Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

344

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

Slow Light Enhanced Nonlinear Optics in Silicon Photonic Crystal Waveguides


Christelle Monat, Member, IEEE, Bill Corcoran, Student Member, IEEE, Dominik Pudo, Member, IEEE, Majid Ebnali-Heidari, Christian Grillet, Mark D. Pelusi, Member, IEEE, David J. Moss, Senior Member, IEEE, Benjamin J. Eggleton, Senior Member, IEEE, Thomas P. White, Liam OFaolain, and Thomas F. Krauss

AbstractWe present a summary of our recent experiments showing how various nonlinear phenomena are enhanced due to slow light in silicon photonic crystal waveguides. These nonlinear processes include self-phase modulation (SPM), two-photon absorption (TPA), free-carrier related effects, and third-harmonic generation, the last effect being associated with the emission of green visible light, an unexpected phenomenon in silicon. These demonstrations exploit photonic crystal waveguides engineered to support slow modes with a range of group velocities as low as c/50 and, more crucially, with signicantly reduced dispersion. We discuss the potential of slow light in photonic crystals for realizing compact nonlinear devices operating at low powers. In particular, we consider the application of SPM to all-optical regeneration, and experimentally investigate an original approach, where enhanced TPA and free-carrier absorption are used for partial regeneration of a high-bit rate data stream (10 Gb/s). Index TermsIntegrated optics, nonlinear optics, silicon, slow wave structures.

I. INTRODUCTION

NTEGRATED nanophotonics, particularly on the silicon on insulator (SOI) platform, has recently enabled the successful realization of various nonlinear optical devices on chip [1]. At the core of these advances lies the tight connement of light within submicrometer-sized silicon nanowires that substantially increases the optical energy density, allowing nonlinear optical effects to occur at greatly reduced input powers. Planar silicon photonic crystal (PhC) waveguides represent another integrated

Manuscript received May 7, 2009; revised August 26, 2009. First published November 6, 2009; current version published February 5, 2010. This work was supported in part by the Australian Research Council, through its Federation Fellow, Centre of Excellence and Discovery Grant programs as well as the International Science Linkages program of the Australian Department of Education, Science and Technology, the European Union EU-FP6 Marie Curie Fellowship project Slow Light Propagation in Photonic cRYstals (SLIPPRY), the EU-FP6 Network of Excellence European Network of Excellence on Photonic Integrated Compnentsa & Circuits (ePIXnet), and the EU-FP6 Slow Photon Activated SwitcH (SPLASH) project. C. Monat, B. Corcoran, D. Pudo, M. Ebnali-Heidari, C. Grillet, M. D. Pelusi, D. J. Moss, and B. J. Eggleton are with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems (CUDOS), Institute for Photonics and Optical Sciences (IPOS), School of Physics, University of Sydney, Sydney, N.S.W. 2006, Australia (e-mail: monat@physics.usyd.edu.au; billc@physics.usyd.edu.au; dpudo@mit.edu; ebnali_majid@yahoo.com; grillet@physics.usyd.edu.au; m.pelusi@physics.usyd.edu.au; dmoss@physics.usyd.edu.au; egg@physics. usyd.edu.au). T. P. White, L. OFaolain, and T. F. Krauss are with the School of Physics and Astronomy, University of St. Andrews, St. Andrews KY169SS, U.K. (e-mail: tom.white@st-andrews.ac.uk; jww1@st-andrews.ac.uk; tfk@standrews.ac.uk). Color versions of one or more of the gures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identier 10.1109/JSTQE.2009.2033019

platform, where light is conned by total internal reection within the vertical silicon/air step index waveguide while experiencing the periodic modulation of the refractive index in the plane of the structure. While this platform provides a similar degree of optical connement to that of silicon nanowires, it offers the additional capability to reduce signicantly the speed of light propagationcreating the so-called slow light [2], [3]. In the context of integrated nonlinear optics, slow light holds the promise for further enhancing optical nonlinear effects [4], [5]. Slow light enhancement of nonlinear effects is, however, often hard to take advantage of due to various challenges. For instance, coupling energy into slow light modes has posed some difculties, which have been addressed recently [6][8]. Another issue lies with propagation loss, which tends to increase in the slow light regime and has been reported to scale with the square of the inverse group velocity [9]. More recently, it has been suggested [10] that slow light modes away from the Brillouin zone edge may not suffer so much loss, with a different dependence on group index than band edge modes. Another signicant challenge has been the inherent high group velocity dispersion (GVD) associated with common approaches to producing slow light in PhCs [11], [12]. Large GVD is generally undesirable, as it reduces the peak intensity of pulses through temporal broadening, effectively compromising the nonlinear enhancement afforded by slow light pulse spatial compression [13]. As a result, reports demonstrating slow light enhancement using dispersive band edges in IIIV semiconductors have been restricted to group velocities no smaller than c/20 [14] and c/8 [15], respectively. Recently, however, the exibility of the PhC geometry has been exploited to create slow light modes with little to no dispersion [16][19]. The ability to dramatically modify the PhC dispersive properties through subtle structural changes makes PhCs a unique and powerful platform for controlling optical dispersion. These PhC waveguides can support guided modes with an engineered low group velocity that is almost constant over a signicant bandwidth (510 nm). This is crucial for realizing nonlinear devices that take full advantage of slow light nonlinear enhancement, as reported in early 2009 by our group [20] and other authors [21]. In this paper, we present the results of our recent studies on the slow light enhancement of nonlinear effects in dispersionengineered silicon PhC waveguides. We report the enhancement of self-phase modulation (SPM), two photon absorption (TPA) and free-carrier effects (Section IV), by comparing the spectral signature of picosecond pulses propagating in short PhC

1077-260X/$26.00 2009 IEEE

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

345

waveguides (80 m) over a range of (low) group velocities (between c/20 and c/50), respectively. These results are supported by numerical simulations that allow us to determine the contribution of the different nonlinear effects (Kerr, TPA, free carriers), and understand their respective dependence on slow light. We then discuss the potential of this platform for realizing compact nonlinear devices, in particular for optical regeneration. We also present an original approach to regenerate a distorted 10 Gb/s optical data signal (Section V) using slow light enhanced nonlinear loss, representing the rst time PhC waveguides have been used for signal processing in system experiments at high bit rates. By analyzing eye diagrams and bit error rates (BERs), we show that the asynchronous RF amplitude distortion imposed on the logical ones is notably reduced by our device. Finally, we investigate another interesting nonlinear process, third-harmonic generation (THG), associated with the emission of visible (green) light in silicon (Section VI). We similarly nd a strong dependence of this nonlinear phenomenon on group velocity. II. SLOW LIGHT ENHANCEMENT OF NONLINEAR PROCESSES Enhancement of optical nonlinear processes through slow light is commonly attributed to two effects, which we explain in the following by using simple arguments. First, light propagating at low group velocity spends longer in a waveguide of given length than if the light was traveling fast. This in turn increases the interaction time between the slow light and the structure it travels through, therefore, increasing all effects relying on lightmatter interaction (such as Kerr effects). To illustrate this, we take the example of SPM. In a lossless medium, the nonlinear phase shift (m ax ) acquired by a pulse over a given length of propagation L is [22] m ax = P0 L (1)

Fig. 1. (a) Schematic presentation of the slow light spatial pulse compression effect for a pulse propagating from a fast medium (region 1blue) into a slow medium (region 2red). (b) Illustration of spatial pulse compression in the case of the transition from a nanowire into a slow light PhC waveguide (S = 5).

with P0 being the peak pulse power and the nonlinear parameter given as n2 0 = (2) cAe where n2 is the nonlinear index, 0 is the radial frequency of light used, c is the speed of light in vacuum, and Ae is the effective area of the guided mode. Clearly, as the propagation length increases, so does the maximum achievable nonlinear phase shift. In PhC waveguides, slow light is generally achieved through coherent interaction between forward and backward propagating modes that are coupled via the periodic lattice [23]. The resulting multiple reections undergone by the slow light mode back and forth along the waveguide effectively increase the interaction length relative to the physical waveguide length, by the slow down factor S nslow vfast = (3) S= vslow nfast where n is the group index of the guided mode and v is the group velocity. In calculating the phase shift given by (1), this effect can be taken into account by considering L as the effective length rather than the physical distance, or by equivalently, increasing by a factor S.

