Open Access Paper
12 July 2019 Acousto-optic interaction model with mercury halides (Hg2Cl2 and Hg2Br2) as AOTF cristals
A. Pierson, C. Philippe
Author Affiliations +
Proceedings Volume 11180, International Conference on Space Optics — ICSO 2018; 1118064 (2019) https://doi.org/10.1117/12.2536139
Event: International Conference on Space Optics - ICSO 2018, 2018, Chania, Greece
Abstract
Acousto-optic tunable filters (AOTF) are used in the development of hyperspectral imagers from the ultra violet (UV) to the long wave infrared (LWIR). They have the advantage to be an all-solid-state and robust device with no-moving parts that can fast tune their filtered frequency carrier. Such a device is developed by bonding a piezoelectric transducer on a specially cut birefringent crystal. In the LWIR there is a global investigation on efficient solutions for AOTF in future spectral filtering applications. Crystal of mercurous chloride (Hg2Cl2) and mercurous bromide (Hg2Br2) are candidates that demonstrate an advantage in this spectral domain, thanks to a broadband transparency and a fairly large value of their figure of merit M2. Industrial development of those crystals has recently started and presents an excellent opportunity to prospect feasibility of such AOTF development. The purpose of our work is to develop a numerical tool to have a better understanding of the acousto-optic interaction inside mercurous halides crystals. Unfortunately, all the characteristics of those mercurous halides based crystals are not well known nowadays. One important component of the photoelastic tensor is always missing. It implies that the unique interaction configuration that optimizes the diffraction efficiency for a wavelength range can’t be computed analytically with the usual techniques. To overcome this difficulty, we propose a numerical tool based on an acousto-optic interaction and an acoustic wave propagation models. We will also include the AOTF considerations and the theoretical background needed to carry out the filter design.

1.

INTRODUCTION

An Acousto-optic tunable filter (AOTF) is a device with no-moving parts that can provide an electronically tuned band pass filter by the application of a radiofrequency (RF) signal. The main emphasis is on the development of all solid state compact field-portable imagers from ultraviolet (UV) to long wave infrared (LWIR). Such a device is developed in a birefringent crystal by bonding a piezoelectric transducer to a specially cut prism1-5. Acousto-optic (AO) interactions diffract a narrowband part of a white broadband light incident to the input facet of a birefringent crystal that is transparent in the optical wavelength region of interest. This transmitted wavelength band can be tuned by varying the applied RF. Compared to other traditional dispersive optical elements like grating or prism, AOTF devices have slow response time that allows having a high-tuning speed optical bandpass filter.

A spectrometer or hyperspectral imager based on an AOTF in front of a two-dimensional focal plan array (2D-FPA) captures a hyperspectral image cube by scanning the wavelengths in a time sequence to cover the spectral range of operations. This image cube is recorded in a much simpler manner without any relative motion of the scene and the imager compared to a dispersive system such as a grating or a prism. Scenes of interest from hyperspectral activities are in the three atmospheric window regions, visible up to short wave infrared (VIS+NIR+SWIR : from 400 nm to 2.5 μm), medium wave infrared (MWIR : from 3 to 5 μm) and the long wave infrared (LWIR : from 8 to 14 μm). The first two regions are used for enhanced vision for various applications. The latter region is preferred to obtain spatial and spectral signatures from a rich number of chemical, biological species and all objects around 300K.

A number of spectral imagers using AOTF operating from the UV to the LWIR using KDP6-8, TeO210-15, MgF29 and Tl3AsSe316-20 crystals to cover different spectral regions have been developed. Commonly, the Tl3AsSe3 crystals also named as TAS are used as AO materials for the LWIR region. Nevertheless, there is a lack of high quality AO materials and global research is going on in the test of new materials. We bring our interest on a recently synthesized class of crystals for having superior properties in AO devices. The mercurous halide crystals21-26, such as Hg2Cl2 (i.e. Calomel), Hg2Br2 and Hg2I2 are highly anisotropic with a high AO figure of merit thanks to a high photoelastic constant and slow acoustic velocity. Their wide spectral range of transparency are 0.35 – 20 μm, 0.4 – 30 μm and 0.45 – 40 μm respectively. Some prototype devices have been fabricated with single crystal of these materials recently grown27-28.

