Ge/SiGe Asymmetric Fabry-Perot Quantum Well .

2y ago
18 Views
2 Downloads
1.49 MB
10 Pages
Last View : 1m ago
Last Download : 3m ago
Upload by : Josiah Pursley
Transcription

Ge/SiGe asymmetric Fabry-Perotquantum well electroabsorptionmodulatorsElizabeth H. Edwards,1, Ross M. Audet,1 Edward T. Fei,1 StephanieA. Claussen,1 Rebecca K. Schaevitz,2 Emel Tasyurek,1 Yiwen Rong,3Theodore I. Kamins,1 James S. Harris,1 and David A. B. Miller11 Departmentof Electrical Engineering, Stanford University, Stanford, CA 94305, USAaddress: Corning West Technology Center, Palo Alto, CA 94304, USA3 Current address: Philips LumiLEDs, San Jose, CA 95131, USA2 Current ehe@alumni.stanford.eduAbstract:We demonstrate vertical-incidence electroabsorption modulators for free-space optical interconnects. The devices operate via thequantum-confined Stark effect in Ge/SiGe quantum wells grown onsilicon substrates by reduced pressure chemical vapor deposition. Thestrong electroabsorption contrast enables use of a moderate-Q asymmetricFabry-Perot resonant cavity, formed using a film transfer process, whichallows for operation over a wide optical bandwidth without thermal tuning.Extinction ratios of 3.4 dB and 2.5 dB are obtained for 3 V and 1.5 Vdrive swings, respectively, with insertion loss less than 4.5 dB. For 60 µmdiameter devices, large signal modulation is demonstrated at 2 Gbps, and a3 dB modulation bandwidth of 3.5 GHz is observed. These devices showpromise for high-speed, low-energy operation given further miniaturization. 2012 Optical Society of AmericaOCIS codes: (230.4110) Modulators; (230.4205) Multiple quantum well (MQW) modulators;(200.4650) Optical Interconnects; (250.0250) Optoelectronics.References and links1. D. A. B. Miller, “Device requirements for optical interconnects to silicon chips,” Proc. IEEE 97, 1166–1185(2009).2. G. T. Reed, G. Mashanovich, F. Y. Gardes, and D. J. Thomson, “Silicon optical modulators,” Nat. Photonics 4,518–526 (2010).3. J. Liu, M. Beals, A. Pomerene, S. Bernardis, R. Sun, J. Cheng, L. C. Kimerling, and J. Michel, “Waveguideintegrated, ultralow-energy GeSi electro-absorption modulators,” Nat. Photonics 2, 433–437 (2008).4. D. Feng, S. Liao, H. Liang, J. Fong, B. Bijlani, R. Shafiiha, B. J. Luff, Y. Luo, J. Cunningham, A. V. Krishnamoorthy, and M. Asghari, “High speed GeSi electro-absorption modulator at 1550 nm wavelength on SOIwaveguide,” Opt. Express 20, 22224–22232 (2012).5. Y.-H. Kuo, Y. K. Lee, Y. Ge, S. Ren, J. E. Roth, T. I. Kamins, D. A. B. Miller, and J. S. Harris, “Quantum-confinedStark effect in Ge/SiGe quantum wells on Si for optical modulators,” IEEE J. Sel. Top. Quantum Electron. 12,1503–1513 (2006).6. S. Ren, Y. Rong, T. I. Kamins, J. S. Harris, and D. A. B. Miller, “Selective epitaxial growth of Ge/Si0.15 Ge0.85quantum wells on Si substrate using reduced pressure chemical vapor deposition,” Appl. Phys. Lett. 98, 151108(2011).7. R. K. Schaevitz, J. E. Roth, S. Ren, O. Fidaner, and D. A. B. Miller, “Material properties of Si-Ge/Ge quantumwells,” IEEE J. Sel. Top. Quantum Electron. 14, 1082–1089 (2008).8. R. Schaevitz, E. Edwards, J. Roth, E. Fei, Y. Rong, P. Wahl, T. Kamins, J. Harris, and D. Miller, “Simple electroabsorption calculator for designing 1310 nm and 1550 nm modulators using germanium quantum wells,” IEEEJ. Quantum Electron. 48, 187–197 (2012).#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29164