The second mechanism for nonlinear enhancement is due to the optical energy density increase associated with slow light [23]. As a pulse travels from a fast to a slow waveguide, the front of the pulse slows down while the tail of the pulsestill in the fast modemoves closer to the front of the pulse, as schematically represented in Fig. 1(a). This induces a spatial compression of the pulse energy, which in turn increases the peak of electric eld intensity carried by the slow light pulse. This can be expressed by considering that the total energy of the pulse is conserved, i.e., U dL is equal in both slow and fast waveguides when integrating the optical energy density per unit length (U) along the propagation length L. This is equivalent to saying that the total ux of energy is conserved, i.e., dt also expressed as U vdt, with v being the group velocity of the pulse. The rst conservation argument allows us to write, to rst approximation Up eak1 L1 = Up eak2 L2 (4)

with L being the pulse spatial length. The second argument similarly gives us Up eak1 v1 1 = Up eak2 v2 2 (5)

with being the pulse temporal duration. Because 1 = 2 (in the absence of dispersion) L1 v1 Up eak2 = = . Up eak1 L2 v2 (6)

346

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

The energy density (proportional to the electric eld intensity) is, therefore, compressed as the ratio of the group velocities, and its peak is increased by the same factor, as illustrated in Fig. 1(a) (b). Referring back to the SPM example (1), this effect can be taken into account either by increasing the peak power P0 by a factor S or equivalently again by a factor S. Note that while all (linear and nonlinear) effects are enhanced through the increase in interaction time, the increase of the energy density will only enhance the nonlinear processes that effectively depend on the electric eld intensity, resulting in an S- and S 2 - dependence for linear and nonlinear processes, respectively. The above analysis leads us to take into account the effect of slow light on the different linear and nonlinear processes that affect, for example, the propagation of picosecond pulses in silicon PhC waveguides, as follows. We use the nonlinear Schr odinger equation (NLSE) that describes the evolution of the slowly varying envelope A(z, T) of the pulse electric eld amplitude [20], [24] i2 2 A 3 3 A A + A+ z 2 2 T 2 6 T 3 TPA = i |A|2 A |A|2 ANc 2Ae

TABLE I PARAMETERS USED IN (7) FOR SILICON AROUND 1550 nm , MOSTLY FROM [24] AND [28]

2kc +i 2 0

A (7)

Fig. 2. Schematic of the silicon chip and coupling arrangement (top). SEM micrograph of nanowire/PhC waveguide coupling region (bottom).

where is the linear loss coefcient, 2 and 3 are the secondand third-order dispersion parameters, TPA the two-photon absorption coefcient, and NC the free-carrier density generated by TPA, with and kc relating to free-carrier absorption and dispersion, respectively. The slow light enhancements are accounted for by multiplying and TPA by S 2 , and the parameters related to free carrier effects ( and kc ) by S. Although the linear loss has been found to vary differently with the group velocity for different slow light modes, and in different regimes [9], [10], [25][27], we consider a linear dependence on S, which gives the best t to experimental results. Note that modeling the propagation effects in PhCs through (7) relies on the approximation that the PhC waveguide behaves as a 1-D effective medium with a specic dispersion. Because PhC waveguides are inherently more complicated than homogeneous waveguides, their dispersion is in practice rigorously calculated through 3-D plane wave expansion method (PWM) that provides the band structure of the waveguide. The parameters 2 and 3 of (7) are then obtained by expanding the dispersion given by the band structure as a Taylor series around the wavelength of interest. Typical values for 2 (up to 1019 s2 /m) and 3 (up to 1032 s3 /m) for standard slow light PhC modes are generally much larger than for standard silicon waveguides (2 1024 s2 /m and 3 1037 s3 /m) and optical bers (2 1026 s2 /m and 3 1040 s3 /m), which led us to include several i orders in (7). Nonetheless, this approach still restricts the validity of (7) to a certain spectral window, with the number of i coefcients needing to be increased to account for more complex PhC dispersion proles. The effective area Ae of the PhC waveguide mode is calculated by slightly modifying the standard expression dened for optical bers [22]. Because the cross-section mode prole [i.e., E(x, y)] is not constant along the PhC waveguide direction z, but

periodic with the period a of the PhC lattice, Ae is averaged over the length a after integrating the electric eld amplitude E(x, y, z) (as calculated through 3-D PWM) over the volume Vol dened by a unit cell of length a using the following: Ae = 1( a
Vol Vol

|E |2 dV )2 |E |4 dV

(8)

This calculation gives us the averaged effective area of an equivalent uniform waveguide, which is consistent with the effective medium approximation that we use for modeling the propagation of pulses through the PhC waveguide in (7). In this way, we effectively simplify the 3-D problem associated with the PhC waveguide as a 1-D problem. The parameters used in the simulations in Section IV are given in Table I. The nonlinear parameter was estimated by using the above denition of Ae , which gives typical areas 3.3 a2 0.55 m2 for the slow PhC modes that we probe, and by replacing n2 by n2e n2 into (2) to take into account the partial overlap (n2e < 1, typically 0.83) of the PhC mode within the silicon material. When including the S 2 enhancement, we nd values ranging between 990 and 6190 W1 m1 for PhC waveguides with group velocities vg between c/20 and c/50, respectively. III. LOW DISPERSION SLOW LIGHT PHC WAVEGUIDES Our devices consist of 80 m long silicon PhC waveguides coupled to input and output silicon ridge waveguides tapered in width over 200 m from 0.7 (at the PhC interfaces) to 3 m at the chip facets (see Fig. 2). The whole length of the chip used in Section IV is 0.9 mm, with 400 m long access ridges on both sides. The PhC is a 2-D triangular periodic lattice of air holes (period a 410 nm, radii 0.29a) that were etched in a silicon

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

347

Fig. 3. (a) Band structure of a dispersion engineered silicon PhC waveguide (corresponding to v g = c/ 40 ) calculated by using 3-D PWM with the atband slow light region of interest highlighted in gray. (b) Measured group index for four of the dispersion engineered slow light PhC waveguides. Illustration of the shift technique used to produce the low-dispersion slow modes (Inset). The four PhC waveguides (with increasing group index) are obtained by shifting the rst row by 56, 52, 50, and 48 nm and the second row by 16, 0, 10, and 16 nm, respectively.

