Concept of a miniature optical spectrometer using integrated optical and micro-optical components Ivan Avrutsky, Kalyani Chaganti, Ildar Salakhutdinov, and Gregory Auner

We describe the concept of a super compact diffractive imaging spectrometer, with optical components a few millimeters across in all dimensions, capable of detecting optical fluorescence spectra within the entire visible spectral range from 400 nm to 700 nm with resolution of the order of 2 nm. In addition, the proposed spectrometer is capable of working simultaneously with multiple, up to 35, independent input optical channels. A specially designed diffractive optical element integrated with a planar optical waveguide is the key component of the proposed device. In the preliminary experimental tests, a uniform waveguide grating with a microlens was used to mimic operation of the diffractive optical element. A microspectrometer with optical components measured below 1 cm in all dimensions covers the spectral range from 450 nm to 650 nm and shows a spectral resolution of 0.5 nm at wavelengths close to 514 nm and 633 nm. © 2006 Optical Society of America OCIS codes: 050.1950, 050.1970, 130.3120, 350.3950.

1. Introduction

Miniature optical spectrometers are required for onchip systems used in general chemical analysis, DNA sequencing, detection of hazardous substances, and other applications in biology and chemistry.1 Also, optical spectroscopy is essential for development of noninvasive methods in medical examination. Fluorescence2,3 time-resolved fluorescence4,5 and Raman6 spectroscopy are promising techniques for identification and characterization of malignant tissues. For such applications the spectrometers have to be really small to enable convenient handheld cordless diagnostic instruments. Currently available microspectrometers are either too bulky for the on-chip integration or do not provide necessary spectral resolution. Since the early days of optical spectroscopy, terminology such as “x-meter spectrometer,” the longer the better, was supposed to indicate the high quality of the instrument. When using a prism or a long-period grating as a dispersive element, a longer focal distance is indeed preferable for achieving larger linear

I. Avrutsky ([email protected]), K. Chaganti, I. Salakhutdinov, and G. Auner are with the Department of Electrical and Computer Engineering, Wayne State University, Detroit, Michigan 48202. Received 27 February 2006; revised 13 June 2006; accepted 15 June 2006; posted 22 June 2006 (Doc. ID 68450). 0003-6935/06/307811-07$15.00/0 © 2006 Optical Society of America

dispersion at the output slit and thus for providing better resolution. These days, spectroscopic diffraction gratings of any desirable period are available, and photodetectors behind the output slit are being replaced with image sensors that typically have pixel size below 10 ␮m. The only real reason to keep a larger distance between the dispersive element and the photodetector array is associated with the light scattering by the grating imperfections7: with the focal distance increased, the stray-light intensity at the detector array is reduced. Stray scattering by diffraction gratings has been substantially reduced due to the improved quality of gratings. For example, replicated holographic gratings on microlenses with focal distance of only several millimeters provide spectral resolution of 2.25–3 nm and stray-light suppression of 25–30 dB.8,9 The origin of light scattering by diffractive optical elements has also been studied, and the design rules for their application in optical microsystems have been formulated.10 An industry benchmark for miniature optical spectrometers is set by companies such as Ocean Optics, Inc.,11 StellarNet, Inc.,12 and Spectro-Solutions.13 The most compact Raman detector, to the best of our knowledge, is manufactured by Ahura Corporation.14 Versatile optical spectrometers can be squeezed into a several-inch-large package. This represents the ultimate degree of miniaturization achievable by scaling down a conventional bulk spectrometer. Inspired by the design of optical multiplexers and demultiplexers in the telecommunication industry, 20 October 2006 兾 Vol. 45, No. 30 兾 APPLIED OPTICS

