You are on page 1of 5

Mo.D2.

1 44 ICTON 2006

Photonic Crystal Sensors


D. Biallo, A. D’Orazio, M. De Sario, V. Marrocco, V. Petruzzelli
Dipartimento di Elettrotecnica ed Elettronica – Politecnico Di Bari, Via Re David 200, 70125 Bari (Italy)
Phone (39) 080 5963268; fax (39) 080 5963410; e-mail: dorazio@poliba.it
F. Prudenzano
Dipartimento di Ingegneria per l’Ambiente e lo Sviluppo Sostenibile – Politecnico di Bari
Viale del Turismo 8, 74100 Taranto (Italy)
ABSTRACT
The paper deals with the possibility of using photonic crystals for sensing purpose. The optical properties of
photonic crystals allow to realize sensing devices characterized by a high degree of compactness and a good
resolution of the quantity to detect. A particular attention is devoted to force/pressure sensors and the design of
a PhC microcavity pressure sensor is reported.

1. INTRODUCTION
Sensing systems, optoelectronics, nanotechnology: these vast fields of research are growing together and new
applications characterized by extremely low dimensions and the combination of new materials and technologies
are proposed in various systems such as nanostructures and nanomachines, microreactors, displays, chemical
biochips, telecommunications, human/machine interfaces, miniaturized medical and surgery equipments, sensor
systems. The interest in microsensors [1] is dramatically increased during the last decade of 20th century, due to
the fact that they guarantee a relatively stable price-performance ratio. An ambitious goal of the Research in last
years has been the fabrication of monolithic or integrated chips that can not only sense (with microsensors) but
also actuate (with microactuators), that is, the fabrication of a microsystem that encompasses the information-
processing triptych. In the last years, micro-electro-mechanical (MEMS) and micro-opto-electro-mechanical
systems (MOEMS) technologies have been considerably developed and the expansion of the application fields
can be expected to grow both in terms of academic and commercial interest [2-4]. Before the development of
micromechanical devices, macroscopic-force sensors were used in many instruments. MEMS devices enabled
a completely new approach to many common instruments and the opportunity of miniaturization enables force
sensing to become important in many new applications [5]. In general, miniaturization of the dimensions of
a force sensor imposes a reduction in the force to be detected. Therefore, miniature sensors are generally
required to detect smaller forces than macrosensors. One of the key components of a force sensor is the
displacement transducer that is used to detect the induced deflection. Many transduction mechanics have been
explored: among them it is important to recall piezoresistive [6] magnetomotive [7], tunneling-based transducers
[8] and so on. MOEMS applications include proximity sensors, force transducers, magnetic/electric field sensors
or temperature or strain or pressure or mass flow accelerometers, chemical sensors and so on.
However the possibilities for new sensor concepts have been hardly explored, so the field of sensing
applications is also expanding scientifically. There is a need for other properties in sensors which are not yet
possible in MEMS devices such as fast responsivity, high sensitivity and wireless features. Integrated optics has
been one of the most interesting research fields due to the clear advantages that it has as compared to the
traditional microelectronics: its capability to resist harsh environments, immunity to electromagnetic
interference, safety in explosive media, electrical passivity, measurement without direct contact, good sensitivity
with high dynamic range and low-cost manufacturing. The importance of optical sensors in future sensing
systems has been widely recognized [9-14]. Different configurations of optical sensors are reported in literature,
based on interferometric waveguiding devices, such as Mach Zehnder [15] or Michelson [16] or which exploit
intermodal interference between modes having different polarization [17]. Recently, a pressure and acceleration
sensor based on anti-resonant waveguides in which the displacement of a mass produce a misalignment of three
waveguides has been proposed [18].
The Photonic Crystal (PhC) technology [19-21] can be useful exploited for sensing purpose. It is well known
that a PhC slab, consisting of a two-dimensional periodic index contrast introduced into a high-index guiding
layer, supports in-plane guided modes that are completely confined by the slab without any coupling to external
radiation. In addition to in-plane waveguiding, photonic crystal slabs can also interact with external radiations in
complex and interesting ways [22]. Of particular importance for sensing applications is the possibility to tune
the operating wavelength of a PhC-based device by changing the refractive index of the materials of the PhC.
It has been demonstrated that it is possible to induce changes in the refractive index of material by means of
thermo-optical effect [23], carrier plasma effect [24], electro-optical effect [25]. Moreover, the modulation of
the optical properties of photonic crystals has been obtained by the infiltration of conducting polymers,
photoresponsive organic materials and liquid-crystals [26]. The optical properties of a photonic crystal structure
can be also changed by modifying its geometrical parameters. This can be accomplished if the PhC materials can

1-4244-0236-0/06/$20.00 ©2006 IEEE


ICTON 2006 45 Mo.D2.1

be bent, deformed or compressed. In this case the unit cell of the PhC structure and its energy dispersion
diagram are modified and the transmittance and reflectance of the device result to be altered.
The possibility to modify dynamically the geometrical parameters of the PhC structure makes thus possible
the realization of sensitive optical device and thus the 2DPhC technology extremely attractive for the fabrication
of physical sensors.

