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

Ultrafast nonlinear effects in hydrogenated amorphous silicon wire waveguide

Open Access Open Access

Abstract

We, for the first time, present the ultrafast optical nonlinear response of a hydrogenated amorphous silicon (a-Si:H) wire waveguide using femtosecond pulses. We show cross-phase and cross-absorption modulations measured using the heterodyne pump-probe method and estimate the optical Kerr coefficient and two-photon absorption coefficient for the amorphous silicon waveguide. The pumping energy of 0.8 eV is slightly lower than that required to achieve two-photon excitation at the band gap of a-Si:H (~1.7 eV). An ultrafast response of less than 100 fs is observed, which indicates that the free-carrier effect is suppressed by the localized states in the band gap.

©2010 Optical Society of America

1. Introduction

Third-order nonlinear effects in semiconductors have been studied for all-optical switching devices. Cross-phase modulation (XPM) and cross-absorption modulation (XAM) are used for ultrafast signal processing where light is controlled by another light. In recent years, the efficient nonlinear interactions in silicon waveguides have been investigated because the strong optical field confinement in a core with a small cross section allows a pulse to have high power density with relatively low injected power [110]. The dominant nonlinear effects in crystalline Si (c-Si) are ultrafast Kerr nonlinearity, two-photon absorption (TPA), and slow effects of free carriers generated by TPA. Therefore, the response time of a nonlinear interaction is limited by the free-carrier lifetime of several hundred picoseconds in c-Si. To avoid the influence of free carriers, ion implantation of the silicon waveguide [11], carrier reduction by reverse-biasing the p-i-n diode in the waveguide [12,13], and using high-bit-rate signals where the free-carrier density is saturated [14] have been reported.

An alternative approach is to employ another material such as amorphous silicon (a-Si), which has a wider band gap and shorter carrier lifetime owing to the localized state in the band gap, as a high-index contrast waveguide. Since hydrogenated amorphous silicon (a-Si:H) has low absorption loss because of the terminating dangling bond in a-Si, it has recently attracted attention for the interconnection or three-dimensional integration of high-index contrast waveguides [1519]. However, there has been little investigation of the optical nonlinearity in a-Si at telecom wavelengths. Ikeda et al. investigated nonlinear optical absorption in a-Si:H films using a Z-scan technique and in a hybrid waveguide of a-Si and c-Si with a short response time of ~30 ps [20].

In this paper, we present ultrafast nonlinear responses in an a-Si:H wire waveguide measured using heterodyne pump-probe methods and femtosecond pulses. This is the first report of Kerr nonlinearity in an a-Si:H waveguide at a telecom wavelength. Ultrafast response of less than 100 fs without slow free-carrier relaxation is observed.

2. Experiment

2.1 Fabrication of amorphous silicon wire waveguide

A hydrogenated amorphous silicon layer was deposited by plasma-enhanced chemical vapor deposition with SiH4 and H2 mixture at 250 °C on a silicon wafer with a 2-μm-thick thermal oxide layer [15,16]. The band gap energy was ~1.7 eV and the absorption coefficient at 0.8 eV (λ = 1550 nm) was measured to be ~10−2 cm−1 using the constant photocurrent method. The refractive index was determined to be 3.516 at a wavelength of 1550 nm by spectroscopic ellipsometry. The waveguide core was formed by electron-beam lithography and reactive ion etching with SF6 gas. After RCA cleaning, 1-μm-thick SiO2 was deposited as over-cladding by electron cyclotron resonance sputter deposition. Each end of the waveguide had a spot-size converter with a narrowed waveguide for field enlargement. Figure 1(a) shows a scanning electron microscope (SEM) image of the cross section of the 400 nm × 250 nm waveguide.

 figure: Fig. 1

Fig. 1 (a) SEM image of cross section of fabricated a-Si:H wire waveguide. (b) Insertion loss as a function of propagation length at λ = 1550 nm.