2.

AOTF CRYSTAL CONSIDERATION

AOTF crystal is an electronically controlled agile device that filters incident white light into narrow spectral band of light at a specific wavelength determined by RF signal. Wavelength of the filtered spectral band can be tuned without any physical movement of the filter by changing the frequency of the applied RF signal, making it a no-moving-parts robust and fast tunable filter. Principles of that filter operation are based on the well-known phenomenon of anisotropic diffraction of light by acoustic wave propagation in a birefringent crystal29-32. As shown in the figure 1, an AOTF crystal is fabricated by bonding a piezoelectric transducer to a specially cut birefringent high quality single crystal. When an RF signal is applied to the transducer, it produces an ultrasonic wave that travels through the crystal with acoustic frequency equal to the RF signal. This establishes a diffraction grating in the crystal with a velocity specific to the material and a grating period equal to the acoustic wave’s period. An acoustic absorber absorbs the sound wave after its transmission into the crystal. The spectral bandwidth of the diffracted light depends upon diffracted wavelength, birefringence of the material, length of the transducer, acoustic cut angle and geometry of the wide-angle AO diffraction.

Figure 1

AOTF crystal operation principles based on a phenomenon of anisotropic diffraction of light by acoustic wave propagation in a birefringent crystal

00248_PSISDG11180_1118064_page_3_1.jpg

So far, TeO2 is the best AO material to design imaging filters due to its large M2 value equal to 1200×10-18 s3/g and its difference between the extraordinary and ordinary refractive indices relatively high. For TeO2, no = 2.26 and ne = 2.41, giving a birefringence Δn = 0.15. Due to its large M2 value, filters designed in TeO2 operate with high efficiency at relatively low applied power. Unfortunately, it cannot be used for AOTF devices operating above 5 μm. The three mercurous halide crystals33 have a tetragonal crystalline structure class with the point group D(4h) and the space group 4/mmm. The mercurous halides have the same uniaxial crystal structure as TeO2 and design considerations can draw upon the TeO2 AOTF designs which have been discussed extensively in the literature10-15.

As seen in Table 1, those materials have really high birefringence and have anomalously slow shear wave acoustic velocities of propagation along the [110] plane24. The lowest velocities are obtained for shear propagation along the [110] plane and are respectively : 347 m/s for Hg2Cl2, 273 m/s for Hg2Br2, and 254 m/s for Hg2I2. The previously lowest recorded velocities (under normal conditions) were 616 m/s for Te02, 520 m/s for Tl3PSe4 and 600 m/s for Tl3AsS4. Shear velocity is significantly anisotropic, which leads to a very large walk-off angle between acoustic wave propagation and energy flow. Birefringence of those materials is high with 0.66 for Hg2Cl2, 0.86 for Hg2Br2 and 1.48 for Hg2I2. Those materials are optically positive uniaxial crystals, i.e. ne is larger than no. Large birefringence and slow velocity properties are attributed to the strong anisotropy of the lattice field and the nature of the heavy Hg2+ ion. In this structure, parallel chains of linear X-Hg-Hg-X molecules are aligned along the direction of the crystallographic C axis. Bond between adjacent molecules is a Van der Waals type, while intramolecular bonding is mainly covalent. This structural configuration generates strong photoelastic, elastic, and optical anisotropies in those crystals. Moreover, components of the photoelastic tensor of the calomel1 are also quite large, p11 = 0.551, p12 = 0.44, and p31= 0.137, as compared to p11 = -0.0074, p12 = 0.187 and p31 = 0.0905 in TeO2. Although various research works had measured and computed mercurous halide crystals main parameters34-40, one component stay even nowadays unknown, the p44 component of the photoelastic tensor. Even by using well-known design methods of TeO2 AO cells, this missing component value doesn’t allow converge numerical computations toward a single optimum solution of diffraction efficiency. Therefore, it is reasonable to conclude that mercurous chloride may be considered as a promising material for a variety of applications in AO devices.

