+ All Categories
Home > Documents > Optical-limiter MEMS dynamic range compression deconvolution

Optical-limiter MEMS dynamic range compression deconvolution

Date post: 03-Oct-2016
Category:
Upload: jed
View: 219 times
Download: 2 times
Share this document with a friend
17
Optical-limiter MEMS dynamic range compression deconvolution Bahareh Haji-saeed, 1, * William D. Goodhue, 2 Charles L. Woods, 1 John Kierstead, 3 and Jed Khoury 1 1 Air Force Research Laboratory, Sensors Directorate, Hanscom Air Force Base, Massachusetts 01731, USA 2 Photonics Center, Physics Department, University of Massachusetts Lowell, Lowell, Massachusetts 01854, USA 3 Solid State Scientific Corporation, Hollis, New Hampshire 03049, USA *Corresponding author: bahareh.haji[email protected] Received 8 October 2008; revised 21 May 2009; accepted 3 June 2009; posted 17 June 2009 (Doc. ID 102425); published 24 June 2009 We propose dynamic range compression deconvolution by a new nonlinear optical-limiter microelectro- mechanical system (NOLMEMS) device. The NOLMEMS uses aperturized, reflected coherent light from optically addressed, parabolically deformable mirrors. The light is collimated by an array of microlenses. The reflected light saturates as a function of optical drive intensity. In this scheme, a joint image of the blurred input information and the blur impulse response is captured and sent to a spatial light modulator (SLM). The joint information on the SLM is read through a laser beam and is Fourier transformed by a lens to the back of the NOLMEMS device. The output from the NOLMEMS is Fourier transformed to produce the restored image. We derived the inputoutput nonlinear transfer function of our NOLMEMS device, which relates the transmitted light from the pinhole to the light intensity incident on the back side of the device, and exhibits saturation. We also analyzed the deconvolution orders for this device, using a nonlinear transform method. Computer simulation of image deconvolution by the NOLMEMS device is also presented. © 2009 Optical Society of America OCIS codes: 230.6120, 070.4340, 190.4360. 1. Introduction The optical limiter is a device whose nonlinear in- putoutput transfer function saturates as the inci- dent light increases [1]. Previously, we introduced a new technique, dynamic range compression decon- volution using two-beam coupling [2], and demon- strated that its inputoutput transfer function saturates as a function of light intensity. Two forms of two-beam coupling deconvolution were introduced: spectrally invariant [2] and spectrally variant [3]. For spectrally invariant two-beam coupling, it was necessary to operate the system at very high beam ratios in order to perform dynamic range compres- sion deconvolution; working in this regime leads to very low grating efficiency. To overcome the high beam ratio problem the spectrally variant two-beam coupling was introduced. Spectrally variant deconvo- lution performed better image restoration than spectrally invariant deconvolution; however, the drawback in spectrally variant deconvolution is that the response time at very high frequencies becomes extremely large (slow processing speed), since the re- sponse time of photorefractive crystals is inversely proportional to the total intensity of light incident on the crystal [46]. Another limitation of both spectrally variant and spectrally invariant two-beam coupling deconvolu- tion is the need for an input spatial light modulator (SLM). This may cause some problems, in particular, for remote sensing applications, in which the light is incoherent. The interface between the SLM and the CCD camera imposes other constraints in systems for correcting supersonic turbulence. One way to overcome this limitation is to use an all-optically 0003-6935/09/193771-17$15.00/0 © 2009 Optical Society of America 1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3771
Transcript
Page 1: Optical-limiter MEMS dynamic range compression deconvolution

Optical-limiter MEMS dynamic rangecompression deconvolution

Bahareh Haji-saeed,1,* William D. Goodhue,2 Charles L. Woods,1

John Kierstead,3 and Jed Khoury1

1Air Force Research Laboratory, Sensors Directorate, Hanscom Air Force Base, Massachusetts 01731, USA2Photonics Center, Physics Department, University of Massachusetts Lowell, Lowell, Massachusetts 01854, USA

3Solid State Scientific Corporation, Hollis, New Hampshire 03049, USA

*Corresponding author: bahareh.haji‑[email protected]

Received 8 October 2008; revised 21 May 2009; accepted 3 June 2009;posted 17 June 2009 (Doc. ID 102425); published 24 June 2009

We propose dynamic range compression deconvolution by a new nonlinear optical-limiter microelectro-mechanical system (NOLMEMS) device. The NOLMEMS uses aperturized, reflected coherent light fromoptically addressed, parabolically deformable mirrors. The light is collimated by an array of microlenses.The reflected light saturates as a function of optical drive intensity. In this scheme, a joint image of theblurred input information and the blur impulse response is captured and sent to a spatial lightmodulator(SLM). The joint information on the SLM is read through a laser beam and is Fourier transformed by alens to the back of the NOLMEMS device. The output from the NOLMEMS is Fourier transformed toproduce the restored image. We derived the input–output nonlinear transfer function of our NOLMEMSdevice, which relates the transmitted light from the pinhole to the light intensity incident on the backside of the device, and exhibits saturation. We also analyzed the deconvolution orders for this device,using a nonlinear transform method. Computer simulation of image deconvolution by the NOLMEMSdevice is also presented. © 2009 Optical Society of America

OCIS codes: 230.6120, 070.4340, 190.4360.

1. Introduction

The optical limiter is a device whose nonlinear in-put–output transfer function saturates as the inci-dent light increases [1]. Previously, we introduceda new technique, dynamic range compression decon-volution using two-beam coupling [2], and demon-strated that its input–output transfer functionsaturates as a function of light intensity. Two formsof two-beam coupling deconvolution were introduced:spectrally invariant [2] and spectrally variant [3].For spectrally invariant two-beam coupling, it was

necessary to operate the system at very high beamratios in order to perform dynamic range compres-sion deconvolution; working in this regime leads tovery low grating efficiency. To overcome the high

beam ratio problem the spectrally variant two-beamcoupling was introduced. Spectrally variant deconvo-lution performed better image restoration thanspectrally invariant deconvolution; however, thedrawback in spectrally variant deconvolution is thatthe response time at very high frequencies becomesextremely large (slow processing speed), since the re-sponse time of photorefractive crystals is inverselyproportional to the total intensity of light incidenton the crystal [4–6].

