You are on page 1of 2

© 2008 OSA / CLEO/QELS 2008

a505_1.pdf
CFH4.pdf

High-speed silicon electro-optical modulator that can be


operated in carrier depletion or carrier injection mode*
S. J. Spector1*, M. W. Geis1, M. E. Grein1, R. T. Schulein1, J. U. Yoon1, D. M. Lennon1, F. Gan2, G.-R. Zhou2,
F. X. Kaertner2, and T. M. Lyszczarz1
1
Lincoln Laboratory, Massachusetts Institute of Technology, Lexington, MA 02420-9108 USA
2
Massachusetts Institute of Technology, Cambridge, MA 02139-4327 USA
spector@ll.mit.edu

Abstract: A high-speed silicon optical modulator has been demonstrated which can operate either
in forward bias using carrier injection, or in reverse bias using carrier depletion. In forward bias,
the device requires less than 10 mW of drive power, but has a low bandwidth of 100 MHz. In
reverse bias, the device has a nearly flat response to 18 GHz, but requires 700 mW for large
modulation depths.
¤2008 Optical Society of America
OCIS codes:

High-speed optical modulators are an important component in the goal of monolithically integrating silicon photonic
devices. The most successful silicon optical modulators are based on using carrier density to control the optical
index of the device. Modulators have used carrier depletion in a reversed biased PN diode [1] or carrier injection in
a forward biased PIN diode [2]. Demonstrated here is a Mach-Zehnder interferometer modulator which can be
operated either as a forward biased PIN diode or as a reverse biased PN diode. Depending on the application,
different modes of operation may be preferable. In addition, combining both types of operation into one device
allows a direct comparison between the two methods of optical modulation.
Although less compact than modulators based on resonant devices [2], the Mach-Zehnder based modulator is
simpler to analyze and is less sensitive to laser wavelength and resonator temperature variations. The two arms of
the modulator are nominally identical with diodes fabricated in both arms. The junctions are connected so that the
modulator can be driven differentially; however, for all the results listed here only a single arm of the modulator is
used.
A diagram of the active part of the modulator is shown in Fig. 1. The sidewalls of the waveguide are moderately
doped to a concentration of 1018 cm-3, and the center of the waveguide is doped n-type to a lighter concentration of
1017 cm-3. These doping concentrations were chosen to produce a diode that would simultaneously cause a
reasonable phase shift without excessive optical loss [3]. When operated in reverse bias, a depletion region forms at
the junction at one side of the waveguide. As the reverse bias voltage is increased, this depletion region becomes
larger and extends into the center of the waveguide. When operated in forward bias, extra carriers are injected into
the waveguide. The carrier concentration when operated in forward bias is greater than 1018 cm-3 and therefore
dominates over the 1017 cm-3 doped level.

Metal Via Si Waveguide: Fig. 1. A cross-sectional diagram of the active part of the
500 x 220 nm modulator fabricated on a silicon-on-insulator substrate.
The center part of the waveguide is doped weakly n-type
Si (1017 cm-3 of phosphorus). The left side of the structure is
1 Pm doped n-type (phosphorus), and the right side of the
structure is doped p-type (boron). The doping
oxide 50 nm concentrations are higher away from the waveguide to
improve conductivity without affecting optical loss.
A highly doped (10 cm ), 50-nm-thick silicon layer connects the waveguide to electrical contacts located 1 Pm
19 -3

from the waveguide. Even higher doping concentrations of 1021 cm-3 are used under the metal contacts to assure
good ohmic contacts. Functionally, this device is similar to vertical PN devices demonstrated [1], but the orientation
of the junction is sideways. This orientation is easier to fabricate because no epitaxial overgrowth is necessary. The
current design does not include traveling wave electrodes. This is expected to limit the high speed operation of our
devices, particularly as the devices become longer.
The modulators were tested under DC operation, and, as expected, the devices had much greater sensitivity when
operated in forward bias than when operated in reverse bias. Under forward bias a 0.25-mm-device can achieve a 
phase shift with a change in input voltage of 0.1 V (with a bias of 1V). This corresponds to an extremely low VL
of 0.0025 V cm. Because this is a low impedance device, the current increases from 3 mA to 10 mA for this  phase
shift. Under reverse bias, a 5 mm device can achieve a  phase shift with a change in input voltage of 10.5 V,

