Download Visualization of optical deflection and switching operations by a

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Gaseous detection device wikipedia , lookup

Optical rogue waves wikipedia , lookup

Dispersion staining wikipedia , lookup

X-ray fluorescence wikipedia , lookup

3D optical data storage wikipedia , lookup

Optical amplifier wikipedia , lookup

Photomultiplier wikipedia , lookup

Rutherford backscattering spectrometry wikipedia , lookup

Birefringence wikipedia , lookup

Laser beam profiler wikipedia , lookup

Night vision device wikipedia , lookup

Surface plasmon resonance microscopy wikipedia , lookup

Silicon photonics wikipedia , lookup

Passive optical network wikipedia , lookup

Nonimaging optics wikipedia , lookup

Optical coherence tomography wikipedia , lookup

Ellipsometry wikipedia , lookup

Photon scanning microscopy wikipedia , lookup

Interferometry wikipedia , lookup

Optical tweezers wikipedia , lookup

Retroreflector wikipedia , lookup

Harold Hopkins (physicist) wikipedia , lookup

Magnetic circular dichroism wikipedia , lookup

Ultraviolet–visible spectroscopy wikipedia , lookup

Anti-reflective coating wikipedia , lookup

Nonlinear optics wikipedia , lookup

Opto-isolator wikipedia , lookup

Transcript
Visualization of optical deflection and switching
operations by a domain-engineered-based
LiNbO3 electro-optic device
Giuseppe Coppola, Pietro Ferraro, and Mario Iodice
Istituto per la Microelettronica e i Microsistemi del CNR, Sezione di Napoli,
via P. Castellinio,111 – 8013, Napoli, Italy
[email protected]; [email protected]; [email protected]
http://www.imm.cnr.it
Sergio De Nicola
Istituto di Cibernetica "E. Caianiello" del CNR, ViaCami Flegri, 34, 80072, Pozzuoli (Napoli), Italy
[email protected]
Simonetta Grilli
Istituto Nazionale di Ottica Applicata, Sezione di Napoli, ViaCampi Flegrei, 34,
80072, Pozzuoli (Napoli), Italy
[email protected]
http://www.ino.it
Davide Mazzotti and Paolo De Natale
Istituto Nazionale di Ottica Applicata
Largo E. Fermi,6 50125-Firenze, Italy
[email protected]; [email protected]
http://www.ino.it
Abstract: An electro-optic device applied as an optical beam deflector and
switch at different wavelengths has been built and tested. The electro-optic
device is based on domain-engineered lithium niobate (LiNbO3). In this
paper, for the first time, its operation has been visualized by an imaging
camera. The device has been characterized both at the visible wavelength
(632.8 nm) and at a typical telecom wavelength (1532 nm). Furthermore,
the device has been tested as an amplitude modulator in the mid-infrared
region as well, at a wavelength of ~4.3 µm, where no Pockels cells are
available. A detailed description of this device is given, and the
experimental results are discussed.
 2003 Optical Society of America
OCIS codes: (230.2090) Electro-optical devices; (160.3730) Lithium niobate; (230.0040)
Detectors; (230.6120) Spatial light modulators
References and links
1.
2.
3.
4.
Y. Chiu, V. Gopalan, M. J. Kawas, T. E. Schlesinger, D. D. Stancil, and W. P. Risk, “Integrated optical
device with second-harmonic generator, electrooptic lens, and electrooptic scanner in LiTaO3,” J.
Lightwave Technol. 17, 462-465 (1999).
K. T. Gahagan, V. Gopalan, J. M. Robinson, Q. X. Jia, T. E. Mitchell, M. J. Kawas, T. E. Schlesinger, and
D. D. Stancil, “Integrated electro-optic lens/scanner in a LiTaO3 single crystal,” Appl. Opt. 38, 1186-1190
(1999).
M. Yamada and M. Saitoh, “Electric-field induced cylindrical lens, switching and deflection devices
composed of the inverted domains in LiNbO3 crystal,” Appl. Phys. Lett. 69, 3659-3661 (1996).
H. Gnewuch, C. N. Pannell, G. W. Ross, P. G. R. Smith, and H. Geiger, “Nanosecond response of Bragg
deflectors in periodically poled LiNbO3,” Phot. Technol. Lett. 10, 1730-1732, (1998).
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1212
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.