tions in the Section IV. Hence, dispersion is not a limiting factor in these slow light enhancement experiments. Considering other length scales associated with the PhC waveguides is useful for understanding the situations where we expect slow light waveguides to be advantageous over their fast counterparts. Propagation loss reduces the length of guide which contributes to nonlinear effects from the physical distance L to an effective length Le that is expressed as (1 e L )/. No matter how long the physical waveguide is, the effective length is therefore bound to a maximum of 1/, i.e., inversely proportional to the loss. With an estimated linear loss of dB = 20 dB/cm for the PhC waveguides, the maximum in Le varies from 2 mm for the fast regime (ng 3) to 150 m for the slow regime (ng 50) when taking the variance of loss as S. Hence, the performance of long slow light PhC waveguides will be compromised compared to fast PhC waveguidesand certainly compared with nanowires, which can exhibit effective lengths as long as 4 cm due to their lower loss [30], [31]. In contrast to long PhC waveguides, a short (80 m) waveguide has an effective length only slightly reduced in the slow regime (Le = 62 m) when compared with the fast regime (Le = 78 m). This highlights the potential interest of slow light silicon PhC waveguides as compact devices, as the effectiveness of long, slow PhC waveguides is limited by linear propagation lossat least with current technology. IV. SPM EXPERIMENTS SPM manifests as the spectral broadening of an optical pulse when launched into a nonlinear waveguide. Our experiments [20] were performed using a gure-of-eight passively mode-locked ber laser as a probe source. This laser generates almost Fourier-transform limited, sech2 -shaped, 1.2 ps pulses at a 4 MHz repetition rate, and with a tunable wavelength around 1555 nm. The input polarization was controlled to preferentially excite the TE-like modes of the waveguides. The light was coupled to the 3-m wide ridge access waveguides via lensed bers, which present a focused spot diameter of 2.5 m (see Fig. 2). The end-to-end linear insertion loss into the waveguides was 15 dB, dominated by end-facet losses. The spectra of the output pulses were measured by an optical spectrum analyzer. We investigated various slow light regimes by probing the series of 80 m long PhC waveguides with group velocities of 20, 30, 40, and 50, respectively. We also probed as a reference, a fast (ng 3.5) waveguide that consists of an 80-m long, 0.7-m wide nanowire with similar access waveguides to the PhC structure (0.9 mm total length for this chip). This nanowire has a roughly similar modal area (Ae 0.15 m2 ) to the PhC waveguide mode. Note that due to their short length (400 m) in combination with their low values (30 and 130 W1 m1 for the 3 m wide ridge and the 700 nm wide nanowire, respectively), the access waveguides are expected to induce a negligible contribution to the total nonlinear effect measured. The L product that dictates the magnitude of the nonlinear phase shift is indeed much larger for the short slow PhC waveguides ( L between 990 80 106 and 6000 80 106 , i.e., between 0.08 and 0.48 W1 ) as

membrane suspended in air. One row of holes in the lattice was omitted to create a line defect in the K direction that serves as the waveguide. The PhC waveguides were fabricated on an SOITEC silicon-on-insulator wafer comprising of a 220 nm layer of silicon on top of a 2 m layer of silica. The structure was dened by e-beam lithography and dry-etched using reactive ion etching. The silica layer was nally underetched below the PhC section with dilute hydrouoric acid to suspend the silicon membrane in air. The chip includes a series of PhC waveguides where the dispersion was engineered [16] by shifting the rst two rows of holes adjacent to the PhC waveguide defect in the direction perpendicular to propagation. Each PhC waveguide supports a slow light mode with a group velocity ranging between c/20 and c/50, and with a low group velocity dispersion over a substantial (>5 nm) bandwidth, as experimentally measured [29] and displayed in Fig. 3. This reduces the group velocity dispersion (2 ) in these regions to 106 ps2 /km (against typically 108 ps2 /km for a band-edge mode in a standard W1 PhC waveguide with similar low group velocity), as well as third-order dispersion 3 ( 106 ps3 /km) yielding associated dispersion lengths for picosecond pulses of 1 mmsignicantly larger than the 80 m long PhC waveguides probed here. This led us to neglect the two dispersion terms in (7) when carrying out the simula-

348

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

Fig. 4. Experimentally measured output spectra, normalized, for different coupled peak powers. Trace (a) corresponds to the reference 0.7 m wide nanowire. Traces (b)(e) correspond to slow PhC waveguides probed in the low-dispersion slow-light region, with group index (n g ) indicated on the plots.

compared with the access waveguides, namely the long and wide ridges ( L 30 800 106 0.02 W1 ) and the short nanowire tapered sections ( L 130 200 106 0.02 W1 ). Fig. 4 shows the experimentally measured output spectra arising from these different waveguides for incident peak powers ranging between 2 and 235 W, respectively. The actual power coupled into the guides was 20% of this value. The spectra associated with the reference nanowire waveguide [see Fig. 4(a)] are broadened at higher input powers, signifying SPM. The spectra at the output of the slow PhC waveguides [see Fig. 4(b) (e)] present a different signature. They are broader than for the nanowire, with the broadening increasing for higher group indices. However, the spectra are not symmetrically broadened, as one would expect from pure SPM, but are shifted preferentially to the blue end of the spectrum. This blue shift is associated with the generation of free carriers in silicon. Fig. 5(a) illustrates how the decrease in the local refractive index caused by the generation of free carriers across the pulse causes a negative nonlinear phase shift . This nonlinear phase shift is added to the symmetric phase shift due to SPM that follows the variation of the pulse envelope. Both of these contributions manifest as the generation of new frequencies across the pulse as described by (t) = (/t) (N L (t)). Fig. 5(b) illustrates the competing processes of Kerr-based SPM and free-carrier dispersion. On the front end of the pulse (t<0), red-shifted frequencies emerge from SPM as the phase shift

is increasing. Simultaneously, free carriers induce an opposite phase shift as they build up over the pulse, countering the net amount of SPM-induced red-shift. On the tail of the pulse (t>0), blue-shifted frequencies are generated through SPM, the effect of which is reinforced by the additional contribution of the free carriers. Therefore, SPM in the presence of free carrier dispersion no longer produces a spectrum symmetrically broadened around the pulse frequency. Instead, it gives rise to an asymmetrical spectral broadening shifted towards the blue. Note that this effect is further compounded by slow light. While TPA (and so the TPA-induced free-carrier density NC ) scales with S 2 , the free-carrier effects are further enhanced by factor S, through the increased interaction between slow light and the generated free carriers. This should impart a more severe blue shift of the spectrum than, for instance, in fast and long waveguides with the same expected nonlinear phase shift. These competing contributions of SPM and free carriers and their dependence on slow light can be explored in a more rigorous manner by solving the NLSE presented in (7) numerically using the split-step Fourier method (SSFM). In particular, we calculate the density of free carriers generated across the pulse by noting that the free carrier lifetime (recomb 100 ps 1 ns in silicon PhC geometries [32]) is much greater than the picosecond pulse duration, but signicantly less than the 250 ns repetition period. In other words, free carriers build up over the pulse, and recombine before the next pulse arrives, avoiding any accumulation of free carriers over subsequent pulses. We,

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

349

Fig. 5. Opposing effects of Kerr-based SPM and free-carrier dispersion. (a) Phase shift and (b) associated spectral frequency shift associated to SPM (blue), and free-carrier effects (red). The envelope of the pulse is represented by the green trace (right axis). The free carrier lifetime is assumed to be > > 5 ps.

