9
Review Chalcogenide glass photonic crystals Darren Freeman a , Christian Grillet b , Michael W. Lee b , Cameron L.C. Smith b , Yinlan Ruan c , Andrei Rode a , Maryla Krolikowska a , Snjezana Tomljenovic-Hanic b , C.Martijn de Sterke b , Michael J. Steel b,e , Barry Luther-Davies a , Steve Madden a , David J. Moss b , Yong-Hee Lee d , Benjamin J. Eggleton b, * a Centre for Ultrahigh-bandwidth Devices for Optical Systems (CUDOS), Laser Physics Centre, Australian National University, Canberra, ACT 0200, Australia b Centre for Ultrahigh-bandwidth Devices for Optical Systems (CUDOS), School of Physics, University of Sydney, Sydney, NSW 2006, Australia c Centre of Expertise in Photonics, School of Chemistry and Physics, University of Adelaide, Adelaide, SA 5005, Australia d Nanolaser Laboratory, Department of Physics, Korea Advanced Institute of Science and Technology (KAIST), Daejeon 305-701, Republic of Korea e RSoft Design Group, Inc., 65 O’Connor Street, Chippendale, NSW 2008, Australia Received 30 July 2007; received in revised form 2 November 2007; accepted 13 November 2007 Available online 19 November 2007 Abstract All-optical switching devices are based on a material possessing a nonlinear optical response, enabling light to control light, and are enjoying renewed interest. Photonic crystals are a promising platform for realizing compact all-optical switches operating at very low power and integrated on an optical integrated circuit. In this review, we show that by making photonic crystals from a highly nonlinear chalcogenide glass, we have the potential to integrate a variety of active devices into a photonic chip. We describe the fabrication and testing of two-dimensional Ge 33 As 12 Se 55 chalcogenide glass photonic crystal membrane devices (waveguides and microcavities). We then demonstrate the ability to post-tune the devices using the material photosensitivity. In one proposal we hope to introduce a double-heterostructure microcavity using the photosensitivity alone. # 2007 Elsevier B.V. All rights reserved. PACS : 42.70.Qs; 42.82.Cr; 42.65.Pc; 42.70.Gi Keywords: Integrated optics; Photonic crystal; Chalcogenide glass; Nonlinear optics; Microcavity; Resonator Contents 1. Introduction ......................................................................... 4 2. Chalcogenide glass photonic crystal platform .................................................. 4 2.1. Exploiting the nonlinearity .......................................................... 4 2.2. Exploiting the photosensitivity ........................................................ 5 www.elsevier.com/locate/photonics Available online at www.sciencedirect.com Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–11 * Corresponding author. E-mail addresses: [email protected] (D. Freeman), [email protected] (C. Grillet), [email protected] (B.J. Eggleton). 1569-4410/$ – see front matter # 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.photonics.2007.11.001

Chalcogenide glass photonic crystals

Embed Size (px)

Citation preview

Page 1: Chalcogenide glass photonic crystals

www.elsevier.com/locate/photonics

Available online at www.sciencedirect.com

ntals and Applications 6 (2008) 3–11

Photonics and Nanostructures – Fundame

Review

Chalcogenide glass photonic crystals

Darren Freeman a, Christian Grillet b, Michael W. Lee b, Cameron L.C. Smith b,Yinlan Ruan c, Andrei Rode a, Maryla Krolikowska a, Snjezana Tomljenovic-Hanic b,C.Martijn de Sterke b, Michael J. Steel b,e, Barry Luther-Davies a, Steve Madden a,

David J. Moss b, Yong-Hee Lee d, Benjamin J. Eggleton b,*a Centre for Ultrahigh-bandwidth Devices for Optical Systems (CUDOS), Laser Physics Centre, Australian National University,

Canberra, ACT 0200, Australiab Centre for Ultrahigh-bandwidth Devices for Optical Systems (CUDOS), School of Physics,

University of Sydney, Sydney, NSW 2006, Australiac Centre of Expertise in Photonics, School of Chemistry and Physics, University of Adelaide, Adelaide, SA 5005, Australia

d Nanolaser Laboratory, Department of Physics, Korea Advanced Institute of Science and Technology (KAIST),

Daejeon 305-701, Republic of Koreae RSoft Design Group, Inc., 65 O’Connor Street, Chippendale, NSW 2008, Australia

Received 30 July 2007; received in revised form 2 November 2007; accepted 13 November 2007

Available online 19 November 2007

Abstract

All-optical switching devices are based on a material possessing a nonlinear optical response, enabling light to control light, and

are enjoying renewed interest. Photonic crystals are a promising platform for realizing compact all-optical switches operating at

very low power and integrated on an optical integrated circuit. In this review, we show that by making photonic crystals from a

highly nonlinear chalcogenide glass, we have the potential to integrate a variety of active devices into a photonic chip. We describe

the fabrication and testing of two-dimensional Ge33As12 Se55 chalcogenide glass photonic crystal membrane devices (waveguides

and microcavities). We then demonstrate the ability to post-tune the devices using the material photosensitivity. In one proposal we

hope to introduce a double-heterostructure microcavity using the photosensitivity alone.