Another limitation of both spectrally variant andspectrally invariant two-beam coupling deconvolu-tion is the need for an input spatial light modulator(SLM). This may cause some problems, in particular,for remote sensing applications, in which the light isincoherent. The interface between the SLM and theCCD camera imposes other constraints in systemsfor correcting supersonic turbulence. One way toovercome this limitation is to use an all-optically

0003-6935/09/193771-17$15.00/0© 2009 Optical Society of America

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3771

Page 2: Optical-limiter MEMS dynamic range compression deconvolution

addressed SLM in which the pixels are addressedindependently, in parallel. One can also design anall-optically addressed SLM with dynamic rangecompression capability. We have already reportedan all-optically addressed microelectromechanicalsystem (MEMS) device [7,8]. This device exhibitsphase saturation with increasing light intensity. Inthis paper, we introduce a nonlinear optical-limiterMEMS (NOLMEMS) [9] whose input–output trans-fer function saturates with light intensity. The dy-namic range compression saturation nonlinearitywas designed using the membrane saturationdescribed in Refs. [7,8]. In contrast to Horner’s work,in which the deformable MEMS device was used asan SLM within an optical correlator system, and thequadratic phase factor within the MEMS deforma-tion was assumed to be a phase error factor [10],this device takes advantage of the phase error factoras an essential requirement for designing aNOLMEMS.NOLMEMS dynamic range compression has sev-

eral advantages over two-beam coupling deconvolu-tion: (1) The membrane reflectivity is ∼100%; thisovercomes the photorefractive crystal’s low gratingefficiency, which is caused by the high beam intensi-ties necessary to achieve dynamic range compres-sion. (2) The processing time is in microseconds(dictated by the membrane response time), comparedto the millisecond scale associated with two-beamcoupling (dictated by the photorefractive material’sresponse time). (3) There is no response time depen-dency on spatial frequency intensities as in thespectrally variant deconvolution. (4) The saturationlimit is approached faster than in two-beamcoupling, which is more effective in extracting theamplitude information for better image recovery.(5) NOLMEMS works with both coherent and inco-herent light. (6) A NOLMEMS device with opticalamplification capability, integrated with photo-detectors, responds to broadband ultralow lightintensities.This new design (NOLMEMS) enhances both the

high-frequency components of the image, restoringedge features, and the signal-to-noise ratio whenthe noise is stronger than the signal, reducingadditive noise. We have already reported electricalcontrol of these devices [7,8]. A specific combinationof AC and DC electronic bias that stabilizes theMEMS optical limiter controls the NOLMEMS de-vice. The deconvolution orders for this device werealso analyzed using a nonlinear transform method[11]. Computer simulations of image deconvolutionusing the NOLMEMS device input–output transferfunction and the deconvolution orders were alsoperformed.

2. Background

To understand the nonlinear dynamic range com-pression deconvolution approach introduced in thispaper, we present background related to processingphase-only information. Images have spectra, just

like signals. Low-frequency components occur whensignal strength changes slowly, while for images, low-frequency components occur when there is littlechange in gray levels. High-frequency componentsin the signal occur when there are rapid changesin signal strength; for images, high-frequency compo-nents are created by sharp edges. The spectrum of animage has two components, the magnitude spectrumand the phase spectrum. Because the spectrum is ob-tained from a two-dimensional input, the spectrumhas frequency elements in two directions, x (rows)and y (columns). If the combined magnitude andphase spectra are inverse Fourier transformed, theoriginal image appears. If the magnitude spectrumis inverse Fourier transformed, the result showsnothing; however, if the inverse Fourier transformis applied to the phase information with the magni-tude set to unity, an edge-enhanced version of theoriginal image is reconstructed [12]. Therefore, oneconcludes that the phase information of an imageis the more important information in reconstructingthe image. The gray scale of the image can be recov-ered by cascading the phase image with a statisticallowpass filter that has a similar spectral distributionto that of the original image. The quality of thisrecovered image is almost as good as that of the ori-ginal. This method was introduced in 1990 by Hornerand Flavin [13]. Figure 1 illustrates the above discus-sion regarding image recovery using only the phaseinformation of the original input.

Figure 1A is the original image, Fig. 1B is the ori-ginal image with no phase information, and Fig. 1C isthe original image with no amplitude information.Figure 1B shows that if the phase information is ex-tracted from the original, the recovered image showsno information from the original, while Fig. 1C de-monstrates the recovered image from the phase-onlyinformation, which is the edge-enhanced version ofthe original input. Applying a generic form of low-pass filter to the image in Fig. 1C recovers a gray-level image with a quality comparable to the original.The filtered image is shown in Fig. 1D. The genericlowpass filter used here was of the following form:

LPFðu; νÞ ¼ 1

1þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiðuþM=2Þ2þðνþN=2Þ2

D0

q ; ð1Þ

where u and v are the frequency coordinates, M ×Npixels is the array size, andD0 is the cutoff frequency.This filter will be used for gray-level image recoveryin this paper.

3. MEMS Optical Limiter

Figure 2 shows the proposed architecture for a singlepixel of a MEMS optical limiter. In the proposedNOLMEMS, a plane wave is incident on the mem-brane mirror of a back-illuminated device and, asthe membrane deflects, its parabolic shape focusesthe light to its diffraction limit. The electronic biasacross the device can be set such that the membrane

3772 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 3: Optical-limiter MEMS dynamic range compression deconvolution

mirror, when it is fully deflected, can focus the beamon the pinhole. Appropriate pinhole and microlensarrays will collimate each of the saturated MEMSoutput beams, as shown in Fig. 2. If the device is ac-curately fabricated, uniform illumination on thephotoconductive layer will produce a plane-wave out-put accompanied by weak off-axis higher-order dif-fraction beams. For an individual MEMS elementwith very little light incident on the photoconductivelayer, the reflected beam from the MEMS mirrorwill be very small, while very strong illuminationof the photoconductive layer produces an intensereflected beam.

4. Background Theory

The input–output nonlinear transfer function of theNOLMEMS device, which shows the relationship be-tween the light transmitted through the pinhole tothe light intensity incident on the back side of the

device, was derived by Haji-saeed [14]. It was foundthat the device has a saturation limit given by

GðyÞ ∝�

AyAyþ B

�2; ð2Þ

where the parameters are defined as

A ¼ 8λTs2; ð2aÞ

B ¼ r61V2L2ω2π3ε30; ð2bÞ

y ¼ ðaþ bðΔnþ nÞ2Þ; ð2cÞ

a ¼ L2ω2π2r41ε20; ð2dÞ

Fig. 1. A, original image; B, image recovery from the amplitude-only information; C, image recovery from the phase-only information;D, gray-level recovery via application of a lowpass filter to C.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3773

Page 4: Optical-limiter MEMS dynamic range compression deconvolution

b ¼ q2μ2nA2ds

2: ð2eÞ

In the above equations, λ is the back-illuminationwavelength, ε0 is the permittivity of free space, 2r1is the diameter of a pixel, s is the depth of the well,T is the membrane tension, V is the total voltageapplied across the MEMS, q is the electronic charge,μn is the carrier mobility, L is substrate thickness, Adis the area of each pixel, and ω is the angularfrequency. n andΔn are the substrate carrier concen-tration and the photogenerated carrier concentra-tion, respectively.Δn can be described as a function of light intensity,

I0, incident on the back of the device:

Δn ¼ ατnI0hνAd

; ð3Þ

where α is the absorption coefficient, τn is the carrierlifetime, h is Planck’s constant, v is the lightfrequency, and Ad is the illuminated area.The optical-limiter nonlinear transfer function,

GðyÞ, as a function of back-illumination intensity,I0, is shown in Fig. 3. The plot shows the device sa-turation (GaAs substrate) as a function of light

intensity. Note that, as the beam intensity increases,the total reflected light intensity saturates to unity.