7811

integrated optical spectrometers for general use have been developed. Nevertheless, the concept of these devices is such that the range remains relatively narrow, simultaneous spectral resolution of multiple optical inputs is often impossible, and these devices are still too large. The state-of-the-art in optical microspectrometers is reviewed in a recent paper.15 For microspectrometers with a size of the optical components of about 1 cm, the best reported resolution in the visible range is 5–10 nm for both planar integrated-optical and free-space micro-optical designs. A single-input planar-waveguide-based microspectrometer16 with a footprint of 11 mm ⫻ 11 mm uses a focusing grating manufactured in a slab waveguide and covers a 350–650 nm spectral range with 10 nm resolution. A micro-optical imaging spectrometer based on a reflection grating17 covers a 510–610 nm spectral range with a resolution of 5 nm. This device uses a 10 mm fiber bundle as an input and thus allows for multiple optical inputs. The grating’s dimensions are 11 mm ⫻ 6.5 mm, and the largest dimension of the optical part of the device is at least 21 mm. Very recently, a microspectrometer based on sensing of a standing wave in front of a mirror18 has been proposed. It is very compact, but its resolution is no better than 6 nm. A highresolution 共0.07 nm兲 single-input-channel microspectrometer based on an arrayed waveguide grating is described in Ref. 19. This spectrometer covers a limited spectral range 共14 nm兲 and presumable has a large footprint to accommodate as many as 250 waveguides in the array. 2. Diffractive Imaging Spectrometer

In this paper we propose a concept of the diffractive imaging spectrometer that combines guided-wave optics with micro-optical elements. It is the unique separation of functions between the two-dimensional (2D) integrated-optical part of the device and its three-dimensional (3D) free-space part that allows for the supercompact design. Such a degree of miniaturization is not available with a pure integrated optical or a pure micro-optical design. The 2D part of the microspectrometer is implemented in a planar waveguide. The 3D part is mounted on top of the waveguide slab. The diffractive imaging spectrometer proposed here includes N input waveguide channels, a planar waveguide expansion section, a diffractive optical element, an aberration correction prism, and an image sensor [Fig. 1(a)]. The input waveguides are facing the diffractive optical element [Fig. 1(b)]. The waveguide ends are located along the circle of radius R centered at the diffractive optical element. To optimize the power efficiency, the angular divergence of the radiation coming out of the input channel waveguides is close to the angular size of the diffractive optical element at the distance R. The expansion section is a uniform planar waveguide. Compared to the free-space optics, a waveguide-based design allows for a more compact design due to light confinement within the waveguide. Using high modal index 7812

APPLIED OPTICS 兾 Vol. 45, No. 30 兾 20 October 2006

Fig. 1. (a) Side and (b) top view of the proposed diffractive imaging spectrometer. Particular dimensions are given to illustrate the concrete design described in the text. A smaller or larger package size is possible depending on the desirable spectral range and resolution.

共nm兲 waveguides also results in a denser structure of the diffractive optical element, which further reduces the system dimensions. The key component of the proposed microspectrometer is the diffractive optical element. It combines functions of an input collimating optics, a diffraction grating, and an output focusing optics. Physically it is a set of curved grooves or areas with a modified refractive index in the planar waveguide. The diffractive optical element is designed to provide ideal focusing of the radiation at wavelength ␭0 coming out of the central waveguide channel to a focal spot located at a distance H right above the diffractive optical element. A normal output direction is a logical choice as long as it provides the largest visible angular size of the diffractive optical element and, thus, the best diffraction-limited convergence of the radiation. Space between the waveguide and the focal spot is filled with high index (n) material, which helps to improve diffraction divergence of the focused radiation. In a practical implementation this is a prism placed between the waveguide and the image sensor [shown as a triangle between the waveguide and the image sensor in Fig. 1(a)]. The prism parameters such as angle, index of refraction, and its spectral

dispersion are (to a certain degree) adjustable factors that help to reduce focusing aberrations of the diffractive optical element. The diffractive optical element has small size, which is just enough for providing necessary spectral resolution. The size L of the diffractive optical element can be estimated as follows. The resolving power of an optimally designed grating-based spectrometer is approximately equal to the number of the grating’s grooves. Thus, if at the wavelength ␭0 the required resolution is ␦␭, the diffractive element should contain approximately M ⫽ ␭0兾␦␭ grooves. Then, the average period of the grooves in the diffractive optical element is estimated using the phase synchronism condition in a planar waveguide grating-based coupler: for outcoupling a guided mode normally to the plane of the waveguide, the grating period must be ⌳ ⫽ ␭兾nm. These estimations result in L⬇M