# 2007 Elsevier B.V. All rights reserved.

PACS : 42.70.Qs; 42.82.Cr; 42.65.Pc; 42.70.Gi

Keywords: Integrated optics; Photonic crystal; Chalcogenide glass; Nonlinear optics; Microcavity; Resonator

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2. Chalcogenide glass photonic crystal platform . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.1. Exploiting the nonlinearity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2.2. Exploiting the photosensitivity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

* Corresponding author.

E-mail addresses: [email protected] (D. Freeman), [email protected] (C. Grillet), [email protected] (B.J. Eggleton).

1569-4410/$ – see front matter # 2007 Elsevier B.V. All rights reserved.

doi:10.1016/j.photonics.2007.11.001

Page 2: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–114

3. Fabrication of planar photonic crystals in a chalcogenide glass using a focused ion beam . . . . . . . . . . . . . . . . . . . 5

4. Tapered fiber coupling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

5. Chalcogenide L3 nanocavities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

6. Photosensitive post-tuning of W1 waveguides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

1. Introduction

Photonics is evolving towards miniaturized optical

functionality, aiming to integrate multiple components

onto the same chip. Nanophotonic devices typically

make use of a high refractive index contrast to achieve

tight optical confinement, allowing wavelength-scale

resonators and waveguide bends. Two-dimensional

(2D) photonic crystals (PhCs) consist of a thin, high

refractive index dielectric slab, perforated with a

periodic lattice of air holes. PhCs, with engineered

defects, are now recognized as a promising platform for

the control of light in a photonic integrated circuit at the

wavelength scale. This has led to the demonstration of

compact photonic devices for integrated optical circuits

[1]. Research is advancing towards more complex and

‘‘active’’ devices, such as modulators and switches. It is

hoped that 2D PhCs will find utility in compact all-

optical processors, incorporating optical logic gates,

switching, pulse regeneration, wavelength conversion,

dispersion management and a variety of other applica-

tions at low power levels [2,3].

To date, most 2D PhCs have been made from Si or III–

V semiconductors, and their active functions have

typically exploited thermal or free-carrier nonlinear

effects, both of which are relatively slow [4–6]. Recently,

all-optical switching has been achieved in a Si PhC

cavity, with a free-carrier lifetime reduced to 70 ps using

recombination at the surfaces and internal dislocations

[7]. Chalcogenide glasses are infrared transmitting

materials containing the chalcogen elements S, Se or

Te, compounded with network forming elements such as

As, Si and Ge. In this paper we review our work on a PhC

platform using these glasses, aiming to exploit their high

third-order Kerr nonlinearity for all-optical ultra-fast

switching at low powers. The Kerr nonlinearity can be

regarded as instantaneous (< 100 fs) with no recovery

time. Chalcogenide glasses can be processed using

conventional lithographic techniques.

In Section 1, we show how the chalcogenide glass

PhC platform [8–11] appears to be a promising

architecture for confining and guiding light at the

wavelength scale, and where the Kerr nonlinearity and

photosensitivity of the material can be exploited to

achieve a new range of compact integrated devices [12–

14]. In Section 2, we will describe a fabrication method

using focused ion beam (FIB) milling [8–10]. Evanes-

cent coupling with tapered fibers is discussed in Section

3, while Section 4 investigates microcavities (0D or 1D

defects) in a 2D PhC membrane [9,12,15]. Finally, in

Section 5, we describe our post-process tuning

technique which utilizes the photosensitivity of

chalcogenide glass to modify the optical properties of

a planar photonic crystal device [13,14]. Our recent

demonstration of a ‘‘double heterostructure’’ micro-

cavity, made by microfluidic infiltration and with

Q� 4� 103, will be described elsewhere [16].

2. Chalcogenide glass photonic crystal platform

2.1. Exploiting the nonlinearity

When an optical cavity (resonator) contains a

nonlinear material, an increase in incident power leads

to an increase in circulating intensity, and via the

nonlinear light–matter interaction there is a small shift

in the resonant frequency which then alters the coupled

power into the resonator. This effect can be exploited to

make all-optical active devices. The transfer function of

such a system, operated at a fixed wavelength, can

display a steep transition between two states (low/high

transmission) or even optical bistability (memory of the

previous state) (Fig. 1). The required power can be

significantly lowered by using a highly nonlinear

material and a high-finesse (or high Q) microcavity

that enhances the intensity. It is well known that the

required incident power scales with V=Q2, where V is

the mode volume. Such devices, potentially using

multiple wavelengths, may serve as all-optical logic

gates (or ‘‘photonic transistors’’), and become the basis

of more complex all-optical circuits.