Download Full Size | PDF

Propagation losses of 14 dB/cm for the transverse electric-like (TE) mode and 7 dB/cm for the transverse magnetic-like (TM) mode were measured using the cut-back method with fiber-to-fiber alignment using continuous-wave light at λ = 1550 nm as shown in Fig. 1(b). The coupling loss between the a-Si:H waveguide and the lensed fiber was estimated to be 8 dB/facet. The power confinement of the TE mode within the a-Si:H core was three times that of the TM mode for the core dimensions used in this study, and therefore, the TE mode was more sensitive to waveguide loss because of the sidewall roughness. This should explain the larger loss in the TE mode.

2.2 Heterodyne pump-probe measurement

The ultrafast optical nonlinearity of the a-Si:H wire waveguide was measured using the heterodyne pump-probe method and femtosecond pulses [10,21]. Our experimental setup is shown in Fig. 2 . An optical parametric oscillator (OPO), which is synchronously pumped by a mode-locked Ti:sapphire laser, provides 150-fs optical pulses at a wavelength of 1550 nm and repetition rate of 80 MHz. The OPO output is split into three beams as pump, probe, and reference pulses. The probe and reference beams pass through acousto-optic modulators (AOMs), where the frequencies are shifted by + 75 and + 85 MHz respectively. The probe and reference pulses were broadened from 150 fs to 240 fs by chromatic dispersion in the AOMs. The three beams are recombined so that the reference pulse leads the pump and probe pulses by ~600 ps. The time delay Δt between the pump and probe pulses is controlled with the motorized translation stage. The pulses are launched into the device under test and they transmit through the 2-mm-long waveguide. A Michelson interferometer induces interference between the output reference and probe pulses with a 10-MHz beat signal, which is detected by a radio-frequency lock-in amplifier. By comparing the beat signals with and without the pump pulse, the changes in amplitude and phase of the probe pulse due to the pump pulse are obtained. The pulse energies of the probe and reference were set at 1.75 pJ. The pulse energy of the pump was varied from 1.25 to 10 pJ. All the beams were polarized so as to excite the TE mode. The coupling loss between the objective lens in this measurement and the waveguide is estimated to be ~6 dB/facet, which is better than the measured loss shown in Fig. 1(b).

 figure: Fig. 2

Fig. 2 Experimental setup of heterodyne pump-probe measurement.

Download Full Size | PDF

Figure 3 shows the results of the measurement. A positive delay time means the pump pulse leads the probe pulse. For comparison, figures in the insets show the experimental results for c-Si, i.e., measurements for a Si-wire waveguide composed of a silicon-on-insulator wafer taken using the same method for the pump pulse energy of 2.0 pJ [10]. The XAM is due to TPA or two-step absorption [20] across the band gap of a-Si:H. The XPM is due to the optical Kerr effect. In the case of c-Si, free carriers generated by TPA affect the loss due to free-carrier absorption (FCA) and a negative refractive index change by free-carrier plasma effect (FCP) with a slow response of several hundred picoseconds, as shown in Figs. 3(a’) and 3(b’), respectively. In the case of a-Si:H, the pumping energy of two-photon excitation for light of 0.8 eV (λ = 1550 nm) is slightly less than the band gap energy of a-Si:H (~1.7 eV). Therefore, the number of excited carriers is less than that in the case of c-Si and the carriers are instantaneously trapped in the localized state (tail state) in the band gap, which prohibits the slow relaxation of free carriers. Consequently, an ultrafast nonlinear response of less than 100 fs was observed.

 figure: Fig. 3

Fig. 3 Experimental results of measured change in (a) amplitude and (b) phase of probe pulse for different pulse energies of pump pulse. Insets show the experimental results for c-Si obtained previously using the same method [10].