␭02 ␭0 ⫽ . nm ␦␭ · nm

(1)

The exact size of diffractive optical element is a subject of the numerical optimization problem. For a given diffractive optical element, which provides focusing of light with wavelength ␭0 to a focal spot right above it, radiation with other wavelengths will be focused into spots located at some surface above the diffractive optical element. A prism with optimized angle ␪ is placed between the diffractive optical element and the image sensor. The prism also provides mechanical support for the image sensor. The exact prism angle ␪ and the prism location (e.g., offset between the ideal focal spot for ␭0 and the prism facet) are a subject of the numerical optimization problem. The image sensor, such as a CCD or a complementary metal-oxide semiconductor (CMOS) detector matrix, is located at the prism surface. Its pixel size, s, must be small enough to resolve diffraction-limited focal spots produced by monochromatic radiation at the image plane. The diffraction-limited spot size can be estimated as follows. For radiation with wavelength ␭0, which is focused above the center of the diffractive optical element, let us assume that the diffractive optical element forms a converging wave with perfectly spherical phase front. At the focus located at the distance H from the waveguide, the diffraction-limited spot size is estimated at d ⬇ H ⫻ 共␭0⬘兾L兲, where ␭0⬘ ⫽ ␭0兾n is the in-material wavelength. As the sensor is set at an angle ␪, the spot size at the center is enlarged by factor 1兾cos ␪. Finally, using Eq. (1), the spot size is estimated at s⬍

H nm ␦␭ ⫻ ⫻ . cos ␪ n ␭0

(2)

In practical implementation, the pixel size of the image sensor should be at least 2–3 times smaller than the size of the diffraction-limited spot so that the resolution of the microspectrometer will not be compromised by the discrete structure of the sensor.

3. Design Example

This section presents an example of the diffractive imaging spectrometer, which was designed to cover the entire visible range from 400 to 700 nm. The optimal design wavelength ␭0 was found to be 480 nm. Choice of ␭0 is not critical. Diffraction-limited resolution was chosen to be ␦␭ ⫽ 2 nm. We assumed the waveguide’s modal index of nm ⫽ 1.55, and the prism index of n ⫽ 1.50 (glass). The material dispersion was neglected in this simplified design example. According to the above design rules, the size of the diffractive optical element (I) becomes L ⫽ 80 ␮m. The size of input waveguide channels was chosen to be close to the typical single-mode fiber core size, that is, 6.5 ␮m. This results in a waveguide separation (center-tocenter) of 13 ␮m and the radius R ⬇ 1.6 mm. The image of the input waveguide facet at the sensor plane defined by geometrical optics is magnified by a factor H兾R. Thus we took H ⫽ 2 mm in order to achieve the geometrical image size of about 8 ␮m, which is approximately 2–3 times larger than the pitch of a dense CMOS matrix. Further numerical optimization of the image size location gives an offset of about ⫺170 ␮m and an optimal angle ␪ ⫽ 53°. The structure of the diffractive optical element is determined by an interference pattern of two diverging waves. One of them has an origin at the facet of the central input waveguide, and the second one is centered at a distance H above the diffractive optical element. The vacuum wavelength of both waves is ␭0, and the refractive indices are nm and n, respectively. The typical period of the interference pattern is ␭0兾nm ⬇ 320 nm. Then, assuming electron-beam fabrication, we created a digitized binary picture with pixel size of 100 nm ⫻ 100 nm. A central 25 ␮m ⫻ 25 ␮m part of the diffractive optical element is shown in Fig. 2(a). The intensity distribution at the image sensor is then calculated using the Fraunhofer approximation. Figure 2(b) illustrates the intensity distribution produced by the radiation with wavelengths 400, 500, 600, and 700 nm coming to the diffractive optical element from the central input waveguide and two other input waveguides located at ⫾250 ␮m with respect to the central one. The dashed curves stretched primarily from left to right show the focal spot locations for the radiation coming out of a given input waveguide. The curves stretched primarily in the vertical direction show the locations of the focal spots for the radiation with a given wavelength. The tiny ovals at the curve intersections show the focal spots corresponding to a given wavelength and a given input waveguide location. The spots are shown at the half-intensity level. The goal for the numerical optimization procedure was to get the spots as small as possible while keeping them well separated. Based on the spot sizes, spectra from up to 35 input channels can be analyzed by this device. Wavelength resolution is found to be close to ⬃2 nm across the image field. The focal spots are located within a 0.8 mm ⫻ 1.7 mm area of image field. The 20 October 2006 兾 Vol. 45, No. 30 兾 APPLIED OPTICS