A. J. Boyland, S. Mailis, J. M. Hendricks, P. G. R. Smith, and R. W. Eason, “Electro-optically controlled
beam switching via total internal reflection at a domain-engineered interface in LiNbO3,” Opt. Commun.
197, 193-200 (2001).
R. W. Eason, A. J. Boyland, S. Mailis, and P. G. R. Smith, “Electro-optically controlled beam deflection
for grazing incidence geometry on a domain-engineered interface in LiNb O3,” Opt. Commun. 197, 201207 (2001).
M. Yamada, N. Nada, M. Saitoh, and K. Watanabe, “First-order quasi-phase matched LiNbO3 waveguide
periodically poled by applying an external field for efficient blue second-harmonic generation,” Appl. Phys.
Lett. 62, 435-436 (1993).
J. Webjörn, V. Pruneri, P. St. J. Russell, J. R. M. Barr, and D. C. Hanna, “Quasi-phase-matched blue light
generation in bulk lithium niobate, electrically poled via periodic liquid electrodes,” Electron. Lett. 30, 894895 (1994).
V. Pruneri, J. Webjörn, P. St. J. Russell, J. R. M. Barr, and D. C. Hanna, “Intracavity second harmonic
generation of 0.532µm in bulk periodically poled lithium niobate,” Opt. Commun. 116, 159-162 (1995).
G. D. Miller, R. G. Batchko, M. M. Fejer, and R .L. Byer, “Visible quasi-phase-matched harmonic
generation by electric-field-poled lithium niobate,” SPIE 2700, 34-36 (1996).
S. Grilli, S. De Nicola, P. Ferraro, A. Finizio, P. De Natale, M. Iodice, and G. Pierattini, “Investigation on
overpoled lithium niobate patterned crystal,” in ICO XIX, 19th Congress of the International Commission
for Optics , Technical Digest, Italy, 25-31 August 2002, Part 2, pp. 735-736.
S. Grilli, S. De Nicola, P. Ferraro, A. Finizio, P. De Natale, G. Pierattini, and M. Chiarini, “Investigation on
poling of lithium niobate patterned by interference lithography,” in Integrated Optical Devices: Fabrication
and Testing, Proc. SPIE 4944 (2002).
A. M. Prokhorov and Y. S. Kuzminov, Physics and Chemistry of Crystalline Lithium Niobate (Hilger,
Bristol, UK, 1990).
D. Mazzotti, P. De Natale, G. Giusfredi, C. Fort, J. A. Mitchell, and L. W. Hollberg, “Difference-frequency
generation in PPLN at 4.25 µm: an analysis of sensitivity limits for DFG spectrometers,” Appl. Phys. B 70,
747-750 (2000).
D. Mazzotti, S. Borri, P. Cancio, G. Giusfredi, and P. De Natale, “Low-power Lamb-dip spectroscopy of
very weak CO2 transitions near 4.25 µm,” Opt. Lett. 27, 1256-1258 (2002).
M. Reich, F. Korte, C. Fallnich, H. Welling, and A. Tunnermann “Electrode geometries for periodic poling
of ferroelectric materials,” Opt. Lett. 23, 1817-1819 (1998).
I. E. Barry, G. W. Ross, P. G. R. Smith, and R. W. Eason, “Ridge waveguides in lithium niobate fabricated
by differential etching following spatially selective domain inversion,” Appl. Phys. Lett. 74, 1487-1488
(1999).
1. Introduction
In the past ten years, the growing need for optical switches, laser scanning, and highbandwidth signal processing systems—which require fast modulation, low cost, and low
power consumption—has resulted in the construction of many novel high-speed photonic
devices. Electro-optical beam deflectors are good potential candidates for these applications,
and in particular domain-engineered samples of ferroelectric materials, such as lithium
niobate (LN), seem to satisfy all the aforementioned specifications [1-6]. In this paper, an
electro-optically addressable device in a sample of LiNbO3, which can be used as either an
optical deflector or an optical switch, is investigated. A similar device was originally
developed by Eason and co-workers at the Optoelectronics Research Center in Southampton
(UK) [5, 6].
The device presents a sharp boundary between two oppositely oriented domain regions
obtained by means of an electric-field poling process. When an external electric field is
applied to this boundary, equal magnitude refractive-index changes of opposite sign will occur
between the adjacent domain regions. In this way, the normal electro-optic-induced refractiveindex change is doubled. Thus, an optical beam crossing the interface is subjected to both