Download Full Size | PDF

The slight rebound of the amplitude for the positive delay time is to be noted. It is known that one-photon absorption due to a defect state results in a background loss over a wide range of wavelengths. We presume the rebound represents the pump-induced saturation of such one-photon resonance. A slightly longer trailing edge in XPM indicates that there is a change in the index with the same sign as the Kerr effect. This is also due to the generated carrier. Although we cannot provide a clear explanation, we believe it is worth mentioning that the absorption saturation and the increase in the index are consistent with the trend that the Kramers-Kronig relation predicts for the range of anomalous dispersion.

3. Discussion

For more quantitative analysis, we show the plots of measured XAM and XPM as functions of the incident pump peak power in Fig. 4 . First, we consider the XAM due to the TPA of the a-Si:H wire waveguide using coupled differential equations [9]:

dIpumpdz=αIpumpβT(Ipump+2Iprobe)Ipump,
dIprobedz=αIprobeβT(Iprobe+2Ipump)Iprobe,
where α is the linear coefficient of loss due to sidewall scattering and linear absorption, βT is the TPA coefficient, and z is the propagation length along the waveguide. Dispersion and Raman effects are ignored because the nonlinear length is much shorter than the dispersive and Raman contribution lengths. I is the intensity of the pulses derived from P/Aeff, where P is the peak power of the pulse and Aeff is the effective mode area. Here, we assume Aeff = 0.0585 μm2 for the TE mode, which we obtained by a mode field calculation. A coupling loss of 6 dB is taken into account for incident pulse energies at the waveguide edge (z = 0). A linear loss of 14 dB/cm based on the cut-back measurement is associated with α. XAM is calculated using Eqs. (1) and (2) as the cross term. When βT = 0.08 cm/GW, the calculation result for z = 2.0 mm agrees with the measurement result as shown by the dashed line in Fig. 4(a).

 figure: Fig. 4

Fig. 4 (a) XAM and (b) XPM as functions of incident pump peak power. The plotted values are the measurement results. Dashed lines show calculation results for z = 2.0 mm where βT = 0.08 cm/GW and n 2 = 0.5 × 10−18 m2/W.

Download Full Size | PDF

XPM induced by the Kerr effect is given by

dφprobedz=2k0n2Ipump,
where k 0 is the wave number and n 2 is the Kerr coefficient. The total phase shift depends on the pump power and its attenuation along the propagation length. When n 2 = 0.5 × 10−18 m2/W, the calculation results obtained using Eqs. (1), (2), and (3) for z = 2.0 mm agree with the measurement results as shown by the dashed line in Fig. 4(b). Estimated nonlinear coefficients of a-Si:H are one-tenth those of c-Si.

In this analysis, we neglect the effects of free carriers because they were observed to be small compared with the Kerr effect and the effect of TPA. Instead, we can say that the contribution is quite different from that for c-Si since free carrier transition occurs in lower energy states than the conduction band state. We speculate that the small mismatch in XAM may be the contribution of free carriers. However, at this time, there is little information that can be used to determine the carrier effect and the precise energies and densities of the localized states, which may strictly depend on the condition of a-Si:H deposition. We leave it to a future study to establish a model that can explain the carrier dynamics.

4. Conclusion

We studied the ultrafast nonlinear effects in an a-Si:H wire waveguide using femtosecond pulses. We fabricated an a-Si:H wire waveguide and characterized the nonlinear effect using the heterodyne pump-probe method. A high-speed response of less than 100 fs without slow free-carrier relaxation was observed for a photon energy of 0.8 eV. Nonlinear coefficients of βT = 0.08 cm/GW and n 2 = 0.5 × 10−18 m2/W were estimated, and thus, this is the first report with regard to Kerr nonlinearity in an a-Si:H waveguide at a telecom wavelength. The ultrafast response will be useful for all-optical signal processing.

Acknowledgments

This study was supported in part by Special Coordination Funds for Promoting Science and Technology from the Ministry of Education, Culture, Sport, Science and Technology, Japan.