9. L. Lever, Z. Ikonic, A. Valavanis, J. Cooper, and R. Kelsall, “Design of Ge/SiGe quantum-confined Stark effect electroabsorption heterostructures for CMOS compatible photonics,” J. Lightwave Technol. 28, 3273–3281(2010).10. J. E. Roth, O. Fidaner, R. K. Schaevitz, Y. Kuo, T. I. Kamins, J. S. Harris, and D. A. B. Miller, “Optical modulatoron silicon employing germanium quantum wells,” Opt. Express 15, 5851–5859 (2007).11. Y. Rong, Y. Ge, Y. Huo, M. Fiorentino, M. Tan, T. Kamins, T. Ochalski, G. Huyet, and J. Harris, “Quantumconfined Stark effect in Ge/SiGe quantum wells on Si,” IEEE J. Sel. Top. Quantum Electron. 16, 85–92 (2010).12. E. H. Edwards, R. M. Audet, S. A. Claussen, R. K. Schaevitz, E. Tasyurek, S. Ren., O. I. Dosunmu, M. S. Ünlü,and D. A. B. Miller, “Si-Ge surface-normal asymmetric Fabry-Perot quantum-confined Stark effect electroabsorption modulator,” in “Proc. IEEE Photonics Society Summer Topical Meetings, Playa del Carmen, Mexico,”211–212 (2010).13. S. Ren, Y. Rong, S. Claussen, R. Schaevitz, T. Kamins, J. Harris, and D. Miller, “A Ge/SiGe quantum well waveguide modulator monolithically integrated with SOI waveguides,” in “2011 8th IEEE International Conference onGroup IV Photonics (GFP),” 11–13 (2011).14. P. Chaisakul, D. Marris-Morini, M.-S. Rouifed, G. Isella, D. Chrastina, J. Frigerio, X. Le Roux, S. Edmond,J.-R. Coudevylle, and L. Vivien, “23 GHz Ge/SiGe multiple quantum well electro-absorption modulator,” Opt.Express 20, 3219–3224 (2012).15. M. Whitehead and G. Parry, “High-contrast reflection modulation at normal incidence in asymmetric multiplequantum well Fabry-Perot structure,” Electron. Lett. 25, 566–568 (1989).16. R. H. Yan, R. J. Simes, and L. A. Coldren, “Surface-normal electroabsorption reflection modulators using asymmetric Fabry-Perot structures,” IEEE J. Quantum Electron. 27, 1922–1931 (1991).17. C. C. Barron, C. J. Mahon, B. J. Thibeault, and L. A. Coldren, “Design, fabrication and characterization of highspeed asymmetric Fabry-Perot modulators for optical interconnect applications,” Opt. Quantum Electron. 25,S885–S898 (1993).18. F. B. McCormick, T. J. Cloonan, F. A. P. Tooley, A. L. Lentine, J. M. Sasian, J. L. Brubaker, R. L. Morrison, S. L.Walker, R. J. Crisci, R. A. Novotny, S. J. Hinterlong, H. S. Hinton, and E. Kerbis, “Six-stage digital free-spaceoptical switching network using symmetric self-electro-optic-effect devices,” Appl. Opt. 32, 5153–5171 (1993).19. A. G. Kirk, D. V. Plant, T. H. Szymanski, Z. G. Vranesic, F. A. P. Tooley, D. R. Rolston, M. H. Ayliffe, F. K.Lacroix, B. Robertson, E. Bernier, and D. F.-Brosseau, “Design and implementation of a modulator-based freespace optical backplane for multiprocessor applications,” Appl. Opt. 42, 2465–2481 (2003).20. M. Haney, M. Christensen, P. Milojkovic, G. Fokken, M. Vickberg, B. Gilbert, J. Rieve, J. Ekman, P. Chandramani, and F. Kiamilev, “Description and evaluation of the FAST-Net smart pixel-based optical interconnectionprototype,” Proc. IEEE 88, 819–828 (2000).21. R. M. Audet, E. H. Edwards, P. Wahl, and D. A. B. Miller, “Investigation of limits to the optical performance ofasymmetric Fabry-Perot electroabsorption modulators,” IEEE J. Quantum Electron. 48, 198–209 (2012).22. A. Nayfeh, C. O. Chui, K. C. Saraswat, and T. Yonehara, “Effects of hydrogen annealing on heteroepitaxial-Gelayers on Si: Surface roughness and electrical quality,” Appl. Phys. Lett. 85, 2815–2817 (2004).23. S. Ren, “Ge/SiGe quantum well waveguide modulator for optical interconnect systems,” Ph.D. thesis, StanfordUniversity (2011).24. P. Zouganeli, P. J. Stevens, D. Atkinson, and G. Parry, “Design trade-offs and evaluation of the performanceattainable by GaAs Al0.3 Ga0.7 As asymmetric Fabry-Perot modulators,” IEEE J. Quantum Electron. 31, 927–943 (1995).25. P. Zouganeli and G. Parry, “Evaluation of the tolerance of asymmetric Fabry-Perot modulators with respect torealistic operating conditions,” IEEE J. Quantum Electron. 31, 1140–1151 (1995).26. M. Schmidt, “Wafer-to-wafer bonding for microstructure formation,” Proc. IEEE 86, 1575–1585 (1998).27. C. L. Mitsas and D. I. Siapkas, “Generalized matrix method for analysis of coherent and incoherent reflectanceand transmittance of multilayer structures with rough surfaces, interfaces, and finite substrates,” Appl. Opt. 34,1678–1683 (1995).28. L. M. Giovane, H.-C. Luan, A. M. Agarwal, and L. C. Kimerling, “Correlation between leakage current densityand threading dislocation density in SiGe p-i-n diodes grown on relaxed graded buffer layers,” Appl. Phys. Lett.78, 541–543 (2001).29. E. Onaran, M. C. Onbasli, A. Yesilyurt, H. Y. Yu, A. M. Nayfeh, and A. K. Okyay, “Silicon-germanium multiquantum well photodetectors in the near infrared,” Opt. Express 20, 7608 (2012).30. C. Barron, C. Mahon, B. Thibeault, G. Wang, W. Jiang, L. Coldren, and J. Bowers, “Millimeter-wave asymmetricFabry-Perot modulators,” IEEE J. Quantum Electron. 31, 1484–1493 (1995).31. D. A. B. Miller, “Energy consumption in optical modulators for interconnects,” Opt. Express 20, A293–A308(2012).32. J. J. Lin, A. M. Roy, A. Nainani, Y. Sun, and K. C. Saraswat, “Increase in current density for metal contacts ton-germanium by inserting TiO2 interfacial layer to reduce Schottky barrier height,” Appl. Phys. Lett. 98, 092113(2011).33. S. A. Claussen, E. Tasyurek, J. E. Roth, and D. A. B. Miller, “Measurement and modeling of ultrafast carrier dynamics and transport in germanium/silicon-germanium quantum wells,” Opt. Express 18, 25596–25607 (2010).#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29165