therefore, calculate the free-carrier density prole in time at the position z along the waveguide, as generated by two-photon absorption through the following integral [20]: NC (z, T ) = TPA 2h0 A2 e
T T0

|A(z, )|4 d

(9)

with T0 being the time leading edge of the optical pulse. Using this method, we numerically simulate the output spectra at different input powers and for the different guides. For clarity, we represent these spectra, as well as the experimental ones, on a 2-D map, with the input power increasing along the vertical axis in Fig. 6. We obtain relatively good agreement between the simulations and experiment, with both clearly showing that the pulses preferentially broaden to the blue side of the spectrum with increasing power. The observed increase in spectral broadening and free carrier blue shift for slower modes are also well reproduced by the simulations, conrming the validity of the slow light enhancement factors used in the model. There is, however, the peculiar signature of the slowest guide (c/50) at high powers that deviates from theory and is not yet fully understood. To compare the pulse signatures more quantitatively, we measure the average absolute deviation of the output pulse spectra ( | 0 | ) from the input central wavelength (0 ) to account for the spectral broadening [see Fig. 7(a)], and calculate the average deviation of the output pulse spectra ( 0 ) from the input central wavelength (0 ) to quantify the free-carrierinduced blue shift in the output spectra [see Fig. 7(b)]. Fig. 7(a)(b) clearly attest that both spectral broadening and free-carrier-induced blue shifting increase with increasing group index, re-emphasizing the slow light enhancement of these effects. For example, the peak power needed for doubling the input spectral width of the pulse drops from 50 W to 8 W when comparing the ng 3 nanowire and ng 50 PhC waveguide. They also reveal that the blue shift and spectral broadening rates are higher for slower modes, at least at low powers. Toward higher powers, in particular for the spectral broadening, a roll-off seems to occur for the slowest guides. The simulations in Fig. 7(c) conrm this effect, with the spectral broadening rates converging towards a similar value for all PhC waveguides. This highlights that much of the benet of exploiting slow light in these structures should occur for low power devices, in a regime where

Fig. 6. 2-D plots of the output pulse spectra for various coupled peak powers for different slow light waveguides: n g = 30 (a) and (d), n g = 40 (b) and (e), and n g = 50 (c) and (f). The gures on the left-hand side (a)(c) are experimentally measured spectra. The gures on the right-hand side (d)(f) are modelled with the NLSE (7, 9) calculated using the SSFM. Color map indicates intensity in arbitrary units. Reproduced from Ref. [20].

all other parasitic effects (in particular nonlinear absorption) are still restricted. While slow light enhances useful effects like SPM induced spectral broadening, we indeed expect from (7) that it similarly reinforces all additional nonlinear losses, i.e., TPA and free carrier absorption. Plotting the average output power versus input power for all guides in Fig. 8 indicates that the clamping associated with nonlinear loss at high powers is enhanced in the slower waveguides, an effect we later utilize for all-optical regeneration (discussed in Section V). Even though we observe a reinforcement of the nonlinear losses through TPA and free carrier absorption, both of which

350

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

Fig. 8. Power transfer function through the different waveguides both experimental (dots) and simulated (dotted lines). Reproduced from Ref. [20].

Fig. 7. Experimental spectral broadening ( = | 0 | ) (a) and average spectral shift ( 0 ) (b) versus coupled peak power for the different waveguides. (c) Simulated spectral broadening ( = | 0 | ).

temper the Kerr induced spectral broadening, remarkably we still obtain a net enhancement of SPM from these short structures, which is especially noticeable at low powers. The nonlinear effects measured in these short slow light PhC waveguides can be compared with those obtained in long silicon nanowires as reported in [33]. Despite the large difference in of more than one order of magnitude between the nanowires (190 W1 m1 ) and the slow PhC waveguides (4000 W1 m1 for ng 40), the 4 mm long nanowires with 3.6 dB/cm propagation loss present slightly larger Le product values as compared with the short (80 m) slow PhC waveguides (0.4 and 0.26/W, respectively). Therefore, we expect a comparable nonlinear phase shift (which increases with Le P ) for close peak powers P, even in the presence of TPA [24], as the gure of merit (equal to n2 /(0 TPA )) is unchanged by the

slow down factor S. However, the output spectra reported in ref [33] are quite different from the slow PhC waveguides for coupled peak pump powers such that Le P is similar (6.85 and 11 W for the nanowire and the c/40-PhC waveguide, respectively). The most striking difference seems to be the larger blue shift observed for the slow PhC waveguide spectra, which look much more asymmetric and make it hard to directly compare the SPM-induced spectral broadening. Both devices show strongly clamped power transfer functions, revealing large nonlinear loss. This comparison re-emphasizes the further enhancement of free-carrier effects in the slow light regime. Next, we consider how SPM in our PhC waveguides can be potentially interesting for Mamyshev-type all-optical regeneration [34]. The goal of signal regeneration is to increase the SNR associated with noisy data. In general, regeneration requires an S-shaped [see Fig. 9(b)] nonlinear transfer function so that the (low power) noise is reduced relative to the (high power) signal when the signal power is high enough to reach the saturation regime of the transfer function. A Mamyshev regenerator exploits SPM induced spectral broadening to obtain such a transfer function [see Fig. 9(a)]. Here, we numerically lter the spectra plotted in Fig. 4 with a Gaussian bandpass lter (bandwidth = 2 nm, centered at 6 nm lower than the central pulse wavelength) to illustrate the potential benet of slow light in this context. We obtain the transfer functions shown in Fig. 9(c). While the transfer functions are linear for the fast nanowire and the PhC waveguide with c/20 group velocity, it presents an abrupt transition for the slower PhC waveguides, the power threshold of which decreases for the slowest PhC waveguide. These results clearly illustrate the potential of short slow light PhC waveguides for realizing nonlinear functions with reduced input powers. V. TPA/FCA-BASED ALL-OPTICAL REGENERATION Although TPA and associated free carrier effects are detrimental to the magnitude of SPM pulse broadening, they may be useful for limited signal processing applications in themselves.

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

351

Fig. 10. Measured nonlinear power transfer function of n g = 40 PhC waveguide for a 10 GHz, 2% duty cycle, PRBS optical data stream. Overlay shows the logical 1 regeneration concept.

Fig. 9. (a) Mamyshev regenerator that relies on the spectral ltering of the side band of spectrally broadened pulses through SPM to obtain the S-shape transfer function schematically represented on (b). (c) Power transfer functions associated with the different waveguides as obtained from numerical spectral ltering (Gaussian, full-width at half maximum = 2 nm, lter center detuned 6 nm lower than the input central wavelength).

The clamping effect associated with the power transfer functions in Fig. 8 have the potential to provide the optical limiting functionality required for regeneration of the logical 1s of a noisy optical data signal. We investigate this effect as a method to regenerate logical 1s in a 10 Gb/s optical data stream distorted by amplitude noise. This represents a rst step in using silicon slow light PhCs in all-optical signal processing at high bit rates. In future, high-bit rate communication systems, all-optical processes such as 2R regeneration and clock recovery will be needed to exceed the 40 Gb/s speed limit of electronics. These processes rely on fast optical nonlinearities to provide nonlinear power transfer functions. Here, we consider a power clamping transfer function, such as exhibited by the high ng waveguides in Fig. 8 to achieve this. Fig. 10 shows the concept of power clamping for regeneration of noisy logical 1s in an optical data stream. The amplitude uctuation on the 1spossibly caused in a real system by the temporal Talbot effect-based clock recovery [35] for exampleoccurs within the power clamping region and so is evened out to this clamping level. We investigated this effect experimentally by distorting the logical 1s of an optical data signal [10 Gb/s, 2% duty cycle,

Fig. 11. BER curves (left) and eye diagrams (right) for 10 MHz amplitude distortion, before and after the PhC device. For the BER curves (i) is for back-toback measurement of the distorted signal, (ii) is for back-to-back measurement of the undistorted signal, (iii) is for an 80 mW average input power signal as measured after the device, and (iv) is for a 20 mW average input power signal as measured after the device.

