Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Mode conversion in tapered submicron silicon

ridge optical waveguides


Daoxin Dai,1,2,* Yongbo Tang,2 and John E Bowers2
1
Centre for Optical and Electromagnetic Research, State Key Laboratory for Modern Optical Instrumentation,
Zhejiang Provincial Key Laboratory for Sensing Technologies, Zhejiang University, Zijingang Campus, Hangzhou
310058, China
2
Department of Electrical and Computer Engineering, University of California, Santa Barbara, California 93106,
USA
*
dxdai@zju.edu.cn

Abstract: The mode conversion in tapered submicron silicon ridge optical


waveguides is investigated theoretically and experimentally. Two types of
optical waveguide tapers are considered in this paper. One is a regular
lateral taper for which the waveguide width varies while the etching depth
is kept the same. The other is a so-called “bi-level” taper, which includes
two layers of lateral tapers. Mode conversion between the TM fundamental
mode and higher-order TE modes is observed in tapered submicron silicon-
on-insulator ridge optical waveguides due to the mode hybridization
resulting from the asymmetry of the cross section. Such a mode conversion
could have a very high efficiency (close to 100%) when the taper is
designed appropriately. This enables some applications e.g. polarizer,
polarization splitting/rotation, etc. It is also shown that this kind of mode
conversion could be depressed by carefully choosing the taper parameters
(like the taper width, the etching depth, etc), which is important for the
applications when low-loss propagation for the TM fundamental mode is
needed.
©2012 Optical Society of America
OCIS codes: (130.0130) Integrated optics; (230.5440) Polarization-selective devices.

References and links


1. Y. Shani, C. Henry, R. Kistler, K. Orlowsky, and D. Ackerman, “Efficient coupling of a semiconductor laser to
an optical fiber by means of a tapered waveguide on silicon,” Appl. Phys. Lett. 55(23), 2389–2391 (1989).
2. R. Smith, C. Sullivan, G. Vawter, G. Hadley, J. Wendt, M. Snipes, and J. Klem, “Reduced coupling loss using a
tapered-rib adiabatic-following fiber coupler,” IEEE Photon. Technol. Lett. 5, 1053–1056 (1993).
3. R. Zengerle, H. Bruckner, H. Olzhausen, and A. Kohl, “Low-loss fiber-chip coupling by buried laterally tapered
InP/InGaAsP waveguide structure,” Electron. Lett. 28(7), 631–632 (1992).
4. K. Kasaya, O. Mitomi, M. Naganuma, Y. Kondo, and Y. Noguchi, “A simple laterally tapered waveguide for
low-loss coupling to single-mode fibers,” IEEE Photon. Technol. Lett. 5(3), 345–347 (1993).
5. T. Schwander, S. Fischer, A. Kramer, M. Laich, K. Luksic, G. Spatschek, and M. Warth, “Simple and low-loss
fiber-to-chip coupling by integrated field-matching waveguide in InP,” Electron. Lett. 29(4), 326–328 (1993).
6. L. Yang, D. Dai, B. Yang, Z. Sheng, and S. He, “Characteristic analysis of tapered lens fibers for light focusing
and butt-coupling to a silicon rib waveguide,” Appl. Opt. 48(4), 672–678 (2009).
7. D. Dai, S. He, and H. K. Tsang, “Bilevel mode converter between a silicon nanowire waveguide and a larger
waveguide,” J. Lightwave Technol. 24(6), 2428–2433 (2006).
8. A. Barkai, A. Liu, D. Kim, R. Cohen, N. Elek, H.-H. Chang, B. H. Malik, R. Gabay, R. Jones, M. Paniccia, and
N. Izhaky, “Double-stage taper for coupling between SOI waveguides and single-mode fiber,” J. Lightwave
Technol. 26(24), 3860–3865 (2008).
9. Y. Shani, C. H. Henry, R. C. Kistler, R. F. Kazarinov, and K. J. Orlowsky, “Integrated optic adiabatic devices on
silicon,” IEEE J. Quantum Electron. 27(3), 556–566 (1991).
10. R. S. Fan and R. B. Hooker, “Tapered polymer single-mode waveguides for mode transformation,” J. Lightwave
Technol. 17(3), 466–474 (1999).
11. K. W¨orhoff, P. V. Lambeck, and A. Driessen, “Design, tolerance analysis, and fabrication of silicon oxynitride
based planar optical waveguides for communication devices,” J. Lightwave Technol. 17(8), 1401–1407 (1999).
12. P. Sewell, T. M. Benson, and P. C. Kendall, “Rib waveguide spot-size transformers: modal properties,” J.
Lightwave Technol. 17(5), 848–856 (1999).
13. K. Sasaki, F. Ohno, A. Motegi, and T. Baba, “Arrayed waveguide grating of 70×60µm2 size based on Si
photonic wire waveguides,” Electron. Lett. 41(14), 801–802 (2005).