reflection and refraction.
A great improvement in the angular deflection sensitivity can be obtained by use of a socalled grazing angle geometry, which is when the angle of the beam impinging on the
interface approaches the total internal reflection (TIR) limit angle. In this configuration it is
possible to achieve relatively large deflection angles by use of a control signal of a few
hundred volts. Moreover, when the incident angle exceeds the limit angle, i.e., the TIR
phenomenon onsets, the device operation is not continuous. In fact, it exhibits an abrupt
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1213
deflection of the light beam. In other words, without application of voltage the light passes
through the boundary practically unchanged, and only a residual poling-induced strain
occurring at the boundary can slightly affect the propagation of the beam. With application of
external voltage, an opposite refractive-index change will be induced to the adjacent domain
regions so that if the beam light arrives from the region whose refractive index has increased
and impinges on the interface at angles larger than the limit, it will be reflected with a
theoretical efficiency of 100%.
The angle necessary for TIR, at an electro-optically controlled interface, is given by the
usual expression:
n − ∆n
(1)
sin θ TIR =
n + ∆n
The main equation relating the induced refractive-index change ∆n, under application of a
voltage V, to the electric field amplitude, is given by:
1
V
(2)
∆n = − r n 3
2
33
e
d
where r33 is the largest electro-optic coefficient accessed by extraordinary s-polarized light, ne
is the wavelength dependent value of extraordinary refractive index, and d is the thickness of
the device. The induced refractive-index change related to p-polarized light can be derived
from Eq. (2) by replacing the electro-optic coefficient r33 with r13 and the extraordinary
refractive index (ne) with the ordinary refractive index (no). Anyway, this induced refractiveindex change is not particularly useful because its value is only about one-third than that
obtained using r33 and ne in the above equation.
The aforesaid device provides several advantages, including an easy fabrication, with the
possibility of achieving high contrast ratios, relatively low electric drive power, and finally a
wavelength dependence that is much less critical than in other electro-optic devices, such as
Pockels cells. In fact, a Pockels cell is characterized by the half wave voltage, i.e.,
Vπ=λ/(2⋅n3e⋅r33), that is directly proportional to the wavelength. In the domain-engineered
device, the switching voltage, i.e. V=(∆n⋅d)/(n3e⋅r33), is effectively wavelength independent,
because intrinsic wavelength dispersion characteristics of both the electro-optic coefficient
(r33) and refractive index (ne) are almost negligible in a relatively small wavelength range.
This feature is particular interesting for the spectral ranges where no Pockels cells are
available and switch performances faster than mechanical chopping are demanded.
The presented device has been tested at visible wavelength (632.8 nm), at typical telecom
wavelength (1532 nm) and, finally, in the mid-infrared region at a wavelength of 4.3 micron,
where no Pockels cells are available and such electro-optic device could be of great interest.
In this paper a detailed description of the design, fabrication and testing of the electrooptic device are reported. In particular, technological steps for room-temperature electric-field
poling are given in Section 2. Performance of the device working like either an optical
deflector or optical switch is detailed in Section 3. Finally, further developments and
cocnclusions are discussed in Section 4.
2. Technological processes
In Fig. 1 the external electrical circuit employed to reverse ferroelectric domains in LN
samples is reported [7-12]. The local reversing of the sample spontaneous polarization is
achieved by applying a high positive voltage slightly exceeding the LN coercive field (21.0