34. R. K. Schaevitz, D. S. Ly-Gagnon, J. E. Roth, E. H. Edwards, and D. A. B. Miller, “Indirect absorption ingermanium quantum wells,” AIP Advances 1, 032164 (2011).35. M. S. Rouifed, P. Chaisakul, D. Marris-Morini, J. Frigerio, G. Isella, D. Chrastina, S. Edmond, X. L. Roux,J.-R. Coudevylle, and L. Vivien, “Quantum-confined Stark effect at 1.3 µm in Ge/Si0.35 Ge0.65 quantum-wellstructure,” Opt. Lett. 37, 3960–3962 (2012).36. O. Dosunmu, D. Cannon, M. Emsley, L. Kimerling, and M. Unlu, “High-speed resonant cavity enhanced Gephotodetectors on reflecting Si substrates for 1550-nm operation,” IEEE Photon. Technol. Lett. 17, 175–177(2005).37. J. Potfajova, B. Schmidt, M. Helm, T. Gemming, M. Benyoucef, A. Rastelli, and O. G. Schmidt, “Microcavityenhanced silicon light emitting pn-diode,” Appl. Phys. Lett. 96, 151113 (2010).1.IntroductionThe substantial communication bandwidth required for short-distance interconnects in nextgeneration high-performance computing systems and large-scale data centers may soon exceedthe capabilities of conventional electrical links, due to pin density constraints, power consumption, and signal integrity issues at high bit rates. Optical links, already the preferred solution atlonger length scales, have the potential to replace electrical interconnects in short distance (cmscale) inter-chip connections as well, provided they are sufficiently energy efficient [1]. Closeintegration of optical components with complementary metal oxide semiconductor (CMOS)circuitry will be required in order to decrease both the cost and the required energy per bit tolevels competitive with electrical links.There has been much recent work on silicon optical modulators for short-distance inter- andintra-chip optical interconnects [2], although the weak nature of electrooptic effects in silicontypically necessitates high-Q resonant devices that require thermal tuning. Electroabsorptionmodulators based on germanium represent another possible path towards CMOS-integrablephotonic links. In addition to promising work on waveguide modulators based on the FranzKeldysh effect in bulk SiGe [3, 4], there has been additional interest in utilizing the strongerelectroabsorption contrast afforded by the quantum-confined Stark effect (QCSE) in Ge quantum well (QW) structures. Progress in the growth [5, 6] and modeling [7–9] of Ge QWs onSi substrates has enabled the fabrication of QCSE modulators using this CMOS-compatiblematerial system [10–14].Many of these recently demonstrated modulators are waveguide-integrated devices. However, there are significant challenges associated with coupling light efficiently into and out ofsilicon waveguides given their small dimensions. An alternative geometry that may be preferable for chip-to-chip links (such as between multiple processors or between a processor andDRAM) is the asymmetric Fabry-Perot modulator (AFPM), in which both the incident andmodulated reflected beams are oriented perpendicular or near-perpendicular to the chip surface [15–17]. The challenge in using a vertical incidence geometry is that the optical interactionlength is limited by the active region thickness (typically a few microns or less). Incorporatingthe active region inside an asymmetric Fabry-Perot (AFP) resonant cavity increases the effective interaction length, and thus enables a large change in the reflected power given only amodest change in the material absorption, yielding potentially high extinction ratios even forsmall voltage swings. Asymmetric Fabry-Perot modulators exhibit low insertion loss, polarization independence, and larger alignment tolerance compared to waveguide devices. They aresuitable for dense 2-D array integration, which could enable spatially multiplexed free-spaceoptical links with thousands of channels. This would provide continued scaling of inter-chipinterconnect bandwidth without the complexity of wavelength division multiplexing schemesthat would be required by waveguide approaches [1]. To date, there have been several experimental demonstrations of highly parallelized free-space optical links for short-distance opticalinterconnects [18–20]. Modeling of AFPMs based on the QCSE in Ge/SiGe QWs indicates#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29166