Table 1

table of key parameters of mercurous halide crystals compared to tellurium dioxide.

MaterialTransparency window (μm)Refractive indicesAcoustic velocity (x105 cm/s)Density (g/cm3)Figure of merit (x10-18 s3/g)
TeO20.35 – 4.5no=2.26; ne=2.410.626.01200
Hg2Cl20.35 – 20no=1.96; ne=2.620.357.181050
Hg2Br20.40 – 30no=2.21; ne=2.980.277.313900
Hg2I20.45 – 40no=2.43; ne=3.910.257.704800

From Table 1 it is clear that mercurous halide crystals have a large transparency region and really high figure of merit. We could fabricate AOTFs operating from the visible to the LWIR using anyone of these halides as Hg2Cl2 by suitably bonding multiple transducers to cover a large spectral region. Although Hg2I2 and Hg2Br2 crystals have really high qualities for AO devices, some development research works are still ongoing to be able to grow longer robust crystals26, 41. At the present time research is ongoing in finding the most suitable bonding media with mercury based crystals which will be stable over a long time42, 43.

3.

NUMERICAL MODEL AND RESULTS

3.1

Refractive indexes

Dispersion formulas of refractive indexes can be written in function of the wavelength filtered. For the calomel crystal, this change of refractive indexes has been reported by a European consortium of MINERVA project44 using a polynomial fitting from experimental data:

00248_PSISDG11180_1118064_page_4_1.jpg
00248_PSISDG11180_1118064_page_4_2.jpg

With the photon energy Eph [eV] defined by:

00248_PSISDG11180_1118064_page_4_3.jpg

Where h is the Planck’s constant, c is the speed of light and λ is the filtered wavelength in the vacuum.

3.2

Acoustic Velocity along tZ plane

The t[110]-Z[001] plane is the preferred plane for the slow shear AO interaction45 and it is selected by rotating the references plan about the Z axis of π/4 from [100] direction. The rotation of the reference axis is applied to obtain the new elastic stiffness constant matrix for this rotation:

00248_PSISDG11180_1118064_page_5_1.jpg

Where:

00248_PSISDG11180_1118064_page_5_2.jpg
00248_PSISDG11180_1118064_page_5_3.jpg
00248_PSISDG11180_1118064_page_5_4.jpg

The velocity of acoustic wave for a given direction of θa, which is defined from [110] direction, for the slow shear mode can be defined as :

00248_PSISDG11180_1118064_page_5_5.jpg

The Acousto-optic interaction plane considered is the t-Z plane. In the literature, the elastic stiffness constant values28 are c11 = 1.8925 x1010 N/m², c12 = 1.7192 x1010 N/m², c13 = 1.563 x1010 N/m², c33 = 8.037 x1010 N/m², c44 = 0.8456 x1010 N/m² and c66 = 1.225 x1010 N/m². Velocities of the slow shear acoustic waves in each direction are equal to Vt = 384 m/s and Vz = 1084 m/s, respectively.

3.3

Phase matching condition

Efficient Acousto-optic interaction is obtained when the phase matching condition between acoustic waves and incident electromagnetic radiation is satisfied and described by equations:

00248_PSISDG11180_1118064_page_5_6.jpg

Where:

00248_PSISDG11180_1118064_page_5_7.jpg
00248_PSISDG11180_1118064_page_5_8.jpg

With nd,i the refractive indexes of incident and deflected beams, V(θa) the acoustic velocity and f the RF applied to the ultrasonic transducer. The phase matching condition described above is graphically represented in the K-space diagram below, where ki is the incident light, kd is the first diffracted order, ka is the acoustic wave momenta.

