Вы находитесь на странице: 1из 10

Photonic crystal structures in ion-sliced

lithium niobate thin films


Frederik Sulser, Gorazd Poberaj, Manuel Koechlin, and Peter Günter
Nonlinear Optics Laboratory, Institute of Quantum Electronics, ETH Zurich,
8093 Zurich, Switzerland
nlo@phys.ethz.ch

Abstract: We report on the first realization of photonic crystal structures in


600-nm thick ion-sliced, single-crystalline lithium niobate thin films bonded
on a lithium niobate substrate using adhesive polymer benzocyclobutene
(BCB). Focused ion beam (FIB) milling is used for fast prototyping
of photonic crystal structures with regular cylindrical holes. Unwanted
redeposition effects leading to conically shaped holes in lithium niobate
are minimized due to the soft BCB layer underneath. A high refractive
index contrast of 0.65 between the lithium niobate thin film and the BCB
underlayer enables strong light confinement in the vertical direction. For
TE polarized light a triangular photonic crystal lattice of air holes with a
diameter of 240 nm and a separation of 500 nm has a photonic bandgap
in the wavelength range from 1390 to 1500 nm. Experimentally measured
transmission spectra show a spectral power dip for the ΓK direction of the
reciprocal lattice with an extinction ratio of up to 15 dB. This is in good
agreement with numerical simulations based on the three-dimensional plane
wave expansion (PWE) and the finite-difference time-domain (FDTD)
method.
© 2009 Optical Society of America
OCIS codes: (130.3730) Lithium niobate; (160.5298) Photonic crystals; (250.5300) Photonic
integrated circuits; (310.6845) Thin film devices and applications.

References and links


1. J.D. Joannopoulos, S.G. Johnson, J.N. Winn, R.D. Meade, “Photonic Crystals: Molding the Flow of Light,”
Princeton University Press 2nd Edition (2008).
2. Y. Akahane, T. Asano, B. S. Song, S. Noda, “High-Q photonic nanocavity in a two-dimensional photonic crystal,”
Nature 425, 944-947 (2003).
3. V. R. Almeida, C. A. Barrios, R. R. Panepucci, M. Lipson, “All-optical control of light on a silicon chip,” Nature
431, 1081-1084 (2004).
4. F. Lacour, N. Courjal, M.-P. Bernal, A. Sabac, C. Bainier, M. Spajer, “Nanostructuring lithium niobate substrates
by focused ion beam milling,” Opt. Mater. 27, 1421-1425 (2005).
5. M. Roussey, M.-P. Bernal, N. Courjal, and F.I. Baida, “Experimental and theoretical characterization of a lithium
niobate photonic crystal,” Appl. Phys. Lett. 87, 241101 (2005).
6. M. Roussey, M.-P. Bernal, N. Courjal, D. Van Labeke, F.I. Baida, R. Salut, “Electro-optic effect exaltation on
lithium niobate photonic crystals due to slow photons,” Appl. Phys. Lett. 89, 241110 (2006).
7. J. Amet, F.I.Baida, G.W. Burr, M.-P. Bernal, “The superprism effect in lithium niobate photonic crystals for
ultra-fast, ultra-compact electro-optical switching,” Photon. Nanostruct. Fundam. Appl. 6, 47-59 (2008).
8. S. Diziain, J. Amet, F.I. Baida, and M.-P. Bernal, “Optical far-field and near-field observations of the strong
angular dispersion in a lithium niobate photonic crystal superprism designed for double passive and active de-
multiplexer applications,” Appl. Phys. Lett. 93, 261103 (2008).
9. C.-H. Hou, M.-P. Bernal, C.-C. Chen, R. Salut, C. Sada, N. Argiolas, M. Bazzan, M.V. Ciampollilo, “Purcell
effect observation in erbium doped lithium niobate photonic crystal structures,” Opt. Commun. 281, 4151-4154
(2008).