In this plot the parameters are as follows:

a. For calculating the Ad parameter, λ ¼ 980nm,s ¼ 5 μm, and T was calculated using

hd ¼ ε0r21V2

32Ts2:

All the parameters for this equation are known fromthe experimental results. The maximum deflectionhd for this particular device was almost 0.4 for50V applied bias with the pixel radius r1 ¼ 1mm,and ε0 is 8:85 × 10−12 F=m.

b. For calculating the B parameter, L ¼ 250 μmand ω ¼ 10kHz.

c. For calculating y, the a and b first need to befound. All parameters for a are known and for b,q ¼ 1:6 × 10−19 C, μn ¼ 8500 cm2=ðV-sÞ, and Ad ¼πr12.

TheΔn can be calculated using Eq. (3), where α forGaAs substrate is 2 cm−1, τn ¼ 10−8 s, the Planck’sconstant h is 6:63 × 10−34 J, v ¼ c=λ, wherec ¼ 3 × 108 m=s, and I0 is from 0 to 10mW. The nparameter was chosen to be an intrinsic carrier con-centration for GaAs by a value of 2 × 106 cm−3.

5. NOLMEMS Joint Fourier Processor

Figure 4 shows a schematic diagram of a NOLMEMSjoint Fourier processor. Reference (R) and signal (S)information are joint Fourier transformed by a lens,L, into an all-optically addressed MEMS deformablemirror located at one facet of a cubic beam splitter.Reference and signal correspond to a very remote ob-ject (a point source) and the target of interest passingthrough the atmosphere, respectively. Therefore, thejoint input image consists of the impulse response ofthe atmosphere (the blur) and the blurred target in-put image. The joint information before addressingthe device is high-pass filtered for emphasizing

Fig. 2. (Color online) Single-pixel MEMS optical-limiterarchitecture.

Fig. 3. (Color online) Plot of NOLMEMS transfer function.

3774 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 5: Optical-limiter MEMS dynamic range compression deconvolution

higher frequencies and avoiding very high beam in-tensity requirements to achieve dynamic range com-pression. Otherwise, the high spatial frequencyintensities will be degraded due to the power eightdependency within the saturation function inEq. (2). This technique makes the joint spectra illu-mination more uniform. The high-pass filter infor-mation is fed to the SLM located right before theNOLMEMS device. The deformable mirror functionsas an array of parabolic mirrors. The focal point ofeach of these parabolic mirrors is dependent onthe joint spectra back-illumination intensity. The sa-turation focal plane is located within distance f s fromthe deformable mirror.A plane wave propagating counter to the direction

of the joint Fourier spectrum of the input images, isincident to the front side of the deformable mirror.The plane wave is reflected out of the parabolic mi-cromirror to a focused array of wavelets. Dependingon the joint spectra back-illumination intensity, eachof the wavelets is focused on a different focal plane.The minimum focal length of each of the microde-formable mirrors is always greater than or equalto f s.A planoconcave lens is located at the other facet of

the beam splitter, imaging the saturation focal planeinto an array of pinholes. The pinhole array isaligned to the center of the converging wavelets.Only a portion of the converging wavelets is trans-

mitted through the pinhole array. An array of wave-let sources is generated after transmission withinthe pinhole array. The array of point sources fromthe pinhole array is collimated to a plane wave usinga lenslet array. The superposed wavefront of plane-wave arrays from the micromirror represents a dy-namic range compressed version of the joint spectraof the reference and the signal information. Thesuperposed plane wave consists of a collimated wave-let array, which is spatially low-pass filtered by usingan SLM. After spatial filtering, the output is Fouriertransformed by a lens, L, into a CCD camera, therebyproducing the required joint Fourier transformsignal processing functionality.

6. Device Theory

In the joint Fourier processor, Δn corresponds to thenumber of carriers generated by the light intensitythat has the joint spectra information and is equal to

Δnðνx; νyÞ ¼ατnI0jRþ Sj2hνADðλf Þ2

; ð4Þ

where I0 is the total beam intensity transmitted withthe reference and the signal information, λ is thewavelength, and f is the lens focal length. Incorpor-ating the joint spectra into Eq. (2) and expanding thenonlinear transfer function to several orders gives us

Fig. 4. (Color online) Schematic diagram of a NOLMEMS joint Fourier processor.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3775

Page 6: Optical-limiter MEMS dynamic range compression deconvolution

GðyÞ ¼X∞k¼0

Hkðνx; νyÞ cos½2kyoνy þ kðϕR − ϕSÞ�; ð5Þ

where y0 is the separation factor between the refer-ence and signal information, ΦR and ΦS are thereference and signal phases, respectively, andHkðvx; vyÞ is the deconvolution order given by

Hkðvx; vyÞ ¼ 1 − εkikþ1

��2mð0ÞJA×

1 − f 2 þffiffiffiffiffif 1

p�

k�Bm2ð0Þ

Ab

�1ffiffiffiffiffif 1

p�−B2m4ð0Þ4ðAbÞ2

�f 2 þ f 3f 1

ffiffiffiffiffif 1

p�

−kB2m4ð0Þ4ðAbÞ2

�2 − f 2

ð1 − f 2Þf 1 þ f 1ffiffiffiffiffif 1

p��

�2mð0Þy

−1 − f 2 þ ffiffiffiffiffie1

p�

k�Bm2ð0Þ

Ab

�1ffiffiffiffiffie1

p�−B2m4ð0Þ4ðAbÞ2

�f 3 − f 2e1

ffiffiffiffiffie1

p�

−kB2m4ð0Þ4ðAbÞ2

�f 2

ð−1 − f 2Þe1 þ e1ffiffiffiffiffie1

p���

;

ð6Þ

where εk is the Neumann factor (εk ¼ 1 for k ¼0 andεk ¼ 2 for k > 0), and

JA−¼ ðR − SÞ2; ð6aÞ

JAþ ¼ ðRþ SÞ2; ð6bÞ

JA×¼ RS; ð6cÞ

JI¼ ¼ R2 þ S2; ð6dÞ

f 2 ¼ imð0ÞðJIþ þ nÞ; ð6eÞ

f 3 ¼ m2ð0ÞðJAþ þ nÞðJA− þ nÞ; ð6f Þ

f 1 ¼ 1 − 2f 2 − f 3; ð6gÞ

e1 ¼ 1þ 2f 2 − f 3; ð6hÞ

m2ð0Þ ¼ AbaAþ B

: ð6iÞ

The joint spectra coefficients in Eq. (6) are complexnumbers and complex coefficients introduce phase

distortion to the output signal. In order to evaluatethe effect of the severity of this phase distortion, theabsolute value, real, and imaginary part of the com-plex numbers and the argument of the joint spectracoefficient [Eq. (6)] are plotted and are shown inFigs. 5–7. In these plots R and S are assumed tobe equal to unity.