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13425
14. D. Dai, L. Liu, L. Wosinski, and S. He, “Design and fabrication of ultra-small overlapped AWG demultiplexer
based on Si nanowire waveguides,” Electron. Lett. 42(7), 400–402 (2006).
15. W. Bogaerts, S. K. Selvaraja, P. Dumon, J. Brouckaert, K. De Vos, D. Van Thourhout, and R. Baets, “Silicon-
on-Insulator Spectral Filters Fabricated With CMOS Technology,” IEEE J. Sel. Top. Quantum Electron. 16(1),
33–44 (2010).
16. H. Fukuda, K. Yamada, T. Tsuchizawa, T. Watanabe, H. Shinojima, and S. Itabashi, “Silicon photonic circuit
with polarization diversity,” Opt. Express 16(7), 4872–4880 (2008).
17. W. Bogaerts, P. Dumon, D. V. Thourhout, D. Taillaert, P. Jaenen, J. Wouters, S. Beckx, V. Wiaux, and R. G.
Baets, “Compact wavelength-selective functions in silicon-on-insulator photonic wires,” IEEE J. Sel. Top.
Quantum Electron. 12(6), 1394–1401 (2006).
18. M. Soltani, S. Yegnanarayanan, and A. Adibi, “Ultra-high Q planar silicon microdisk resonators for chip-scale
silicon photonics,” Opt. Express 15(8), 4694–4704 (2007).
19. O. Boyraz and B. Jalali, “Demonstration of a silicon Raman laser,” Opt. Express 12(21), 5269–5273 (2004).
20. C. Li, L. Zhou, and A. W. Poon, “Silicon microring carrier-injection-based modulators/switches with tunable
extinction ratios and OR-logic switching by using waveguide cross-coupling,” Opt. Express 15(8), 5069–5076
(2007).
21. H. Rong, R. Jones, A. Liu, O. Cohen, D. Hak, A. Fang, and M. Paniccia, “A continuous-wave Raman silicon
laser,” Nature 433(7027), 725–728 (2005).
22. Q. Xu, B. Schmidt, S. Pradhan, and M. Lipson, “Micrometre-scale silicon electro-optic modulator,” Nature
435(7040), 325–327 (2005).
23. C. A. Barrios, V. R. Almeida, R. Panepucci, and M. Lipson, “Electrooptic modulation of silicon-on-insulator
submicrometer-size waveguide devices,” J. Lightwave Technol. 21(10), 2332–2339 (2003).
24. Y. Tang, H.-W. Chen, S. Jain, J. D. Peters, U. Westergren, and J. E. Bowers, “50 Gb/s hybrid silicon traveling-
wave electroabsorption modulator,” Opt. Express 19(7), 5811–5816 (2011).
25. K. Mertens, B. Scholl, and H. Schmitt, “New highly efficient polarization converters based on hybrid
supermodes,” J. Lightwave Technol. 13(10), 2087–2092 (1995).
26. L. Liu, Y. Ding, K. Yvind, and J. M. Hvam, “Silicon-on-insulator polarization splitting and rotating device for
polarization diversity circuits,” Opt. Express 19(13), 12646–12651 (2011).
27. D. Dai, Z. Wang, N. Julian, and J. E. Bowers, “Compact broadband polarizer based on shallowly-etched silicon-
on-insulator ridge optical waveguides,” Opt. Express 18(26), 27404–27415 (2010).
28. R. S. Tummidi, T. G. Nguyen, A. Mitchell, and T. L. Koch, “An ultra-compact waveguide polarizer based on
“anti-magic widths”,” 2011 8th IEEE International Conference on Group IV Photonics (GFP), London, UK, pp.
104–106, 14–16 Sept. 2011.
29. D. Vermeulen, S. Selvaraja, W. A. D. De Cort, N. A. Yebo, E. Hallynck, K. De Vos, P. P. P. Debackere, P.
Dumon, W. Bogaerts, G. Roelkens, D. Van Thourhout, and R. Baets, “Efficient tapering to the fundamental
Quasi-TM mode in asymmetrical waveguides,” ECIO 2010 (2010).
30. D. Dai and J. E. Bowers, “Novel concept for ultracompact polarization splitter-rotator based on silicon
nanowires,” Opt. Express 19(11), 10940–10949 (2011).
31. M. Kohtoku, T. Hirono, S. Oku, Y. Kadota, Y. Shibata, and Y. Yoshikuni, “Control of higher order leaky modes
in deep-ridge waveguides and application to low-crosstalk arrayed waveguide gratings,” J. Lightwave Technol.
22(2), 499–508 (2004).
32. D. Dai, J. He, and S. He, “Elimination of multimode effects in a silicon-on-insulator etched diffraction grating
demultiplexer with bi-level taper structure,” IEEE J. Sel. Top. Quantum Electron. 11(2), 439–443 (2005).
33. J. H. Schmid, B. Lamontagne, P. Cheben, A. Delâge, S. Janz, A. Densmore, J. Lapointe, E. Post, P. Waldron,
and D.-X. Xu, “Mode Converters for coupling to high aspect ratio silicon-on-insulator channel waveguides,”
IEEE Photon. Technol. Lett. 19(11), 855–857 (2007).
34. FIMMWAVE/FIMMPROP, Photon Design Ltd, http://www.photond.com.
35. D. Dai, Z. Wang, J. Peters, and J. E. Bowers, “Compact polarization beam splitter using an asymmetrical Mach-
Zehnder Interferometer based on silicon-on-insulator waveguides,” IEEE Photon. Technol. Lett. 24(8), 673–675
(2012).