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20291
10. G.W. Burr, S. Diziain, M.-P. Bernal, “The impact of finite-depth cylindrical and conical holes in lithium niobate
photonic crystals,” Opt. Express 16, 6302-6316 (2008).
11. S. Diziain, S. Harada, R. Salut, P. Muralt, M.-P. Bernal, “Strong improvement in the photonic stop-band edge
sharpness of a lithium niobate photonic crystal slab,” Appl. Phys. Lett. 95, 101103 (2009).
12. P. Rabiei, P. Günter, “Optical and electro-optical properties of submicrometer lithium niobate slab waveguides
prepared by crystal ion slicing,” Appl. Phys. Lett. 85, 4603-4605 (2004).
13. G. Poberaj, M. Koechlin, F. Sulser, A. Guarino, J. Hajfler, P. Günter, “Ion-sliced lithium niobate thin films for
active photonic devices,” Opt. Mater. 31, 1054-1058 (2009).
14. P. Rabiei, W.H. Steier, “Lithium niobate ridge waveguides and modulators fabricated using smart guide,” Appl.
Phys. Lett. 86, 161115 (2005).
15. A. Guarino, G. Poberaj, D. Rezzonico, R Degl’Innocenti, P. Günter, “Electro-optically tunable microring res-
onators in lithium niobate,” Nat. Photonics 1, 407-410 (2007).
16. M. Koechlin, G. Poberaj, P. Günter, “High-resolution laser lithography system based on two-dimensional
acousto-optic deflection,” Rev. Sci. Instrum. 80, 085105 (2009).
17. S.G. Johnson, J.D. Joannopoulos, “Block-iterative frequency-domain methods for Maxwell’s equations in a
planewave basis,” Opt. Express 8, 173-190 (2001).
18. S.G. Johnson, S. Fan, P.R. Villeneuve, J.D. Joannopoulos, “Guided modes in photonic crystal slabs,” Phys. Rev.
B 60, 5751-5758 (1999).
19. T. Weng, G.Y. Guo, “Band structures of honeycomb photonic crystal slabs,” J. Appl. Phys. 99, 093102 (2006).
20. J.-P. Berenger, “A perfectly matched layer for the absorption of electromagnetic waves,” J. Computat. Phys. 114,
185-200 (1994).
21. S.D. Gedney, “An Anisotropic Perfectly Matched Layer-Absorbing Medium for the Truncation of FDTD Lat-
tices,” IEEE Trans. Antennas Propag. 44, 1630-1639 (1996).

1. Introduction
Photonic crystal (PhC) structures have a great potential to reduce the size and increase the func-
tionality of integrated optics devices [1]. Up to know, a remarkable progress in the realization of
high quality PhC structures has been achieved mainly in silicon [2, 3] and some other semicon-
ductor materials. These materials possess high refractive indices, a good optical transmission in
the telecom wavelength range, and their micro- and nano-structuring is strongly supported by
the advanced semiconductor technology. On the other hand, there is a strong interest in devel-
opment of high density integrated photonic devices based on nonlinear optical materials, which
offer the possibility of electro-optical modulation and switching. In this paper we present new
developments of photonic crystal structures in lithium niobate (LiNbO3 ) as one of very attrac-
tive materials of choice due to its excellent nonlinear- and electro-optical properties.
Previous reports on the fabrication and optical characterization of photonic crystal structures
in LiNbO3 are limited mainly to annealed proton exchanged (APE) waveguides nanostructured
by focused ion beam (FIB) milling [4, 5]. Using this approach, the first electro-optically tunable
photonic bandgap structures in LiNbO3 have been realized [6]. Also other effects, such as the
superprism effect in LiNbO3 [7, 8] and Purcell effect in Er:LiNbO3 photonic crystals [9] have
been studied and demonstrated. However, APE waveguides have a very small refractive index
contrast with respect to the bulk crystal (∆n ∼ 0.03), resulting in a weak light confinement and
a broadly distributed mode with its centroid located 2.5 µm below the crystal surface. Thus,
air holes with high aspect ratio are required. It turned out that FIB milling of holes with high
aspect ratio in LiNbO3 is difficult because of redeposition effects leading to conically shaped
holes with a limited depth of < 1.5µm. Several limitations in performance of photonic crystal
structures arising due to the finite-depth and conically shaped holes, as well as the weak vertical
confinement of guided modes, are discussed in [10].
Consequently, much stronger vertical confinement of guided light in planar photonic crys-
tal structures is needed. This can be achieved using LiNbO3 thin films on low-index substrate
materials. Following this approach, different fabrication techniques have been studied in par-
allel. Very recently, the fabrication method and optical properties of a FIB-milled photonic
crystal structure in a polycrystalline 380-nm thick LiNbO3 film deposited on a MgO substrate