Figures 5A, 5B, 5C, and 5D show plots of the abso-lute, the real, the imaginary, and the argumentvalues of H1ðvx; vyÞ, respectively. The same para-meters that have been used for Fig. 3 plot are usedfor these plots. The only unknown parameters arethe joint spectra parameters. Since R ¼ S ¼ 1, theseparameters are found to be JA− ¼ 0, JAþ ¼ 4,JIþ ¼ 2, and JA× ¼ 1. As is evident from Fig. 5A,the H1ðvx; vyÞ value has been increased by simplyincreasing the intensity I0 from 3mW to 9mW.Comparing almost identical absolute and real plotsof the H1ðvx; vyÞ in Fig. 5 shows that the main contri-bution of the H1ðvx; vyÞ magnitude is from its realvalue component. This also can be verified from al-most identical imaginary and phase plots. Eventhough the phase has been increased by 2 ordersof magnitude (from 10−11 to 10−9) by an increase from1mW light intensity to 10mW, yet the imaginarycomponent has an insignificant contribution to theH1ðvx; vyÞ magnitude. The phase and the imaginaryfactors can be identical if and only if the imaginarycomponent of a complex number is extremely small.Since the phase of the first-order joint spectra coeffi-cient (H1ðvx; vyÞ) is insignificant, it is possible toignore the phase distortion and utilize this coefficientfor both correlation and convolution.

Figure 6 shows the same sequence of plots as Fig. 5except for theH2ðvx; vyÞ coefficient. It is evident fromthese plots that the magnitude of the imaginary com-ponent of this coefficient is much larger than its realcomponent. The real values have a small declinefrom 1 to 0.99985 as the intensity increases from1mW to 8mW, and the imaginary values increaseslightly from −85 to −83 as the intensity increasesfrom 2mW to 10mW; hence, the imaginary compo-nent has the most contribution of the H2ðvx; vyÞ

3776 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 7: Optical-limiter MEMS dynamic range compression deconvolution

absolute value rather than the real component. Thisis evident by comparing Figs. 6A and 6C.The insignificant changes in theH2ðvx; vyÞ real and

imaginary component plots lead to nearly a constantphase. According to Fig. 6D, the H2ðvx; vyÞ phase de-creases slightly from 1.5591 to 1.5587 as the inten-sity increases from 1mW to 10mW. Since thephase is almost a constant, the phase distortioncan be negligible for utilizing this coefficient in cer-tain signal processing criteria. In contrast to thefirst-order coefficient H1ðvx; vyÞ, where its absolutevalue was dominated from its real component, herethe absolute value of the second-order H2ðvx; vyÞ isdominated by its imaginary component.Figure 7 shows the same sequence of plots as

Figs. 5 and 6 for the third-order coefficientH3ðvx; vyÞ. Similar to the first-order coefficient theabsolute value of the H3ðvx; vyÞ is mainly dominatedby its real component. This is evident when compar-ing plots of H3ðvx; vyÞ absolute values (Fig. 7A) andthe real values (Fig. 7B). Both curves almost areidentical and drop from 1:44 × 107 to 1:38 × 107 asthe intensity increases from 1mW to 10mW. Theimaginary values drop from −1 to −25 as the inten-

sity increases from 1mW to 10mW, which are insig-nificant compared to the real values. Although theH3ðvx; vyÞ phase factor increases from 0 to1:8 × 10−6, it is insignificant enough to cause anyphase distortion.

7. Computer Simulation

The effectiveness of the proposed image recovery wastested with computer simulation. A 128 × 128 pixelLena face located in the middle of a 512 × 512 nullarray was used for the input image. The blurred im-age was generated by convolving the original inputwith common blur functions: motion, atmosphericturbulence, and misfocusing. Random noise was alsoadded to the blurred image. The noise was created bya random-number generator that produced whiteGaussian noise with zero mean and a variance of1, and was added to the blurred joint input.

The motion impulse response was a 2 × 50 pixelrectangle. The atmospheric turbulence was [15]

Hðνx; νyÞ ¼ e−βðν2xþν2yÞ5=6 ð7Þ

Fig. 5. (Color online) H1ðvx; vyÞ plots: A, absolute value; B, real value; C, imaginary value; and D, phase.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3777

Page 8: Optical-limiter MEMS dynamic range compression deconvolution

for β ¼ 0:0025. The misfocusing impulse responsewas a circle of ones with a diameter of 20 pixels.The nearly hard-clipping nature of the NOLMEMS

device saturation nonlinearity fully clips the jointspectra high frequencies, even for low lightintensities.Figures 8–15 show the simulation results based on

the first-order nonlinear transform method expan-sion presented in Eq. (6). Figures 8, 10, 12, and 14represent the results of deconvolution for the follow-ing blur functions respectively: motion, atmospheric,misfocusing, and phase, while A is the noisy blurredinput with SNR ¼ 5; A′ is the corresponding recov-ered image. The B and C rows are sequentially thesame as the A row for SNR ¼ 1 and SNR ¼ 0:1, re-spectively. Figures 9, 11, 13, and 15 represent a com-parative study of the deconvolution via a NOLMEMSdevice with both of theWiener and the inverse filters.We used the following equation for the Wiener filter:

Wðu; vÞ ¼�

H�ðu; vÞjFðu; vÞj2jHðu; vÞj2jFðu; vÞj2 þ jNðu; vÞj2

�; ð8Þ

where Hðu; vÞ is the blur function, H�ðu; vÞ is thecomplex conjugate of Hðu; vÞ, jHðu; vÞj2, jFðu; vÞj2,

and jNðu; vÞj2 are the power spectrum of the distor-tion function, the undegraded image, and the noise,respectively. In this approach, the goal is to find anestimated recovered image of the uncorrupted imagein such a way that the mean square error betweenthem is minimized.

In all of these figures, A represents the recoveredimage with spectrally variable joint information forSNR ¼ 1, B is the recovered image using the Wienerfilter, with the expected value of F (an estimated re-covered image), C is the recovered image using in-verse filtering, and D is the recovered image usingthe Wiener filter, with the exact value of F (theuncorrupted image).

As it is evident in the C figures, the inverse filterwas incapable of either recovering or even detectingthe image. In these results noise is covering thewhole array. By comparing the As with the Ds, itis clear that, for all the blur functions tested here,the image recovery of a noisy blurred signal withSNR ¼ 1 via the spectrally variant NOLMEMS de-convolution was always better (with a generic shapeof the lowpass filter or the expected value of F) thanthe Wiener filter with the exact value of F. When theexact value of F was replaced by a generic shape of

Fig. 6. (Color online) H2ðvx; vyÞ plots: A, absolute value; B, real value; C, imaginary value; D, phase.