Ultra-small, high-Q cavities can be created by

introducing a carefully designed defect into a PhC

lattice [17–19], which can be used to create active

Page 3: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–11 5

Fig. 1. Schematic of an all-optical switch, a ‘‘photonic transistor’’. In

the off-state (A), the input signal does not penetrate the gate; in the on-

state (B), a control signal opens the gate through nonlinear interaction

and allows the input signal to be transmitted. One signal can serve both

roles.

devices with predicted power levels of only a few mW

[20,21]. Most experimental work has focused on PhCs

made from crystalline semiconductors [4–6], with much

less research into PhCs made from Kerr-type highly

nonlinear materials. For a purely third-order (xð3Þ)nonlinearity, the effective refractive index is a first-

order function of the instantaneous intensity, according

to nðIÞ ¼ n0 þ n2I, where n0 is the linear refractive

index and n2 is the nonlinear Kerr coefficient.

Chalcogenides have generated a great deal of interest

because of their attractive properties: glasses can be

formed over a wide range of compositions; the

refractive index is high, typically between 2.4 and 3

(allowing a 2D photonic band-gap); linear absorption

losses are low over a wide wavelength range (near- to

mid-infrared); they possess a large xð3Þ nonlinearity (n2

from 100 to 1000� that of silica, i.e. from 3� 10�14 to

3� 10�12 cm2/W, comparable to AlGaAs below half-

bandgap [22]); and low two-photon absorption (b). Just

as important as the large nonlinearity is the large figure

of merit, FOM¼ n2=bl, which should be > 1 to enable

all-optical processing, and is often > 5 (� 12 for

As2S3). In addition to reducing the switching power

requirement, the pure Kerr-like nonlinearities offer

instantaneous response times and the switching speed is

expected to be only limited by the resonator Q factor.

2.2. Exploiting the photosensitivity

Chalcogenide glasses are also known to exhibit a

wide range of photostructural effects. Photosensitivity

is known to arise from structural rearrangements

induced by the absorption of light at photon energies

near the electronic band-edge of the material, leading to

changes in the refractive index and the density (volume)

[23]. Photosensitivity has been used for the creation of

directly written waveguides [24–27], strong Bragg

gratings [28] and for post-tuning of optical components

such as distributed feedback lasers [29] and quantum

cascade lasers [30]. The possibility of post-trimming the

properties of individual components is highly attractive,

relaxing the fabrication tolerances and allowing novel

devices to be more easily created. For example, devices

which use strong coupling between a cavity and a two-

level system (atom or quantum dot), such as ‘‘single

photon’’ sources, require precise tuning of the cavity to

the emission peak. The ability to tune the cavity

resonance via a photosensitive chalcogenide film

applied to the PhC device appears to be promising.

We have proposed the following device to directly

exploit the chalcogenide photosensitivity. ‘‘Double

heterostructure’’ cavities suggested by Ref. [17] rely

on modification of the lattice constants along a line

defect. Light is confined to the central region (larger

lattice constant) due to differences in the mode-gap

frequencies along this line defect. In the same way, we

recently proposed a novel scheme for creating high-Q

cavities in PhCs of photosensitive material. Spatially

selective post-exposure to light in a photosensitive PhC

membrane alters the refractive index (Fig. 2a), which in

simulations [13] was predicted to yield nanocavities

with Q� 106 for a refractive index change of Dn ¼ 0:04

(Fig. 2b), consistent with photosentitivities seen in

chalcogenide glasses. In Section 5 we present the first

experimental demonstration of post-tuning the disper-

sion of a 2D PhC waveguide made from a chalcogenide

glass using the material photosensitivity.

3. Fabrication of planar photonic crystals in achalcogenide glass using a focused ion beam

In high index contrast structures, especially PhCs,

the optical fields are strong at the interfaces and the

etched surfaces must be extremely smooth to avoid

scattering losses. Compared to Si, fabrication processes

using alternative high-index materials, such as chalco-

genide glasses, have had less time to mature and suffer

from increased roughness, poorer sidewall profiles, and

higher optical losses. At CUDOS we have elected to

fabricate our first prototype devices using a focused ion

beam (FIB), because directly milling out the structures

is a single-step, maskless process, resulting in smooth

sidewalls. A scanning electron microscope (SEM) is

incorporated into the dual-beam instrument to provide

immediate feedback after fabrication. (A parallel effort,

involving KAIST, has subsequently demonstrated that

electron-beam lithography followed by chemically

assisted ion beam etching (CAIBE) is another promis-

ing process.)

Page 4: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–116

Fig. 2. (a) Refractive index distribution in the plane of the chalcogenide-based PhC waveguide and (b) calculated Q of the defect mode as a function

of the refractive index perturbation for a step profile (squares) and for a Gaussian profile (triangles).