231 1 pseudorandom bit sequence (PRBS)] through modulating it with an asynchronous RF sinusoid. The distorted signal was passed through the PhC waveguide device, and the output eye diagrams and BER measured. When distorting the optical data stream with a 10 MHz RF sinusoid, we see an improvement in the output eye diagram (shown in Fig. 11), indicating that the device reduces the distortion of the signal. Comparing curves (i) and (ii) in Fig. 11, the amplitude distortion incurs a 1.85 dB power penalty for achieving error-free (BER < 109 ) operation. Comparing curves (i) and (iii) shows that operating the device in the saturation region (80 mW average power) reduces the receiver sensitivity requirement for error-free operation by 1.3 dB. Operating in the linear transfer region (iv) (out of the saturation region) does not improve the BER [as compared with (i)]. However, we expect the performance of this device to be limited by free-carrier recombination dynamics. This will not set an upper limit on the regeneration of higher bit-rate signals, but

352

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

it will limit the frequency of amplitude distortion that may be compensated for. Further work is needed to investigate this quantitatively and may also provide another useful (indirect) method of determining the free-carrier lifetime within our silicon PhC waveguides. Note that various approaches have been reported to reduce free carrier lifetimes in silicon, such as reverse-biased p-i-n junctions [40] or ion implantation [41], [42] in order to improve nonlinear device speeds limited by free carriers. The two methods for regeneration presented above are fundamentally different. The clamping-based regeneration scheme is limited by free-carrier dynamics within the guide and so for this scheme to work, the free-carrier density needs to be able to follow the distortion in the time domain. This technique is also limited to only reducing noise on the logical 1s. However, an advantage is that this method has proven to work at much lower peak powers (800 mW coupled peak power) than that required for SPM-based regeneration (20 W coupled peak power), and does not rely on spectral ltering that shifts the optical spectra. By contrast, SPM regeneration should allow for noise reduction of the logical 0s, if the average noise power is much less than the signal peak power. As opposed to the clamping scheme, the SPM scheme is not limited by free-carrier recombination for the suppression of logical 0s. This scheme may also allow for noise reduction on the logical 1s with sufcient operating poweralthough the proposed operating power requirement is already quite high. Note that in the SPM-based regeneration scheme, where spectral broadening is limited by TPA and free carriers [20], it may prove advantageous to move to a different material platform with lower TPAsuch as amorphous chalcogenide glasses [35][38]although these platforms also present their own challenges.

Fig. 12. Group index versus wavelength for the waveguide investigated. n g varies from 10 to 40 from 1550 to 1559 (blue shaded region).

expressed as [46], [47] I3 = (3 )2 sinc2 2 nc


4 3 2 I L (3) 2

k ( )L 2

f (A3 , A )
3 I3 I

(10)

VI. THG EXPERIMENTS THG is a frequency up conversion process, combining three photons at a frequency into one photon at frequency 3 . The third harmonic of infrared (C-band) light is in the visible spectrum around 520 nm, producing green light. THG is, like SPM that arises from the Kerr nonlinearity, a result of the thirdorder optical susceptibility (3) [43]. THG has been numerically studied in 2-D PhC waveguides made of cylindrical rods and including Kerr defects [44]. There has been signicant interest in producing light from silicon through nonlinear processes due to its resistance to doping with rare earths such as erbium. In particular, Raman gain has been explored [45], leading to the creation of Raman lasers on a planar silicon platform, aimed at extending the functionality of the silicon photonic integrated platform. There has, however, been no practical demonstration of the generation of visible light in passive silicon. Part of the problem is that silicon absorbs visible light on the order of 3 dB/m, so any visible light generatedunless coupled out immediatelyis absorbed. Despite this, we recently reported [46] the emission of visible green light from silicon PhC waveguides, through slow light enhanced THG. The generation of third-harmonic light from a fundamental pump beam and in a lossless medium can be

where I3 and I are the electric eld intensity of the third harmonic and fundamental respectively, L the length of the waveguide, n the refractive index, (3) the third-order susceptibility of the material, k the phase mismatch and f (A3 , A ) the mode overlap of the fundamental and third harmonic modes. It follows from (10) that increasing the generated thirdharmonic power requires an increase of the electric eld intensity of the fundamental light. This has been achieved in silica microtoroids [48], although these structures tend to suffer from narrow bandwidths associated with extremely high Q resonances. As explained in Section II, slow light enhances the electric eld intensity through I P S A (11)

in addition to the tight connement of light afforded by the submicrometer cross section of the PhC waveguide. This allowed us to demonstrate THG in PhC waveguides with 10 W of peak pump power, in contrast to megawatt levels typically employed in previous studies in PhCs, where the pump light was coupled through free-space congurations [47][50]. To probe the slow light dependence of THG, we exploit the group velocity dispersion of dispersion engineered PhC waveguides, as shown in Fig. 12, and tune the wavelength of the pump light over a 10 nm window. This provides us with a variation of the group velocity between c/10 and c/40 in a single waveguide. The laser used as a pump in these experiments was the same as that discussed in Section IV. The third-harmonic light was observed to be emitted out of plane of the waveguide and was localized on top of the PhC area (see Fig. 13) with its power decreasing along the waveguide (at least in part) due to the pump nonlinear loss [44]; it was collected by a microscope objective

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

353

Fig. 13. Green light emission from a PhC waveguide as captured by microscope (a) with and (b) without external illumination. The dashed white line denotes the approximate position of the PhC waveguide. Note the lack of emission from the nanowire access waveguides.

Fig. 15. Coupled peak power density needed to achieve 0.2 pW THG emission for varied group indices (probe wavelength varied from 1550to 1559 nm). The solid curve is a 1 /n g t as expected from intensity enhancement given in (11).

The PhC plays another crucial role in this process aside from slow light enhancement. It supports an emissive mode at 3 , allowing at least a fraction of the visible light generated to be coupled out of the opaque silicon waveguide. The maximum measured third-harmonic conversion efciency ( ), which scales as the square of the pump power, is approximately 107 (at 10 W). This represents a near 100fold increase compared to fully quasi-phase-matched processes in KTP/PPLN waveguides [53]. Our efciency scales down to = 5 1010 for 1 W peak power which, although small, compares favorably to the best available gure [52] of = 1015 for 1 W peak power (inferred from = 105 for 10 MW peak power).
Fig. 14. Dots represent measured green emission with changing coupled IR power for probe light at 1556.5 nm (left axisblack). The line is a cubic t to this data. (right axisred) Crosses represent throughput of fundamental IR light through the guide. The line is a guide to the eye.