3778 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 9: Optical-limiter MEMS dynamic range compression deconvolution

the lowpass filter or the expected value of F, theWiener filter failed, as is evident in B for Figs. 9,11, 13, and 15.As it was discussed previously, this large difference

in the power requirement is attributed to the powereight dependency of the input–output nonlineartransfer function of this device on the incident opticalintensity. For blurred signals, from the signal proces-sing point view, power eight dependency leads forinfinitesimally small high frequency intensitieswithin the input–output nonlinear transfer function[Eq. (2)]. This requires a high power to obtain outputsaturation. Therefore, the use of a highpass filter forpreprocessing is essential to release the constraintsfor unrealistically high power beam intensities.The dynamic range compression deconvolution

technique is associated with noise conversion fromlow frequencies to high frequencies, which is the ran-dom distribution of delta functions (salt and peppernoise) [16–19]. This frequency conversion leads tofurther enhancement of the signal-to-noise ratiothrough spreading the noise over the entire array,while in image recovery via the Wiener filter, thenoise remains nearly concentrated in the originalnoisy image area.

It is clear from the simulation results that dynamicrange compression deconvolution via NOLMEMS iscapable of both noise reduction and image restora-tion embedded in a severe noise environment fromaberrations such as motion, atmospheric, misfocus-ing, and phase. A signal-to-noise ratio of 1 meansthat the probability of signal detection is ∼50% witha 0.1 average false alarm rate (FAR).

8. Conclusion

This paper proposes dynamic range compressiondeconvolution using a newly designed NOLMEMSdevice, and shows that it has several advantages overtwo-beam coupling deconvolution. This device candefeat the low grating efficiency problem in two-beam coupling deconvolution. The processing timefor the NOLMEMS device is in the microsecondrange. Unlike the two-beam coupling deconvolutionthat processes only coherent light, this device canalso work with incoherent light. It is possible to de-sign a MEMS optical limiter that can detect and pro-cess broadband ultralow light intensities. Computersimulation results showed excellent image recoveryfrom three different blur functions in a high noiseenvironment.

Fig. 7. (Color online) H3ðvx; vyÞ plots: A, absolute value; B, real value; C, imaginary value; and D, phase.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3779

Page 10: Optical-limiter MEMS dynamic range compression deconvolution

Fig. 8. Motion aberration indirect simulation results: A, noisyblurred input with SNR ¼ 5; A′, corresponding recovered image.The B and C rows are sequentially the same as the A row forSNR ¼ 1 and SNR ¼ 0:1, respectively.

Fig. 9. Deconvolutionresult for themotionaberrationcomparedtoWiener and inverse filters result: A, recovered image for SNR ¼ 1;B,recoveredimageusingtheWienerfilter,withtheexpectedvalueofF;C,recoveredimageusinginversefiltering,andD, recoveredimageusing Wiener filter, with the exact value of F.

Fig. 10. Atmospheric turbulence indirect simulation results:A, noisy blurred input with SNR ¼ 5; A′, corresponding recoveredimage. The B and C rows are sequentially the same as the A rowfor SNR ¼ 1 and SNR ¼ 0:1, respectively.

Fig. 11. Deconvolution result for the atmospheric turbulencecompared to theWiener and inverse filters result: A, recovered im-age for SNR ¼ 1; B, recovered image using the Wiener filter, withthe expected value of F; C, recovered image using inverse filtering;D, recovered image using the Wiener filter, with the exact valueof F.

3780 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 11: Optical-limiter MEMS dynamic range compression deconvolution

Fig. 12. Misfocusing aberration indirect simulation results:A, noisy blurred input with SNR ¼ 5; A′ the corresponding recov-ered image. The B and C rows are sequentially the same as the Arow for SNR ¼ 1 and SNR ¼ 0:1, respectively.

Fig. 13. Deconvolution result for the misfocusing aberration com-pared to the Wiener and inverse filters result: A, recovered imagefor SNR ¼ 1; B, recovered image using the Wiener filter, with theexpected value of F; C, recovered image using inverse filtering; andD, recovered image using Wiener filter, with the exact value of F.

Fig. 14. Phase-only aberration indirect simulation results:A, noisy blurred input with SNR ¼ 5; A′, the corresponding recov-ered image. The B and C rows are sequentially the same as the Arow for SNR ¼ 1 and SNR ¼ 0:1, respectively.

Fig. 15. Deconvolution result for the phase-only aberration com-pared to the Wiener and inverse filters result: A, recovered imagefor SNR ¼ 1; B, recovered image using the Wiener filter, with theexpected value of F; C, recovered image using inverse filtering; andD, recovered image using Wiener filter, with the exact value of F.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3781

Page 12: Optical-limiter MEMS dynamic range compression deconvolution

Appendix A

The nonlinear transformmethod was used to analyzethe nonlinear transfer function in Eq. (2). For thisanalysis, the nonlinear transfer function was ex-panded to a superposition of four infinitesimally dif-ferent complex negative saturation nonlineartransfer functions as follows:

f ðEÞ ¼ 11� iGpðEþ E0Þ

; ðA1Þ

whereGp is a generic parameter, E is the energy, andE0 is a constant energy parameter.In the analysis of any nonlinear system, generally

any generic function in the form of f ðEÞ is expandedin terms of its frequency spectrum FðωÞ, where ω isits complement frequency variable.The relationship between the input nonlinear

transfer function and its frequency spectrum canbe expressed as

FðωÞ ¼Z

−∞

f ðEÞ expð−iωEÞdE; ðA2Þ

f ðEÞ ¼ 12π

Z∞

−∞

FðωÞ expðiωEÞdω: ðA3Þ

In a joint Fourier processor, let us assume that thescene transmittance function is sðx; yþ y0Þ and thereference transmittance function is rðx; y − y0Þ. yo isthe separation between the reference and the signalinformation. In the Fourier plane, the joint intensityspectrum is

E ¼ jSðνx; νyÞ þ Rðνx; νyÞj2¼ S2ðνx; νyÞ

þ Sðνx; νyÞRðνx; νyÞ exp½iϕSðνx; νyÞ− iϕRðνx; νyÞ� expð−i2y0νyÞþ Sðνx; νyÞRðνx; νyÞ exp½−iϕSðνx; νyÞþ iϕRðνx; νyÞ� expðþi2y0νyÞ þ R2ðνx; νyÞ; ðA4Þ

where R and S are the Fourier transforms of r and s,respectively. ΦR and ΦS are the reference and thesignal phases and vx and vy are the spatial frequencycoordinates.Substituting Eq. (A3) into Eq. (A4) yields

f ðEÞ ¼ 12π

Z∞

−∞

FðωÞ exp½iωðR2

þ S2Þ� exp½i2ωRS cosð2y0νy þ ϕR − ϕSÞ�dω:ðA5Þ

The exponential factor containing the cosine termis expanded using the Jacobi–Anger formula:

f ðEÞ ¼ 12π

Z∞

−∞

FðωÞ exp½iωðR2

þ S2Þ��X∞

k¼−∞

ikJkð2ωRSÞ cos½2ky0νy

þ kðϕR − ϕSÞ��dω; ðA6Þ

where Jk is the Bessel function of the first kind,order k.

The integral form of f ðEÞ in Eq. (A3) is thenconverted into a summation series:

f ðEÞ ¼X∞k¼0

Hkðνx; νyÞ cos½2ky0νy þ kðϕR − ϕSÞ�; ðA7Þ

where the weighting factors Hk are given by

Hkðνx; νyÞ ¼εk2π i

k

Z∞

−∞

FðωÞ exp½iωðR2

þ S2Þ�Jkð2ωRSÞdω; ðA8Þ

and εk is the Neumann factor (εk ¼ 1 for k ¼ 0 andεk ¼ 2 for K > 0).