To summarise, free-standing PhC membranes were

created in the following manner: thin free-standing

membranes of Si3N4 (silicon nitride) were fabricated on

a Si substrate; the chalcogenide glass Ge33As12Se55

(AMTIR-1) was coated onto the underside, through

window openings in the substrate; a thin conductive

layer of C was applied to the top; the lattice of holes was

milled through the membrane using a beam of Ga+ ions

scanned by the FIB; and finally the C was removed

before optical testing. Each of these steps will now be

discussed.

To create the support membranes, low-pressure

chemical vapor deposition (LPCVD) was used to coat

the Si wafers on both sides with 100 nm of Si3N4.

Square openings were lithographically defined in the

backside Si3N4 layer with inductively coupled plasma

(ICP) etching. The backside layer then acted as a mask

for anisotropic wet etching with an aqueous KOH

solution. The resulting wafer had truncated pyramid

holes, with Si3N4 membranes under tension on the top

Fig. 3. Chalcogenide glass PhC W1 waveguide, carbon still present, viewed f

cold-cathode field-emission SEM.

side. ICP etching was then used to thin the top side, in

order to reduce the milling time and the optical

perturbation due to the Si3N4 (having a refractive index

of around 2).

The Si3N4 membranes, 30 nm thick, were vacuum

coated with 300 nm of AMTIR-1 glass, supplied by

Amorphous Materials Inc. (Garland, TX), using ultra-

fast pulsed laser deposition [31]. Energy-dispersive X-

ray analysis indicated that the films had the same

stoichiometry as the bulk glass, but measurements using

an SCI Filmtek 4000 metrology tool indicated a higher

refractive index than the bulk, 2.69 versus 2.54 at

1:55mm. This is most likely due to differences in the

bond structure, and annealing the films would have

lowered the index [32]. For our application, the higher

index and higher photosensitivity of unannealed films

was considered advantageous. The top side was

vacuum-coated with around 50 nm of amorphous C

using a thermal gun, to eliminate charging when FIB

milling and SEM imaging.

rom the Si3N4 side at normal incidence and 30�, using a Hitachi S-4500

Page 5: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–11 7

An Orsay Physics Canion FIB column was used to

mill the structure of Fig. 3, using Ga+ ions at an energy

of 30 keV and probe current of 95 pA, with the beam

scanned by an in-house custom-developed pattern

generator. We employed vector scanning rather than

the more commonly available raster scanning, in an

attempt to minimize the surface roughness inside the

holes. By milling through the Si3N4 before the AMTIR-

1, rounding of the entry surface of the PhC was

significantly reduced. AMTIR-1 is much softer than

Si3N4, which served to mask out the beam pedestal and

improve the sidewall profile. The holes were milled

over multiple passes, by first scanning the beam in a

small circle at every lattice site, then in a slightly larger

circle, progressively widening the holes with each pass

over the lattice. Since the structure is a membrane, there

is no scattering and redeposition from the bottom of the

hole, as would otherwise be the case. Each circle was

traced at a constant linear speed of 1 mm/s for 170

repetitions, with a radius increment of 20 nm for the

next circle. This amounts to 0.57 s per hole (compared

to just 0.20 s with Au substituting for C). The lattice

constant was 550 nm and the outer circle had a diameter

of 270 nm, producing an actual diameter of � 330 nm

(measured from the SEM images relative to the lattice

constant).

The C film was difficult to observe in SEM images of

the final milled structures, because it is electron

transparent for energies 0 2 keV. Before optical

testing, the C layer was easily removed by exposure

to an Ar/O2 microwave-excited plasma, unlike earlier

work with Au which required wet etching to remove.

Fig. 4 presents SEM images aquired after carbon

removal (using frame integration to reduce charging).

The visible roughness on the top surface was present

before FIB milling, and was probably created when the

Fig. 4. PhC W1 waveguide of Fig. 3, after the removal of carbon, viewed fr

integrated at TV-S scan speed to mitigate the specimen charging, followed

nitride was ICP thinned. It serves to demonstrate the

polishing effect of the FIB, near the edges of the holes.

To assess the quality of the PhCs produced using this

method, one may measure the transmission spectrum of

a large and uniform lattice, as a function of angle,

position and polarization, for comparison with numer-

ical simulations [10]. It was found that the wavelengths

of the ‘‘Fano’’ resonances would shift slightly as the

collection fiber was scanned across the structure,

revealing an unwanted distortion in the lattice [33],

due to an unavoidable slow drift in the FIB system

during milling. This has subsequently been corrected,

using active feedback of the beam position [33,34], to

create the devices presented here. Based on the high

quality seen in SEM images (low roughness, near-

vertical side-walls), and the high level of unifomity

across large regions, we expect to achieve low-loss in-

plane waveguiding, and loss measurements are in

progress. The optical loss due to FIB damage

(amorphization and Ga contamination) has so far not

been determined for these structures, but fortunately the

chalcogenide is already amorphous.

