throbber
Optical Phased Array Technology
`
`PAUL F. MCMANAMON, SENIOR MEMBER, IEEE, TERRY A. DORSCHNER, MEMBER, IEEE,
`DAVID L. CORKUM, LARRY J. FRIEDMAN, MEMBER, IEEE, DOUGLAS S. HOBBS,
`MICHAEL HOLZ, SERGEY LIBERMAN, HUY Q. NGUYEN, DANIEL P. RESLER,
`RICHARD C. SHARP, MEMBER, IEEE, AND EDWARD A. WATSON
`
`Optical phased arrays represent an enabling new technology
`that makes possible simple, affordable, lightweight, optical sensors
`offering very precise stabilization, random-access pointing, pro(cid:173)
`grammable multiple simultaneous beams, a dynamic focus/defocus
`capability, and moderate to excellent optical power handling ca(cid:173)
`pability. These new arrays steer or otherwise operate- on an
`already formed beam, as compared to modem microwave phased
`arrays which both generate a beam and direct it in a specific
`direction. A phase profile is imposed on an optical beam as it
`is either transmitted through or reflected from the phase shifter
`array. The imposed phase profile steers, focuses, fans out, or
`corrects phase aberrations on the beam. The array of optical
`phase shifters is realized through lithographic patterning of an
`electrical addressing network on the superstrate of a liquid crystal
`waveplate. Refractive index changes sufficiently large to realize
`full-wave differential phase shifts can be effected using low ( < 10
`V) voltages applied to the liquid crystal phase plate electrodes.
`High efficiency large-angle steering with phased arrays requires
`phase shifter spacing on the order of a wavelength or less; con(cid:173)
`sequently addressing issues make 1-D optical arrays much more
`practical than 2-D arrays. Orthogonal oriented 1-D phased arrays
`are used to deflect a beam in both dimensions. Optical phased
`arrays with apertures on the order of 4 em by 4 em have been
`fabricated for steering green, red, 1.06 f1m, and 10.6 {1m radiation.
`Steering efficiencies of about 60% at 4° and 85% at about 2°
`have been achieved to date with switching times as short as
`a few milliseconds in the visible. Fluences of several hundred
`W/cm 2 have been demonstrated at 10.6 11m with nonoptimally
`engineered devices. Higher fiuences can be handled at shorter
`wavelengths. Larger apertures are feasible, as is operation at
`other wavelengths and significantly faster switching times. System
`concepts that include a passive acquisition sensor as well as a
`laser radar are presented.
`
`Manuscript received June 30, 1995; revised November 14, 1995. Tills
`work was supported in part by Raytheon internal funds, and in part by
`the Air Force Wright Laboratory at Wright Patterson AFB, Dayton, OH.P.
`F. McManamon and E. A. Watson are with Wright Laboratory, Wright
`Patterson AFB, OH 45433 USA.
`D. L. Corkum is with Texas Instruments, Att1eborough, MA 02703
`USA.
`T. A. Dorschner, L. J. Friedman, D. S. Hobbs, M. Holz, D. P. Resler, and
`R. C. Sharp are with Raytheon Company, Electronic Systems, Lexington,
`MA 02173 USA.
`H. Q. Nguyen is with Kopin Corp., Taunton, MA 02173 USA.
`S. Liberman is with SemiTest, Billerica, MA 01821 USA.
`Publisher Item Identifier S 0018-9419(96)01390-4.
`
`I.
`
`INTRODUCTION
`Currently optical sensor systems, including laser radar,
`are often limited in performance and cost by mechanical
`beam directing and stabilization mechanisms. The requisite
`pointing and stabilization usually requires precise, rapid,
`mechanical motion, and is often associated ·with substan(cid:173)
`tial masses. Submicroradian steering precisions are often
`desired, but are usually impractical for available, afford(cid:173)
`able, mechanical beam directing systems. Most mechanical
`systems do not facilitate rapid random pointing. Further(cid:173)
`more, the rapid steering of a large aperture optical sensor
`often requires a prohibitive amount of power. Despite the
`considerable accumulated manufacturing experience in this
`field, mechanical beam steering for optical sensors remains
`complex, precise, and expensive.