2. PHOTONIC BAND GAP STRUCTURES FOR SENSING APPLICATIONS


The application of photonic crystals for optical sensing purposes is an open research field. The key advantage of
a photonic crystal sensing device is the extreme high degree of compactness and the obtainable resolution of the
quantity to detect. Optical fiber based sensors offer clear advantage such as immunity to electromagnetic
interference, small size, low cost and the possibility for distributed measurements. Different photonic band gap
fiber design have been reported in literature to detect gas or for stress, temperature and pressure measurements
[27-34].
One- and two-dimensional photonic crystal slabs have been also employed in sensing applications. In [35]
a quasi-one-dimensional photonic crystal consisting of a wide ridge-type channel waveguide of silicon nitride
with a grating etched into the core layer was proved to operate as a refractometric optical sensor.
In [36] a class of composite polymer films, consisting of 1024 alternating nanolayers of an elastomer and
a glassy polymer is reported in which the refractive index is varied by a simple mechanical force. The
ultracompact biochemical sensor built with two-dimensional photonic crystal microcavity, fabricated on silicon
on insulator substrate, can detect the change of the refractive index of 0.002 by measuring the resonant
wavelength of microcavity [37]. An interesting application is the tunable MEMS PhC, reported in [38], in
which a micro-bridge serves as substrate for the photonic crystal structure, consisting of two sets of mirrors
(holes) separated by a microcavity. A smart MOEMS, fabricated using a multi-layer, metal dielectric PBG
structure with electrically controllable optical properties was proposed in [39] which can be used for sensing
applications, in harsh or hostile environments, as pressure sensor, micro gravitometer or micro accelerometer.
Exploiting the guided resonance phenomenon, it is possible to fabricate structures that are highly sensitive to
longitudinal displacements, for example by using the photonic crystal slabs as mirrors to form Fabry Perot cavity
[40]. In this case, the highest achievable sensitivity is directly related to the maximum reflectivity expected in
a single slab.
The described structures have been realized by using multilayered structures such as commercial wafers or
heterostructures deposited by both epitaxial growth techniques, e-beam evaporation, thermal evaporation,
sputtering and by also plasma enhanced chemical vapour deposition (PECVD) techniques. Suspended structures
have been also considered, obtained by first performing an high resolution e-beam lithographic patterning
followed by a deep isotropic etching and by a final selective etching. This latter process makes possible to
remove the bottom sacrificial layer, allowing the realization of a membrane or a cantilever.
The choice of materials is often dictated by a specific need, or by required properties of the device meant to
realize, or by available fabrication technology. The materials usually used for sensing PhC planar devices are
heteroepitaxial layers such as AlGaAs/GaAs, silicon on insulator wafers (SOI), III-nitride compound layers or
dielectric layers such as Si3N4, TiO2, SiO2. Organic compounds, both small molecules and polymers, have
attracted in the last few years an increasing interest, due to their low cost, easy functionality and possibility to
finely tune their optical and electrical properties. In spite of their low refractive index contrast, several
application of photonic crystal devices based on polymers have been reported [41-43]. Since the organics are
typically not resistant to wet solutions and solvents used in micro and nanotechnological processes, the
fabrication of organic photonic crystal slabs requires the use of soft litographies such as hot embossing, injection
molding or replica molding [44], the latter being particularly suitable for the realization of elastomeric 2D
photonic crystal structures for the realization of force and pressure sensors.
Furthermore the possibility to realize a photonic crystal structure inside a soft and elastic materials would be
a straightforward way to realize pressure-sensitive 2D photonic crystal devices. This approach makes not
necessary the realization of a suspended structure since the applied force would directly modify the thickness,
the geometry of the unit cell of the PhC slab.