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20292
by pulsed laser deposition have been reported [11]. The authors have demonstrated strong im-
provement in the photonic stop-band edge sharpness compared to their previous work based on
APE waveguides. However, the fabricated polycrytsalline LiNbO3 thin film lacks the excellent
electro-optical and nonlinear optical properties of bulk LiNbO3 crystals.
In our study we have overcome these problems using a new type of photonic crystal structures
based on ion-sliced, single-crystalline LiNbO3 films which we developed in our laboratory
[12, 13]. These LiNbO3 thin films show electro-optical properties comparable to bulk LiNbO3
crystals as demonstrated by the realization of electro-optical Mach-Zehnder modulators [14]
and microring resonators [15]. In our approach we use 600-nm thick ion-sliced LiNbO3 films
bonded to a LiNbO3 handling substrate using adhesive polymer benzocyclobutene (BCB) to
achieve a high index contrast (∆no = 0.67 and ∆ne = 0.60 at λ = 1.5 µm) as a prerequisite for
high-density optical integration. Such LiNbO3 films enable even stronger confinement of light
in the vertical direction and are very well suited for the fabrication of photonic crystal slabs.
We combine conventional ridge optical waveguides fabricated by laser lithography patterning
and Ar-ion etching with photonic crystal structures fabricated by FIB milling.
This paper is structured as follows. Section 2 describes the fabrication of ion-sliced thin films
and the nanostructuring of photonic crystal lattices of air holes using the focused ion beam
miling. In Section 3 we present transmission measurements of the fabricated photonic crystal
structures to reveal the existence of photnic bandgaps. Finally, in Section 4 the experimental
results are compared with numerical calculations based on the three-dimensional plane wave
expansion (PWE) and the finite-difierence time-domain (FDTD) method.

2. Fabrication of photonic crystal slabs


In this section we first describe the fabrication of ion-sliced LiNbO3 thin films serving as a
platform for photonic crystal slabs. Then we focus on the nanostructuring of airholes by means
of focused ion beam (FIB) milling.

2.1. Ion-sliced LiNbO3 thin films


In order to produce single-crystalline thin films of LiNbO3 , we use the ion slicing and wafer
bonding technique. In the following we describe shortly the fabrication procedure which we
developed in our laboratory [13, 15]. Figure 1 schematically shows the fabrication steps and
the structure of a thin film.

2.
He+
6 00
nm
LN
1. 2µ
m
BCB
510
300°C µm
LN Substrate
3. 4.
Fig. 1. Fabrication steps and configuration of a LiNbO3 thin film. 1. He+ implantation, 2.
Bonding, 3. Heating process, 4. Lift off

In the first step a sacrificial layer is created in a LiNbO3 wafer by He+ implantation. This
layer enables a separation of the thin film from the donor wafer. The film thickness depends

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20293
on the implantation energy and yields around 670 nm for 195-keV He+ ions in our case. In
the second step, the implanted wafer is bonded on a LiNbO3 substrate which is covered with a
2-µm layer of adhesive polymer benzocyclobutene (BCB). Afterwards the sample is gradually
heated up to 220°C. The thermally induced stress enables the sacrificial layer to split off the
thin film from the donor wafer. The transeferred film is then annealed at a temperature of 300°C
for several hours. Finally, the surface of the thin film is smoothened by Ar+ ion etching. This
results in a 600-nm thick LiNbO3 film with an aerea of > 1 cm2 , a surface roughness of 4
nm (rms), and a high index contrast of ∆n = 0.65. Alternatively, mechanical polishing can be
applied to reduce the surface roughness below 1 nm rms. The described procedure enables the
fabrication of both x- and z-cut LiNbO3 thin films, which are needed for different integrated
optics devices.