(a)(b)2500600 nm n-type Si0.13Ge0.8760 nm undoped Si0.16Ge0.8430 nm undoped Si0.16Ge0.84400 nm p-type Si0.12Ge0.88200 nm p-type Si0.12Ge0.88, annealed 750 C200 nm p-type Si0.12Ge0.88, annealed 800 C 1Absorption (cm )15 x 10 nm Ge QWs with17 nm Si0.19Ge0.81 barriers0V1V2V3V4V5V6V7V8V200015001000500200 nm p-type Si0.12Ge0.88, annealed 850 Cp-type Si wafer0140014501500Wavelength (nm)1550Fig. 1. (a) Epitaxial layer structure consisting of fifteen Ge/SiGe quantum wells grown ona p-type silicon wafer with a fully relaxed SiGe buffer layer grown using a three-stagehydrogen annealing process. (b) Absorption spectrum of the epitaxial structure, deducedfrom photocurrent measurements. The effective absorption coefficient is calculated fromthe absorption per pass divided by the thickness of the epitaxial region.that these structures can be made sufficiently small to have attractive switching energies andmodulation rates while maintaining good extinction ratios and low insertion losses [21], andfirst demonstrations of Ge/SiGe AFPMs have shown low-speed (DC) reflectance modulation indevices based on both a double silicon-on-insulator (SOI) and film transfer approach [12].In this paper, we present results from Ge/SiGe QCSE AFPMs fabricated using a film transferprocess. The moderate-Q resonant cavity enables substantial reflectance contrast with low insertion loss over a large optical bandwidth. We obtain open eye diagrams at 2 Gbps and a 3 dBmodulation bandwidth of 3.5 GHz on 60 µm diameter modulators, suggesting that smaller devices will be capable of low-power, high-speed operation at tens of gigahertz.2.2.1.Design and fabricationEpitaxyThe Ge/SiGe QW epitaxial structure from which the modulators were processed was growndirectly on p-type Si wafers using an Applied Materials Centura Epi reduced pressure chemicalvapor deposition (RPCVD) system, operated at a chamber pressure of 40 Torr and a growthtemperature of 400 C. The layer structure is shown in Fig. 1(a).The modulator uses a p-i-n structure, with the QW active region situated inside the intrinsicregion of the reverse-biased diode. The structure is grown on top of a fully relaxed Si0.12 Ge0.88p-type buffer layer that reduces the propagation of crystal defects arising from the 4% latticemismatch between Si and Ge. The boron-doped buffer is grown using several intermediate hydrogen annealing steps, as has been previously demonstrated for the growth of bulk Ge onSi [22]. In contrast to graded buffers grown via low-energy plasma-enhanced chemical vapordeposition [14], buffers grown using the multiple hydrogen anneal process can be made extremely thin while preserving strong QCSE [23].The active region consists of fifteen 10 nm Ge QWs with 17 nm Si0.19 Ge0.81 barriers. Thinundoped spacer regions above and below the active region prevent dopants from migrating intothe QWs. At the top of the structure is an arsenic-doped n-type layer for making electrical#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29167