1. Introduction
Optical waveguide taper is a fundamental element for photonic integrated circuits (PIC’s). It
is often used to change the light spot size in order to have better coupling efficiency between
two sections with different cross sections (e.g., a planar optical waveguide and a singlemode
or lens fiber) [1–8]. In order to achieve a low-loss taper, one usually makes the taper long
enough to be adiabatic so that higher-order modes are not excited [9–12]. This design rule
works well usually especially for low index-contrast (∆) optical waveguides (e.g., SiO2-on-Si
buried waveguides). However, the situation becomes complicated for small-sized high-∆
optical waveguides, e.g., submicron silicon-on-insulator (SOI) waveguides, which have been
used widely for ultra-compact CMOS-compatible PIC’s in the recent years [13–24]. In a
high-∆ optical waveguide, the mode hybridization is significant at some special waveguide
widths [25–30] and consequently mode conversion may happen in a tapered structure [29,30].
In Ref [29,30], the authors give a discussion on the mode conversion in a tapered SOI strip

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13426
nanowire. When a SOI nanowire has a SiO2 under-cladding and an air upper-cladding, which
makes the SOI nanowire asymmetrical in the vertical direction, the mode conversion between
the TM fundamental mode and the first-order TE mode is observed when light propagates
along a taper structure. Such a mode conversion is not desired usually because it introduces
some serious excess loss as well as crosstalk due to the excited higher order modes, e.g., in
AWG (arrayed-waveguide grating) demultiplexer [31]. Such undesired mode-conversion
could be minimized by using several kinds of modified tapered structures suggested in Ref
[29]. A simple and easy way to depress such a mode conversion in a SOI-nanowire taper is to
introduce a SiO2 upper-cladding (instead of air) to make the SOI nanowire symmetrical in the
vertical direction [30]. On the other hand, such a kind of mode conversion could be very
useful. For example, in our previous paper a SOI-nanowire taper was designed to have an
almost 100% mode conversion efficiency from the TM fundamental (TM0) mode to the first
higher-order TE (TE1) mode so that polarization splitter-rotators could be realized with a very
simple design and easy fabrication process [30].
In this paper, we focus on the mode conversion in submicron SOI rib waveguides (other
than SOI nanowires), which is also very popular for silicon-based integrated optoelectronics
[19–24]. One should note that there is a significant difference between an SOI rib waveguide
and a SOI strip nanowire. A SOI strip nanowire could be symmetrical or asymmetrical in the
vertical direction by simply choosing an appropriate material for the upper-cladding so that
the mode conversion could be eliminated or enhanced accordingly [30]. In contrast, for a SOI
rib waveguide, it is still asymmetrical in the vertical direction even when having the same
material for the upper-cladding and the under-cladding. Therefore, when it is desired to have
a mode conversion between the TM0 mode and the higher-order TE mode for the case of
using a SOI rib waveguide, it is not necessary to choosing different materials for the upper-
cladding and the under-cladding. On the other hand, such an asymmetry also makes that one
cannot yet avoid the mode conversion in a taper section due to the mode hybridization by
simply choosing the same material for the upper-cladding and the under-cladding. Such a
mode conversion will introduce a significant excess insertion loss as well as some channel
crosstalk due to the excited higher-order modes (e.g., in AWG demultiplexers [31]).
In the following section, we give a detailed analysis for light propagation in SOI rib
waveguide tapers and present the mode conversion numerically. The experimental
observation for the mode conversion has also been presented. Two types of taper structures
are considered here. One is a regular lateral taper for which the waveguide width varies while
the etching depth is kept the same. Since the structure and the fabrication are very simple, a
regular lateral taper is very popular for modifying the waveguide mode size in the lateral
direction. The other is a so-called “bi-level” taper, which includes two layers of lateral tapers
and consequently a double-etching process is needed for the fabrication. A bi-level taper is
often used to connect two sections with different etching depths, e.g., from a shallowly-etched
rib waveguide to a deeply-etched rib waveguide [7–8, 32–33]. For example, bi-level taper are
very useful for the case when a singlemode rib waveguide is needed at the input/output ends
of a chip while a strong confinement is desired for e.g., sharp bending. Our experimental and
theoretical results show that one should be very careful when designing an adiabatic lateral
taper or bi-level taper with a small-sized high-∆ optical waveguide, e.g., submicron SOI ridge
waveguides considered in this paper.
2. Structure and analysis
In this paper, we consider tapered submicron SOI rib waveguides, which has been used very
widely for silicon optoelectronics [19–24]. Two types of taper structures are analyzed here.
The first one is a regular lateral taper, and the other is the so-called bi-level taper [7–8, 32–
33]. In the present example, the SOI wafer has a 400nm-thick top Si layer and the refractive
indices of Si and SiO2 are nSi = 3.455, and nSiO2 = 1.445, respectively. A finite-difference
method (FDM) mode-solver (from Fimmwave) is used to calculate the mode field profiles
and the effective indices for all eigenmodes.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13427
A. Regular lateral taper
Figure 1(a) and 1(b) show the 3D-view for the regular lateral taper and the cross section for
the SOI rib waveguide. In this taper section, the waveguide width varies while the etching
depth is kept the same. Such a taper is often used when it is needed to modify the mode size,
e.g., at the input/output ends of a silicon photonic integrated chip in order to enhance the
coupling efficiency between fibers and the chip. Regarding that the spot size of a
commercialized lens fiber is usually around 3µm, in our calculation we give a modal analysis
for a SOI rib waveguide whose core width varies from 3µm to 0.5µm in order to characterize
the mode conversion in a waveguide taper to match the lens fiber [6].