Figure 2

K-space diagram of an AO interaction respecting the phase matching condition between incident and diffracted electromagnetic waves and acoustic wave.

00248_PSISDG11180_1118064_page_5_9.jpg

From Eq. 9, it is possible to define the tuning relation, which is the relation between RF frequency applied to the transducer and the wavelength of the diffracted order for broadband incident light. This relation is defined by:

00248_PSISDG11180_1118064_page_6_1.jpg

An infinite number of configurations of the AO interaction exist, but in practice some constraints are applied in order to optimize one of the parameters in real devices such as: resolution, field of view, RF drive power versus diffraction efficiency, etc. The parallel tangent matching condition29 is defined as:

00248_PSISDG11180_1118064_page_6_2.jpg

Where tangents refractive index surfaces are parallel for the incoming and diffracted light, thus a constraint is set between θi and θd. It’s possible to determine input direction for an arbitrary propagation of the phase velocity of the acoustic wave. Phase matching condition is derived from a given θa and the input angle of electromagnetic radiation:

00248_PSISDG11180_1118064_page_6_3.jpg

From the solution of the above equations, it’s possible to determine θi and consequently θd. The phase matching condition could also be solved by defining θi to determine the range of θa where the parallel tangent matching condition is satisfied for a given wavelength.

3.4

AO interaction’s efficiency

Diffraction efficiency of the AO interaction is proportional to the RF power applied to the ultrasonic transducer, when the phase matching condition is satisfied, and theoretically it can be estimated45 by :

00248_PSISDG11180_1118064_page_6_4.jpg

Where L/H is the dimensional ratio of the transducer, P is the RF Power, λ is the wavelength of diffracted order and M2 is the Acousto-optic figure of merit defined as:

00248_PSISDG11180_1118064_page_6_5.jpg

With peff the effective photoelastic constant, V(θa) the phase velocity of acoustic wave, ρ the density and ni,d the refractive indexes of incident and diffracted beams respectively. Effective photoelastic constant is obtained from tonsorial equation; therefore it is related to the geometry of the AO interaction. Calomel photoelastic constant matrix46-47 is determined by the tetragonal crystal structure as:

00248_PSISDG11180_1118064_page_6_6.jpg

In the literature28, the photoelastic constant values are p11 = 0.551, p12 = 0.44, p13 = 0.137, p33 = 0.01, p44 = unknown and p66 = 0.047. Effective photoelastic constant in the case of an interaction occurring in the plane tZ can be estimated by equation:

00248_PSISDG11180_1118064_page_6_7.jpg

This general expression of peff for tetragonal crystals is used to estimate the Acousto-optic figure of merit for a given AO interaction.

3.5

Model principles and results

AO interaction model objectives are in a first hand to be able to make an evaluation of configurations that allow highest value of diffraction efficiency. In a second hand, to be able to make the criticality control points of the change of several main parameters. Those optima configurations are found for a certain crystal material. Calomel crystal (Hg2Cl2) has a serious flaw, its photo-elastic tensor is not complete and the p44 component is unknown from the literature. Our model takes account of this flaw. Overview schema below presents main nodes of calculation with main decisions criteria to eliminate unsatisfactory configurations. From the material’s parameters, it evaluate diffraction efficiency for each AO configuration (θi, θd, θa, L/H, PRF, fRF), sweeping the p44 value. Thus is made for an electromagnetic wave with a wavelength from 0.4 to 20 μm. To achieve this computation, the model sweeps value of incident angle of broadband light, validate for each the RF frequency range to access all wavelengths. Then it starts the diffraction efficiency calculation in function of interaction dimension ratio between the two waves and the acoustic power used. That research tries to point out impact of several parameters of AO configuration design on diffraction efficiency. In particular, understand the impact of the unknown value of p44, the impact of the acoustic power signal used and the interaction size ratio between the two waves.