`Optical phased arrays appear to have the potential to
`overcome many of the limitations of mechanical beam
`steering. Liquid-crystal-based phased arrays require very
`little prime power, even for large apertures, thereby opening .
`up application areas such as missile interceptors, satellite
`communications, and portable sensors of all types. Phased
`arrays are inherently random-access devices, a distinct
`advantage when regions of interest are distributed widely
`across a sensor field of regard (FOR). Unlike mechanical
`systems, liquid crystal devices are generally insensitive
`to accelerations, and their costs can drop rapidly with
`volume production, as is the general case for the electronic
`devices they resemble. Flat panel displays, fabricated using
`technologies that are similar to those required for liquid(cid:173)
`crystal optical phased arrays, are now inexpensive enough
`to be in every notebook computer.
`In the related microwave radar arena, phased arrays are
`rapidly displacing conventional horn antennas. The clear
`benefits of random-access, rapid beam pointing with no
`moving parts have made phased arrays the technology of
`choice, despite their high cost. Fortunately, cost trades for
`optical radars using optical phased arrays promise to be
`more favorable since the optical arrays are monolithically
`fabricated with no discrete elements, consist of an array
`of phase shifters rather than individual transmiUreceive
`
`PROCEEDINGS OF THE IEEE, VOL. 84, NO. 2, FEBRUARY 1996
`
`0018-9219/96$05.00 © 1996 IEEE
`
`268
`
`FINISAR 1009
`
`

`

`modules, and are designed to use low-cost addressing
`electronics. The optical phased arrays discussed here are
`passive arrays, consisting soley of phase shifters, and are
`operated as space-fed arrays, meaning that an already
`formed beam is fed to the array of phase shifters, which
`then effects steering of that beam. This contrasts to an active
`phased array in which individual transmit modules form a
`beam as it exits a large array of transmitters.
`There are many application areas that can benefit from
`the performance/cost benefits made possible by optical
`phased arrays. Inexpensive, reliable laser radar for target
`detection, wind profiling, and gas cloud identification are
`examples of high interest. Laser communication, whether
`effected with directed beams in free space or by switching
`of guided beams within fiber links, is another application
`area. Defense against infrared guided missiles benefits from
`directed laser energy, and is another potential optical phased
`array application area. Later in the paper issues associ(cid:173)
`ated with steering broadband optical energy are addressed.
`Passive infrared sensors for imaging or point detection
`applications can also benefit from phased array optical beam
`steering, but to a more limited degree at this time; however,
`future applications are expected to expand as techniques for
`reducing the influence of dispersion are developed.
`Optical beam steering by means of phased elements is
`a rich area heavily researched by prior workers. As early
`as 1971, Meyer [1] had developed a 1-D optical phased
`array using bulk, lithium tantalate phase shifters. The
`array comprised 46 phase shifters on one-half millimeter
`spacings. The number of addressable beam positions, beam
`widths, scan angles, and beam spacings all were shown to
`agree with theory as developed for microwave phased array
`antennas. Shortly thereafter Ninomiya [2] demonstrated a
`1-D array of lithium niobate electrooptic prism deflectors.
`The resolving power of the array was shown to be N
`times that of a single prism, where N is number of arrayed
`prisms. The array successfully demonstrated 50 resolvable
`spots with 600 V applied. Both discrete and continuous
`steering were demonstrated. The power required was noted
`to be similar to that for acousto-optic deflectors. Although
`these early phased arrays clearly demonstrated the concept,
`they were neither developed for high performance nor
`were intended for practical application. Large phase shifter
`spacings of hundreds of wavelengths were unavoidable,
`given the state of the technology, and precluded achieving
`efficient large angle beam steering. The small aperture fill
`factors also guaranteed large insertion losses. However,
`many of the key advantages of the phased array approach to
`beam steering were well appreciated by these early workers.
`Ninomiya pointed out that a phased array offers random
`access, that the resolving power of a phased array is high,
`that the steering angle is very accurate, and that there is no
`shift of optical frequency as with acousto-optic deflectors.