Fig. 1. (a) The schematic configuration of a regular lateral taper; (b) the cross section for a SOI
rib waveguide.

Fig. 2. The calculated effective indices for the eigen modes of SOI rib waveguide with
different etching depths. (a) het = 0.4H; (b) het = 0.5H; (c) het = 0.6H. Here the total height of
the Si layer is H = 400nm.

Figure 2(a)-2(c) show the effective indices for SOI rib waveguides with different etching
depths het as the core width wco increases from 0.5µm to 3µm. Here the etching depth is
chosen as het = 0.4H, 0.5H, and 0.6H, respectively. Particularly, for the case of het = 0.4H, one
should note that the TM0 mode becomes leaky and is to be cutoff in the range of wco<0.95µm

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13428
and thus the curve for the TM0 mode in Fig. 2(a) stops at w0 = 0.95µm. Here het is given with
a ratio in respect to H just to understand that the case considered here is with an etching depth
around half of the total height (which is used very often).
Since a SOI rib waveguide is asymmetrical in the vertical direction, mode hybridization is
observed in some special ranges of the rib width, e.g., around wco0 = 1µm, and 2.45µm, as
shown by the circles labeled in Fig. 2(a)-2(c). Due to the mode hybridization around wco =
wco0, mode conversion between the two hybridized modes will happen when the light
propagates along an “adiabatic” (long) taper structure whose end-widths (w1, and w2) satisfy
the condition: w1<wco0<w2. In Fig. 2(a)-2(c), the arrowed curves indicate that the mode
conversions between the TM0 mode and the higher-order TE mode as the core width varies.
Such a mode conversion is harmful when one expects to have a low-loss adiabatic taper [29].
In order to avoid the undesired mode conversion, one can choose the taper widths (w1, w2) so
that there is no mode hybridization in the width range of w1<w<w2. In this way, there will not
be mode conversion when light propagates along a long taper. From Fig. 2(a)-2(c), it can be
seen that the mode hybridization region shifts when choosing different etching depths het. One
has a smaller wco0 (where the mode hybridization region locates) when the optical waveguide
is etched less. This indicates that the mode conversion due to the mode hybridization could be
modified by slightly adjusting the etching depth het, which makes the design flexible. For
example, when reducing the etching depth from 0.6H to het = 0.4H, the first and second mode
hybridization regions shift from around wco0 = 1.1µm and 2.55µm to around wco0 = 0.9µm and
2.35µm, respectively, as shown in Fig. 2(a). Then one can choose the taper end-widths in the
range of 0.90µm<(w1, w2)<2.35µm so that no mode conversion happens in the designed taper
for the case of het = 0.4H. Particularly, regarding that the TM0 mode becomes leaky when
wco<0.95µm, one should choose the taper end-widths (w1, and w2) to be larger than 0.95µm,
i.e., (w1, w2)>0.95µm. Finally the end-widths of the low-loss taper should be 0.95µm<(w1,
w2)<2.35µm. On the other hand, it is also possible to utilize such kind of mode conversion to
obtain a polarization rotation, which is similar to the case of tapered SOI nanowires [30].

Fig. 3. The field profiles (Ex and Ey) for modes #1 and #2 of a SOI ridge waveguide with wco =
2.45µm, (a) mode #1; (b) mode #2. The total height of the Si core layer is H = 400nm, and the
etching depth het = 0.5H. Here modes #1 and #2 are the two hybridization modes in the region
around w = 2.45µm.

In order to show the mode hybridization which causes the mode conversion in a tapered
SOI rib waveguide, we consider the case of het = 0.5H as an example. From Fig. 2(b), it can
be seen that there are two regions (i.e., wco0 = 2.45µm, and 1.0µm) where mode hybridization
happens. In the region around wco = 2.45µm, the mode hybridization happens between the
TM0 and the third-order TE (TE3) mode. The mode profiles for these two modes are shown in
Fig. 3(a)-3(b), respectively. It can be seen that the minor-component (Ex or Ey) is comparable
to the corresponding major-component (Ey or Ex). In this case, it is hard to distinguish these
two modes. When wco = 1.0µm, the mode hybridization is similar while it happens between
the TM0 mode and the first-order TE mode (TE1), as shown in Fig. 4(a)-4(b).