Figure 3

principle schema of numerical model, in square shapes the calculation nodes and in diamond shapes the feedback loops.

00248_PSISDG11180_1118064_page_7_1.jpg

In figure 4, we first compare the AO diffraction efficiency in function of output diffracted wavelength sweeping the p44 value from -0.5 to 0.5. Repeating those measurements for several incident light angle values shows that we start to be really affected by the unknown p44 value with incident angle of broadband light superior to 5 degrees with a Δη ~ 10% for wavelength superior at 12 μm. This sensitivity to p44 value increases drastically with the incident angle. And Δη reach 50% when θi = 10° and 100% when θi = 20°. With those results, we clearly see that the unknown value of p44 have a serious impact on broadband performance of AOTF crystal of calomel when the incident angle is high.

Figure 4

Impact of p44 value on diffraction efficiency for several angles of incident broadband light.

00248_PSISDG11180_1118064_page_8_1.jpg

For next studies, we apply the atmospheric transmission in the model to be able to see and evaluate impact that would have the interaction size ratio and acoustic power on the atmospheric windows. The atmospheric transmission factor dataset was used from the planetary spectrum generator of NASA website48.

In figure 5, we compare the AO diffraction efficiency in function of output diffracted wavelength for several interaction size ratio values. From the equations, increase of L/H ratio will be reciprocal to an increase of diffraction efficiency. This is true for high values of wavelength beyond 5 μm. Unfortunately, increase of the size ratio will also cause a translation movement in the diffraction efficiency peaks that are important in low wavelength below 5μm. In particular, the important decrease of diffraction efficiency visible in Figure 4 at 3.8 μm that will move from 3 to 4 μm for a respective variation of L/H from 10 to 30. We see this impact in this region on the transmission windows that decrease instead of increase with L/H value.

Figure 5

Impact of size interaction ratio on AO diffraction efficiency

00248_PSISDG11180_1118064_page_8_2.jpg

Represented in figure 6, we compare the AO diffraction efficiency in function of output diffracted wavelength for several acoustic powers. As expected, increase of Acoustic power will also increase the AO diffraction efficiency for every wavelength, the previous effect of peaks shifting doesn’t occur with this parameter.

Figure 6

Impact of acoustic power on AO diffraction efficiency

00248_PSISDG11180_1118064_page_9_1.jpg

4.

DISCUSION

An AOTF is a device with no-moving parts that can provide an electronically tuned bandpass filter by the application of a RF signal. AOTF are used in the development of hyperspectral imagers from the UV to the LWIR. In the LWIR there is a global investigation on efficient solutions for AOTF. Crystals of mercurous halides are candidates that demonstrate an advantage in this spectral domain, thanks to a broadband transparency and a fairly large value of their figure of merit. In our works, we develop a numerical tool to compute the diffraction efficiency of AO configurations in those crystals, in particular in the calomel crystal (Hg2Cl2). Mercurous halide crystals have the flaw that their photoelastic tensor are not complete, the component p44 is unknown from the literature. Our research reports that diffraction efficiency is less sensitive to this p44 value in a AO configuration with low value of incident light angle. We also report that the choice of increasing interaction size ratio doesn’t mean a constant increase of diffraction efficiency in certain wavelengths. In particular in the region below 5μm where a shift of diffraction efficiency peaks exist. Industrial development of those crystals has recently started and presents an excellent opportunity to prospect feasibility of such AOTF. Depending of the use case for AOTF crystals and the spectral region of interest, it will be a compromise between acoustic power and dimension size ratio increase to find an optimal AO configuration.