The Fourier transform of the complex negativesaturation nonlinear transfer functions describedin Eq. (A1) is

FðωÞ ¼ −i2π�Gp

e−ω=�GpeiωE0 : ðA9Þ

Substituting Eq. (A9) into Eq. (A8) yields theweighting factors Hkðvx; vyÞ as

Hkðvx;vyÞ¼εk2πi

k

Z∞

0

−i2π�Gp

e−ω=�GpeiωðJIþÞeiωE0Jkð2ωJA×Þdω

¼−εkikþ1

�Z∞

0e−rcosðrGpðJIþ

þE0ÞÞJkð2rGpJA×Þdr

þiZ

0e−rsinðrGpðJ2

þE0ÞÞJkð2rGpJA×Þdr�;

ðA10Þ

where

� r ¼ ω�Gp

; ðA10:1Þ

and JA−, JAþ, JA×, and JIþ are the joint spectra mo-mentums defined as

3782 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 13: Optical-limiter MEMS dynamic range compression deconvolution

JA− ¼ ðR − SÞ2; JAþ ¼ ðRþ SÞ2; JA×

¼ RS; JIþ ¼ R2 þ S2:

ðA10:2Þ

The above integral has a pair of Laplace transformsin the form of sinðbxÞJvðaxÞ and cosðbxÞJvðaxÞ. TheLaplace transforms of the above pair are

�R∞

0 e−px sin bxR∞

0 e−px cosbx

�JνðaxÞdx ¼ 1

2

�i1

�ðuþv−νþ ∓u−v−ν− Þ;

ðA11Þ

where

u� ¼ ½ðp� ibÞ2 þ a2�−12; ðA11:1Þ

ν� ¼ a−1½p� ibþ ððp� ibÞ2 þ a2Þ�: ðA11:2Þ

Accordingly the u� and v� of the Laplace transformswithin Eq. (A10) are

u� ¼ 1ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið1� iGpðJIþ þ E0ÞÞ2 þ 4G2

pJ2A×

q ; ðA12Þ

ν� ¼1� iGpðJIþ þ E0Þ þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið1� iG2

pðJIþ þ E0ÞÞ2 þ 4G2pJ2

q2GpJA×

: ðA13Þ

Through using the Laplace transform in Eq. (A11), Hkðvx; vyÞ in Eq. (A10) yields

HCkðvx; vyÞ ¼ εkikþ1

�1

fð1 − ið�GpÞðJAþ þ E0ÞÞð1 − ið�GpÞðJA− þ E0ÞÞg1=2

�

2ð�GpÞJA×

1 − ið�GpÞðJIþ þ x0Þ þ fð1 − ið�GpÞðJAþ þ E0ÞÞð1 − ið�GpÞðJA− þ E0ÞÞg1=2�

k�: ðA14Þ

The zero order in this case is

HC0ðvx; vyÞ ¼1

fð1 − ið�GpÞðJAþ þ E0ÞÞð1 − ið�GpÞðJA− þ E0ÞÞg1=2; ðA15Þ

PCkðvx; vyÞ ¼�

2ð�GpÞJA×

1 − ið�GpÞðJIþ þ x0Þ þ fð1 − ið�GpÞðJAþ þ E0ÞÞð1 − ið�GpÞðJA− þ E0ÞÞg1=2�

kðA16Þ

HCkðvx; vyÞ ¼ −εkikþ1HC0ðvx; vyÞPCkðvx; vyÞ: ðA17Þ

The input–output nonlinear transfer function ofthe optical limiter as described in Eq. (2) is

Iout ¼A2y2

ðAyþ BÞ2 Iin ¼ GðyÞIin; ðA18Þ

where

y ¼ aþ bðxþ x0Þ2; ðA18:1Þ

a ¼ L2ω2π2r41ε20; ðA18:2Þ

b ¼ q2μ2nA2s2; ðA18:3Þ

A ¼ 8λTs2; ðA18:4Þ

B ¼ r61V2L2ω2π3ε30: ðA18:5Þ

For simplicity it was assumed that Δnþ n ¼ xþ x0.

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3783

Page 14: Optical-limiter MEMS dynamic range compression deconvolution

In above equations, ε0 is the permittivity of freespace, 2r1 is the size of one pixel pattern, s is thedepth of the well, T is the membrane tension, V isthe total voltage applied across the MEMS, q isthe electronic charge, μn is the carrier mobility, nis the carrier concentration under dark illuminationconditions, L is the substrate thickness, A is the areaof each pixel, Δn is the increase in carrier concentra-tion due to light illumination, and ω is the frequency.For subsequent presentation of the joint spectra

coefficient in a general standard presentation, it isrelevant to write Eq. (A18) in different format withnew parameters. Equation (A18) can be rearrangedin the following format:

haA

½AaþB� þ bA½AaþB� ðxþ x0Þ2

i2

½Aaþ B�2h1þ bA

AaþB ðxþ x0Þ2i2

¼h

α1βð0Þ þ γ2ð0Þðxþ x0Þ2

i2

β2ð0Þ½1þ γ2ð0Þðxþ x0Þ2�2

¼ ½γ1ð0Þ þm2ð0Þðxþ x0Þ2�2β2ð0Þ½1þm2ð0Þðxþ x0Þ2�2

; ðA19Þ

α1 ¼ aA; ðA19:1Þ

α2 ¼ Ab; ðA19:2Þ

βð0Þ ¼ aAþ B; ðA19:3Þ

γ1ð0Þ ¼α1βð0Þ ¼

AaaAþ B

; ðA19:4Þ

γ2ð0Þ ¼α2βð0Þ ¼

AbaAþ B

¼ m2ð0Þ: ðA19:5Þ

In a joint Fourier processor, x corresponds to thenumber of carriers that is generated by the light in-tensity that has the joint spectra information and isequivalent to

xðνx; νyÞ ¼ατnI0jRþ Sj2hνADðλf Þ2

; ðA20Þ

where α is the absorption coefficient, τn is the carrierlifetime, I0 is the total beam intensity transmittedwith the reference and the signal information, h isPlanck’s constant, v is the light frequency, AD isthe detector area, λ is the wavelength, and f is thelens focal length.Incorporating the joint spectra into Eq. (A19)

yields that

hγ1ð0Þþm2ð0Þ

�ατn

hνADðλf Þ22�I0jRþSj2þx0

hνADðλf Þ2ατn

2i2

β2ð0Þh1þm2ð0Þ

�ατn

hνADðλf Þ22�I0jRþSj2þx0

hνADðλf Þ2ατn

2�2

¼hγ1ð0Þþm2ð0ÞB2

f

�I0jRþSj2þ x0

Bf

2i2

β2ð0Þh1þm2ð0ÞB2

f

�I0jRþSj2þ x0

Bf

2i2

¼ ½γ1ð0ÞþCmðI0jRþSj2þ IeqÞ2�2β2ð0Þ½1þCmðI0jRþSj2þ IeqÞ2�2

; ðA21Þ

where

Bf ¼ατn

hνADðλf Þ2; ðA21aÞ

Cm ¼ mð0ÞBf ; ðA21bÞ

Ieq ¼ x0Bf

: ðA21cÞFor simplicity of the analysis at this stage, theanalysis using Eq. (A18) is considered. Throughthe analysis, the parameters were defined inEqs. (A19)–(A21) will be replaced.