4. Tapered fiber coupling

Coupling light into and out of waveguides and

cavities that have very small mode field dimensions, for

example in 2D PhCs, has proven to be challenging. One

effective approach is evanescent coupling via silica

nanowires [9,19,35,36]. When the propagation constant

of the mode in the taper and the PhC are matched, power

can couple from the fiber to modes of the PhC structure.

This is observed as a dip in the transmission spectrum

through the tapered fiber. The silica fiber diameter must

be reduced to the mm-scale in order to obtain an

evanescently extended mode field [9] that can

om the Si3N4 side at 45�. One thousand and twenty-four frames were

by deconvolution with a model of the detector’s impulse response.

Page 6: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–118

Fig. 5. Scheme used to couple light from a curved tapered fiber to a

PhC nanocavity

Fig. 6. (a) Experimental measurement setup for evanescent coupling

from a silica nanowire to a PhC structure; (b) schematic showing the

backward-coupling from a tapered fiber to a PhC waveguide; (c)

transmission spectra through the tapered fiber as a function of fiber to

PhC waveguide separation.

efficiently interact with the PhC structure. The fiber

taper is manufactured using a flame brushing and

tapering process, that heats and stretches conventional

single-mode fiber. Taper waist lengths are typically a

few mm, with outer diameters demonstrated down to

800 nm. To restrict the interaction to the PhC cavity and

avoid coupling to the surrounding lattice, either a bow

or a loop was introduced into the taper waist (Fig. 5).

The tapered fiber ends were moved closer together by

3 mm and one end was twisted to create the loop, which

normally forms at the taper waist where the fiber

diameter is the smallest. Once the loop is formed, the

ends are then separated, tightening the loop to achieve a

circumference of approximately 0.4 mm. This also

reinforces the mechanical stability of the taper.

The taper is then brought into close proximity with

the PhC structure. Light from a broadband Er source, at

1550 nm, is launched into the single-mode fiber via a

polarizer, see Fig. 6a. We selected TE-like polarization

for the chalcogenide glass devices (with the electric

field mainly lying in the plane of the membrane). In the

taper region, light is adiabatically converted into the

fundamental evanescent taper mode. Transmission to

the output of the fiber is measured using an optical

spectrum analyzer (OSA).

We have optically probed various chalcogenide

PhC structures with the evanescent coupling techni-

que. These include: W1 waveguides formed by

leaving out a row of holes from the lattice; and L3

nanocavities formed by removing three adjacent holes

from the lattice and shifting the positions of the

remaining holes at each end of the nanocavity to

optimize the Q [37].

In the W1 waveguides, Fig. 6b, strong coupling

could be obtained when the fiber was in contact with the

PhC, leading to a transmission dip of down to �18 dB,

which corresponds to � 98% coupling efficiency.

Fig. 6c presents the transmission spectra and the

variation in coupling strength with the fiber to PhC

separation. We also found that the coupling to the W1

waveguide modes is highly dependent on precise lateral

positioning of the taper [9].

5. Chalcogenide L3 nanocavities

Coupling to L3 nanocavities was achieved using the

looped nanowires of Fig. 5. Numerical simulations

predicted Q> 104 for the optimal geometry. Unlike the

previous structures, these cavities were fabricated using

e-beam lithography followed by CAIBE [38]. Fig. 7

presents experimental measurements of a cavity with

both a side-hole shift and side-hole diameter reduction.

Page 7: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–11 9

Fig. 7. Transmission spectra through the tapered fiber for coupling to

a modified L3-type nanocavity, as a function of fiber to PhC separa-

tion.

Q values as high as 104 were measured for a fiber

separation of 800 nm. As this separation decreased and

the loading of the cavity increased, the measured Q

decreased to 2� 103, and the depth of the transmission

increased to 1.5 dB.

These data indicate that, although the transmission

depth is in this case was restricted to a few dB, limiting

the contrast ratio of a switching device, simple

chalcogenide PhC resonators might exhibit sufficiently

high Q to make all-optical switching feasible.

6. Photosensitive post-tuning of W1 waveguides

We have recently demonstrated a novel post-

process tuning technique which utilizes the photo-

sensitivity of a chalcogenide glass to modify the

optical properties of a planar photonic crystal device.

Fig. 8. A schematic diagram showing the principle of in situ monitoring

A change in the dispersion of a W1 waveguide was

measured in situ, using evanescent coupling as

described in Section 3.

Fig. 8 shows the principle of the photosensitive post-

tuning experiment. The resonant coupling wavelength

was monitored by measuring the transmission spectrum

through the taper with an OSA. The dips in the

transmission spectrum are associated with coupling to

the modes of the PhC waveguide. The photoinduced

change in the PhC was observed by monitoring the shift

(in wavelength) of these dips, during the exposure of the

PhC sample to 633 nm light at an intensity of 1.3 W/

cm2.