REFERENCES

[1] 

I. C. Chang, “Tunable acousto-optic filters: an overview,” in Proc. SPIE, 12 –22 (1976). Google Scholar

[2] 

J. Xu and R. Stroud, “Acousto-Optic Devices,” Wiley, New York, N.Y. (1992). Google Scholar

[3] 

A. Goutzoulis and D. Pape, “Designing and Fabrication of Acousto-Optic Devices,” Marcel Dekker, New York, N.Y. (1994). Google Scholar

[4] 

N. J. Berg and J. M. Pellegrino, “Acousto-Optic Signal Processing: Theory and Implementation,” Marcel Dekker, New York, N.Y. (1996). Google Scholar

[5] 

V. Voloshinov, A. Tchernyatin, E. Blomme and O. Leroy, “A dozen of Bragg effects in tellurium dioxide single crystal,” in Proc. SPIE, 141 –152 (1998). Google Scholar

[6] 

N. Gupta, and V. Voloshinov, “Deep UV spectral imaging characterization of an improved KDP acousto-optic tunable filter,” J. Opt., 16 035301 (2014). https://doi.org/10.1088/2040-8978/16/3/035301 Google Scholar

[7] 

V. B. Voloshinov and N. Gupta, “Ultraviolet/Visible Imaging Acousto-Optic Tunable Filters in KDP,” Appl. Opt., 43 3901 –3909 (2004). https://doi.org/10.1364/AO.43.003901 Google Scholar

[8] 

V. B. Voloshinov and N. Gupta, “Investigation of magnesium fluoride crystals for imaging acousto-optic tunable filter applications,” Appl. Opt., 45 3127 –3135 (2006). https://doi.org/10.1364/AO.45.003127 Google Scholar

[9] 

N. Gupta, “Acousto-optic tunable filters for Infrared Imaging,” in Proc SPIE, 1 –10 (2005). Google Scholar

[10] 

N. Gupta, R. Dahmani, and S. Choy, “Acousto-optic tunable filter based visible-to-near-infrared spectropolarimetric imager,” Opt. Eng., 41 1033 –1038 (2002). https://doi.org/10.1117/1.1467936 Google Scholar

[11] 

N. Gupta and D. R. Suhre, “AOTF imaging spectrometer with full Stokes polarimetric capability,” Appl. Opt., 46 2632 –2037 (2007). https://doi.org/10.1364/AO.46.002632 Google Scholar

[12] 

V. B. Voloshinov and N. Gupta, “Acousto-optic imaging in the mid-infrared region of the spectrum,” in Proc. SPIE, 62 –73 (1999). Google Scholar

[13] 

N. Gupta and V. B. Voloshinov, “Development and Characterization of Two-Transducer Imaging Acousto-Optic Tunable Filters with Extended Tuning Range,” Appl. Opt., 46 1081 –1088 (2007). https://doi.org/10.1364/AO.46.001081 Google Scholar

[14] 

Javier Calpe-Maravilla, Javier Calpe-Maravilla, Joan Vila-Frances, “400- to 1000-nm imaging spectrometer based on acousto-optic tunable filters,” in Proc. SPIE 5570, Sensors, Systems, and Next-Generation Satellites VIII, (2004). https://doi.org/10.1117/12.565587 Google Scholar

[15] 

Rula Tawalbeh; David G. Voelz; David A. Glenar, “Infrared acousto-optic tunable filter point spectrometer for detection of organics on mineral surfaces,” Optical Engineering, 52 (6), 063604 (2013). https://doi.org/10.1117/1.OE.52.6.063604 Google Scholar

[16] 

D. R. Suhre, N. Gupta and M. Gottlieb, “LWIR spectral imager with an 8-cm-1 passband acousto-optic tunable filter,” Optical Engineering, 44 (9), 094601 (2005). https://doi.org/10.1117/1.2048755 Google Scholar

[17] 

N. B. Singh, D. Suhre, N. Gupta, W. Rosch, and M. Gottlieb, “Performance of TAS crystal for AOTF Imaging,” Jour. Crystal Growth, 225 124 –128 (2001). https://doi.org/10.1016/S0022-0248(01)00833-8 Google Scholar