3. PHOTONIC CRYSTAL PRESSURE SENSOR DESIGN


The operation principle of the PhC pressure sensor is based on the fact that the force/pressure application
provokes, due to the photoelastic, piezoelectric and electrooptic properties of the materials constituting the
structure, a change of the refractive index which modifies the transmission spectrum of the regular photonic
crystal or that of the localized state when a defected photonic crystal is considered.
A preliminary analysis of the PhC has been performed using the commercial code by RSoft (FullWave),
considering the equivalent 2D structure obtained following the refractive effective index method [45].
A proprietary parallel 3D-FDTD computer code, running on a Linux Cluster CLX CINECA, having 1024 CPUs,
Mo.D2.1 46 ICTON 2006

has been used to analyze the nature of the resonant modes and to evaluate the Q factor of the PhC resonant
cavity. For all simulations, a grid size has been chosen equal to a/20, whereas the temporal step size is
3.178.10-17 s, corresponding to 90% of the Courant limit, and uniaxial perfectly matched layers (UPML) are
used to limit the computational domain. At first, an input pulse having a Gaussian temporal behavior,
characterized by a wide spectrum centered at λ = 1310 nm, is launched towards the photonic crystal in order to
excite the modes supported by the cavity. Then a Gaussian temporal shaped pulse, having smaller spectrum
centered at λ = 1310 nm, is launched onto the structure in order to analyze the field distribution and the
symmetry of each supported resonant mode of the cavity lying in the wavelength range 1300 – 1400 nm.
The PhC pressure sensor under consideration consists of a PhC cavity, realized in a GaAs/AlGaAs
waveguide, shown in Fig. 1a. The starting waveguide is constituted by a 300 nm thick GaAs core grown onto
a substrate of Al0.70GaAs0.30. The modal analysis of the waveguide at the wavelength λ = 1310 nm reveals that
the waveguide is single mode for both TE and TM polarization.

Monitor value [a.u.]

Wavelength [µm]
Fig.1. (a) Sketch of the PhC microcavity pressure sensor; (b) transmission spectrum
of the unperturbed microcavity.
Curva di calibrazione
1335
Monitor value [a.u.]

1330
Spectral position of resonant pick [nm]

1325

1320

1315

1310

1305
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5
Applied pressure [GPa]
Wavelength [µm]
Figure.2. (a) Transmission spectrum of the resonant cavity in presence of an applied force,
(b) Resonant peak position as a function of the applied pressure.
A triangular lattice of air holes, having period a = 400 nm and radius r = 0.35a is patterned in the
GaAs/Al0.70GaAs0.30 waveguide. The height of the air columns has been fixed, by 3D simulations, equal to
800 nm, value that guarantees a good resolution of the band gap and smaller losses. The PhC exhibits a large
band gap for TM polarization (magnetic field component parallel to air hole axis) between 1150 – 1800 nm
whereas a smaller stop band is observed for TE polarization between 1650 – 1850 nm. Therefore the following
analysis only concerns the TM polarization. An opportunely designed cavity, made by reducing the radius of
the hole in the PhC center till to r = 0.2a and varying the radius of the six surrounding holes from r = 0.35a to
r = 0.325a, is introduced in the regular lattice. The cavity, in absence of applied pressure, shows a localized
mode (see Fig. 1b) resonant at the wavelength λ = 1308 nm.
ICTON 2006 47 Mo.D2.1

To model the pressure action, the change of the dielectric impermeability tensor b of the GaAs and AlGaAs,
accounting for the strain and stress effects acting on the lattice of the PhC, has been evaluated by the following
expression:
∆b = (pE:s – rT . gT): Τ, (1)
where p is the photoelastic fourth-order tensor at constant electric field; s is the fourth-rank elastic compliance
sensor, rT = rS + p: dS is the electrooptic third-rank tensor at constant stress in which rS is the third-order
electrooptic tensor at constant strain u and dS is the third-rank piezoelectric tensor at constant strain; gT is the
third-rank piezoelectric tensor at constant stress, and T is the stress tensor.
By applying a force normal to the plane of photonic crystal, the refractive index change is principally due to
the photoelastic effect.
Fig. 2a shows the localized mode in the microcavity, evaluated when a pressure varying between
0.25 – 5 GPa is applied. The resonant mode changes its spectral position, following a linear law as it can be
noticed in Fig. 2b showing the shift of the resonant peak as a function of the applied force. By considering an
area of application of the force equal to 1 µm2, the numerical simulations indicate that the detectable minimum
force is equal to 0.25 mN.

4. CONCLUSIONS
In this paper, a detailed, even if not exhaustive, state of the art of the sensing applications is reported.
The design of a photonic crystal microcavity, etched in a GaAs/AlGaAs waveguide, for force/pressure
measurements is illustrated. The device is under fabrication.