2.2. Focused ion beam milling


Photonic crystal structures were fabricated by focused ion beam milling using a Zeiss Nvision
40 device in combination with a RAITH Elphy Quantum lithography system. It is important
to minimize unwanted charging effects during focused ion beam milling, which would dete-
riorate the quality of milled structures. Therefore, the samples were covered with 5 to 10-nm
thick chromium layer using a Leybold Univex 450 e-gun deposition system and grounded to
the sample holder. After completion of the nanostructuring process, the chromium layer was
removed by chemical etching.

(a) (b)

Fig. 2. (a) Scanning electron micrograph of triangular lattice of holes in a z-cut, 600-nm
thick LiNbO3 film milled by focused ion beam (FIB). The lattice constant is 500 nm and
the holes have a diameter of 240 nm. (b) FIB milled cross section showing holes with a
regular cylindric shape in the LiNbO3 thin film. The holes with a depth of 1µm penetrate
into the BCB underlayer.

Numerical calculations pointed out that a bandgap for TE-polarized light around telecom
wavelength region can be achieved by a triangular hole structure with a lattice constant of 500
nm and a hole radius of 120 nm. To mill such delicate structures all parameters such as aperture,
focus position, and astigmatism of the ion beam must be configured very carefully in order to
achieve a completely round beam shape in the focal plane. All our work on the Zeiss Nvision
40 was performed by focussing gallium ions onto the sample surface with a 120-µm aperture
in the high current mode and an acceleration voltage of 30 kV. For precise milling of hole
lattices RAITH Elphy Quantum soft- and hardware is used to control the Nvision 40 externally.
This enables the use of semiconductor’s industry standard format GDS2 to define and control
the milling process. Several experiments pointed out, that a triangular structure with a lattice
constant of 500 nm and a hole diameter of 240 nm is milled with a beam current of 300 pA

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20294
Fig. 3. Cross section of FIB milled holes in a bulk LiNbO3 crystal. The milling parameters
were the same as in case of the thin film shown in Fig. 2. A pronounced conical shape of
holes is due to unwanted redeposition effects.

and a milling time of 700 ms per hole. A single pass milling strategy is used, which enables
shorter processing times and is less prone to slight drifts of focused ion beam. Figure 2 shows
a scanning electron micrograph of such a lattice.
In contrast to previously reported photonic crystal structures in annealed proton exchanged
planar waveguides, where FIB milled holes have a very pronounced conical shape due to rede-
position effects [10], our approach avoids this problem. On the one hand, the required aspect
ratio of holes is substantially smaller due to the strong confinement of guided modes in 600-
nm thin films. On the other hand, the redeposition effects in the milling process are minimized
because of the relatively soft BCB underlayer, which facilitates the milling of cylindrical holes
with vertical walls. For a comparison, Fig. 3 shows a cross-section of a hole array on the sur-
face of a bulk LiNbO3 crystal. The same FIB milling parameters were used as in the case of
LiNbO3 thin films. The holes have a strongly pronounced conical shape as a consequence of
the redeposition effects.