`Beam steering of visible light has recently been reported
`using a liquid crystal television panel as an elementary
`phased array [3]. Although liquid crystal displays are usu(cid:173)
`ally configured to effect intensity modulation, when the po(cid:173)
`larizers are removed the accompanying phase shift becomes
`
`observable. The display pixels are programmed to effect
`a discrete blazed-grating phase ramp across the aperture.
`However, the relatively large pixel spacing (several hundred
`waves), the nonunity array fill factor, and the limited
`available phase modulation depth ( 1.3?r) have severely
`limited the achievable steering efficiency and angle ( <0.1 °).
`Other workers in the field have attempted to develop
`higher performance optical phased arrays by greatly re(cid:173)
`ducing the phase shifter spacings. Vasey et al. [ 4] have
`developed an integrated optics approach comprising a 1-
`D phased array based on a linear array of closely spaced
`AlGaAs waveguides, the relative phases of which can be
`electrically adjusted using the electrooptic effect in the
`waveguiding material itself. Beams are coupled into the
`guided structure and launched into free space using grating
`couplers. Continuous steering is achieved by electrically
`imposing a linear phase ramp of adjustable slope across
`the aperture. Addressing is accomplished via a fine/coarse
`architecture, somewhat similar to the approach discussed
`in Section IV. Continuous steering of a 900 nm beam over
`a ±7.5 mrad field has been reported. Element spacings
`are orders of magnitude less than those in earlier bulk
`demonstrations, but remain multiple ( 13-14) wavelengths.
`Consequently, maximum steering angles are limited and
`efficiencies are low due to the large number of radiated
`diffraction orders (so-called grating lobes). Although the
`electrooptic effect used in this approach is inherently fast
`(ns), achieving the steering angles and efficiency levels
`required for laser radar is expected to be difficult, as is
`scaling to required aperture sizes and obtaining steering in
`two-dimensions.
`Another approach reported recently [5] uses a thin, 2-D
`array of liquid crystal phase shifting elements configured to
`operate as coherent microprisms with a relatively high fill
`factor. The individual elements are multiple wavelengths
`in extent and spacing, but are constructed to produce a
`linear phase ramp of adjustable slope across each element
`face, thereby simulating a discrete blazed grating with
`programmable blaze angle. The current device steers in
`one dimension only, although in principal two devices
`could be cascaded to steer in both dimensions. Unlike
`most preceding optical arrays, this device was specifically
`designed for laser radar application and is, in principle,
`capable of high efficiency at large angles (20°) and of being
`fabricated with large apertures. However, to date, only
`small apertures (2 mm square), moderate steering angles
`(5°), and low efficiencies (1 %-9%) have been achieved
`owing to fabrication difficulties inherent to the approach.
`One of the difficulties is that this approach requires using
`only the linear portion of the liquid crystal versus voltage
`curve, resulting in a limited use of available birefringence.
`A more classic approach to optical phased arrays has been
`under development by the authors. This work has resulted in
`development of a true optical phased array with 1-D phase
`shifter spacings smaller than a single free-space wave(cid:173)
`length, 100% aperture fill factors over significant apertures,
`and performance approaching theoretical predictions for
`small to moderate angles. The unity fill factor, small phase
`
`MCMANAMON et al.: OPTICAL PHASED ARRAY TECHNOLOGY
`
`269
`
`

`

`shifter spacings, and careful fabrication techniques used
`result in very low sidelobe levels. These are the first optical
`phased arrays which are capable of redirecting a single input
`beam into essentially a single, diffraction limited, output
`beam with negligible sidelobes. Using this approach, high
`performance optical phased array based steering of carbon
`dioxide laser beams (10.6 p,m) was first demonstrated in
`1989, with demonstrations of Nd:YAG steering (1.06 p,m)
`following soon thereafter [6], [7]. That work will soon
`appear in the open, reviewed, literature [8], [9].