kV/mm). A conventional Voltage Generator (VG) drives an High Voltage Amplifier (HVA 2000x), provided by Trek, Inc., with a series current limiting resistor, R S = 50 MΩ , in order
to produce a positive voltage of 12kV. A diode rectifier was connected to the output of the
HVA to prevent backswitch current flowing in the circuit.
The polarization reversal has been carried out on 500-µm-thick lithium niobate single
domain crystal samples (Crystal Technology Inc). The samples were z-cut 3-inch diameter
wafers with both sides polished. After solvent cleaning, the z+ face was spin coated with a
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1214
1.3-µm-thick photoresist layer (Shipley S1813-J2) and then exposed through a standard mask.
The positive high voltage was applied over the z+ patterned crystal face by using liquid
electrolyte (LiCl in deionized water). The liquid electrode fixture consists of two-electrolyte
containing chambers which squeeze the LiNbO3 sample between O rings.
Vpol
RS
D
V+
VG
HAV
(2000x)
HVP
LN sample
Ipol
Z
Rm
Osc
Fig. 1. Electrical circuit employed for poling lithium niobate samples.
The LN conductivity at room temperature is low enough that the poling current Ipol
flowing in the circuit can be readily monitored. This current was measured by recording the
voltage drop across the resistance Rm=10 kΩ, while the poling voltage Vpol was measured
using a conventional High Voltage Probe (HVP). Both the current and voltage waveforms
were visualized on the oscilloscope during the poling process.
Figure 2 illustrates an optical micrograph of the interface between the two reversed
ferroelectric domains. The interface has been revealed by an etching process of one hour in a
HF:HNO3=1:2 acid mixture at room temperature. This acid mixture etches the z– face much
faster than the z+ face. On a zoomed scale, the interface shows evident lack of homogeneities,
but locally it seems to be smooth enough for the construction of our devices.
z+ domain
z- domain
Interface
Fig. 2. Optical micrograph of the interface between reversed domains as revealed by an acid
mixture (HF:HNO3).
The electro-optic device was fabricated by cutting a 20 mm × 15 mm large sample from
the obtained domain-engineered LN wafer and by polishing its edges to have optical quality
facets for light coupling. Finally, gold electrodes were sputtered on both z– and z+ faces
across the domain interface region. Thus, applying an electrical field Ez over these electrodes,
a refractive-index change, equal in magnitude and opposite in sign, can be achieved between
the adjacent anti-parallel domain regions.
3. Device characterizations
The aforementioned technological steps were used for fabricating a device which is able to
work as both beam deflector and Total Internal Reflection-based switch depending on the setup employed. In order to characterize both the operation modes, the device has been mounted
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1215
on an insulating goniometric support, which allows a fine control of the angle of incidence for
the light beam (Fig. 3).
goniometric
support
gold electrode
∆x
half-wave
plane
LASER
Electro-optic
device
CCD
d
Fig. 3. Schematic setup employed to characterize the domain engineered device.
For the characterization in the visible range, experiments have been performed using a
He-Ne polarized laser source at λ=632.8 nm, with an output optical power of about 1.0 mW.
A half-wave plane has inserted at the output of the laser to select the required linear light
polarization. Then, the laser beam has been focused onto the LN sample edge, and the output
beam has been collected by means of a 560x800 pixels silicon CCD array with a pixel
dimension of 10 µm. Finally, the electrical CCD output was acquired by a standard PCI frame
grabber card. For the visible spectral range, the electro-optic coefficient (r33) and the
extraordinary refractive index (ne) are assumed to be equal to 33⋅10-12 m/V and 2.14,