ACKNOWLEDGMENT
This work was supported by MIUR – PRIN 2005, in the framework of the National research project “Photonic
band gap nanosensors”.

REFERENCES
[1] J.W. Gardner, Microsensors, Wiley, Chichester, 1994.
[2] N. Maluf, An introduction to Microelectromaechanical Systems Engineering, Artech House Publishers,
Boston, London, 2000.
[3] Issue on MOEMS, IEEE Journ. Select. Topics of Quantum Electronics, 5, (1999).
[4] Issue on Optical Microsystems, IEEE Journ. Select. Topics of Quantum Electronics, 10, (2004)
[5] T. Kenny, “Nanometer-scale Force Sensing with MEMS Devices”, IEEE Sensors Journal, 1, 148-157
(2001).
[6] J.A. Harley, “High-sensitivity piezoresistive cantilevers under 1000 A thick”, Appl.Phys. Lett., 75, 289-
291 (1999).
[7] A.N. Cleland and M.L. Roukers, “A nanometer-scale mechanical electrometer”, Nature, 392, 160-162
(1998).
[8] S.B. Waltman and W.J. Kaiser, “An electron tunneling sensor”, Sensors and Actuators, 19, 201-210
(1989).
[9] M. Izutzu, A. Enokihara, T. Sueta, “Integrated Optical Temperature and Humidity Sensors”, J. of
Lightwave Techn. LT-4, pp. 883-836.
[10] C.H. Bulmer, “Integrated Optical Sensors in Lihium Niobate”, Optics News, 20-23, (1988).
[11] M. Tabib-Azar, Integrated Optics, Microstructures and sensors, Kluwer Academic Publishers, MA, 1995.
[12] D. Rugar, H.K. Mamin and P. Guethehner, “Improved fiber-optic interferometer for atomic force
microscope”, Appl.Phys.Lett., 55, 2588-2590 (1989).
[13] G. Calò, A. D’Orazio, M. De Sario, V. Petruzzelli, F. Prudenzano, “Stability analysis of Ti:LiNbO3
integrated optical sensors for electromagnetic compatibility measurements”, in “Recent Research
Developments in Optics”, S.G. Pandalai Editor, vol.3, 563-574, (2003).
[14] A. D’Orazio, M. De Sario, C. Giasi, L. Mescia, V. Petruzzelli, F. Prudenzano, “Design of planar optic
sensors for hydrocarbon detection”, Optical and Quantum Electronics, 36, 507-526, (2004).
[15] R.G. Heideman, R.P.H. Kooyman, J. Greve, “Performance of a highly sensitive optical waveguide Mac-
Zehnder interferometer immunosensor”, Sensors and Actuators B, 10, 209-217, (1993).
[16] A. D’Orazio, M. De Sario, G. Ficarella, D. Grando, V. Petruzzelli, F. Prudenzano, “Design and
demonstration of interferometric integrated-optic sensors in Ti:LiNbO3 Waveguides”, Fiber and
Integrated Optics, 16, 369-386, (1997).
[17] M. Ohkawa, K. Hasebe, C. Nishikawi, S. Sekine, T. Sato, “Integrated optic pressure sensor using
intermodal interference between two mutual orthogonal guided modes”, Optical Review, 7, 144-148,
(2000).
Mo.D2.1 48 ICTON 2006