(a)(b)Incident light /Modulated reflected light100 μmTop DBRp SiGeGe/SiGe QWsContactsn SiGeBack DBRPyrex handle waferFig. 2. (a) Schematic illustration of the asymmetric Fabry-Perot modulator. Light entersfrom the top. A voltage is applied across a p-i-n diode containing the quantum wells insidethe intrinsic region. Field-dependent absorption in the QWs modulates the intensity of thereflected light. The asymmetric Fabry-Perot cavity is formed by DBR mirrors surroundingthe SiGe. The device is bonded to a Pyrex handle wafer. (b) Microscope image of deviceshowing the AFP modulator, electrically contacted by a high-speed probe.contact. Following the epitaxial growth, a 15-second, 750 C post-growth rapid thermal annealwas performed.Absorption spectra derived from photocurrent measurements are shown in Fig. 1(b). Thesespectra are taken from test structures fabricated on a second, nominally identically grown wafer.The built-in field in the p-i-n diodes ensures that even at small reverse bias voltages, nearly all ofthe photogenerated carriers are collected through the contacts. Thus, the photocurrent measurement can be directly correlated with the optical absorption. An absorption coefficient contrastapproaching 5 dB is obtained for a 1 V drive swing across a broad range of wavelengths, giventhe proper choice of bias voltage.2.2.Asymmetric Fabry-Perot cavity designThe asymmetric Fabry-Perot modulator, whose basic structure is shown in Fig. 2(a), consists oftwo mirrors surrounding a QW region, in which the absorption can be altered by application ofan electric field. At normal incidence, the fraction of reflected optical power on resonance, RT ,is R f Rb,eff 2(1)RT , 1 R f Rb,eff 2where R f is the front mirror reflectance, and the effective back mirror reflectance Rb,eff is givenby Rb,eff Rb exp ( 2α L) [16]. Here, Rb is the reflectance of the back mirror (ideally nearunity), α is the effective power absorption coefficient inside the cavity, and L is the cavitylength. Changing α and hence Rb,eff will alter RT . A critically coupled condition with zeroreflectance is achieved when Rb,eff R f , or, equivalently, when the effective absorption is α ln Rb /R f /(2L).Design considerations for AFPMs have been discussed in depth previously for III-V baseddevices. Many of these results also apply to Ge/SiGe devices, including investigations of designoptimization [17, 24] and studies of sensitivity to cavity length variations across a wafer aswell as temperature [25]. Additionally, simulation work has investigated the effects of beamdiffraction as well as both lateral and angular misalignments for Ge/SiGe modulators similar to#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29168