VII. CONCLUSION Dispersion engineered slow light PhC waveguides provide a suitable platform for enhancing nonlinear effects in order to realize compact, low power, all-optical signal processing devices. In particular, we have demonstrated signicant enhancement of SPM-induced spectral broadening in 80-m long PhC silicon waveguides, an effect that could be exploited for all-optical signal regeneration. In addition, this platform is compatible with high bit rate signals, as attested by the system experiments that we have performed on these waveguides at 10 Gb/s. However, in the same way that the increased propagation loss restricts the benet of the slow light approach to short device lengths, the observed simultaneous reinforcement of nonlinear losses (associated with TPA and free-carrier absorption) highlights the interest of using slow light at low incident powers. This latter restriction may be less severe in materials other than silicon, such as amorphous material systems like chalcogenide glasses, which have a better nonlinear gure of merit (proportional to the ratio between the nonlinear parameter over the TPA factor). The role of free-carriers in realizing nonlinear functions in silicon is signicant, as it generally limits the effectiveness (speed and power consumption) of the devices. We have shown here that although this effect may be harnessed to provide useful functionality, the management of photo-generated free-carriers

and imaged onto a charge-coupled device (CCD) camera. Using bandpass lters in front of the CCD camera, the wavelength of the emitted visible light was found to lie within 520 5 nm, as expected for the third-harmonic associated with the 1560 nm infrared pump light. The camera response was linearized and calibrated against a low relative intensity noise (RIN) doubled Nd:YAG laser diode in order to use the CCD array as a detector. A transfer function of fundamental input to third-harmonic emission could then be measured. The third-harmonic transfer curve shown in Fig. 14 follows the cubic trend predicted in (10) at low powers. At higher powers, the behavior deviates from this trend, due to nonlinear loss as shown in the throughput curve in Fig. 14. In order to isolate the effect of slow light enhancement on THG, while avoiding the saturation regime, we measured the input power required to achieve a xed green light power as a function of wavelength (and hence group index). Fig. 15 shows that we do indeed see slow light enhancement of THG, with the power needed to excite a given level of thirdharmonic light scaling as 1/ng , as expected from (11).

354

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

generally presents a challenge that must be met for efcient all-optical device operation in this material platform. Slow light PhC waveguides also enable another interesting functionality, associated with the emission of green visible light through the THG process, which adds to the already impressive toolbox of functions exhibited by integrated silicon photonics. The work presented here represents an initial foray into exploiting slow light effects for improving the nonlinear functions of this toolbox. Other useful phenomena, such as cross phase modulation, four-wave mixing and Raman processes, which have all been used successfully in silicon waveguides, are likely to similarly benet from slow light enhancement in PhC geometries. As such, slow light nonlinear enhancement has the potential to push the eld of integrated nonlinear optics to lower operating powers and more compact devices. ACKNOWLEDGMENT The authors would like to thank the EU-FP6 Network of Excellence ePIXnet, for the fabrication, which was carried out in the framework of the ePIXnet nanostructuring platform for photonic integration (www.nanophotonics.eu). REFERENCES
[1] M. A. Foster et al., Nonlinear optics in photonic nanowires, Opt. Exp., vol. 16, pp. 13001320, 2008. [2] T. F. Krauss, Why do we need slow light?, Nat. Photon., vol. 2, pp. 488 449, 2008. [3] T. Baba, Slow light in photonic crystals, Nat. Photon., vol. 2, pp. 465 473, 2008. [4] M. Soljacic et al., Photonic-crystal slow-light enhancement of nonlinear phase sensitivity, J. Opt. Soc. Amer. B, vol. 19, pp. 20522059, 2002. [5] N. A. R. Bhat and J. E. Sipe, Optical pulse propagation in nonlinear photonic crystals, Phys. Rev. E, vol. 64, pp. 056604-1056604-16, 2001. [6] Y. A. Vlasov et al., Coupling into the slow light mode in slab-type photonic crystal waveguides, Opt. Lett., vol. 31, pp. 5052, 2006. [7] J. P. Hugonin et al., Coupling into slow-mode photonic crystal waveguides, Opt. Lett., vol. 32, pp. 26382640, 2007. [8] T. P. White et al., Efcient slow-light coupling in a photonic crystal waveguide without transition region, Opt. Lett., vol. 33, pp. 26442646, 2008. [9] S. Hughes et al., Extrinsic optical scattering loss in photonic crystal waveguides: role of fabrication disorder and photon group velocity, Phys. Rev. Lett., vol. 94, pp. 03390310339034, 2005. [10] L. OFaolain et al., Dependence of extrinsic loss on group velocity in photonic crystal waveguides, Opt. Exp., vol. 15, pp. 13129131138, 2007. [11] X. Letartre et al., Group velocity and propagation losses measurement in a single-line photonics-crystal waveguide on InP membranes, Appl. Phys. Lett., vol. 79, pp. 23122314, 2001. [12] M. Notomi et al., Extremely large group-velocity dispersion of linedefect waveguides in photonic crystal slabs, Phys. Rev. Lett., vol. 87, pp. 25390212539024, 2001. [13] R. J. P. Engelen et al., The effect of higher-order dispersion on slow light propagation in photonic crystal waveguides, Opt. Exp., vol. 14, pp. 16581672, 2006. [14] K. Inoue et al., Enhanced third-order nonlinear effects in slowlight photonic-crystal slab waveguides of linedefect, Opt. Exp., vol. 17, pp. 72067216, 2009. [15] A. Baron et al., Light localization induced enhancement of third order nonlinearities in a GaAs photonic crystal waveguide, Opt. Exp., vol. 17, pp. 552557, 2009. [16] J. Li et al., Systematic design of at band slow light in photonic crystal waveguides, Opt. Exp., vol. 16, pp. 62276232, 2008. [17] L. H. Frandsen et al., Photonics crystal waveguides with semi-slow light and tailored dispersion properties, Opt. Exp., vol. 14, pp. 94449450, 2006.

[18] S. Kubo et al., Low-group-velocity and low-dispersion slow light in silicon photonic crystal waveguides, Opt. Lett., vol. 32, pp. 29812983, 2007. [19] M. Ebnali-Heidari et al., Dispersion engineering of slow light photonic crystal waveguides using microuidic inltration, Opt. Exp., vol. 17, pp. 16281635, 2009. [20] C. Monat et al., Slow light enhancement of nonlinear effects in silicon engineered photonics crystal waveguides, Opt. Exp., vol. 17, pp. 2944 2953, 2009. [21] H. Hamachi et al., Slow light with low dispersion and nonlinear enhancement in a lattice-shifted photonic crystal waveguide, Opt. Lett., vol. 34, pp. 10721074, 2009. [22] G. Agrawal, Nonlinear Fibre Optics. San Diego, CA: Academic, 1995. [23] T. F. Krauss, Slow light in photonics crystal waveguides, J. Phys. DAppl. Phys., vol. 40, pp. 26662670, 2007. [24] L. H. Yin and G. P. Agrawal, Impact of two-photon absorption on selfphase modulation in silicon waveguides, Opt. Lett., vol. 32, pp. 2031 2033, 2007. [25] R. J. P. Engelen et al., Two regimes of slow-light losses revealed by adiabatic reduction of group velocity, Phys. Rev. Lett., vol. 101, pp. 1039011103901-4, 2008. [26] N. Le Thomas et al., Grating-assisted super resolution of slow waves in Fourier space, Phys. Rev. B, vol. 76, pp. 035103-1035103-5, 2007. [27] M. Lee et al., Characterizing photonic crystal waveguides with an expanded k-space evanescent coupling technique, Opt. Exp., vol. 16, pp. 1380013808, 2008. [28] M. Dinu et al., Third-order nonlinearities in silicon at telecom wavelengths, Appl. Phys. Lett., vol. 82, pp. 29542956, 2003. [29] A. Gomez-lglesias et al., Direct measurement of the group index of photonic crystal waveguides via Fourier transform spectral interferometry, Appl. Phys. Lett., vol. 90, pp. 261107-1261107-3, 2007. [30] M. Gnan et al., Fabrication of low-loss photonic wires in silicon-oninsulator using hydrogen silsequioxane electron-beam resist, Electron. Lett., vol. 44, pp. 115116, 2008. [31] W. Ding et al., Solitons and spectral broadening in long silicon-oninsulator photonics wires, Opt. Exp., vol. 16, pp. 3310319, 2008. [32] T. Tanabe et al., All-optical switches on a silicon chip realized using photonic crystal nanocavities, Appl. Phys. Lett., vol. 87, pp. 151112-1 151112-3, 2005. [33] E. Dulkeith et al., Self-phase modulation in submicron siliconon-insulator photonic wires, Opt. Exp., vol. 14, pp. 55245534, 2006. [34] P. V. Mamyshev et al., All-optical data regeneration based on self-phase modulation effects, in Proc. Eur. Conf. Opt. Commun. (ECOC 1998), p. 475. [35] D. Pudo and L. Chen, Simple estimation of pulse amplitude noise and timing jitter evolution through the temporal Talbot effect, Opt. Exp., vol. 15, pp. 63516357, 2007. [36] G. Lenz et al., Large Kerr effect in bulk Se-based chalcogenide glasses, Opt. Lett., vol. 25, pp. 254256, 2000. [37] H. C. Nguyen et al., Dispersion in nonlinear gure of merit of As2 Se3 chalcogenide bre, Electron. Lett., vol. 42, pp. 571572, 2006. [38] C. Grillet et al., Efcient coupling to chalcogenide glass photonic crystal waveguides via silica optical ber nanowires, Opt. Exp., vol. 14, pp. 10701078, 2006. [39] D. Freeman et al., Chalcogenide glass photonic crystals, Photon. Nanostruct. Fundam. Appl., vol. 6, pp. 311, 2008. [40] R. Jones, H. Rong, A. Liu, A. Fang, M. Paniccia, D. Hak, and O. Cohen, Net continuous wave optical gain in a low loss silicon-on-insulator waveguide by stimulated Raman scattering, Opt. Exp., vol. 13, pp. 519 525, 2005. [41] A. Chin, K. Y. Lee, B. , C. Lin, and S. Horng, Picosecond photoresponse of carriers in Si ion-implanted Si, Appl. Phys. Lett., vol. 69, pp. 653655, 1996. [42] T. Tanabe et al., Fast all-optical switching using ion-implanted silicon photonic crystal nanocavities, Appl. Phys. Lett., vol. 90, pp. 1031115-3, 2007. [43] D. J. Moss et al., Band-structure calculation of dispersion and anisotropy in (3) for third-harmonic generation in Si, Ge, and GaAs, Phys. Rev. B, vol. 41, pp. 15421560, 1990. [44] M. Bahl, N.-C. Panoiu, and R. M. Osgood, Jr, Nonlinear optical effects in a two-dimensional photonic crystal containing one-dimensional Kerr defects, Phys. Rev. E, vol. 67, pp. 056604-1056604-9, 2003. [45] B. Jalali, Teaching silicon new tricks, Nat. Photon., vol. 1, pp. 193195, 2007.