7813

Fig. 3. Schematic image of the grooves forming the diffractive optical element (shadowed) and appropriate phase mask (dashed curves).

Fig. 2. (a) Central part 共25 ␮m ⫻ 25 ␮m兲 of the diffractive optical element and (b) simulated intensity distribution at the image sensor. The scale is in micrometers.

optical part of the spectrometer has a rather small footprint. For example, its width is 0.8 mm defined by the required image field width. The expansion section and the projection of the image field on the waveguide plane partly overlap, so the overall length becomes 1.6 mm ⫹ 0.7 mm ⫻ cos共53°兲 ⫽ 2 mm (the expansion section is estimated to be equal to the radius R ⬇ 1.6 mm, and the part of the image field that does not overlap with the expansion section, as follows from Fig. 2 at the right, is extended by approximately 700 ␮m). To the best of our knowledge, the performance characterized by (a) 2 nm spectral resolution over a 400–700 nm spectral range (b) in a package with the size of the optical part below 2 mm in all dimensions and (c) a functionality of an imaging spectrometer that provides simultaneous spectral resolution of up to 35 independent input channels cannot be achieved within any other microspectrometer concept known so far. 4. Manufacturability of the Diffractive Optical Element

There are several technologies that may be used for fabrication of the diffractive optical element required for the proposed microspectrometer. Direct writing by electron-beam lithography and focused ion-beam milling are likely to be prohibitively expensive and thus may only be used at the research stage of device development. 7814

APPLIED OPTICS 兾 Vol. 45, No. 30 兾 20 October 2006

The diffractive element is essentially a chirped grating with curved grooves. Thus it can in principle be recorded holographically using diverging laser beams. Such an approach, however, would require additional analysis in order to understand the level of accuracy achievable with the holographic recording and its effect on the performance of the device in terms of spectral resolution and number of input channels. As often happens for the holographically recorded structures, because the writing wavelength is shorter than the wavelengths at which the device operates, the recorded hologram never reproduces exactly the designed pattern. The holographic approach may be used with limited success for laboratory studies, but it will hardly allow for manufacturing perfectly optimized microspectrometer. A much more practical approach would be to fabricate an appropriate phase mask first, and then perform deep-UV contact lithography through the phase mask. The diffractive optical element resembles a grating. Moreover, the pitch and orientation of the grooves are only slightly and smoothly varying across the diffractive optical element (the pitch variation being within several percent, and the groove bending not exceeding several degrees). It is expected that the phase mask, which essentially has the same structure as the diffractive optical element but with exactly twice the distance between the grooves (Fig. 3), will enable easy reproduction of diffractive optical elements. The mask would have to be manufactured using electron-beam lithography or focused ion-beam milling similar to the phase masks for fiber Bragg gratings. The most promising technology for manufacturing diffractive optical elements would be the nanoimprint lithography. Currently, up to 8-in 共200 mm兲 nanostamps with minimal feature size to 20 nm are available on commercial basis. As the duty cycle of the structure is very close to 50 percent at any location

across the diffractive optical element, the nanoimprinting is expected to work perfectly. 5. Preliminary Results