The structure under test consists of a 70 mm W1

waveguide, i.e. a missing row of holes along the G–K

direction of a triangular lattice of air holes in a

chalcogenide membrane manufactured using e-beam

lithography followed by CAIBE [38].

The sample was exposed to light for a period of 5 h

and the transmission spectrum through the taper was

recorded at 1 min intervals during this time. Subse-

quently the sample was monitored for a period of 5 d to

verify the stability of the change. The experiment was

conducted at room temperature with the ambient room

light switched off. Fig. 9a shows the transmission

spectrum of the fiber taper due to coupling to the

fundamental PhC waveguide mode for different

exposure times. The resonance associated with the

TE0 mode was found to shift to longer wavelengths

with increasing exposure. Fig. 9b shows a plot of the

resonant wavelength versus exposure, which clearly

displays saturation at higher fluences. The circles in the

graph are experimental data points, whereas the curve

is a fit using an exponential model. The maximum

wavelength shift was 5:2� 0:4 nm, whilst the max-

imum increase in the resonance depth was 2 dB.

the photosensitive changes in a chalcogenide glass PhC waveguide.

Page 8: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–1110

Fig. 9. (a) Photosensitive tuning of the TE0mode during the exposure and (b) shift in coupling wavelength to the W1 waveguide, versus exposure

time at 633 nm.

Preliminary investigations into the photosensitivity

of unpatterened AMTIR-1 films [39] at 633 nm have

shown a decrease in the material refractive index and a

volume expansion, while Ref. [23] reported photo-

expansion in a range of As–Se–Ge glasses. For the PhC

waveguide, a refractive index decrease results in a shift

of the waveguide modes to shorter wavelengths. Our

calculations indicate that a wavelength shift of 5 nm is

obtained with an index change of Dn ¼ �0:01, and that

this change occurs linearly with index over the region of

interest. Conversely, expansion of the PhC causes a shift

towards longer wavelengths, and a 5 nm wavelength

shift of the TE0 mode is obtained for 0:31% material

expansion. Thus we attribute the observed wavelength

shift to a combination of these two competing effects.

However, the material expansion has the bigger effect in

this case leading to the observed resonance shift to

longer wavelengths.

7. Conclusion

We have given an overview of our proposed photonic

crystal platform, as a promising architecture for all-

optical switching on a chip. Using a focused ion beam,

or electron-beam lithography followed by chemically

assisted ion beam etching, we fabricated chalcogenide

glass photonic crystal waveguides and microcavities.

We then demonstrated post-tuning of these components

using photoinduced changes in the glass, a relatively

straightforward technique that is amenable to in situ

monitoring. According to our simulations, it should

even be possible to introduce microcavities using post-

fabrication light exposure.

Acknowledgements

The authors gratefully acknowledge the assistance of

the Australian Research Council under the ARC

Federation Fellowship and Centres of Excellence

programs. CUDOS (the Centre for Ultrahigh-bandwidth

Devices for Optical Systems) is an ARC Centre of

Excellence. In addition, the ANU group acknowledges

the assistance of the ANU Department of Engineering

for use of their LPCVD facility, as well as the ANU

Electron Microscopy Unit for use of their FIB. Y.

Ruan’s visit to KAIST was supported by a Young

Endeavour travel scholarship.

References

[1] M. Notomi, A. Shinya, S. Mitsugi, E. Kuramochi, H.-Y. Ryu,

Waveguides, resonators and their coupled elements in photonic

crystal slabs, Opt. Express 12 (8) (2004) 1551–1561.

[2] V.G. Ta’eed, M. Shokooh-Saremi, L. Fu, I.C.M. Littler, D.J.

Moss, M. Rochette, B.J. Eggleton, Y. Ruan, B. Luther-Davies,

Self-phase modulation-based integrated optical regeneration in

chalcogenide waveguides, IEEE J. Sel. Top. Quant. Electron. 12

(3) (2006) 360–370.

[3] R.E. Slusher, B.J. Eggleton, Nonlinear Photonic Crystals,

Springer, Berlin, 2003.

[4] T. Tanabe, M. Notomi, S. Mitsugi, A. Shinya, E. Kuramochi, All-

optical switches on a silicon chip realized using photonic crystal

nanocavities, Appl. Phys. Lett. 87 (15) (2005) 151112.

[5] P.E. Barclay, K. Srinivasan, O. Painter, Nonlinear response of

silicon photonic crystal microresonators excited via an integrated

waveguide and fiber taper, Opt. Express 13 (3) (2005) 801–820.

[6] F. Raineri, C. Cojocaru, P. Monnier, A. Levenson, R. Raj, C.

Seassal, X. Letartre, P. Viktorovitch, Ultrafast dynamics of the

third-order nonlinear response in a two-dimensional InP-based

photonic crystal, Appl. Phys. Lett. 85 (11) (2004) 1880–1882.