those presented here [21].Based on the absorption properties of the Ge/SiGe QW epitaxy, as shown in Fig. 1(b), aresonator structure with 93% top mirror reflectance and 99.8% bottom mirror reflectance waschosen here, such that the AFP matching condition RT 0 in Eq. (1) could be satisfied in themodulator’s ”off” state.2.3.Device fabricationIn order to form the AFP resonant cavity with DBR mirrors surrounding the QW structure, afilm transfer process is used. Fabrication begins by chemical mechanical polishing (CMP) of thetop surface of the SiGe epitaxy to yield less than 1 nm root mean square (RMS) roughness, asdetermined by atomic force microscopy. A three-layer distributed Bragg reflector (DBR) consisting of alternating quarter-wavelength layers of amorphous silicon (a-Si) and silicon dioxide(SiO2 ) forms the high-reflectance back mirror of the device. The a-Si layers are approximately98 nm thick and are deposited by electron beam evaporation, while the SiO2 layers are approximately 243 nm thick and are formed using plasma-enhanced chemical vapor deposition(PECVD).The top surface of this DBR mirror is then bonded to a Pyrex 7740 carrier wafer using ananodic bonding process [26]. The majority of the silicon substrate is removed by wafer grinding, and the remaining 25-50 µm of material is removed using a potassium hydroxide wetetch that stops at the lower Si/SiGe interface. Most of the Si0.12 Ge0.88 buffer layer, where theepitaxy’s crystal defects are concentrated, is then removed via an iterative CMP process untilspectrophotometer measurements indicate the cavity resonance is positioned near the wavelength corresponding to maximum electroabsorption contrast in the quantum wells.To provide electrical isolation, mesas are dry etched using an SF6 chemistry through theepitaxy into the n-doped layer (closest to the Pyrex carrier wafer). A two-layer a-Si/SiO2 DBRtop mirror is then deposited, with the top a-Si layer thickness determined by transfer matrixsimulations [27] such that the top mirror reflectance target of 93% is achieved. A full 2-perioda-Si/SiO2 DBR, deposited on SiGe, has a reflectance of 98% in air. By depositing a 35 nm topa-Si layer instead of the full 98 nm quarter-wave a-Si layer, the reflectance is reduced to 93%at 1440 nm.Following deposition of the mirror, vias are dry etched through the insulating top mirror layers to reach the p- and n- doped regions. Electrical contacts are formed by e-beam evaporationof Ti/Pt/Au followed by a liftoff step. The structures are electrically contacted using standard100-um pitch ground-signal-ground high-speed probes.3.3.1.Device characterizationDC measurementsThe fabricated modulators exhibit good electrical performance, with IV curves of the p-i-ndiodes showing a reverse breakdown voltage near 15 V. The dark current is 4 mA/cm2 at 1 Vreverse bias and 17 mA/cm2 at 5 V reverse bias, indicating low defect density in the epitaxiallayers, as well as good surface passivation [28]. These values compare favorably to a recentlyreported Ge/SiGe MQW photodetector fabricated using a similar growth method, which had19 mA/cm2 dark current at 1 V reverse bias [29].The basic experimental setup (used for both DC and high-speed measurements) is illustratedin Fig. 3. To perform DC photocurrent and reflection measurements, light from a 1369-1481 nmfiber-coupled tunable laser source is sent through a polarization controller then collimated usinga pigtailed fiber collimator. The linearly polarized light then passes directly through a polarizingbeam splitter and a quarter-wave plate. The beam is focused onto the surface of the deviceat normal incidence using a Mitutoyo 10x long working distance near-infrared microscope#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29169

Fiber collimatorMirrorPolariza illoscopeDC biassupply10xobjec velensPickoffTBiasT3.5 GHzpa erngeneratorAFPM chipFig. 3. Simplified diagram of the experimental setup. The instrumentation shown is forlarge-signal high-speed measurements. DC measurements and small-signal high-speedmeasurements use the same optical train but different measurement equipment, as describedin Sec. 3.1 and 3.2, respectively.objective. The reflected light travels back through the microscope objective and quarter-waveplate. By passing through the quarter-wave plate twice, the reflected beam polarization is rotated90 degrees relative to the incident polarization. The polarizing beam splitter thus deflects thereflected light, and it is focused onto a germanium detector.To measure the DC behavior of the device, a voltage is applied across the modulator. Thetunable laser’s internal modulation is used to generate an incident optical signal modulatedat 1.5 kHz. Photocurrent corresponding to optical absorption in the modulator under test iscollected using a current preamplifier then measured using a lock-in detection scheme. Thereflected optical signal intensity is measured by attaching the output of the Ge photodetector toa second lock-in amplifier.Figure 4(a) shows the collected photocurrent as a function of applied reverse bias voltage.Around the cavity resonance at 1430 nm, absorption in the QWs decreases as the applied reversebias is increased, and thus the collected photocurrent decreases. The full width half maximum(FWHM) of the cavity resonance is 20 nm, corresponding to a Q 70.The measured reflectance of a 100 µm diameter device (referenced to a 99.9 % reflective broadband dielectric mirror) is shown in Fig. 4(b). As the absorption inside the AFP cavity (measured by photocurrent) decreases, the device reflectance increases. The minimum reflectance of approximately 19% at 1429 nm is achieved for a reverse bias of 0.5 V, where thecollected photocurrent is maximized. The nonzero reflectance indicates that the AFP matchingcondition has not been achieved. This is because the front mirror reflectance is lower than optimal given the amount of electroabsorption at this wavelength. Nonetheless, a useful change inreflectance is still obtained.Figure 4(c) shows the extinction ratio (ER) corresponding to modulation between differentapplied voltages. For a 3 V swing, between 0.5 and 3.5 V reverse bias, an ER of 3.4 dB isachieved. A 2V swing, between 0.5 and 2.5 V, yields an ER of 2.8 dB, while a 1.5 V swingbetween 1V and 2.5V results in an ER of 2.5 dB. There is no observable dependence of thecontrast ratio on the incident optical power, up to a maximum tested power of approximately3 mW (the maximum obtainable from the tunable laser), suggesting that these power levels#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29170