[18] 

J. D. Feichtner, M. Gottlieb, and J. J. Conroy, “Tl3AsSe3 noncollinear acousto-optic filter operation at 10 μm,” Appl. Phys.Lett., 34 1 –3 (1979). https://doi.org/10.1063/1.90583 Google Scholar

[19] 

D. Suhre and E. Villa, “Imaging spectroradiometer for the 8–12 μm region with 3cm–1 passband acousto-optic tunable filter,” Appl. Opt., 37 (12), 2340 –2345 (1998). https://doi.org/10.1364/AO.37.002340 Google Scholar

[20] 

N. B. Singh, D. Kahler, D. J. Knuteson, M. Gottlieb, D. Suhre, A. Berghmans, B. Wagner, J. Hedrick, T. Karr, and J. J. Hawkins, “Operational characteristics of a longwavelength IR multispectral imager based on an acousto-optic tunable filter,” Opt. Eng., 47 013201 (2008). https://doi.org/10.1117/1.2829768 Google Scholar

[21] 

M. Silvestrova, C. Barta, G. F. Dobrzhanskii, L. M. Belyaev, and Y. V. Pisarevskii, “Acousto-optical properties of calomel crystals Hg2C12,” Soy. Phys. Crystallogr., 20 1062 –1069 (1975). Google Scholar

[22] 

N. B. Singh, R. H. Hopkins, R. Mazeisky, M. Gottlieb, and E. Metz, “Growth of large mercurous chloride single crystals by sublimation,” J. Cryst. Growth, 74 667 –669 (1986). https://doi.org/10.1016/0022-0248(86)90215-0 Google Scholar

[23] 

M. Gottlieb, A. P. Goutzoulis, and N. B. Singh, “Mercurous chloride (Hg2Cl2) acousto-optic devices,” in Proc. IEEE Ultrasonics. Sym., 423 –427 (1986). Google Scholar

[24] 

M. Gottlieb, A. Goutzoulis, and N. Singh, “High-performance acousto-optic materials: Hg2Cl2 and PbBr2,” Opt. Eng., 31 2110 –2117 (1992). https://doi.org/10.1117/12.58879 Google Scholar

[25] 

N. B. Singh and R. Mazelsky, “Development of mercurous halides for acousto-optic devices,” Curr. Top. Cryst. Growth Res., 2 435 –444 (1995). Google Scholar

[26] 

J. Kim, S. B. Trivedi, J. Soos, N. Gupta, and W. Palosz, “Development of mercurous halide crystals for acoustooptic devices,” in Proc. SPIE, 66610B (2007). Google Scholar

[27] 

D. J. Knuteson, N. B. Singh, M. Gottlieb, D. Suhre and N. Gupta, “Crystal Growth, Fabrication and Design of Mercurous Bromide Acousto-Optic Tunable Filters,” Opt. Eng., 46 (6), 064001 –064006 (2007). https://doi.org/10.1117/1.2744369 Google Scholar

[28] 

J. Kim, S. Trivedi, B., J. Soos, W. Palosz, and N. Gupta, “Development of Mercurous Halide Crystals for Acousto-Optic Devices,” in Proc. SPIE, 1 –12 (2007). Google Scholar

[29] 

Chang, I. C., “Tunable acousto-optic filters: an overview,” in Proc. SPIE, 12 –22 (1977). Google Scholar

[30] 

Xu, J. and Stroud, R., “Acousto-Optic Devices,” Wiley, New York (1992). Google Scholar

[31] 

Goutzoulis, A. and Pape, D., “Design and Fabrication of Acousto-Optic Devices,” Marcel Dekker Inc., New York, NY (1994). Google Scholar

[32] 

Gupta, N., “Chapter 32 : Acousto-Optics,” Optical Engineer’s Desk Reference, Optical Society of America, Washington D.C. (2003). Google Scholar

[33] 