respectively [13].
The device has been characterized also in the infrared range. A first experiment has been
performed at a telecommunication wavelength by employing a He-Ne laser source emitting at
λ=1532 nm with an output power of about 5.0 mW, and replacing the CCD array with an
infrared Vidicon camera. Moreover, a Glan-Thompson filter has been used instead of the halfwave plane. Additionally, also radiation at a wavelength of about 4.3 µm, generated by a
nonlinear difference-frequency process in a periodically poled lithium niobate crystal [14],
has been used to characterize the device. For this experiment an InSb liquid-N2-cooled
photodiode and CaF2 lenses have been employed.
The experimental results, obtained for both operation modes, are illustrated in the
following paragraphs.
3.1. Beam deflector
In Fig. 4 is shown the simple geometry for a beam deflection at the interface between the two
regions for a domain engineered LN scanner, in the case of incident angle (θi) at interface
smaller than the limit angle (θTIR).
n+|D n|
q
q
inp
i
q
t
q
out
n-|D n|
Fig. 4. Schematic geometry for the beam deflector.
We should note the values of the transmitted angle (θt) as function of the applied voltage
and for different incident angles at the visible wavelength of 632.8 nm (Fig. 5).
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1216
89.8
θi=89.6°
89.6
θi=89.0°
Trasmitted angle θt [deg]
89.4
θi=89.2°
89.2
89.0
θi=88.8°
88.8
88.6
θi=88.4°
88.4
88.2
θi=88.0°
88.0
87.8
-600
-400
-200
0
200
400
600
Applied Voltage [V]
Fig. 5. Transmitted angle as function of the applied voltage, for different incident
angles at the wavelength of 632.8 nm.
It can be noted that, unlike the case for different kind of electro-optic scanner geometry,
as the prism based one, the grazing incidence geometry is extremely sensitive to small
changes of the local refractive index. Moreover, it is evident that approaching the limit angle
substantial deflection can be obtained even for low control voltage amplitude. Although for
small values of the incident angle (θi), as 88.0 and 88.4 degrees, the relationship between the
transmitted angle and the applied voltage is approximately linear, whereas when the incident
angle approaches the value of 89.0 degrees, the TIR phenomenon occurs and the relationship
is non-linear. Thus, it is very advantageous to use the beam deflector with an angle of
incidence approaching, but not equal to the limit angle, in order to obtain maximum deflection
for a few hundreds of applied voltage. Obviously, the advantage from applying across the
interface a voltage from negative to positive values is to get a doubled total deflection angle.
In Fig. 6 both the simulated and measured shift of the transmitted optical beam, in the visible
range, are reported, for three values of the input angle (θinp).
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1217
4.0
Displacement [mm]
2.0
0
q inp=3.8°
-2.0
q inp=2.0°
q inp=3.3°
-4.0
-6.0
-600
-400
-200
0
200
400
600
Applied Voltage [V]
Fig. 6. Comparison between the simulated and measured shift of the transmitted
optical beam for three different values of the input angle in the visible range.
The sign of the displacement is related to the initial position (i.e., without any applied
voltage) of the light beam; thus the minus (plus) sign reflects a separation from the initial
position towards right (left). Simulated values are obtained considering the following
relationship:
∆x = tg (θ out ) ⋅ d ≅ θ out ⋅ d
(3)
where d=20cm is the distance between the sample and the CCD array (see Fig. 3), and θout is
evaluated taking in account the change of the refracted beam direction at the electro-optically
addressed interface and refractions due both to input and output interfaces between LiNbO3
and air (see Fig. 5). So, applying the Snell’s law to the three interfaces, and bearing in mind
some easy trigonometrical relationships, the output angle is given by:
2
2 