3. Experimental confirmation of a photonic bandgap


3.1. Sample preparation
In order to experimentally demonstrate the existence of a photonic bandgap, we fabricated var-
ious photonics crystal lattices with different number of holes in the middle of ridge optical
waveguides to facilitate the in- and out-coupling of probe laser light. Figure 4 shows a central
region of a ridge optical waveguide with an enlarged tapered section containing a triangular ar-
ray of 15 ×36 air holes between the input and output optical waveguide. The tapered waveguide
structure was chosen to recollect the diffracted light passing through the photonic crystal lat-
tice. The fabricated sample with a length of 3 mm contained several parallel optical waveguides
with and without photonic crystal structures for comparative transmission measurements.
Ridge optical waveguides were fabricated by lithographic patterning and Ar+ ion etching. We
used a negative tone photoresist SU-8 and our home built laser lithography system operating
at 375 nm [16] for the illumination of structures. Using two acousto-optical deflectors and
a xyz-translation stage, this system enables a high flexibility and fast prototyping of various
structures. After the illumination, the standard developing process for the SU-8 was used. The
lithographic patterned structures were then transferred to a z-cut LiNbO3 thin film by Ar+ ion
etching using an Oxford PlasmaLab 80+ device. Different photonics crystal lattices in tapered
waveguide sections were then nanostructured by focused ion beam milling as described in the
previous section. Finally, the sample was cut and waveguide end-facets were smoothened by

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20295
K
M
Γ

Light
Fig. 4. Scanning electron micrograph of a ridge optical waveguide with enlarged tapered
region containing a PBG structure. The triangular array of 15 × 36 air holes was structured
using the FIB milling. Light is propagating along symmetry direction Γ → K. The inset
shows an input facet of the coupling optical waveguide smoothened by FIB milling.

focused ion beam milling as shown in the inset of Fig. 4.


It is worth to note that the total milling time for the array with 15 × 36 holes amounts to
around 6 min, only. Thus, focused ion beam milling is a very suitable method for fast prototyp-
ing of high-qulity photonic crystal structures in ion-sliced LiNbO3 thin films.

3.2. Optical transmission measurements


An optical setup for the characterization of photonic crystal structures in transmission measure-
ments is shown in Fig. 5. The setup for end-fire coupling consists mainly of a white light laser
source, beam focusing and collimating optics, and a spectrometer to analyse the transmitted
spectrum. The laser source was a commercially available supercontinuum fibre laser (Fianium
SC450) delivering 2 W of average output power within a broad spectral range between 450 and
1950 nm. In order to adjust the input beam polarization and to attenuate the beam power, two
rotatable Glan-Tayler polarizers were inserted into the beam path. As the PhC structures were
designed for the telecom wavelength range, the visible part was strongly attenuated using two
specially coated bending mirrors acting as long-pass wavelength filters. A small transmitted
portion of the visible light was exploited for aligning the optics and sample position. The light
was coupled in and out of the sample by means of a 100× and a 20× microscope objective with
appropriate antireflection coatings, respectively. The objectives and the sample were mounted
on precise xyz-translation stages with sub-micrometer positioning capability. An ANDO AQ-
6315A spectrum analyzer was used to measure the spectrum of the transmitted light.
Figure 6 depicts normalized transmission spectra of three triangular lattices of air holes (lat-
tice period a = 500 nm, hole radius r = 120 nm → r/a = 0.24) with 5, 10, and 15 rows, re-
spectively. The measurements were performed with TE-polarized light (polarization parallel
to the LiNbO3 slab) over the wavelength range from 1040 to 1600 nm. In each case, the nor-
malized transmission spectrum was obtained by dividing two transmission spectra of adjacent
waveguides with and without a photonic crystal structure in the middle. As it can be seen, the
photonic crystal stucture with 15 rows shows the most pronounced transmission dip between
1300 and 1550 nm with the maxiumum extinction ratio of 15 dB at a wavelength of 1470 nm.
We attribute this feature to the existence of a photonic bandgap in the fabricated structure. In
contrast to the TE polarization, no photonic gap was observed for TM polarized light.

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20296
Long-pass Collimator
Mirror
White Light Laser
Polarizer

Spectrum
Analyzer

Sample
M.O. M.O.
100x 20x

Fig. 5. Optical setup with a white light laser source and a spectrum analyzer for the char-
acterization of photonic crystal bandgaps in fabricated samples.
Normalized Transmitted Specturm [dB]

-5

-10

-15 5 rows
10 rows
15 rows

-20
1200 1400 1600
Wavelength [nm]

Fig. 6. Normalized transmission spectra (TE-polarized light) of triangular lattices of air


holes (lattice period a = 500 nm, hole radius r = 120 nm) with 5, 10, and 15 rows. The
structure with 15 rows exhibits a broad transmission dip between 1300 and 1550 nm with
the maximum extinction of 15 dB at a wavelength of 1470 nm.