1Photocurrent (arb. u.)0.8(a)0.60.4(b)0.60.5 V1.5 V2.5 V3.5 V0.40.20.2140014201440Wavelength (nm)Extinction Ratio (dB)013800.8Reflectance0.5 V1.5 V2.5 V3.5 V143(c)146001380140014201440Wavelength (nm)14600.5 / 1.5 V0.5 / 2.5 V0.5 / 3.5 V210 1140014201440Wavelength (nm)1460Fig. 4. DC modulator performance. (a) Photocurrent spectra for different applied reversebias voltages. (b) Corresponding reflection spectra for the same set of applied voltages.(c) Spectra showing the extinction ratio versus wavelength for 1, 2, and 3 V swings, withstarting reverse bias of 0.5 V.are below the saturation point of the excitonic absorption. Furthermore, these values are unchanged for electrical modulation frequencies ranging from 10 Hz - 10 MHz, indicating thatthermal effects do not substantially impact the modulation depth at these moderate incidentpower levels.For practical devices, the absolute change in reflectance, ΔR, can be just as important as theextinction ratio. Due to the relatively low insertion loss of these modulators, a large absolutereflectance modulation on resonance ΔR 22.9% (from 42.3% to 19.4%) is achieved for a 3 Vswing, and ΔR 15.4% is obtained for a 1.5 V swing. The insertion loss in the high reflectancestate is 3.7 dB and 4.4 dB for the 3 V and 1.5 V swings, respectively.Because the AFP cavity has Q 70, the optical bandwidth over which substantive modulation can be achieved is large. For a 3 V swing (between 0.5 and 3.5 V), a 2 dB extinction ratiois observed over an optical bandwidth of 1.6 THz (1423-1434 nm), as can be seen in Fig. 4(c).For a 1.5 V swing (between 1 V and 2.5 V), a 2 dB ER is maintained over 1.0 THz (1426-1433nm).3.2.High-speed operationTo characterize the high-speed performance of the modulators, measurements were made usingthe setup depicted in Fig. 3. Large signal modulation was demonstrated using an HP 8133A3.5 GHz pulse generator producing a non-return to zero (NRZ) 223 1 pseudo-random binary#177465 - 15.00 USD(C) 2012 OSAReceived 1 Nov 2012; revised 7 Dec 2012; accepted 10 Dec 2012; published 17 Dec 201217 December 2012 / Vol. 20, No. 27 / OPTICS EXPRESS 29171