The input–output nonlinear transfer function inEq. (A1) is a positive saturation function. For the pur-pose of the analysis, it was found from prior workthat, in general, it is simpler to analyze a negativesaturation function. Therefore, this function can bewritten as

GðyÞ ¼ A2y2

ðAyþ BÞ2 ¼ 1 −2AByþ B2

ðAyþ BÞ2 : ðA22Þ

The second term in the right-hand side of Eq. (A22) isa negative saturation function. This function is stillnot easy to analyze using the nonlinear transformmethods. The fractional expansion is done in twostages. The first stage is for the second-order expan-sions (i.e., y expansion) and the second stage is for thefirst-order fraction expansion (i.e., x expansion).After expansion to the first order,. the nonlineartransform method is used to derive the contributionof each of the expansion to the correlation and theconvolution factors. Then the contribution from allthe fractional orders was superimposed.

At first, the second term in the right-hand side ofEq. (A22) is fractionally expanded as follows:

2AByþ B2

ðAyþ BÞ2 ¼ limε→0

2AByþ B2

ðAyþ Bþ εÞðAyþ B − εÞ : ðA23Þ

The right-hand side of Eq. (A23) can be fractionallyexpanded to

1 −g1ðεÞ

Ayþ Bþ εþg1ð−εÞ

Ayþ B − ε ; ðA24Þ

where

3784 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 15: Optical-limiter MEMS dynamic range compression deconvolution

g1ð�εÞ ¼ B

�1þ B

�2ε

�ðA24:1Þ

Since y ¼ aþ bðxþ x0Þ2 Eq. (A24) can be expandedfurther to the first order as

g1ð�εÞAyþ B� ε ¼

g1ð�εÞβð�εÞ ·

1

1þ γ2ð�εÞðxþ x0Þ2; ðA25Þ

where

βð�εÞ ¼ aAþ B� ε; ðA25:1Þ

γ2ð�εÞ ¼ α2βð�εÞ ¼

AbaAþ B� ε ¼ m2ð�εÞ: ðA25:2Þ

Using partial fraction method, this equation can beexpanded further into two first-order components:

−g1ð�εÞ2βð�εÞ ·

1ffiffiffiffiffiffiffiffiffiffiffiffiffiffiγ2ð�εÞp

xþ iþ ffiffiffiffiffiffiffiffiffiffiffiffiffiffiγ2ð�εÞp

x0þ g1ð�εÞ2βð�εÞ

·1ffiffiffiffiffiffiffiffiffiffiffiffiffiffi

γ2ð�εÞpx − iþ ffiffiffiffiffiffiffiffiffiffiffiffiffiffi

γ2ð�εÞpx0

¼ −ig2ð�εÞ1 − imð�εÞðxþ x0Þ

þ ig2ð�εÞ1þ imð�εÞðxþ x0Þ

;

ðA26Þ

where

g2ð�εÞ ¼ g1ð�εÞ2βð�εÞ : ðA26:1Þ

Equation (A26) consists of the summation ofnonlinear transfer functions described in Eq. (A1).Therefore, combining the contribution from bothmð�εÞ and −mð�εÞ and through replacing x to its cor-respondence equation [Eq. (A21)] gives us

HCkðvx; vy;mð�εÞÞ ¼ εkikþ1g2ð�εÞ�

1

fð1 − imð�εÞðJAþ þ E0ÞÞð1 − imð�εÞðJA− þ E0ÞÞg1=2

�

2mð�εÞJAx

1 − imð�εÞðJIþ þ x0Þ þ fð1 − imð�εÞðJAþ þ E0ÞÞð1 − imð�εÞðJA− þ E0ÞÞg1=2�

−1

fð1þ imð�εÞðJAþ þ E0ÞÞð1þ imð�εÞðJA− þ E0ÞÞg1=2

�

2mð�εÞJAx

−1 − imð�εÞðJIþ þ x0Þ þ fð1þ imð�εÞðJAþ þ E0ÞÞð1þ imð�εÞðJA− þ E0ÞÞg1=2�

k�;

ðA27Þwhere

JA−¼ ðR − SÞ2; JAþ ¼ ðRþ SÞ2; JA×

¼ RS; JI¼ ¼ R2 þ S2: ðA27:1ÞCombining all terms from mðεÞ and mð−εÞ. the approximate joint spectra gives us

HCkðvx; vy;mð�εÞÞ ¼ 1 −

�g1ðþεÞεkikþ1

2βðþεÞ�

1

fð1 − imðþεÞðJAþ þ E0ÞÞð1 − imðþεÞðJA− þ E0ÞÞg1=2

�

2mðþεÞJAx

1 − imðþεÞðJIþ þ x0Þ þ fð1 − imðþεÞðJAþ þ E0ÞÞð1 − imðþεÞðJA− þ E0ÞÞg1=2�

k

−1

fð1þ imðþεÞðJAþ þ E0ÞÞð1þ imðþεÞðJA− þ E0ÞÞg1=2

�

2mðþεÞJAx

−1 − imðþεÞðJIþ þ x0Þ þ fð1þ imðþεÞðJAþ þ E0ÞÞð1þ imðþεÞðJA− þ E0ÞÞg1=2�

k�

þ g1ð−εÞεkikþ1

2βð−εÞ�

1

fð1 − imð−εÞðJAþ þ E0ÞÞð1 − imð−εÞðJA− þ E0ÞÞg1=2

�

2mð−εÞJAx

1 − imð−εÞðJIþ þ x0Þ þ fð1 − imð−εÞðJAþ þ E0ÞÞð1 − imð−εÞðJA− þ E0ÞÞg1=2�

k

−1

fð1þ imð−εÞðJAþ þ E0ÞÞð1þ imð−εÞðJA− þ E0ÞÞg1=2

�

2mð−εÞJAx

−1 − imð−εÞðJIþ þ x0Þ þ fð1þ imð−εÞðJAþ þ E0ÞÞð1þ imð−εÞðJA− þ E0ÞÞg1=2�

k��

:ðA28Þ

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3785

Page 16: Optical-limiter MEMS dynamic range compression deconvolution

The actual value of HCkðvx; vyÞ can be found fromtaking the limit when ε approaches zero:

HCkðvx; vyÞ ¼ limε→0

½HCkðvx; vy;mðεÞÞ

þHCkðvx; vy;mð−εÞÞ�: ðA29Þ

A Taylor expansion equation for the secondorder yields that

Hkðvx; vyÞ ¼ 1 − εkikþ1

��2mð0ÞJA×

1 − f 2 þffiffiffiffiffif 1

p�

k

×�Bm2ð0Þ

Ab

�1ffiffiffiffiffif 1

p�−B2m4ð0Þ4ðAbÞ2

�f 2 þ f 3f 1

ffiffiffiffiffif 1

p�

−kB2m4ð0Þ4ðAbÞ2

�2 − f 2

ð1 − f 2Þf 1 þ f 1ffiffiffiffiffif 1

p��

�2mð0Þy

−1 − f 2 þ ffiffiffiffiffie1

p�

k�Bm2ð0Þ

Ab

�1ffiffiffiffiffie1

p�

−B2m4ð0Þ4ðAbÞ2

�f 3 − f 2e1

ffiffiffiffiffie1

p�

−kB2m4ð0Þ4ðAbÞ2

�f 2

ð−1 − f 2Þe1 þ e1ffiffiffiffiffie1

p���

; ðA30Þ

where

f 2 ¼ imð0ÞðJIþ þ x0Þ; ðA30:1Þ

f 3 ¼ m2ð0ÞðJAþ þ x0ÞðJA− þ x0Þ; ðA30:2Þ

f 1 ¼ 1 − 2f 2 − f 3; ðA30:3Þ

e1 ¼ 1þ 2f 2 − f 3: ðA30:4Þ

Substituting parameters from Eqs. (A19)–(A21)into Eq. (A30), the weighting factors equation yieldsto

Hkðvx; vyÞ ¼ 1 − εkikþ1

��2CmJA×I0

Bf ð1 − f 2 þffiffiffiffiffif 1

p Þ

�k

�BC2

m

B2f Ab

�1ffiffiffiffiffif 1

p�−

B2C4m

4B4f ðAbÞ2

�f 2 þ f 3f 1

ffiffiffiffiffif 1

p�

−kB2C4

m

4B4f ðAbÞ2

�2 − f 2

ð1 − f 2Þf 1 þ f 1ffiffiffiffiffif 1

p��

�2CmJA×I0

Bf ð−1 − f 2 þ ffiffiffiffiffie1

p Þ�

k�BC2

m

B2f Ab

�1ffiffiffiffiffie1

p�

−B2C4

m

4B4f ðAbÞ2

�f 3 − f 2e1

ffiffiffiffiffie1

p�

−kB2C4

m

4B4f ðAbÞ2

�f 2

ð−1 − f 2Þe1 þ e1ffiffiffiffiffie1

p���

; ðA31Þ

where

f 2 ¼ iCm

BfðJIþ þ x0Þ; ðA31:1Þ

f 3 ¼�Cm

Bf

�2ðJAþ þ x0ÞðJA− þ x0Þ; ðA31:2Þ

f 1 ¼ 1 − 2f 2 − f 3; ðA31:3Þ

e1 ¼ 1þ 2f 2 − f 3; ðA31:4Þ

Cm ¼ mð0ÞBf : ðA31:5Þ

References1. Y. Huang, G. Siganakis, M. G. Moharam, and S-T. Wu, “Broad-

band optical limiter based on nonlinear photoinduced aniso-tropy in bacteriorhodopsin film,” Appl. Phys. Lett. 85,5445 (2004).

2. B. Haji-saeed, S. K. Sengupta, W. Goodhue, J. Khoury, C. L.Woods, and J. Kierstead, “Nonlinear dynamic range compres-sion deconvolution,” Opt. Lett. 31, 1969–1971 (2006).

3. B. Haji-saeed, S. K. Sengupta, W. Goodhue, J. Khoury, C. L.Woods, and J. Kierstead, “Spectrally variable two-beamcoupling nonlinear deconvolution,” Appl. Opt. 46, 8244–8249 (2007).

4. H. J. Eichler, P. Gunter, and D. W. Pohl, Laser InducedDynamic Gratings (Springer-Verlag, 1986).

5. T. J. Hall, R. Jaura, L. M. Connors, and P. D. Foote, “The photo-refractive effect—a review,” Prog. Quantum Electron. 10, 77–146 (1985).

6. P. Gunter and J.-P. Huignard, Photorefractive Materials andTheir Applications, Vols. I and II (Springer-Verlag, 1988).

3786 APPLIED OPTICS / Vol. 48, No. 19 / 1 July 2009

Page 17: Optical-limiter MEMS dynamic range compression deconvolution

7. B. Haji-saeed, R. Kolluru, D. Pyburn, R. Leon, S. K. Sengupta,M. Testorf, W. Goodhue, J. Khoury, A. Drehman, C. L. Woods,and J. Kierstead, “Photoconductive optically driven deform-able membrane under high frequency bias: fabrication, char-acterization and modeling,” Appl. Opt. 45, 3226–3236 (2006).

8. J. Khoury, C. L. Woods, B. Haji-saeed, S. K. Sengupta, W.Goodhue, and J. Kierstead, “Optically driven microelectrome-chanical-system deformable mirror under high-frequency ACbias,” Opt. Lett. 31, 808–810 (2006).

9. J. Khoury, B. Haji-saeed, W. D. Goodhue, C. L. Woods, and J.Kierstead, “MEMS-based optical limiter,” Appl. Opt. 47,5468–5472 (2008).

10. J. L. Horner and P. D. Gianino, “Effects of quadratic phase dis-tortion on correlator performance,” Appl. Opt. 31, 3876–3878 (1992).

11. W. B. Davenport andW. L. Root, An Introduction to the Theoryof Random Signal and Noise (McGraw-Hill, 1958), Chaps. 12–13, pp. 255–311

12. J. Van de Vegte, Fundamentals of Digital Signal Processing(Prentice-Hall, 2002), Chap. 15, pp. 659–660.

13. J. L. Horner and M. A. Flavin, “Average amplitude matchedfilter paper,” Opt. Eng. 29, 31–37 (1990).

14. B. Haji-saeed, “Development of novel device assemblies andtechniques for improving adaptive optics imaging systems,”Ph.D. dissertation (University of Massachusetts at Lowell,2006).

15. R. E. Hufnagel and N. R. Stanley, “Modulation transfer func-tion associated with image transmission through turbulentmedia,” J. Opt. Soc. Am. 54, 52–61 (1964).

16. R. C. Gonzalez and R. E. Woods, Digital Image Processing(Prentice-Hall, 2002).

17. G. Asimellis, J. Khoury, and C.Woods, “Effects of saturation onthe nonlinear incoherent-erasure joint-transform correlator,”J. Opt. Soc. Am. A 13, 1345–1356 (1996).

18. J. Khoury, G. Asimellis, P. D. Gianino, and C. L. Woods“Nonlinear compansive noise reduction in joint transformcorrelators,” Opt. Eng. 37, 66–74 (1998).

19. J. Khoury, P. D. Gianino, and C. L.Woods “Engineering aspectsof the two-beam coupling correlator,” Opt. Eng. 39, 1177–1183 (2000).

1 July 2009 / Vol. 48, No. 19 / APPLIED OPTICS 3787


Recommended