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13429
As mentioned above, there are two regions (around wco0 = 1.0µm, and 2.45µm) where
mode conversions happen when the core width is tapered from 3µm to 0.5µm. Therefore, here
we examine two types of tapers. For the first one, the taper end-width is chosen as w1 = 2µm,
and w2 = 2.7µm (w1<2.45µm<w2). The second one has taper end-widths w1 = 0.8µm, and w2 =
1.5µm (w1<1µm<w2).

Fig. 4. The field profiles (Ex and Ey) for modes #1 and #2 of a SOI ridge waveguide with wco =
1.0µm, (a) mode #1; (b) mode #2. The total height of the Si core layer is H = 400nm, and het =
0.5H. Here modes #1 and #2 are the two hybridization modes in the region around w = 1.0µm.

A commercial software (FIMMPROP, Photon Design, UK) employing an eigenmode


expansion and matching method [34] is then used to simulate the light propagation in the
defined taper structure. Figure 5 shows the mode conversion efficiencies coupled to the TM0
mode and the TE3 mode after the launched TM0 mode propagates along the linear lateral taper
with w1 = 2.7µm, and w2 = 2.0µm. From this figure, it can be seen that one could realize a
very high efficiency (>90%) from the TM0 mode to the TE3 mode when choosing the taper
length Ltp appropriately.

Fig. 5. The mode conversion efficiency η as the taper length Ltp varies when the TM0 mode is
launched. The parameters are het = 0.5H, w1 = 2.7µm, and w2 = 2µm.

For example, here we choose Ltp = 1500µm. When the launched field is chosen as the TE0
or TM0 modal field, the simulated light propagation along the designed taper are shown in
Fig. 6(a) and 6(b), respectively. It can be seen that the mode conversion happens as predicted
when the input filed is the TM0 mode, while there is no mode conversion for the case with the
TE0-mode input. When one choose a shorter taper (e.g., 350µm), the launched TM0 mode is
then converted to the TE3 mode partially. Consequently two-mode interference happens,
which introduces some undesired ripples when measuring the wavelength dependence of the
output power.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13430
From Fig. 5, one can also see that the mode conversion could be very slight by choosing a
very short non-adiabatic taper. For example, for a 10µm -long taper, the mode conversion
from TM0 to TE3 is about 5% only and such a low loss is acceptable for some applications.
Figure 7(a) and 7(b) show the simulation results for light propagating along a 10µm-long
taper for the case with the TE0 and TM0 modes launched, respectively. From these two
figures, it can be seen that there are some small ripples due to the multimode-interference
effect. For the case when the TE0 mode is launched, the TE2 mode is excited slightly because
the taper is not adiabatic. In this case, the dominant mode is the TE0 mode which has a power
ratio of 99.66% while the excited TE2 mode has a low power ratio of about 0.26%.

Fig. 6. The light propagation in the designed long taper when the launched field is TE
polarization (a), and TM polarization (b), respectively. The parameters are: H = 400nm, het =
0.5H, w1 = 2.7µm, w2 = 2µm, Ltp = 1500µm.

Fig. 7. The light propagation in the designed short (non-adiabatic) taper when the launched
field is TE polarization (a), and TM polarization (b), respectively. The parameters are: H =
400nm, het = 0.5H, w1 = 2.7µm, w2 = 2µm, and Ltp = 10µm.

Figure 8 shows the mode conversion efficiencies to the TM0 and the TE1 mode after the
launched TM0 mode propagates along a linear lateral taper with w1 = 1.5µm, and w2 = 0.8µm,
between which there is a mode hybridization region round wco0 = 1µm (see Fig. 2 (b)). From
this figure, it can be seen that the mode conversion efficiency from the TM0 mode to the TE1
mode is close 100% when choosing the taper length Ltp appropriately.
For example, here we choose Ltp = 215µm. Figure 9(a) and 9(b) show the simulation
results for light propagating along the designed taper when the launched field is chosen as the
TE0 and TM0 modal field, respectively. It can be seen that the mode conversion happens for
the input TM0 modal field, while there is no mode conversion for the input TE0 modal field,
as predicted. When one chooses a shorter taper, the launched TM0 mode is converted to the
TE1 mode partially. For example, when Ltp = 22.4µm, the mode conversion from the TM0
mode to the TE1 mode is about 50%, and a significant multimode-interference effect is
observed as shown in Fig. 10(a)-10(b). Particularly, when choosing Ltp = 0 (i.e., no taper), the
mode conversion from the TM0 mode to the TE1 mode is about 25% and more power (~75%)

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13431
is preserved to the TM0 mode as shown in Fig. 8. One could design a discontinuous taper
structure by optimizing the widths w1 and w2 further to improve the preservation efficiency
for the TM0 mode, in a way similar to that suggested for the design of SOI strip nanowire
tapers in Ref [29].

Fig. 8. The mode conversion efficiency η as the taper length Ltp varies when the TM0 mode is
launched. The parameters are het = 0.5H, w1 = 1.5µm, and w2 = 0.8µm.

Fig. 9. The light propagation in the designed taper when the input is the TE0 modal field (a),
and the TM0 modal field (b), respectively. The parameters are: H = 400nm, het = 0.5H, w1 =
1.5µm, w2 = 0.8µm, Ltp = 215µm.