MONAT et al.: SLOW LIGHT ENHANCED NONLINEAR OPTICS IN SILICON PHOTONIC CRYSTAL WAVEGUIDES

355

[46] B. Corcoran et al., Green light emission in silicon through slow-light enhanced third-harmonic generation in photonic crystal waveguides, Nat. Photon., vol. 3, pp. 206210, 2009. [47] R. Boyd, Nonlinear Optics. New York: Academic, 1992, ch. 2. [48] T. Carmon and K. J. Vahala, Visible continuous emission from a silica microphotonic device by third-harmonic generation, Nat. Phys., vol. 3, pp. 430435, 2007. [49] M. G. Martemyanov et al., Third-harmonic generation in silicon photonic crystals and microcavities, Phys. Rev. B, vol. 70, p. 073311, 2004. [50] T. V. Dolgova et al., Giant third-harmonic in porous silicon photonic crystals and microcavities, JETP Lett., vol. 75, pp. 1519, 2002. [51] D. Coquillat et al., Enhanced second- and third-harmonic generation and induced photoluminescence in a two-dimensional GaN photonic crystal, Appl. Phys. Lett., vol. 87, pp. 101106-1101106-3, 2005. [52] P. P. Markowicz et al., Dramatic enhancement of third-harmonic generation in three-dimensional photonic crystals, Phys. Rev. Lett., vol. 92, pp. 083903-1083903-4, 2004. [53] M. Rusu et al., Efcient generation of green and UV light in a single PP-KTP waveguide pumped by a compact all-ber system, Appl. Phys. Lett., vol. 88, pp. 121105-1121105-3, 2006.

Majid Ebnali-Heidari (M07) was born in Faridan (Isfahan), Iran, in 1981. He received the B.S. degree from Isfahan University of Technology, Faridan, in 2003, and the M.S. degree from Tarbiat Modares University, Tehran, Iran, in 2005, both in electrical engineering. He is currently working toward the Ph.D. degree in electrical engineering from Tarbiat Modares University. From May 2008 to February 2009, he was a Visiting Student at the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, School of Physics at the University of Sydney, Sydney, N.S.W., Australia. He is currently with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, Institute for Photonics and Optical Sciences, School of Physics, University of Sydney. His research interests include applications of nonlinear optics in photonics and optouidic devices.

Christelle Monat (M08) received the Ph.D. degree in electronic integrated devices from Ecole Centrale de Lyon, Lyon, France, in 2003. She was engaged on InAs/InP quantum dots and planar IIIV photonic crystals, in order to realize a low threshold microlaser. She was with the Ecole Polytechnique F ed erale de Lausanne, Lausanne, Switzerland, for 2 years, as a Research Associate, where she was engaged on single quantum dot LEDs for single photon source applications. She joined the Centre for Ultrahigh-Bandwidth Devices for Optical Systems (CUDOS), Institute for Photonics and Optical Sciences, School of Physics, University of Sydney, Sydney, N.S.W., Australia, in late 2005, where she was engaged in the optouidics research area.. She was a Project Manager of the slow light CUDOS agship project in 2007, in which she coordinated the slow light research activity within the center. Her current research interests include slow light, photonic crystals, and nonlinear optics. Dr. Mount was awarded a fellowship by the Australian Research Council in 2007 to undertake research onto slow light in photonic crystals. Bill Corcoran (S08) received the Bachelors degree in electronic engineering and applied physics from Royal Melbourne Institute of Technology University, Melbourne, Vic., Australia, in 2006. Since 2007, he has been working toward the Ph.D. degree in physics with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, Institute for Photonics and Optical Sciences, School of Physics, University of Sydney, Sydney, N.S.W., Australia. Mr. Corcoran is a Student Member of the IEEE Laser and Electro-Optics Society, in 2008. He is also the current President of the Optical Society of America student chapter, University of Sydney. Dominik Pudo (S01M08) received the B.Eng. and Ph.D. degrees in electrical engineering from McGill University, Montreal, QC, Canada, in 2003 and 2007, respectively. He was a Postdoctoral Fellow with the Massachusetts Institute of Technology, Cambridge, where he was engaged on the integrated femtosecond laser sources, and was a Visiting Researcher with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, University of Sydney, Sydney, N.S.W., Australia, in 2005 and 2008, focusing on nonlinear optics. He is currently with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, Institute for Photonics and Optical Sciences, School of Physics, University of Sydney. His research interests include the temporal Talbot effect for clock recovery, pulse repetition rate multiplication as well as noise and timing jitter ltering.