4. Numerical simulations
4.1. Plane wave expansion (PWE) calculations
For a better understanding of the experimental results, the structures were simulated using the
plane wave expansion (PWE) method. The PWE calculations were performed with the MIT
Photonic-Bands (MPB) package, which is a free-software for computing the band structures
and electromagentic modes of periodic dielectric structures in three dimensions [17].
The photonic crystal slabs fabricated in the frame of this study are asymmetric with respect
to the bisecting horizontal plane since the substrate (BCB) differs from the cladding (air). In
general, asymmetric photonic crystal slabs do not support pure even or odd propagation modes
and strictly speaking a photonic bandgap does not exist. However, if the propagation modes
are strongly localized within the slab, they may still be approximated as even-like or odd-like

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20297
modes and band gap effects are still possible [18, 19]. In our case, this approximation is justified
due to the high refractive index contrast between the LiNbO3 slab and the surrounding mater-
ial. Figure 7 shows the band structure of the symmetric (BCB/LiNbO3 /BCB) and asymmetric
(air/LiNbO3 /BCB) photonic crystal slab with a triangular lattice of air holes (r/a = 0.24). In the
symmetric photonic crystal slab [Fig. 7(a)], the propagation modes are classified as even or odd
modes with respect to the mirror plane bisecting the slab. Only in the mirror plane the modes
are purely TE or TM polarized and therefore we can roughly regard even modes as TE-like and
odd modes as TM-like. The graph reveals a bandgap only for even modes extending from f’ =
0.333 to f’ = 0.359 (with normalized frequency f’=ωa/2πc) corresponding to the wavelength
range between 1390 and 1500 nm for a lattice constant of 500 nm. The asymmetric slab [Fig.
7(b)] shows only some minor differences in the band structure and by comparison with the
symmetric slab we can indeed easily ascribe the branches to even-like and odd-like modes. The
photonic gap in ΓK-direction extends from f’ = 0.333 to f’ = 0.387 (1290 to 1500 nm), which
agrees well with our transmission measurements and thus confirms the presence of a photonic
bandgap.

a) b)
Normalized Frequency f‘ (ωa/2πc)

Even light cone MK light cone


Odd
0.4 0.4
Γ

0.3 0.3

0.2 0.2

0.1 BCB 0.1 air


LN LN
BCB BCB
0 0
Γ M K Γ Γ M K Γ
Fig. 7. Photonic band diagrams of the symmetric (a) BCB/LiNbO3 /BCB and asymmetric
(b) air/LiNbO3 /BCB photonic crystal slab with a triangular lattice of air holes (r = 120 nm,
a = 500 nm, r/a = 0.24). The z-cut LiNbO3 slab has a thickness of 600 nm. The bandgap
exists only for even modes and ranges from 1390 nm to 1500 nm as depicted by the green
area between the dotted lines in band diagram (a).

4.2. Finite-difference time-domain (FDTD) simulations


In addition the light propagation was modelled using a commercially available finite-difference
time-domain (FDTD) software (OptiFDTD, Optiwave Systems Inc.). The software enables two-
and three-dimensional computations on workstations with multi-core 64-Bit processors, which
results in reasonable calculation times. In the calculations, ten layers of anisotropic perfectly
matched layer absorbing medium (APML) truncate the FDTD lattice in order to avoid un-

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20298
wanted reflections from the edges. APML absorbing medium is mathematically equivalent to
the perfectly matched layer medium published by Berenger [20], but is based on a Maxwellian
formulation and is more computationally efficient [21]. In general, 3D FDTD calculations are
very time consuming. In order to achieve reasonable calculation times, the simulation area and
resolution were optimized.

a) Detection APML b)
600nm
ΓΚ LiNbO3 airholes
ΓM
1 µm
2µm
Y
BCB
X
Z
APML Injection