 n + ∆n    sin θ inp   
(4)
 
 1 − 
θ out = sin −1 (n − ∆n ) 1 − 
   n + ∆n   

n
n
−
∆
 
  



where ∆n is defined in Eq. (2). As is possible to observe, there is a good agreement between
simulated and theoretical values of the displacement. The slight discrepancy mainly arises in
the non-ideal profile of the domains interface region and in the uncertainty measure both of
the absolute value of the input angle (θinp) and of distance between the device and the CCD
array (d). Moreover, for the input angle of 2.0 degrees and for applied voltage larger than
300V, the output angle has not been measured, because the deflection was so large to leave
the sensitive area of the CCD array. From the theoretical prediction relating to the value of the
input angle of 2.0°, can be observed that for applied voltage larger than 400 V the Total
Internal Reflection effect arises and the deflected beam is not more present, so the shift cannot
be defined.
In Fig. 7 are shown the optical field distribution taken from the CCD camera at a distance
of 20 cm far from the sample, setting an input angle of 3.3 degrees, without applied voltage
and for the two extreme values of the applied voltage V=±650 V.
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1218
0.9
V=-650 Volt
V=+650 Volt
0.8
V=0 Volt
Intensity [a.u.]
0.7
0.6
0.5
0.4
0.3
0.2
0.1
-4
-3
-2
-1
0
1
Lateral shift [mm]
2
3
4
Fig. 7. Optical field distribution at a distance of 20 cm from the device, for three different
values of the applied voltage at visible range.
The resulting lateral shift is about 3.5mm for positive value of the applied voltage and 3.0
mm for the negative one. The difference can be justified observing that, when a positive
voltage is applied, the deflector approaches the TIR condition, moving on the non linear
region of the curves in Fig. 5, while when a reversed voltage is applied, it goes away from the
more efficient non-linear region, moving towards the less efficient linear part of the
characteristic. In Fig. 8 is reported a movie, obtained by successive acquisitions of the
deflected beam applying a voltage from -650 V to +650 V. The shown behavior points up the
possibility to employ the device as an optical scanner.
Fig. 8. (262 KB) Movie to illustrate the scanner functionality of the domain-engineered-based
device.
The same device has been characterized in near infra-red range with the set-up described
at beginning of the paragraph, and in Fig. 9 the shift obtained placing the Vidicon camera at a
distance of 27.5 cm from the device, is reported. These data, obtained with a large input angle
(θinp), confirm the linear relationship between the applied voltage and the beam shift.
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1219
0.6
0.4
Displacement [mm]
0.2
0
-0.2
-0.4
-0.6
-0.8
-800
-600
-400
-200
0
200
400
600
800
Applied Voltage [V]
Fig. 9. Shift of the transmitted optical beam evaluated at telecom wavelength (λ=1532 nm).
3.2. TIR-based switch
The same device illustrated in the previous paragraph can be used as a digital optical switch
exploiting the Total Internal Reflection effect in order to obtain an abrupt deflection of the
light beam from the transmitted path to the reflected one. This effect occurs when the angle of
incidence (θi) is larger than the limit angle (θTIR). This angle can be evaluated from the Eq. (1)
for both crystal polarizations, and its value becomes handy only for applied voltage larger
than 300 V.
The schematic of the geometry of the interaction between the beam and the interface is
reported Fig. 10. The schematic characterization of the switch has performed at an angle of
incidence of about 89.0 degrees, determined with an error of ±0.1°.
q
TIR On
inp
q i> q
TIR
TIR Off
Fig. 10. Schematic of the interaction between the optical beam and the interface for the switch
device.
For the visible range (λ=632.8 nm), optical reflected power for s- and p-polarization are
illustrated in Fig. 11(a), as function of the applied voltage. As expected, considering the
values of the respective electro-optic coefficient for each polarization orientation, the
switching of s-polarized light occurs for smaller applied voltage than of p-polarized. For this
polarization, the observed extinction ratio between the TIR-on and the TIR-off state is greater
then 20 dB. Moreover, in Fig. 11(b) is reported the optical transmitted power for spolarization (the more efficient configuration). Even in this case the contrast reachable in