Page 9: Chalcogenide glass photonic crystals

D. Freeman et al. / Photonics and Nanostructures – Fundamentals and Applications 6 (2008) 3–11 11

[7] T. Tanabe, K. Nishiguchi, A. Shinya, E. Kuramochi, H. Inokawa,

M. Notomi, K. Yamada, T. Tsuchizawa, T. Watanabe, H. Fukuda,

H. Shinojima, S. Itabashi, Fast all-optical switching using ion-

implanted silicon photonic crystal nanocavities, Appl. Phys.

Lett. 90 (3) (2007) 031115.

[8] D. Freeman, S. Madden, B. Luther-Davies, Fabrication of planar

photonic crystals in a chalcogenide glass using a focused ion

beam, Opt. Express 13 (8) (2005) 3079–3086.

[9] C. Grillet, C. Smith, D. Freeman, S. Madden, B. Luther-Davies,

E.C. Magi, D.J. Moss, B.J. Eggleton, Efficient coupling to

chalcogenide glass photonic crystal waveguides via silica optical

fiber nanowires, Opt. Express 14 (3) (2006) 1070–1078.

[10] C. Grillet, D. Freeman, B. Luther-Davies, S. Madden, R. McPhe-

dran, D.J. Moss, M.J. Steel, B.J. Eggleton, Characterization and

modeling of Fano resonances in chalcogenide photonic crystal

membranes, Opt. Express 14 (1) (2006) 369–376.

[11] K. Paivasaari, V.K. Tikhomirov, J. Turunen, High refractive

index chalcogenide glass for photonic crystal applications,

Opt. Express 15 (5) (2007) 2336–2340.

[12] C. Monat, C. Grillet, P. Domachuk, C. Smith, E. Magi, D.J.

Moss, H.C. Nguyen, S. Tomljenovic-Hanic, M. Cronin-Golomb,

B.J. Eggleton, D. Freeman, S. Madden, B. Luther-Davies, S.

Mutzenich, G. Rosengarten, A. Mitchell, Frontiers in micro-

photonics: tunability and all-optical control, Laser Phys. Lett. 4

(3) (2007) 177–186.

[13] S. Tomljenovic-Hanic, M.J. Steel, C.M. de Sterke, D.J. Moss,

High-Q cavities in photosensitive photonic crystals, Opt. Lett. 32

(5) (2007) 542–544.

[14] M.W. Lee, C. Grillet, C.L.C. Smith, D.J. Moss, B.J. Eggleton, D.

Freeman, B. Luther-Davies, S. Madden, A. Rode, Y. Ruan, Y.-H.

Lee, Photosensitive post tuning of chalcogenide photonic crystal

waveguides, Opt. Express 15 (3) (2007) 1277–1285.

[15] C. Smith, C. Grillet, S. Tomljenovic-Hanic, E.C. Magi, D. Moss,

B.J. Eggleton, D. Freeman, S. Madden, B. Luther-Davies,

Characterisation of chalcogenide 2D photonic crystal wave-

guides and nanocavities using silica fibre nanowires, Phys. B

394 (2) (2007) 289–292.

[16] C.L.C. Smith, D.K.C. Wu, M.W. Lee, C. Monat, S. Tomljenovic-

Hanic, C. Grillet, B.J. Eggleton, D. Freeman, Y. Ruan, S.

Madden, B. Luther-Davies, H. Giessen, Y.-H. Lee, Microfluidic

photonic crystal double heterostructures, Appl. Phys. Lett. 91

(12) (2007) 121103.

[17] B.-S. Song, S. Noda, T. Asano, Y. Akahane, Ultra-high-Q

photonic double heterostructure nanocavity, Nat. Mater. 4 (3)

(2005) 207–210.

[18] Y. Akahane, T. Asano, B.-S. Song, S. Noda, High-Q photonic

nanocavity in a two-dimensional photonic crystal, Nature 425

(6961) (2003) 944–947.

[19] K. Srinivasan, P.E. Barclay, M. Borselli, O. Painter, Optical-

fiber-based measurement of an ultrasmall volume high-Q photo-

nic crystal microcavity, Phys. Rev. B 70 (8) (2004) 081306.

[20] E. Centeno, D. Felbacq, Optical bistability in finite-size non-

linear bidimensional photonic crystals doped by a microcavity,

Phys. Rev. B 62 (12) (2000) 7683–7686 (R).

[21] M. Soljacic, M. Ibanescu, S.G. Johnson, Y. Fink, J.D. Joanno-

poulos, Optimal bistable switching in nonlinear photonic crys-

tals, Phys. Rev. E 66 (5) (2002) 055601.

[22] G.A. Siviloglou, S. Suntsov, R. El-Ganainy, R. Iwanow, G.I.