Remarkably, some attractive features of the proposed concept could be demonstrated without the use of expensive tools such as electron-beam lithography. The primary purpose of this experiment was to verify experimentally that the subnanometer spectral resolution is achievable in a centimeter-scale device. A compact spectrometer 共1 cm兲 with an excellent resolution 共⬃0.5 nm兲 in the visible range has been fabricated and tested. The planar waveguide structure HfO2 共160 nm兲兾SiO2 共75 nm兲 was commercially fabricated at ThinFilm Labs. Pure materials without any intentional doping have been use. Hafnium oxide has a high refractive index, approaching 2.0, and the refractive index of silica is close to 1.45. Laser light butt-coupled to the waveguide revealed a several millimeter long waveguide track, indicating acceptably low losses in the waveguide. A uniform period grating (2 mm ⫻ 4 mm, 370 nm period) was fabricated onsite using a 244 nm (2nd harmonic Ar⫹-ion laser) holography. The grating was transferred to the SiO2 layer by dry-etching in Plasma Term 770 Deep Trench Etcher. A plano-convex lens with focal distance of f ⫽ 1 cm was glued to the waveguide surface. A CCD camera was installed on an xyz-␸ stage to capture light out-coupled by the grating and focused by the lens. The functionality of the entire system as a microspectrometer was verified experimentally. Light from a He–Ne and an Ar⫹-ion laser was coupled through a microscope objective lens. The scheme of the experiment is shown in Fig. 4. To avoid the image sensor saturation, intensity of the laser beams was greatly attenuated by passing light through polarizers with almost orthogonal polarization planes. In such a setup, the multiple optical inputs are, unfortunately, impossible due to astigmatism of the output beam: the divergence point in the plane of the waveguide is at the waveguide’s input facet, while in the plane normal to the waveguide the output beam right above the grating is essentially parallel. As a result, at the image sensor the light out-coupled by the grating is focused into arc-shape spots rather than into sharp diffraction-limited spots as in the case of a perfectly designed diffractive optical element. The CCD chip used in the preliminary experiments was also relatively large 共752 ⫻ 480 pixels, 6 ␮m ⫻ 6 ␮m grid). A Visual C⫹⫹-based code was used to capture and process the images.

Fig. 4. Scheme of the experiment to test the performance of the microspectrometer.

Fig. 5. (a) Intensity distribution along the central row of pixels. The inset shows a snapshot taken by the image sensor. (b) The noise floor due to the stray-light scattering.

The intensity distribution along the central row of the image sensor’s pixels is shown in Fig. 5(a). The arcs are about 1.5–2 pixels wide at the half-maximum level. The inset in the graph shows a snapshot taken by the image sensor. The spectral dispersion at the image sensor is estimated to be 0.26 nm兾pixel, and the resolution was estimated at 0.5 nm at both 514.5 nm and 632.8 nm laser lines. The additional picks close to pixel numbers 300, 400, and in the 600 –700 range were found to be associated with the emission coming from a discharge tube of the Ar⫹-ion laser. Also, measurements done by an independent spectrometer confirmed that the He–Ne laser used in the experiment, together with the 632.8 nm lines, also generates at 640.0 nm. An appropriate pick at pixel number 670 is also seen in Fig. 5(a). Beyond these additional peaks [Fig. 5(b), pixel numbers 1–500 except for the areas around pixel numbers 300 and 400] the stray scattering was about 1兾245 of the pick value corresponding to the green line, which corresponds to the signal-to-noise ratio of ⫺23.9 dB. We also demonstrated applicability of the microspectrometer for measuring a spectrum of a fluorescent dye (Rhodamine-575). A cuvette with the dye solution was places in the conjugated point of the microscope objective, as shown in Fig. 4. In this ge20 October 2006 兾 Vol. 45, No. 30 兾 APPLIED OPTICS