`These new optical phased arrays are direct functional
`analogs of the well known microwave phased array anten(cid:173)
`nas [10] that make possible the agile, inertialess steering of
`microwave beams. The underlying fundamental concepts
`are identical to those for a microwave array. However,
`due to the orders-of-magnitude difference in wavelengths
`between the microwave and optical worlds, these new opti(cid:173)
`cal phased arrays have been implemented quite differently.
`The current optical devices are 1-D, space-fed, passive,
`phase-only, apertures. This differs from modern microwave
`phased arrays with which a beam is usually both formed
`and steered in two dimensions by a 2-D array of active
`elements. The field intensity across the aperture of an active
`microwave array is generally tapered at the edges in order
`to achieve low sidelobe levels. This is not an option with
`a passive, phase-only array. However, being space-fed, if
`the input optical beam is Gaussian in spatial profile, as is
`the usual case, additional tapering is not needed. The 1-D
`phase-only array steers an optical beam in one dimension
`only. Unlike modern microwave arrays, and most other
`optical phased arrays to date, these new arrays are designed
`to be easily cascaded. This allows simple mounting of
`orthogonal1-D arrays to steer the beam in two dimensions.
`Microwave arrays are built using discrete phase shifters,
`as have been most early optical phased arrays. However,
`since a vast number of phase shifters is needed to realize a
`high performance optical array, distributed liquid crystal
`phase shifters have been implemented, as described by
`Huignard et al. [11]; however, it has proven essential to
`implement additional innovative addressing means to avoid
`the otherwise impractical numbers of interconnects.
`The organization of this paper is as follows. In Section II
`liquid crystal optical phased array technology is summa(cid:173)
`rized. In Section III we briefly discuss alternative can(cid:173)
`didates to optical phased arrays for eliminating complex
`and expensive mechanical motion from laser radar optical
`systems. A more detailed description is presented in section
`IV. Section V summarizes performance levels achieved
`and predicted performance potential. Section VI considers
`the pointing of an acquisition sensor, often a passive in(cid:173)
`frared (IR) sensor. Section VII discusses laser radar system
`concepts that incorporate target acquisition and tracking
`capabilities. Section VIII contains conclusions.
`
`II. OVERVIEW OF LIQUID CRYSTAL OPTICAL
`PHASED ARRAY CONCEPTS
`A prism inserted into the aperture of an optical system
`introduces a linear gradient of optical path delay (OPD)
`
`across the aperture which tilts the phase front and thereby
`steers the optical beam. For a given wavelength a phase
`shift of 21r (corresponding to an OPD of one wavelength)
`can be subtracted periodically from the phase front without
`influencing the far-field pattern produced by the phase front
`[12]. The "folded" phase profile represents a blazed grating.
`The phase ramp, or its equivalent modulo-21r sawtooth
`phase profile, further can be approximated by a series of
`discrete phase steps, as long as the steps are small.
`Fig. 1 illustrates the use of nematic liquid crystal cells
`as phase shifters. With no applied fields, the liquid crystal
`molecules align with an average orientation parallel to
`the substrates, according to the liquid crystal alignment
`layer applied at the substrate interface. Application of a
`relatively low voltage, on the order of 1-10 V, reorients
`the liquid crystal molecules and changes the effective index
`of refraction as seen by light polarized along the direction
`of quiescent molecular orientation. The maximum phase
`shift available is proportional to the thickness of the liquid
`crystal layer. The case of a 21r phase retarder is illustrated.
`The switching speed of a nematic liquid crystal phase
`shifter is generally inversely proportional to the square of
`the thickness of the nematic liquid crystal layer [13]. For
`steering angle/aperture size combinations that require phase
`resets, the minimum thickness of the liquid crystal layer
`to produce efficient steering requires a liquid crystal layer
`sufficiently thick to produce a full wavelength of OPD and
`allow modulo 21r operation. Only a combination of very
`small angles, or very small aperture size, allows practical
`beam steering without the use of resets. The liquid crystal
`layer thickness, t, for a 21r phase shift is given by
`
`(1)
`
`where An= (ne- no) is the birefringence of the material
`and A is the free space wavelength. As an example, the ne(cid:173)
`matic liquid crystal E7 has a birefringenceof approximately
`0.2 in the visible and near infrared spectrum. It requires a 5
`p,m layer thickness to achieve a relative phase delay of 21r
`radians at a 1 p,m wavelength. If a reflective-mode design is
`used, allowing two passes through the liquid crystal layer,
`a full wave OPD is created using only half that thickness,
`or 2.5 p,m.