978-1-55752-859-9/08/$25.00 ©2008 IEEE


© 2008 OSA / CLEO/QELS 2008
a505_1.pdf
CFH4.pdf

corresponding to a VL of 5.3 V cm. This is comparable to vertical diode devices reported in the literature where a
VL of 4 V cm was achieved [1].
Fig 2 shows the small signal response of a 0.5-mm-long modulator under forward bias operation and 1.0-mm-
long modulator under reverse bias operation. The response curve for forward bias shows a relatively small
bandwidth of about 100 MHz, likely limited by carrier lifetime. However the device is orders of magnitude more
sensitive in forward bias than in reverse bias at low frequencies. When operated in reverse bias the device shows a
relatively flat response to 18 GHz. There is about 10 dB of roll-off between 1-18 GHz, but 4 dB of this roll-off is
attributable to the cables used to drive the modulator. At all frequencies below 15 GHz the response of the device is
greater in forward bias than in reverse bias. Above 20 GHz the responses of both modes of operation are roughly
equal and show a significant drop. The poor response of the device above 20 GHz, as well as the roll-off at lower
frequencies under reverse bias operation, is likely due to the lumped element electrode design.
The required power to drive the modulator at different frequencies was also measured. When operated in
forward bias at frequencies up to 1 GHz, the modulator can be turned fully on and off with less than 10 mW of RF
input, which is substantially less than other Mach-Zehnder based modulators. Despite the low bandwidth of the
device, reasonable modulation depths can be achieved at high frequencies, with reasonable input powers. At 18
GHz, a 40% modulation depth is achieved with 700 mW of RF input, (where modulation depth is defined as Vmax -
Vmin of the output waveform divided by the Vmax - Vmin achieved when the device is fully turned off and on at low
frequencies). Similar results were achieved for the reverse biased device at high frequencies, which reached a
modulation depth of 60% at 18 GHz with 700 mW of RF input power, as shown in Fig. 3. This performance is
comparable to that achieved in [1], despite the simple electrode design.

1
-10 Intensity (arb. units)
Forward Bias 0.8
-20 Reverse Bias
-30 0.6 18 GHz
S21 (dB)

-40 0.4 6 GHz


-50
0.2
-60
-70 0
-80 0 0.1 0.2 0.3 0.4
0.01 0.1 1 10 100 Time (ns)
GHz
Fig 3. Output of the modulator at 6 and 18 GHz under reverse bias.
Fig. 2. Frequency response of the modulator under At 18 GHz, a modulation depth of 60% is achieved with 700 mW of
reverse bias and forward bias operation. RF input. At 6 GHz the same amount of RF input saturates the
modulator, distorting the waveform.

In conclusion, a modulator has been demonstrated which functions in forward bias (using carrier injection) or in
reverse bias (using carrier depletion). This device is simpler to fabricate than other reverse bias based devices
demonstrated, and this modulator design performs similarly. In the forward bias, the bandwidth of the device is
likely limited by the carrier lifetime. In reverse bias, the bandwidth is limited by electrode design. Despite the
lower bandwidth of the forward bias operation, it demonstrates greater sensitivity up to 15 GHz. Therefore,
depending on the application, it may be desirable to operate the device in either mode. In addition, a small change
in the fabrication of the device will cause it to function as photodetector [4] thus further demonstrating the flexibility
of this design.
Acknowledgment
The Lincoln Laboratory portion of this work was sponsored by the EPIC Program of the Defense Advanced
Research Projects Agency under Air Force Contract FA8721-05-C-0002. Opinions, interpretations, conclusions,
and recommendations are those of the authors, and do not necessarily represent the view of the United States
Government.
References
[1] A Liu, L. Liao, D. Rubin, H. Nguyen, B. Ciftcioglu, Y. Chetrit, N. Izhaky, and M. Paniccia, “High-speed optical modulation based on
carrier depletion in a silicon waveguide," Opt. Express 15, 660-668 (2007).
[2] Q. Xu, S. Manipatruni, B. Schmidt, J. Shakya, and M. Lipson “12.5 Gbit/s carrier-injection-based silicon micro-ring silicon modulators ,"
Opt. Express 15, 430-436 (2007).
[3] F. Gan and F. X. Kärtner, “Low insertion loss, high-speed silicon electro-optic modulator design,” in Integrated Photonics Research and
Applications/ Nanophotonics 2006 Technical Digest (OSA, Washington, DC, 2006), ITuB2.
[4] M. W. Geis, S. J. Spector, M. E. Grein, R. T. Schulein, J. U. Yoon, D. M. Lennon, S. Deneault, F. Gan, F. X. Kaertner, and T. M.
Lyszczarz, “CMOS-compatible all-Si high-speed waveguide photodiodes with high responsivity in near-infrared communication band,"
IEEE Photon. Technol. Lett. 19, 152-154.

You might also like