[18] A. Lloblera et al., IEEE Photonics Techonology Letters 16, 233, (2004).
[19] E. Yablonovitch, “Photonic band-gap structures”, J. of Optical Society of America B, 19, 283-295, (1997).
[20] J.D. Joannopoulos, R.D. Meade, J.N. Winn, Photonic Crystals - Molding the flow of light, Princeton
University press, (1995).
[21] S. Noda, A. Chutinan, M. Imada, Nature, 407, 608, (2000).
[22] S. Fan and J.D. Joannopoulos, “Analysis of guided resonances in photonic crystal slabs”, Phys. Rev. B, 65,
235112-1/8, (2002).
[23] E.A. Camargo, H.M.H. Chong,and R.M. De La Rue, Optic. Express, 12, 588, (2004).
[24] T. Baba, M. Shiga, K. Inoshita and F. Koyama, Electronics Letters, 39, 1516, (2003).
[25] A. Sharkawi, S. Shi, D.W. Prather and R.A. Soref, Optics Express, 10, 1048, (2002).
[26] S.W. Leonard, J.P. Mondia, H.M. van Driel, O. Toader, S. John, K. Bush, A. Birner, U. Gosele,
V. Lehmann, Phys. Rev. B, 61, 2389, (2000).
[27] Y.L. Hoo, W. Jin, H.L. Ho, D.N. Wang, “Measurements of gas diffusion coefficient using photonic crystal
fiber”, IEEE Photonics Technology Letters, 15, 1434-1436, (2003).
[28] G. Pickrell, W. Peng, A. Wang, “Random-hole optical fiber evanescent-wave gas sensing”, Optics Letters,
29, 1476-1478, (2004).
[29] T. Ritari, J. Tuominen, H. Ludvigsen, J.C. Petresen, T. Sorensen, T.P. Hansen: “Gas sensing using air-
guiding photonic bandgap fibers”, Optics Express, vol.12, n.17, 4080-4087, (2004).
[30] T.M. Monro, D.J. Richardson, P.J. Bennett: “Developing holey fibres for evanescent field devices”,
Electronics Letters, 35, 1188-1189, (1999).
[31] G. Humbert, J.K. Knight, G. Bouwmans, P.S.J. Russell, D.P. Williams, P.J. Roberts, B.J. Mangan:
“Hollow core photonic crystal fibers for beam delivery”, Optics Express, 12, 1477-1484, (2004).
[32] L. Jin, W. Zhang, H. Zhang, B. Liu, J. Zhao, Q. Tu, G. Kai, X. Dong: “An embedded FBG Sensor for
simultaneous measurement of stress and temperature”, IEEE Photonics Technology Letters, 18, 154-156,
(2006).
[33] S.O .Konorov, A.M. Zheltikov, M. Scalora: “Photonic-crystal fiber as a multifunctional optical sensor and
sample collector”, Optics Express, 13, 3454-3459, (2005).
[34] W.N. MacPherson, E.J. Rigg, J.D.C. Jones, V.V. Ravi Kanth Kumar, J.C. Knight, P.St.J. Russell: “Finite-
element analysis and experimental results for a microstructured fiber with enhanced hydrostatic pressure
sensistivity, IEEE J. of Lightwave Technology, 23, 1277-1231, (2005).
[35] W. Hopman, P. Pottier, D. Yudistra, J. van Lith, P. Lambeck, R. De La Rue, A. Driessen, J.W.M. Hoestra,
R. de Ridder: “Quasi-one-dimensional photonic crystal as compact building block for refractometric
optical sensors”, IEEE J.of Selected Topics in Quantum Electronics, 11, 11-16, (2005).
[36] M. Sandrock, M. Wiggins, J. Shirk, H. Tai, A. Ranade, E. Baer, H. Hiltner, “A widely tunable refractive
index in a nanolayered photonic material”, Applied Physics Letters, 84, 3621-3623, (2004).
[37] E. Chow, A. Grot, L.W. Mirkarimi, M. Sigalas, G. Girolami, “Ultracompact biochemical sensor built with
two-dimensional photonic crystal microcavity”, Optics Letters, 29, 1093-1095, (2004).
[38] S. Rajic, J.C. Corbeil, P.G. Datskos, “Feasibility of tunable MEMS photonic crystal devices”,
Ultramicroscopy, 97, 473-479, (2003).
[39] S. Baglio, M. Bloemer, N. Savalli, M. Scalora: “Development of novel optoelectromechanical systems
based on “Transparent metals” PBG structures”, IEEE Sensors Journal, 1, 288-295, (2001).
[40] W. Suh, M.F. Yanik, O. Solgaard and S. Fan, “Displacement-sensitive photonic crystal structures based on
guided resonance in photonic crystal slabs”, Appl.Phys.Lett., 82, 1999-2001 (2003).
[41] C. Kee, S. Han,K.B. Yoon, C. Choi, H.K. Sung, S.S. Oh, H.Y. Park, S. Park, H.S chift, Appl.Phys.Lett. 86,
051101 (2005).
[42] C. Liguda et al., Appl. Phys.Lett. 78, 17 (2001), 2434.
[43] J. Li, et al., Chemical Physics Letters, 390,285–289, (2004).
[44] D. Pisignano, G. Gigli, P. Visconti, T. Stomeo, M. De Vittorio, G. Barbarella, L. Favaretto and
R. Cingolani, Nanotechnology 15, 766–770, (2004).
[45] M. Qiu: “Effective index method for heterostructure-slab-waveguide-based-two-dimensional photonic
crystals, Appl.Phys. Lett., 81, 1163-1165 (2002).

You might also like