Fig. 10. Light propagation in the taper when the input is the TE0 modal field (a), and the TM0
modal field (b), respectively. The parameters are: H = 400nm, het = 0.5H, w1 = 1.5µm, w2 =
0.8µm, Ltp = 22.4µm.

In a short summary, for a regular lateral taper whose widths ranges from w1 to w2 (w1<w2),
the mode conversion between the TM0 mode and higher-order TE modes (e.g., TE1, TE3)

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13432
happens when there is a mode hybridization region in the range of w1<w<w2 according to the
simulation given above. Such a mode conversion could be very efficient (close to 100%)
when the taper length is long enough, which is very useful some applications, e.g.,
polarization rotation. On the other hand, it is also possible to design a taper to minimize the
mode conversion by choosing the etching depth (het) or the range for the taper end-widths (w1,
and w2) when it desired to achieve a low-loss waveguide taper. For example, one can choose
the taper end-widths (w1, and w2) so that there is no a mode hybridization region in the range
of w1<w<w2. In this way, there will not be mode conversion when light propagates along a
long taper. Generally speaking, the mode evolution in a gradually-varying taper structure is
insensitive to the variation of the taper dimension (e.g., the height, the width) when the taper
is long enough. However, when there are mode-hybridization regions as discussed here, one
should choose the taper end-widths carefully to be tolerant to the taper dimension variation
because the mode hybridization region shifts slightly as the taper dimension changes. For
example, one can choose the taper end-widths not to be close to the mode hybridization
regions (around wco0), which can make the taper tolerant to the dimension variation.
Bi-level taper is another type of taper structure used often to connect two sections with
different etching depth [7–8, 32–33, 35]. Usually it is assumed that no higher-order mode is
excited when light propagates along a long bi-level taper, so that one achieves a low-loss
smooth transition between the fundamental modes of the SOI ridge waveguide and the silicon
strip waveguide. However, our simulation shows that higher-order modes might be generated
even in a long bi-level taper for submicron SOI ridge waveguides. It is very essential to
understand this issue when designing the waveguide taper. In the following part, we give a
detailed analysis for the mode conversion in bi-level taper.
B. Bi-level taper
Figure 11(a) and 11(b) show the 3D-view for the bi-level taper and the cross section for the
SOI double-rib waveguide. Such a taper is often used to connect two sections with different
etching depths previously [7–8, 32–33].

Fig. 11. (a) The schematic configuration of a bi-level lateral taper; (b) the cross section for an
SOI double-rib waveguide in the taper section.

Figure 12(a)-12(c) show the effective indices for an SOI double-rib waveguide with het =
0.5H as the side-rib width wside decreases from 3µm to 0 when the central-rib width is chosen
as wco = 0.85, 1, and 1.2µm, respectively, so that the SOI rib waveguide is quasi-singlemode.
Since a SOI double-rib waveguide is not symmetrical in the vertical direction, mode
hybridization might happen. The mode hybridization in a SOI double-rib waveguide depends
a lot on the central rib width wco. For the present case (het = 0.5H), it is found that mode
hybridization and conversion happen as the side-rib width wside varies from 3µm to 0 when the
central rib width wco = 1.0µm according to Fig. 12(b) and the mode profiles e.g. shown in Fig.
13(a)-(b) below. In contrast, when choosing wco = 0.85µm, and 1.2µm, there is no mode
hybridization and conversion according to Fig. 12(a)-(c), and Fig. 14(a)-(b) below. The mode
hybridization and conversion can be also avoided by choosing a deeper etching depth het. For
example, when choosing het = 0.6H (see Fig. 12(d)), there is no mode conversion observed.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13433
Fig. 12. The calculated effective indices for the eigen modes of SOI double-ridge waveguides
with different rib widths wco: (a) wco = 0.85µm, and het = 0.5H; (b) wco = 1.0µm, and het = 0.5H;
(c) wco = 1.2µm, and het = 0.5H; (d) wco = 1.0µm, and het = 0.6H. Here H = 400nm.

Fig. 13. The field profiles (Ex and Ey) for modes #1 and #2 of a double ridge waveguide with:
(a) wside = 0.5µm; (b) wside = 0.5µm. The parameters are: wco = 1µm, H = 400nm, and het =
0.5H. Here modes #1 and #2 are the two lowest order modes except the TE0 mode.

Figure 13(a)-13(b) shows the field profiles (Ex and Ey) for mode #1 and #2 in the case of
wco = 1.0µm when wside = 0.5µm and 0µm, respectively. Here modes #1 and #2 are the two
lowest order modes except the TE0 mode. When wside = 0, the double-rib waveguide becomes

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13434
a rectangular waveguide, which is symmetrical in the vertical direction and consequently no
mode hybridization is observed. In contrast, for the case of wside = 0.5µm, it can be seen that
the eigen-modes have significant major as well as minor components (Ex and Ey) due to the
hybridization as shown in Fig. 13(a)-13(b). This mode hybridization between the TM0 mode
and the TE1 mode makes a mode conversion between them when light propagates along an
adiabatic taper. In order to check the mode hybridization for the cases of wco = 0.85µm and
1.2µm, we also consider the waveguide with wside = 0.5µm and show the filed profiles (Ex and
Ey) for mode #1 and #2 in Fig. 14(a)-14(b), respectively. Note that the color scale are
different for Ex and Ey. According to the field profiles, it can be seen that both mode #1 and
#2 have a major components (Ex or Ey), which indicates the mode hybridization is not
significant.