7815

design. Namely, the issues of the light coupling to the waveguide, CCD image acquisition, calibration of the spectrometer (pixel-wavelength correspondence), and integration of multiple frames to improve the sensitivity have been experimentally verified. 6. Conclusions

Fig. 6. Fluorescence of Rhodamine-575 measured using the microspectrometer.

ometry only a tiny fraction of fluorescent light is captured by the objective. However, the spectrum of fluorescence was clearly detected by the CCD (Fig. 6). A stream of images (30 frames per second) was integrated and averaged in real time to improve the sensitivity. The CCD camera was relocated between the experiments so that the peak corresponding to the 514.5 nm line of the Ar⫹-ion laser is at slightly different position compared to the spectrum in Fig. 5. The beam from the He–Ne laser was intentionally blocked. The peak corresponding to the green radiation from the Ar⫹–ion laser was greatly attenuated due to the absorption by the dye. It is noticeable also that the fluorescence spectrum is rather smooth and the gas discharge lines seen in Fig. 5 are effectively blocked by absorption in the dye. The intensity is shown as it was measured by the CCD without any correction for the spectral sensitivity of the camera. The overall uniform background was found to be coming from the dark response of the CCD accumulated from multiple images. Even in complete darkness, without any input either from the laser or from the fluorescing dye, the signal generated by the camera was nonzero. Fortunately, it does not vary too much from pixel to pixel so that it can easily be accounted for simply by shifting down the entire measured spectrum. Figure 6 shows the integrated and averaged signal without correction for the camera’s dark response. By subtracting the dark response integrated and averaged similar to the fluorescence signal, the uniform background virtually disappears. It is worth noting that 0.5 nm spectral resolution for a ⬃1 cm size of the optical scheme is quite an achievement. Nevertheless, a microspectrometer that uses a lens for focusing is a crude device compared to the elegant design based on the diffractive optical element. The main drawback of the lens-based design is the astigmatism of the outcoupled radiation. While it is possible to correct it with cylindrical optics, such a solution is not appealing due to increased complexity and necessity for additional alignment. The experiments with the lens-based design pursued a goal of demonstrating overall functionality of the proposed 7816