Fig. 8. (a) Schematic drawing of the structure for the FDTD simualtions. The injection
direction is along the ΓK axis of the photonic crystal. As in the experiment, the crystal-
lographic z-axis is perpendicular to the crystal slab plane. The structure is surrounded by
APM layers in order to avoid unwanted reflections from the edges of the simulation area.
(b) Schematics of the crossection of the photonic crystal simulated by 3D FDTD simulation
showing the 1-µm deep air holes in the LiNbO3 and the BCB layer underneath.
Normalized Transmitted Specturm [dB]

-5

-10

-15
Simulation
Experiment
-20
1100 1200 1300 1400 1500 1600
Wavelength [nm]

Fig. 9. Comparison of the calculated and measured transmission spectra. The black curve
shows the calculated normalized transmitted spectrum through a structure as shown in
Fig. 8 with lattice constant a = 500 nm, radius r = 120 nm and 15 rows of holes simulated
by the 3D FDTD method. The dotted red curve shows the measured optical transmission
for such a structure.

As shown in Fig. 8, the calculation area was reduced to the photonic crystal lattice only,
assuming that the spectrum of the transmitted light is not substantially influenced by the tapered
waveguide. Based on several simulations with varying calculation resolution, a spatial step size
of 15 nm for X and Y, and 80 nm for the Z direction was proved to be appropriate for our

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20299
calculations. Accordingly a time step size of 2.5*10−17 s was chosen.
The incident light was a Gaussian pulse propagating in the ΓK direction with its wavelength
centered at 1550 nm and a spectral bandwith of about 1700 nm. The inital spatial profile of
this pulse corresponds to a TE00 mode in 4 × 0.6 µm LiNbO3 ridge waveguides, which were
used as in and output waveguides in the experimental part. Finally the computed transmission
spectrum was normalized by the spectral power of the incident light.
Figure 9 shows a comparison of the calculated and measured transmission spectrum. A power
dip in the calculated transmission, which ranges from 1280 to 1500 nm reveals a good agree-
ment between 3D FDTD simulation and the experimentally measured data. We contribute the
small shift in wavelength mainly to little differences between the designed and fabricated struc-
ture.

5. Conclusions
In conclusion, we have fabricated photonic crystal structures in ion-sliced, single-crystalline
LiNbO3 thin films with a BCB barrier layer. SEM pictures show the good quality of the struc-
tures and the advantages of using ion-sliced thin films with large refractive index contrast over
annealed proton exchanged waveguides. Using the PWE method we show that in the fabricated
asymmetric LiNbO3 PhC slab the modes are localized enough to define them as odd-like and
even-like. Consequently this results in the existence of a photonic bandgap around 1470 nm.
Simulation of light propagation by 3D FDTD calculations generates similar results showing a
power dip in the same wavelength range with an extinction ratio of about 20 dB. Experimentally
measured data confirms these results and reveals an exctinction ratio of about 15 dB. Our fur-
ther work will focus on the realization of electro-optically tunable photonic bandgap structures
in combination with other integrated optics elements.

Acknowledgments
This work was supported by ETH Research Grant TH-21/07-1 and the Swiss National Science
Foundation (200020-119961). We acknowledge support by the Electron Microscopy Center of
ETH Zurich (EMEZ), especially Philippe Gasser for help with the focused ion beam device.
Further we wish to thank Steffen Reidt from the Nonlinear Optics Laboratory of ETH Zurich
for his valuable technical support. We are also grateful to the AIM team at the Research Center
Dresden-Rossendorf, Germany, for performing the He+ implantation of LiNbO3 wafers in the
frame of the RITA Program, Contract No. 025646.

#116878 - $15.00 USD Received 8 Sep 2009; revised 8 Oct 2009; accepted 19 Oct 2009; published 22 Oct 2009
(C) 2009 OSA 26 October 2009 / Vol. 17, No. 22 / OPTICS EXPRESS 20300

Вам также может понравиться