Fig. 14. The field profiles (Ex and Ey) for modes #1 and #2 of a double ridge waveguide with
the following parameters: (a) wco = 0.85µm, (b) wco = 1.2µm. The parameters are: wside =
0.5µm, H = 400nm, and het = 0.5H. Here modes #1 and #2 are the two lowest-order modes
except the TE0 mode.

In order to show the mode conversion in a tapered SOI double-rib waveguide, the case of
het = 0.5H is considered as an example. We use a commercial software (FIMMPROP, Photon
Design, UK) [34] to simulate the light propagation in the present structure. Figure 15 shows
the mode conversion efficiencies to the TM0 mode and the TE1 mode after the launched TM0
mode propagates along the bi-level taper. Here the side-rib width at the input end of the bi-
level taper is chosen as wside1 = 3.0µm (see the inset in Fig. 15). From this figure, it can be
seen that the mode conversion efficiency from the TM0 mode to the TE1 mode is close 100%
when choosing the taper length Ltp appropriately (e.g., >300µm). For example, here we
choose Ltp = 300µm. Figure 16(a)-16(b) show the simulation results for light propagating
along the designed taper when the launched field is chosen as the TE0, and TM0 modal field,
respectively. It can be seen that there is a very efficient mode conversion observed between
the TM0 mode and the TE1 mode as predicted, while there is no mode conversion when the

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13435
TE0 modal field is launched. The efficient mode conversion between the TM0 mode and the
TE1 mode enables the realization of a polarization splitter-rotator with the assistance of an
asymmetrical directional coupler as proposed in Ref [30]. Note that the taper length could be
shortened greatly by choosing a smaller side-rib width wside for the input end of the bi-level
taper (e.g., wside1 = 1.0µm) since the mode conversion happens at the region very close to the
output end of the bi-level taper (see Fig. 16(b)).

Fig. 15. The mode conversion efficiency η as Ltp varies when TM0 modal field is launched. The
parameters are wco = 1.0µm, wside = 3.0µm, and het = 0.5H.

Fig. 16. The light propagation in the taper when the input field is TE polarization (a), and TM
polarization (b), respectively. The parameters are: H = 400nm, het = 0.5H, wco = 1.0µm, Ltp =
300µm.

In order to give a comparison, we also calculate the mode conversion efficiencies for the
cases of wco = 0.85, and 1.2µm, as shown in Fig. 17(a) and Fig. 17 (b), respectively. From
these figure, it can be seen that the mode conversion could be eliminated almost by choosing
the taper length appropriately when the rib width is chosen as 0.85 or 1.2µm. As an example,
Fig. 18(a)-(b) show the simulated light propagating in the taper with wco = 1.2µm when the
launched field is the TE0, and TM0 modal fields, respectively. Here the taper length is Ltp =
100µm. It can be seen that there is no mode conversion observed between the TM0 mode and
the TE1 mode as predicted. Therefore, one has to be careful when design a bi-level taper for
TM polarization. It also indicates that the mode conversion between the TM0 mode and the
higher-order TE mode could be cancelled or enhanced by choosing the rib width or etching
depth appropriately.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13436
Fig. 17. The mode conversion efficiency η after light propagating the taper section as Ltp varies
when the TM0 mode is launched. (a) wco = 0.85µm; (b) wco = 1.2µm. Here het = 0.5H.

Fig. 18. The light propagation in the designed adiabatic taper when the input field is TE
polarization (a), TM polarization (b), respectively. The parameters are: H = 400nm, het = 0.5H,
wco = 1.2µm, Ltp = 100µm. Here het = 0.5H.

2. Experimental observation
We fabricated some straight waveguides with tapered structures, as shown in Fig. 19(a), by
using a simple fabrication process including UV lithography, and ICP (inductively coupled
plasma) etching. The etching depth of the fabricated SOI rib waveguide is hrib = 0.5H, and H
= 400nm. The taper structure has a 1µm-wide straight waveguide in the middle while there
are tapers at both ends, which have been used often to obtain better coupling to singlemode
fibers. All the parameters for the widths of the tapers are: w1 = 1µm, w2 = 1.5µm, and w3 =
3µm, as shown in Fig. 19(a). The taper length Ltp = 100µm. Our simulation in Fig. 5 indicates
that the mode conversion in the section tapering from w2 = 3µm to w3 = 1.5µm is not
significant since the taper length is not long. Furthermore, since the TE3 mode is going to be
cutoff (see Fig. 2(b)), it is reasonable to assume that there is only TM0 mode left in the
1.5µm-wide section after light goes through the taper section at the input end. In contrast, as
shown in Fig. 8, the mode conversion between the TM0 mode and the TE1 mode is expected
to be quite significant in the second taper from w2 = 1.5µm to w1 = 1µm. Therefore, two-mode
interference happens between the TM0 mode and the TE1 mode in the 1µm-wide straight
section, which has been observed theoretically (see Fig. 10(a)-10(b)).
Figures 19(b) and 19(c) shows the measured light transmissions of a series of taper
structures when the TM0 and TE0 modes are launched respectively. The length L1 of the 1µm-
wide straight section in the middle varies from 0 to 1020µm. From these figures, it can be
seen that the spectral response is quasi-periodical due to the two-mode interference as
expected when the launched field is the TM0 mode. In contrast, there is no quasi-periodical
responses when the TE0 modal field is launched, which is also consistent with our prediction.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13437
Fig. 19. (a) The structure of the lateral taper in our experiments; (b) the measured spectral
responses for taper structures when the TM0 modal field is launched; (c) the measured spectral
responses for taper structures when the input TE0 modal field is launched; (d) the measured
and calculated quasi-FSR. H = 400nm, and het = 0.5H.