References and links

1. O. Boyraz, P. Koonath, V. Raghunathan, and B. Jalali, “All optical switching and continuum generation in silicon waveguides,” Opt. Express 12(17), 4094–4102 (2004). [CrossRef]   [PubMed]  

2. T. K. Liang, L. R. Nunes, T. Sakamoto, K. Sasagawa, T. Kawanishi, M. Tsuchiya, G. Priem, D. Van Thourhout, P. Dumon, R. Baets, and H. Tsang, “Ultrafast all-optical switching by cross-absorption modulation in silicon wire waveguides,” Opt. Express 13(19), 7298–7303 (2005). [CrossRef]   [PubMed]  

3. H. Yamada, M. Shirane, T. Chu, H. Yokoyama, S. Ishida, and Y. Arakawa, “Nonlinear-optic silicon-nanowire waveguides,” Jpn. J. Appl. Phys. 44(No. 9A), 6541–6545 (2005). [CrossRef]  

4. C. Manolatou and M. Lipson, “All-optical silicon modulators based on carrier injection by two-photon absorption,” J. Lightwave Technol. 24(3), 1433–1439 (2006). [CrossRef]  

5. R. Dekker, A. Driessen, T. Wahlbrink, C. Moormann, J. Niehusmann, and M. Först, “Ultrafast Kerr-induced all-optical wavelength conversion in silicon waveguides using 1.55 mum femtosecond pulses,” Opt. Express 14(18), 8336–8346 (2006). [CrossRef]   [PubMed]  

6. K. Yamada, T. Tsuchizawa, T. Watanabe, J. Takahashi, H. Fukuda, M. Takahashi, T. Shoji, S. Uchiyama, E. Tamechika, S. Itabashi, and H. Morita, “Silicon wire waveguiding system: Fundamental characteristics and applications,” Electron. Commun. Jpn. 89, 42–55 (2006).

7. I.-W. Hsieh, X. Chen, J. I. Dadap, N. C. Panoiu, R. M. Osgood Jr, S. J. McNab, and Y. A. Vlasov, “Cross-phase modulation-induced spectral and temporal effects on co-propagating femtosecond pulses in silicon photonic wires,” Opt. Express 15(3), 1135–1146 (2007). [CrossRef]   [PubMed]  

8. C. Koos, L. Jacome, C. Poulton, J. Leuthold, and W. Freude, “Nonlinear silicon-on-insulator waveguides for all-optical signal processing,” Opt. Express 15(10), 5976–5990 (2007). [CrossRef]   [PubMed]  

9. Q. Lin, O. J. Painter, and G. P. Agrawal, “Nonlinear optical phenomena in silicon waveguides: modeling and applications,” Opt. Express 15(25), 16604–16644 (2007). [CrossRef]   [PubMed]  

10. S. Suda, T. Ogasawara, Y. Shoji, K. Kintaka, H. Kawashima, T. Hasama, and H. Ishikawa, “Femtosecond index dynamics in silicon wire waveguides,” in Proceedings of IEEE Conference on Group IV photonics (Institute of Electrical and Electronics Engineers, New York, 2009), pp. 98–100.

11. Y. Liu and H. K. Tsang, “Time dependent density of free carriers generated by two photon absorption in silicon waveguides,” Appl. Phys. Lett. 90(21), 211105 (2007). [CrossRef]  

12. R. Jones, H. Rong, A. Liu, A. W. Fang, M. J. Paniccia, D. Hak, and O. Cohen, “Net continuous wave optical gain in a low loss silicon-on-insulator waveguide by stimulated Raman scattering,” Opt. Express 13(2), 519–525 (2005). [CrossRef]   [PubMed]  

13. A. C. Turner-Foster, M. A. Foster, J. S. Levy, C. B. Poitras, R. Salem, A. L. Gaeta, and M. Lipson, “Ultrafast free-carrier lifetime for low nonlinear loss in silicon waveguides,” in Conference on Lasers and Electro-Optics/ International Quantum Electronics Conference, Technical Digest (CD) (Optical Society of America, 2009), paper CThCC1.