A. P. Goutzoulis and M. Gottlieb, “Characteristics and design of mercurous halide Bragg cells for optical signal processing,” Opt. Eng., 27 (2), 157 –163 (1988). https://doi.org/10.1117/12.7976661 Google Scholar

[34] 

C. Barta, “Preparation of mercurous chloride monocrystals,” Krist. Tech., 5 541 –549 (1970). https://doi.org/10.1002/(ISSN)1521-4079 Google Scholar

[35] 

I. M. Sylvestrova, C. Barta, G. F. Dobrzhansbii, L. M. Belyaev, and Y. V. Pisarevskii, “Elastic properties of Hg2C12 crystals,” Soy. Phys. Crystallogr., 20 221 –224 (1975). Google Scholar

[36] 

I. M. Silvestrova, C. Barta, G. F. Dobrzhansbii, L. M. Belyaev, and Y. V. Pisarevskii, “Acousto-optical properties of calomel crystals Hg2Cl2,” Soy. Ploys. Crystallogr., 20 649 –65 (1975). Google Scholar

[37] 

G. Barta and J. Trnka, “New types of polarizers made from crystals of the calomel group,” Cryst. Res. Technol., 17 431 –438 (1982). https://doi.org/10.1002/(ISSN)1521-4079 Google Scholar

[38] 

M. Gottlieb, A. P. Goutzoulis, and N. B. Singh, “Fabrication and characterization of mercurous chloride acousto-optic devices,” Appl. Opt., 26 4681 –4687 (1987). https://doi.org/10.1364/AO.26.004681 Google Scholar

[39] 

C. Barta and C. Barta, Jr., “Physical properties of single crystals of the calomel group,” in Growth and Characterization of Acousto-Optical Materials, Materials Science Forum, 61 Trans Tech Publications, Zurich (1990). Google Scholar

[40] 

G. B. Brandt, N. B. Singh, and M. Gottlieb, “Mercurous halides for long time delay Bragg cells,” in Proc. SPIE, 336 –343 (1991). Google Scholar

[41] 

J. Kim, S. B. Trivedi, J. Soos, N. Gupta, and W. Palosz, “Growth of Hg2Cl2 and Hg2Br2 single crystals by physical vapor transport,” J. Cryst. Growth, 310 2457 –2463 (2008). https://doi.org/10.1016/j.jcrysgro.2007.12.067 Google Scholar

[42] 

Valle, Stefano, “Design and application of high performance Acousto-Optic Tunable Filters,” EngD thesis, University of Glasgow, (2017). Google Scholar

[43] 

Neelam Gupta, “Materials for imaging acousto-optic tunable filters,” in Proc. SPIE, 91000C (2014). Google Scholar

[45] 

Xu J, Stroud R., “Acousto-optic devices: principles, design, and applications,” Wiley series in pure and applied optics., Wiley;1992). Google Scholar

[46] 

J. Xu and R. Stroud, “Acousto-Optic Devices,” Wiley,1992). Google Scholar

[47] 

J. F. Nye, “Physical Properties of Crystals,” Masson,1974). Google Scholar
© (2019) COPYRIGHT Society of Photo-Optical Instrumentation Engineers (SPIE). Downloading of the abstract is permitted for personal use only.
A. Pierson and C. Philippe "Acousto-optic interaction model with mercury halides (Hg2Cl2 and Hg2Br2) as AOTF cristals", Proc. SPIE 11180, International Conference on Space Optics — ICSO 2018, 1118064 (12 July 2019); https://doi.org/10.1117/12.2536139
Lens.org Logo
CITATIONS
Cited by 4 scholarly publications.
Advertisement
Advertisement
RIGHTS & PERMISSIONS
Get copyright permission  Get copyright permission on Copyright Marketplace
KEYWORDS
Crystals

Acousto-optics

Acoustics

Diffraction

Mercury

Long wavelength infrared

Photoelasticity

Back to Top