When the TM0 mode is launched, the beat length Lπ of the two-mode interference in the
1µm-wide straight section is then given by Lπ = π/[(neff_TM0-neff_TE1)k0], where neff_TM0 and
neff_TE1 are the effective indices of the TM0 mode and the TE1 mode, respectively, k0 is the
wavenumber in vacuum. Correspondingly, the free-spectral range (FSR) of the spectral
response of the taper structure is given by

λ2
λFSR = ,
(ng_TE1 − ng_TM0 ) L1 + λ

where ng_TM0 and ng_TE1 are the group indices for the TM0 mode and the TE1 mode,
respectively and they are given by ng_TM0 = neff_TM0–λ(∂neff_TM0/∂λ), and ng_TE1 = neff_TE1–
λ(∂neff_TE1/∂λ), respectively.
With this formula, the calculated quasi-FSR of the spectral response for the taper structure
as the length L1 varies, is shown in Fig. 19(c). The quasi-FSR extracted from the measured
spectral responses in Fig. 19(b) is also shown in order to give a comparison. From Fig. 19(c),
it can be seen that the theoretical and experimental results agree with each other very well.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13438
From the measurement results shown here, one should realize that the mode conversion in
a taper structure might be very serious and influence the performances of optical waveguides
and devices. It is necessary to design the taper very carefully to avoid the undesired mode
conversions.
For bi-level tapers, the mode conversion could be removed by choosing a relatively deep
rib, e.g., het = 0.6H, as indicated in Fig. 12(d). Figure 20(a) and 20(b) show the measured light
transmissions of straight waveguides with bi-level taper structures at the input/output ends
when the TM0 and TE0 mode are launched, respectively. Here het = 0.6H. The central-rib
width is wco = 1µm. From these figures, it can be seen that the spectral response is quite
smooth, which indicates that no mode conversion is observed for the case with a TM0
launched field, as the theoretical calculation predicted. Such a design has been used in a MZI-
based PBS successfully in our previous paper [35].

Fig. 20. (a) The measured spectral responses for bi-level taper structures when the TE0 modal
field is launched; (b) the measured spectral responses for taper structures when the TM0 modal
field is launched; The parameters are: H = 400nm, wco = 1µm, H = 400nm, and het = 0.6H.

3. Conclusions
In this paper, the mode conversion in tapered submicron SOI rib optical waveguides has been
studied. We have considered two typical optical waveguide tapers (i.e., regular lateral tapers,
and bi-level tapers) for submicron SOI rib optical waveguides. For a SOI rib waveguide, it is
still asymmetrical in the vertical direction even when choosing the same material for the
upper-cladding and the under-cladding. Therefore, in a tapered SOI rib optical waveguide,
mode conversion between the TM0 mode and higher-order TE modes might happens due to
the mode hybridization in some waveguide width ranges, which has been observed for both
types of waveguide tapers in our simulation. Our experimental results have also been
demonstrated to give the evidence of mode conversions. Such a mode conversion is not
desired usually because some excess loss and crosstalk is introduced in some photonic
integrated circuits. It has also been shown that such harmful mode conversion effect can be
removed almost for both types of waveguide tapers by carefully designing the taper
parameters (e.g., the width and length of the taper, the etching depth, etc). On the other hand,
our simulation results have also shown that a very high mode-conversion efficiency (close to
100%) could be achieved in both SOI rib optical waveguide tapers. Such an efficient mode
conversion could be useful for some applications, e.g., polarization rotation [30].
Acknowledgments
This research is supported by DARPA MTO under the CIPhER contract No: HR0011-10-1-
0079, the National Nature Science Foundation of China (No. 61077040), a 863 project
(Ministry of Science and Technology of China, No. 2011AA010301), Zhejiang provincial
grant (Z201121938) of China, and also supported by the Fundamental Research Funds for the
Central Universities. The authors thank Dr. Di Liang for useful discussions and Jon Peters for
fabrication.

#166272 - $15.00 USD Received 9 Apr 2012; revised 19 May 2012; accepted 22 May 2012; published 31 May 2012
(C) 2012 OSA 4 June 2012 / Vol. 20, No. 12 / OPTICS EXPRESS 13439

You might also like