APPLIED OPTICS 兾 Vol. 45, No. 30 兾 20 October 2006

In conclusion, we have proposed the concept of a microspectrometer that uses both integrated optical and free-space components. Numerically optimized design shows that the optical part of the device can be smaller than 2 mm in every dimension, while the spectrometer covers the entire visible spectral range with approximately 2 nm spectral resolution and can potentially work with multiple independent input channels. Overall functionality of the proposed scheme is verified using a lens-based prototype. Resolution achieved with a 1 cm focal length lens is about 0.5 nm at wavelengths close to 514.5 nm and 632.8 nm. This project has been supported in part by the Photonics Technology Access Program (PTAP-2004) and Michigan Universities Commercialization Initiative (MUCI-2006). References 1. C. P. Bacon, Y. Mattley, and R. DeFrece, “Miniature spectroscopic instrumentation: application to biology and chemistry,” Rev. Sci. Instrum. 75, 1–16 (2004). 2. U. Gustafsson, S. Palsson, and S. Svanberg, “Compact fiberoptic fluorosensor using a continuous-wave violet diode laser and integrated spectrometer,” Rev. Sci. Instrum. 71, 3004 – 3006 (2000). 3. S. G. Demos, R. Gandour-Edwards, R. Ramsamooj, and R. D. White, “Near-infrared autofluorescence imaging for detection of cancer,” J. Biomed. Opt. 9, 587–592 (2004). 4. C. Klinteberg, M. Andreasson, O. Sandstrom, A. AndreassonEngels, and S. Svanberg, “Compact medical fluorosensor for minimally invasive tissue characterization,” Rev. Sci. Instrum. 76, 034303 (2005). 5. L. Marcu, J. A. Jo, P. V. Butte, W. H. Yong, B. K. Pikul, K. L. Black, and R. C. Thompson, “Fluorescence lifetime spectroscopy of glioblastoma multiforme,” Photochem. Photobiol. 80, 98 –103 (2004). 6. L. Notingher, G. Jell, P. L. Notingher, I. Bisson, O. Tsigkou, J. M. Polak, M. M. Stevens, and L. L. Hench, “Multivariate analysis of Raman spectra for in vitro noninvasive studies of living cells,” J. Molec. Struc. 744, 179 –185 (2005). 7. T. N. Woods, R. T. Wrigley III, G. J. Rottman, and R. E. Harig, “Scattered-light properties of diffraction gratings,” Appl. Opt. 33, 4273– 4385 (1994). 8. S. Traut and H. P. Herzig, “Holographically recorded gratings on microlenses for a miniaturized spectrometer array,” Opt. Eng. 39, 290 –298 (2000). 9. S. Traut, M. Rossi, and H. P. Herzig, “Replicated arrays of hybrid elements for application in a low-cost micro-spectrometer array,” J. Mod. Opt. 47, 2391–2397 (2000). 10. M. Rossi and T. Hessler, “Stray-light effects of diffractive beam-shaping elements in optical microsystems,” Appl. Opt. 38, 3068 –3076 (1999). 11. Descripton of compact optical spectrometers developed by Ocean Optics, Inc., is available online at http://www.oceanoptics.com/ products.asp. 12. Description of compact optical spectrometers developed by StellarNet, Inc., is available online at http://www.stellarnetinc.com/products.htm.

13. Description of compact optical spectrometers developed by Spectro-Solutions is available online at http://www. spectrosolutions.com. 14. Description of compact optical spectrometers developed by Ahura Corporation is available online at http://www.ahuracorp.com. 15. R. F. Wolffenbuttel, “State-of-the-art in integrated optical microspectrometers,” IEEE Trans. Instrum. Meas. 53, 197–202 (2004). 16. D. Sander and Jorg Muller, “Self-focusing phase transmission grating for an integrated optical microspectrometer,” Sensors Actuators A 88, 1–9 (2001).

17. S. Ura, F. Okayama, K. Shiroshita, K. Nishio, T. Sasaki, H. Nishihara, T. Yotsuya, M. Okano, and K. Satoh, “Planar reflection grating lens for compact spectroscopic imaging system,” App. Opt. 42, 175–180 (2003). 18. H. Stiebig, D. Knipp, S. R. Bhalotra, H. L. Kung, and D. A. B. Miller, “Interferometric sensor for spectral imaging,” Sensors Actuators A 120, 110 –114 (2005). 19. P. Cheben, I. Powell, S. Janz, and D. X. Xu, “Wavelengthdispersive device based on a Fourier-transform Michelsontype arrayed waveguide grating,” Opt. Lett. 30, 1824 –1826 (2005).

20 October 2006 兾 Vol. 45, No. 30 兾 APPLIED OPTICS

7817

Concept of a miniature optical spectrometer using ...

and Computer Engineering, Wayne State University, Detroit,. Michigan 48202. Received 27 February 2006; revised 13 June 2006; accepted 15. June 2006 ...

469KB Sizes 2 Downloads 258 Views

Recommend Documents

Concept of a miniature optical spectrometer using ...
design rules for their application in optical microsys- tems have been formulated. .... of the diffractive optical element can be estimated as follows. The resolving ...

A simple miniature optical spectrometer with a planar ...
May 1, 2006 - small as 0.2 cubic cm, the spectral resolution varies from 0.3 nm to 4.6 nm within the ... Applications and miniaturization go hand-in-hand.

