rev. 6(4), 549–563 (2012). [CrossRef]  

23. B. Shi, N. Calabretta, and R. Stabile, “Deep neural network thorough an InP SOA-based photonic integrated cross-connect,” IEEE J. Sel. Top. Quantum Electron. 26(1), 7701111 (2020). [CrossRef]  

24. J. F. Morizur, L. Nicholls, P. Jian, S. Armstrong, N. Treps, B. Hage, M. Hsu, W. Bowen, J. Janousek, and H. A. Bachor, “Programmable unitary spatial mode manipulation,” J. Opt. Soc. Am. A. 27(11), 2524–2531 (2010). [CrossRef]  

25. G. Labroille, B. Denolle, P. Jian, P. Genevaux, N. Treps, and J.-F. Morizur, “Efficient and mode selective spatial mode multiplexer based on multiplane light conversion,” Opt. Express 22(13), 15599–15607 (2014). [CrossRef]  

26. N. F. Fontaine, R. Ryf, H. Chen, D. T. Neilson, K. Kim, and J. Carpenter, “Laguerre-Gaussian mode sorter,” Nat. Commun. 10(1), 1865 (2019). [CrossRef]  

27. R. Tang, T. Tanemura, and Y. Nakano, “Integrated reconfigurable unitary optical mode converter using MMI couplers,” IEEE Photonics Technol. Lett. 29(12), 971–974 (2017). [CrossRef]  

28. M. Bachman, P. A. Besse, and H. Melchior, “Overlapping-image multimode interference couplers with a reduced number of self-images for uniform and nonuniform power splitting,” Appl. Opt. 34(30), 6898–6910 (1995). [CrossRef]  

29. K. Cooney and F. H. Peters, “Analysis of multimode interferometers,” Opt. Express 24(20), 22481–22515 (2016). [CrossRef]  

30. G. Dominic and F. Thomas, “Eigenmode expansion methods for simulation of optical propagation in photonics: Pros and cons,” Proc. SPIE 4987, 69–82 (2003). [CrossRef]  

31. F. Mezzadri, “How to generate random matrices from the classical compact groups,” Notices Amer. Math. Soc. 54(5), 592–604 (2007).

32. T. Tanemura, I. M. Soganci, T. Oyama, T. Ohyama, S. Mino, K. A. Williams, N. Calabretta, H. J. S. Dorren, and Y. Nakano, “Large-capacity compact optical buffer based on InP integrated phased-array switch and coiled fiber delay lines,” J. Lightwave Technol. 29(4), 396–402 (2011). [CrossRef]  

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 (5)

Fig. 1.
Fig. 1. (a) Schematic of the MPLC-based OUC using cascaded MMIs. (b) Schematic of general N×N MMI.
Fig. 2.
Fig. 2. Numerical analyses of 4×4 MMI on InP. (a) Top and (b) cross-sectional structures. (c, d) The transmittance to all output ports (Out 1, 2, 3, 4) and their sum (Total) when the light is input from port 1 (c) and port 2 (d), calculated for various MMI length L. At L = 252 µm and 504 µm, the light is split without excess loss, corresponding respectively to the non-uniform and uniform splitters. (e-h) The wave propagation inside the MMI for L = 504 µm (e, f) and 252 µm (g, h) from the input port 1 (e, g) and port 2 (f, h).
Fig. 3.
Fig. 3. Simulated performance of 4 × 4 OUC with uniform (blue) and half-integer (red) MMIs. Average mean-square error (MSE) of the obtained unitary operation after optimizing the OUCs with different number of stages K. The bars represent the distributed results for 100 random unitary matrixes, tested as the target unitary operations.
Fig. 4.
Fig. 4. (a) Fabricated OUC using half-integer MMI on InP platform. (b-d) The SEM image of (b) cross-sectional image at passive waveguide, (c) top-view of the 4×4 MMI, and (d) cross-sectional image of the phase shifter.
Fig. 5.
Fig. 5. Measured transmittance at the wavelength 1550 nm from each input and output port (a) before optimization, after optimization from {In1, In2, In3, In4} to (b) {Out1, Out2, Out3, Out4}, (c) {Out1, Out3, Out2, Out4}, (d) {Out1, Out4, Out2, Out3}, (e) {Out2, Out1, Out3, Out4}, and (f) {Ou41, Out3, Out2, Out1}. The crosstalk and MSE after optimization in the respective cases are (b) (-6.1 dB, -15.8 dB), (c) (-6.1 dB, -14.9 dB), (d) (-6.1 dB, -14.8 dB), (e) (-6.2 dB, -14.7 dB), and (f) (-5.9 dB, -14.6 dB).

Equations (3)

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

U = Φ K M Φ K 1 M Φ 0
E ( x , L 0 q ) = 1 C m = 0 q 1 e i φ m [ E ( x m 2 W q , 0 ) E ( 2 W x m 2 W q , 0 ) ] ,
MSE = 1 N 2 i = 1 , j = 1 N , N | U i j U i j | 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.