3(a)(b)Response (dB)0 3 6 9 120.21Frequency (GHz)10Fig. 5. High-speed measurement results for a 60 µm diameter device, with 2.2 V DC reversebias, at a wavelength of 1436 nm. (a) Open eye diagram at 2 Gbps, 1 V swing. (b) Smallsignal measurement of the optical modulation showing a 3 dB modulation bandwidth of3.5 GHz, and measurable response beyond 10 GHz.sequence (PRBS). Both the pulse generator and a DC power supply (to provide a bias voltage)were connected to the device using a bias-T. An Agilent 86100A Infiniium Digital Communications Analyzer (DCA), connected via a 20 dB pickoff-T, was used to monitor the voltageapplied at the modulator. The laser was operated in continuous wave, set to a fixed wavelength,with an optical power of approximately 2 mW incident upon the modulator. The reflected lightwas focused onto a New Focus 1 GHz photoreceiver, which was connected to one of the electrical input channels of the DCA.For this measurement, 60 µm diameter devices were tested. An open eye diagram at 2 Gbpsdata rate with 2.2 V reverse bias and 1 V peak-to-peak swing is shown in Fig. 5(a). The slowrise and fall times are partly due to the limited bandwidth of the photoreceiver used in thismeasurement.The small-signal response of the devices was also characterized. The electrical output of anHP 8703A 20 GHz Lightwave Component Analyzer (LCA) provided the modulating signal.The reflected beam was coupled into a single mode fiber, and was fed into the optical inputof the LCA. The response curve is shown in Fig. 5(b)

Ge/SiGe asymmetric Fabry-Perot quantum well electroabsorption modulators Elizabeth H. Edwards,1, Ross M. Audet,1 Edward T. Fei,1 Stephanie A. Claussen,1 Rebecca K. Schaevitz,2 Emel Tasyurek,1 Yiwen Rong,3 Theodore I. Kamins,1 James S. Harris,1 and David A. B. Miller1 1Department of Electrical Engineering, Stanford University,

Related Documents:

Theory Resonator losses, outcoupled light, resonance frequencies, . The spectral response of a Fabry-Pérot resonator is based on interference between the light launched into it and the light circulating in the resonator. . The decaying electric field at frequency is represented by a damped harmonic oscillation with an initial amplitude of .

3D SCM dummy gate recess SiN SiN OX OX OX SiN recess Selective SiNx removal for 3D-NAND fabrication 16 Si/SiGe, GAA Fin reveal (SiO 2 /SiN etch) Si, SiGe, Si/SiGe fins STI GAA inner spacer EB [SiN(OC) etch] SiN liner CESL removal (SiN etch) Mertens et al., IEDM (2017). Pacco et al., SPCC (2018) Isolation recess (SiO 2 /SiN etch) Bottom .

volatile memories,6,7 multiple gate field effect transistors (including FinFETs),8-11 and nano-wires.12,13 Achieving a direct bandgap in SiGe core-shell nanowires depends on, among other things, the nanowire's orientation and shell thickness.14-16 Local oxidation of SiGe has long been pro-posed as a method to manipulate the Ge content in .

1.4.3 Enantiopure p-aminoalcohols in the asymmetric Michael 27 addition. 1.4.4 Cz-symmetric ligands in the catalytic enantioselective 32 Michael addition. 1.5 Asymmetric epoxidations 35 1.5.1 The Katsuki-Sharpless asymmetric epoxidation 35 1.5.2 The Oxone asymmetric epoxidation 38 1.5.3 The Julia-Colonna asymmetric epoxidation 41

2.6 Chiral Catalyst-Induced Aldehyde Alkylation: Asymmetric Nucleophilic Addition 107 2.7 Catalytic Asymmetric Additions of Dialkylzinc to Ketones: Enantioselective Formation of Tertiary Alcohols 118 2.8 Asymmetric Cyanohydrination 118 2.9 Asymmetric a-Hydroxyphosphonylation 124 2.10 Summary 127 2.11 References 127 3 Aldol and Related Reactions 135

A surprise attendee at the museum’s opening reception was Ross Perot and his wife Margot. Perot was an independent presidential candidate in 1992, with Admiral Stockdale as his running mate. Perot has always been an inspiring supporter of

chiral cation Q of a PTC and an enolate and subsequent asymmetric α-functionalization with an electrophile (E). A lot of different examples for such asymmetric α-functionali-zation reactions of prochiral nucleophiles under asymmetric chiral cation-based phase-transfer catalysis have been reported so far [9-22].

Department of Aliens LAVRIO (Danoukara 3, 195 00 Lavrio) Tel: 22920 25265 Fax: 22920 60419 tmallod.lavriou@astynomia.gr (Monday to Friday, 07:30-14:30) Municipalities of Lavrio Amavissos Kalivia Keratea Koropi Lavrio Markopoulo . 5 Disclaimer Please note that this information is provided as a guide only. Every care has been taken to ensure the accuracy of this information which is not .