A simple miniature optical spectrometer with a planar waveguide ...
May 1, 2006 - and grating couplers for chemical analysis,” Appl. Opt. 29, 4583-4589 (1990). .... After soft-baking for 15 minutes in oven at 90oC, exposure was.

A Pilot Study of a Thermal Display Using a Miniature ...
limited functional role. In other ... the peltier device and heat sink are size-limited in our situ- ation, we .... sify a thermal stimulus as having one of 7 levels of sen-.

Controlling a Semiconductor Optical Amplifier Using a State-Space ...
the controller designed with the linear model is seen to work well even with 100% modulations ... semiconductor optical amplifiers (SOAs), state space methods.

Rapid communication A time-of-flight spectrometer for ... - Springer Link
Feb 2, 2008 - mination in single hit limit. The electron cloud e from the MCP is collected on the signal wire such that two current pulses u1 and u2 propagate.

Highly Sensitive Monitoring of Ru Etching Using Optical ...
Highly Sensitive Monitoring of Ru Etching Using Optical. Emission .... Another application of Ru emission is related to the chamber cleaning process applied ...

Highly Sensitive Monitoring of Ru Etching Using Optical ...
the gate electrode. Metals have low resistivity and high electron ... source of the emission is Ru atoms.5 Knowledge of the emission spectra is necessary to ...

Force Unfolding Kinetics of RNA using Optical ...
introduce a mesoscopic model to analyze the hopping kinetics of RNA hairpins in an optical tweezers setup. ..... analyzing the experimental data that result in different values ...... otherwise distortion effects are big and the force-feedback.

5. Optical Investigation of Brain Networks Using ...
First, we will briefly summarize the basic princi- ples of SLM operation and then provide a detailed description of the different hardware configurations in which SLMs can be integrated. Finally, we will present an overview of results, recently obtai

Bandwidth compression optical processor using ...
The resolution and sampling rate of today's best analog-to-digital converters are limited by ... analog-to-digital converter (ADC), impresses an electronic signal onto a broadband chirped optical field ... segments off line in a digital storage and p

Detection of Runway and Obstacles using Electro-optical and ieee.pdf ...
Jan 23, 2014 - Detection of Runway and Obstacles using Electro-optical and ieee.pdf. Detection of Runway and Obstacles using Electro-optical and ieee.pdf.

Design and development of a portable miniature ECG ...
to get the benefits of such modern methods of healthcare. ..... Terminal the data can be send to PC, curves can displayed in Microsoft Excel, Origin or other graph.

Design and development of a portable miniature ECG monitor
ABSTRACT: A portable battery powered miniature ECG monitor with built in graphic display has been ... Necessary hardware and software were designed and developed in the ... The ADC has two modes of operation, Free Running or.

A CONCEPT FOR A ROLE OF SEROTONIN AND ...
administration of reserpine show no change in bleeding or clotting time.6 It ... nervous system came from its discovery in the brain.6* The observation by.

A CONCEPT FOR A ROLE OF SEROTONIN AND ...
OF NORMAL. AND RESERPINE-TREATED RABBITS*. Drug. 1 Brain serotonin. I. Efiect .... animals were pretreated with i roniaaid (100 mg./kg.) .... Am. J. Physiol.

Video Concept Detection Using Support Vector Machine with ...
Video Concept Detection Using Support Vector Machine with Augmented. Features. Xinxing Xu. Dong Xu. Ivor W. Tsang ... port Vector Machine with Augmented Features (AFSVM) for video concept detection. For each visual ..... International Journal of Comp

Optical Image Stabilizing System using Multirate Fuzzy ...
Computer Science, ASRI Seoul National University, Seoul, Korea (e-mail: ..... integrator calculates an integral value of e, and the differentiator calculates a ...

Real Time Patients Monitoring System using Optical ...
IJRIT International Journal of Research in Information Technology, Volume 2, ... 1PG student, Electronics Department, Shri Sant Gajanan Maharaj College of ...