Stegeman, D.N. Christodoulides, R. Morandotti, D. Modotto, A.

Locatelli, C. De Angelis, F. Pozzi, C.R. Stanley, M. Sorel,

Enhanced third-order nonlinear effects in optical AlGaAs nano-

wires, Opt. Express 14 (20) (2006) 9377–9384.

[23] T. Igo, Y. Noguchi, H. Nagai, Photoexpansion ‘‘thermal con-

traction’’ of amorphous chalcogenide glasses, Appl. Phys. Lett.

25 (4) (1974) 193–194.

[24] A. Zakery, Y. Ruan, A.V. Rode, M. Samoc, B. Luther-Davies,

Low-loss waveguides in ultrafast laser-deposited As2S3 chalco-

genide films, J. Opt. Soc. Am. B 20 (9) (2003) 1844–1852.

[25] V. Lyubin, A. Klebanov, A. Feigel, B. Sfez, Films of chalco-

genide glassy semiconductors: new phenomena and new appli-

cations, Thin Solid Films 459 (1–2) (2004) 183–186.

[26] R.G. DeCorby, N. Ponnampalam, M.M. Pai, H.T. Nguyen, P.K.

Dwivedi, T.J. Clement, C.J. Haugen, J.N. McMullin, S.O. Kasap,

High index contrast waveguides in chalcogenide glass and

polymer, IEEE J. Sel. Top. Quant. Electron. 11 (2) (2005)

539–546.

[27] S. Ramachandran, S.G. Bishop, Photoinduced integrated-optic

devices in rapid thermally annealed chalcogenide glasses, IEEE

J. Sel. Top. Quant. Electron. 11 (1) (2005) 260–270.

[28] M. Shokooh-Saremi, V.G. Ta’eed, I.C.M. Littler, D.J. Moss, B.J.

Eggleton, Y. Ruan, B. Luther-Davies, Ultra-strong, well-apo-

dised Bragg gratings in chalcogenide rib waveguides, Electron.

Lett. 41 (13) (2005) 738–739.

[29] T.K. Sudoh, Y. Nakano, K. Tada, Wavelength trimming technol-

ogy for multiple-wavelength distributed-feedback laser arrays by

photo-induced refractive index change, Electron. Lett. 33 (3)

(1997) 216–217.

[30] S. Song, S.S. Howard, Z. Liu, A.O. Dirisu, C.F. Gmachl, C.B.

Arnold, Mode tuning of quantum cascade lasers through optical

processing of chalcogenide glass claddings, Appl. Phys. Lett. 89

(4) (2006) 041115.

[31] A.V. Rode, B. Luther-Davies, E.G. Gamaly, Ultrafast ablation

with high-pulse-rate lasers. Part II. Experiments on laser deposi-

tion of amorphous carbon films, J. Appl. Phys. 85 (8) (1999)

4222–4230.

[32] R.P. Wang, A.V. Rode, S.J. Madden, C.J. Zha, R.A. Jarvis, B.

Luther-Davies, Structural relaxation and optical properties in

amorphous Ge33As12Se55 films, J. Non-Cryst. Solids 353 (8–10)

(2007) 950–952.

[33] D. Freeman, B. Luther-Davies, S. Madden, Real-time drift

correction of a focused ion beam milling system. Presented at

Inter. Conf. Nanosci. Nanotechnol., Brisbane, Australia, July,

2006.

[34] D. Freeman, B. Luther-Davies, S. Madden, S. Stowe, Drift

correction of a focused ion beam for nano-fabrication of photo-

nic crystals. Presented at Inter. Microscopy Congress 16, Sap-

poro, Japan, September, 2006.

[35] I.K. Hwang, S.K. Kim, J.K. Yang, S.H. Kim, S.H. Lee, Y.-H.

Lee, Curved-microfiber photon coupling for photonic crystal

light emitter, Appl. Phys. Lett. 87 (13) (2005) 131107.

[36] C. Grillet, C. Monat, C.L.C. Smith, B.J. Eggleton, D.J. Moss, S.

Frederick, D. Dalacu, P.J. Poole, J. Lapointe, G. Aers, R.L.

Williams, Nanowire coupling to photonic crystal nanocavities

for single photon sources, Opt. Express 15 (3) (2007) 1267–

1276.

[37] C. Sauvan, P. Lalanne, J.P. Hugonin, Slow-wave effect and

mode-profile matching in photonic crystal microcavities, Phys.

Rev. B 71 (16) (2005) 165118.

[38] Y. Ruan, M.-K. Kim, Y.-H. Lee, B. Luther-Davies, A. Rode,

Fabrication of high-Q chalcogenide photonic crystal resona-

tors by e-beam lithography, Appl. Phys. Lett. 90 (7) (2007)

071102.

[39] S. Madden, Investigation into the photosensitivity of AMTIR-1

films, personal communication, 2006.