14. S. Suda, S. Namiki, Y. Shoji, K. Kintaka, H. Kawashima, and T. Hasama, H, Ishikawa, “High-bit-rate optical switchi based on XPM in silicon waveguides,” in Proceedings of IEEE Conference on Photonics in Switching (Institute of Electrical and Electronics Engineers, New York, 2009), paper ThI3–4.

15. T. Kamei, N. Hata, A. Matsuda, T. Uchiyama, S. Amano, K. Tsukamoto, Y. Yoshioka, and T. Hirao, “Deposition and extensive light soaking of highly pure hydrogenated amorphous silicon,” Appl. Phys. Lett. 68(17), 2380–2382 (1996). [CrossRef]  

16. T. Kamei, B. M. Paegel, J. R. Scherer, A. M. Skelley, R. A. Street, and R. A. Mathies, “Integrated hydrogenated amorphous Si photodiode detector for microfluidic bioanalytical devices,” Anal. Chem. 75(20), 5300–5305 (2003). [CrossRef]  

17. K. Shiraishi, H. Yoda, A. Ohshima, H. Ikedo, and C. S. Tsai, “A silicon-based spot-size converter between single-mode fibers and Si-wire waveguides using cascaded tapers,” Appl. Phys. Lett. 91(14), 141120 (2007). [CrossRef]  

18. A. Harke, T. Lipka, J. Amthor, O. Horn, M. Krause, and J. Ö. Muller, “Amorphous silicon 3-D tapers for Si photonic wires fabricated with shadow masks,” IEEE Photon. Technol. Lett. 20(17), 1452–1454 (2008). [CrossRef]  

19. R. Sun, M. Beals, A. Pomerene, J. Cheng, C. Y. Hong, L. Kimerling, and J. Michel, “Impedance matching vertical optical waveguide couplers for dense high index contrast circuits,” Opt. Express 16(16), 11682–11690 (2008). [CrossRef]   [PubMed]  

20. K. Ikeda, Y. Shen, and Y. Fainman, “Enhanced optical nonlinearity in amorphous silicon and its application to waveguide devices,” Opt. Express 15(26), 17761–17771 (2007). [CrossRef]   [PubMed]  

21. A. J. Zilkie, J. Meier, P. W. E. Smith, M. Mojahedi, J. S. Aitchison, P. J. Poole, C. N. Allen, P. Barrios, and D. Poitras, “Femtosecond gain and index dynamics in an InAs/InGaAsP quantum dot amplifier operating at 1.55 microm,” Opt. Express 14(23), 11453–11459 (2006). [CrossRef]   [PubMed]  

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 (a) SEM image of cross section of fabricated a-Si:H wire waveguide. (b) Insertion loss as a function of propagation length at λ = 1550 nm.
Fig. 2
Fig. 2 Experimental setup of heterodyne pump-probe measurement.
Fig. 3
Fig. 3 Experimental results of measured change in (a) amplitude and (b) phase of probe pulse for different pulse energies of pump pulse. Insets show the experimental results for c-Si obtained previously using the same method [10].
Fig. 4
Fig. 4 (a) XAM and (b) XPM as functions of incident pump peak power. The plotted values are the measurement results. Dashed lines show calculation results for z = 2.0 mm where βT = 0.08 cm/GW and n 2 = 0.5 × 10−18 m2/W.

Equations (3)

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

d I p u m p d z = α I p u m p β T ( I p u m p + 2 I p r o b e ) I p u m p ,
d I p r o b e d z = α I p r o b e β T ( I p r o b e + 2 I p u m p ) I p r o b e ,
d φ p r o b e d z = 2 k 0 n 2 I p u m p ,
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.