Christian Grillet received the Ph.D. degree in electronic integrated devices from Ecole Centrale de Lyon, Lyon, France, in 2003. He joined the Centre for Ultrahigh-Bandwidth Devices for Optical Systems (CUDOS), University of Sydney, Sydney, N.S.W, Australia, in 2004, where he has been involved in projects combining integrated microphotonics and microuidics. His research included the design, nanofabrication and characterization of InP planar photonic crystal devices for optical interconnections and telecommunications. He was a Project Manager of the Chalcogenide Photonic Crystal All-Optical Switch CUDOS agship project, in 2007, in which he coordinates the research activity within the center and with international collaborators. He is currently with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, Institute for Photonics and Optical Sciences, School of Physics, University of Sydney. His research interests include photonic crystals, nonlinear optics, and slow light. Dr. Grillet was awarded a fellowship by the Australian Research Council to work on all-optical switching in photonic crystal microcavities, in 2007.

Mark D. Pelusi (M00) received the B.Eng. (Hons.) and Ph.D. degrees in electrical engineering, both from the University of Melbourne, Melbourne, Australia, in 1994 and 1998, respectively. From 1997 to 2001, he was a Research Fellow at the Femtosecond Technology Research Association, Tsukuba, Japan, where he was engaged on shortpulse, ultrafast optical communications. He then joined Corvis Corporation, Columbia, MD, during 20012003, as a Senior Hardware Engineer, where he was engaged in the research and development of broadband optical systems. He is currently a Research Fellow, with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems, Institute for Photonics and Optical Sciences, School of Physics, University of Sydney, Sydney, N.S.W., Australia, where he has been engaged in research in areas of nonlinear optical signal processing and high-speed optical communications.

356

IEEE JOURNAL OF SELECTED TOPICS IN QUANTUM ELECTRONICS, VOL. 16, NO. 1, JANUARY/FEBRUARY 2010

David J. Moss (S83M88SM09) received the B.Sc. degree in physics from the University of Waterloo, Waterloo, ON, Canada, in 1981, and the M.Sc. and Ph.D. degrees in nonlinear optics from the University of Toronto, Toronto, ON, in 1983 and 1988, respectively. From 1988 to 1992, he was a Researcher with the Institute for Microstructural Sciences, National Research Council of Canada, Ottawa, ON, working with the solid-state optoelectronics consortium on IIIV optoelectronic devices. From 1992 to 1994, he was a Senior Visiting Scientist with the Hitachi Central Research Laboratories, Tokyo, Japan, working on high-speed IIIV modulators and detectors for 10 Gb/s systems, as well as fundamental studies of quantum well tunneling. From 1994 to 1998, he was a Senior Research Fellow with the Optical Fiber Technology Center, University of Sydney, Sydney, N.S.W., Australia, working mainly in the area of silica planar waveguide devices. From 1998 to 2002, he was a Manager and Senior Scientist with JDS Uniphase, Ottawa, working in a number of areas such as ber Bragg gratings and tunable dispersion compensation devices, where he led a team of engineers and scientists who developed a highly successful commercial prototype for tunable dispersion compensation for 10 and 40 Gb/s transmission systems. Since 2003, he has been with the Institute for Photonics and Optical Sciences, School of Physics, University of Sydney, and is currently an Associate Professor with the Centre for Ultrahigh-Bandwidth Devices for Optical Systems working on all optical signal processing, integrated nonlinear photonic circuits, and photonic crystal devices. He is also an adjunct Professor with the Institut National de la Recherch Scientique, Universite du Quebec, Montreal, QC, Canada. He is the author or coauthor of more than 250 journal and conference papers and 3 book chapters. Dr. Moss is a Senior Member of the IEEE Photonics Society and a Fellow of the Optical Society of America. He has chaired or participated in many conference committees including Optical Fiber Communications during 2007 2009, the Lasers and Electro-Optic Society annual meeting during 20052009, Conference for Lasers and Electro-Optics during 20052007, and is General Program Chair of the Australian Optical Fibre Technology Conference, Melbourne, Vic., Australia, in December 2010.

Thomas P. White received the B.Sc. and Ph.D. degrees in physics from the University of Sydney, Sydney, N.S.W., Australia, in 2000 and 2005, respectively. He was engaged in the research on modeling of microstructured bers and photonic crystal devices with the University of Sydney. Since 2006, he has been a Member of the microphotonics group with the School of Physics and Astronomy, University of St. Andrews, St. Andrews, KY, U.K. His current research interests include dispersion engineering and slow light effects in photonic crystals. Dr. White holds an EU-FP6 Marie Curie International Fellowship.

Liam OFaolain was born in Cork, U.K. He received the Bachelors degree in physics from the University College, Cork, in 2001, and the Ph.D. degree in model-locked semiconductor lasers from the University of St. Andrews, St. Andrews, KY, U.K., in 2005. He is currently a Postdoctoral Research Assistant with the School of Physics and Astronomy, University of St. Andrews, St. Andrews,, where he was engaged in research on silicon photonics. He also coordinates NanoPIX (http://www.nanophotonics.eu), a fabrication service provider for nanophotonics.

Benjamin J. Eggleton (SM00) received the Bachelors degree (Honors) in science and the Ph.D. degree in physics from the University of Sydney, Sydney, N.S.W., Australia, in 1992 and 1996, respectively. In 1996, he joined Bell Laboratories, Lucent Technologies, Murray Hill, NJ, as a Postdoctoral Member of Staff, and was then transferred to the Department of Optical Fiber Research. In 2000, he became the Research Director within the Specialty Fiber Business Division, where he was engaged in forward-looking research supporting Lucent Technologies business in optical ber devices. He is currently an Australian Research Council (ARC) Federation Fellow and Professor of physics at the University of Sydney and the Research Director of the ARC Centre for Ultrahigh-Bandwidth Devices for Optical Systems (CUDOS), Institute for Photonics and Optical Sciences, School of Physics, University of Sydney. He is the author or coauthor of more than 240 journal publications and numerous conference papers. His research interests include nonlinear optics, all-optical signal processing, optical communications, photonic crystals, optouidics, and supercontinuum. Dr. Eggleton is a Fellow of the Optical Society of America. He was also the recipient of the Pawsey Medal from the Australian Academy of Science, the 2004 Malcolm McIntosh Prize for Physical Scientist of the Year, the 2003 International Commission on Optics Prize, the 1998 Adolph Lomb Medal from the Optical Society of America, the Distinguished Lecturer Award from the IEEE/Lasers and Electro-Optics Society, and a Research and Development 100 Award. He is currently the President of the Australian Optical Society.

Thomas F. Krauss received the Dipl.-Ing. degree in photographic engineering from FH Koeln, K oln, Germany, in 1989, on the basis of a diploma thesis on excimer laser ablation from IBM, Yorktown Heights, NY, and the Ph.D. degree in semiconductor ring lasers from the University of Glasgow, Glasgow, U.K., in 1992. He is currently with the School of Physics and Astronomy, University of St. Andrews, St. Andrews, KY, U.K. He initiated the research in the eld of semiconductor photonic crystals in U.K. His research interests include optical nanostructures and its use to control the emission and propagation of light. This includes slowing down light to enhance nonlinear effects, using diffractive optics for enhanced emission from LEDs as well as optically trapping bioparticles in photonic nanoresonators. Dr. Krauss was a Fellow of the Institute of Physics, in 2001 and the Royal Society of Edinburgh, in 2002. He gained SERC in 1993, and Royal Society Research Fellowships in 1995 in support of the present work and has since established a reputation worldwide, as evidenced by the large number of invited talks he presents at international level (1012 such presentations per year). He is a grant holder of several Engineering and Physical Sciences Research Council (EPSRC), European Union, and industrially sponsored research projects, coordinated EU-FP5 Photonic Integrated Circuits using Crystal Optics (PICCO) and leads the current EU-FP6 Slow Photon Activated SwitcH (SPLASH), both studying fundamental and applied aspects of photonic crystals.

You might also like