greater than 20 dB. From Fig. 11(a) and Fig. 11(b) can be noted that when TIR is approached
there is not a sharp variation both of reflected and transmitted power. In fact, unless the beam
is very well collimated, each portion of the beam experiences a different angle of incidence
and, as consequence, the voltage required for TIR is slightly different for each portions of the
beam. For this reason near TIR, a smooth transition to the reflected beam has been achieved.
In Fig. 12 is reported a movie, obtained by successive acquisitions of both the deflected
and reflected beam for an s-polarized light at the visible wavelength, applying a voltage from
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1220
0 to +650 V. Though the image of the optical beam is not well defined owing to a no perfect
optical alignment, it’s evident that as long as the TIR condition is not satisfied, increasing the
applied voltage, both the transmitted angle and optical reflected power increase; when the TIR
condition is reached, the optical transmitted power becomes null, whereas the reflected power
gets to the maximum.
6
5
Transmitted power [a.u.]
Reflected power [a.u.]
2.5
s- polarization
2.0
1.5
p - polarization
1.0
4
3
s- polarization
2
0.5
1
0
0
100
300
400
200
Applied voltage [V]
(a)
500
600
0
0
100
200
300
400
500
600
Applied voltage [V]
(b)
Fig. 11. (a) Optical reflected power both for s- and p- polarization as function of the applied
voltage. (b) Optical transmitted power for s- polarization function of the applied voltage.
The lithium niobate TIR switch has been tested with a 4.3 µm optical radiation too. This
radiation, generated in a periodically-poled lithium niobate crystal, is useful for spectroscopy
applications [15]. A positive sinusoidal voltage, of 500 V peak-to peak, has been applied
between the upper and lower golden electrodes of the device. A knife edge stopped the beam
transmitted by the internal interface between the two domains. A maximum modulation depth
of 70% on the reflected beam has been measured, up to a 10 kHz frequency, limited by the
bandwidth of employed high-voltage amplifier. This experiment has demonstrated the
possibility to perform amplitude modulation faster than mechanical chopping, in a wavelength
region where no Pockels cells are available.
Fig. 12. (235 KB) Movie to illustrate the TIR functionality of the domain-engineered-based
device at the visible wavelength of 632.8 nm.
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1221
4. Further developments and conclusions
The deflection and switching characteristics mentioned above could be significantly improved
in an optimized version of the poled region, where a lateral deep wet etch can be used to
produce a local narrowing [16] of the device in correspondence with the interface between the
inverted domains, thus reducing the needed applied voltage. This approach is currently under
development in our laboratory. Moreover, the device could be integrated in a guided-wave
structure, which could be fabricated by three different methods of proton exchange, ion-beam
implantation, and titanium indiffusion [17]. These optical waveguides can be designed and
realized to illuminate the active domain-engineered region of the device and to collect
refracted and/or reflected light from it. This kind of design could make the device efficient
and easy to couple to external optical components and circuits.
In conclusion, an electrooptic device able control a light beam spatially by means of
induced deflection and/or reflection has been presented. When used as a deflector, it permits a
wide angular scan range. When used as a TIR switch, the device exhibits digital behavior,
with abrupt switching of light from direct to reflected path. In the visible range it has shown a
good contrast ratio, greater than 20 dB. Experiments for visualizing the two operations have
been conducted, and results have been discussed. Moreover, the possibility of performing
amplitude modulation has also been demonstrated in the mid-infrared range. In particular, an
experiment at λ=4.3 µm, where no Pockels cells are available, has been carried out.
Acknowledgments
The authors thank M. Varasi of Alenia Marconi System and A. Zani of IMM-CNR for their
help in performing this research.
#2140 - $15.00 US
(C) 2003 OSA
Received February 14, 2003; Revised April 29, 2003
19 May 2003 / Vol. 11, No. 10 / OPTICS EXPRESS 1222