`The diffraction efficiency, TJ, of a grating' with a stair(cid:173)
`step blaze designed to maximize energy in the. first order
`is given by [23]:
`
`17
`
`(2)
`
`= (sin(n) q)) 2
`7r I q
`where q is the number of steps in the blaze profile. From
`(2) it can be seen that an eight-step approximation gives a
`theoretical efficiency of approximately 95%. Fig. 2 shows
`a step approximation to the wavefront deflected by a prism,
`including the 21r phase resets. Note that a 21r phase reset has
`that value only for the design wavelength. Fig. 3 shows the
`deviation from a straight· line in the unfolded phase profile
`when a wavelength other than the design wavelength is
`used. This variation in phase reset values causes dispersion
`
`270
`
`PROCEEDINGS OF THE IEEE, VOL. 84, NO. 2, FEBRUARY 1996
`
`

`

`Polarization )
`
`Optical Beam
`
`OFF
`
`~~~
`~
`~~~~~~~~~
`~~~
`~~~
`~~~
`~~~
`~48~
`~48~
`~~~
`~~~
`~48~
`~~~
`~48~
`~~~
`~48~
`~41D~
`
`~~ ~ 48 41D~
`
`= :c
`
`0
`
`.c
`=
`a.
`
`RMS Voltage
`
`5
`
`---------
`,,,,,,,,,
`---------
`,,,,,,,,,
`ON "' V 111111111
`.,.,.,.,,.,,.,.,
`--------(cid:173)
`
`~~~48~~~~
`
`Fig. 1. Nematic liquid crystal phase shifters. The liquid crystal molecules are birefringent. Light
`polarized along the long axis of the molecule will experience a different index of refraction than light
`polarized along the short axis of the crystal. The molecules will rotate when a voltage is applied,
`producing an effective index change for light polarized perpendicular to the long axis of the crystal.
`
`Unfolded Phase
`Profile
`
`Design
`Wavelength
`
`Steered
`Wavelength
`
`Undisturbed
`Phase Front
`
`Individual
`Phase Shifters
`
`Fig. 2. Optical phased array agile beam steering. The optical
`phase delay introduced by a prism in an aperture can be approx(cid:173)
`imated by a series of stair-step ramp phase delays. When a ramp
`has an optical path difference equal to or larger than the design
`wavelength one design wavelength of optical path difference is
`subtracted from the ramp. At the design wavelength, the phased
`array effectively reproduces the steering caused by a prism.
`
`[14], which will be discussed further in Section VI. As
`shown in Fig. 3, the unfolded OPD is in error by (>.->.d)
`after each reset, and the l,lnfolded phase is in error by
`21r ( >. - >.d)/>. after each reset, where >. is the actual
`wavelength and >.d is the design wavelength.
`Practical factors can cause the measured efficiency of
`an actual phased array beam steerer to deviate from the
`theoretical value given by (2). One such factor, evident in
`liquid crystal phased arrays currently being developed, is a
`spatial "ftyback" in the molecular orientation of the liquid
`crystals which results from the minimum spatial extent
`required to change from the orientation for a phase shift
`of 21r to that for a phase shift of zero. The actual ftyback
`transition is a complex function of device design and liquid
`crystal visco-mechanical properties. Fig. 4 depicts phase
`versus position for a simple ftyback model. As a result of
`
`Unfolded
`
`Fig. 3. Unfolded phase profile. This figure shows the influence
`on the unfolded phase profile of operation at a wavelength other
`than the design wavelength.
`
`ftyback, only a portion of the grating imposes the correct
`phase distribution to steer a beam in the design direction.
`That portion of the grating over which ftyback occurs can
`be thought of as steering the beam in a different direction.
`The resulting diffraction efficiency 'T/ into the desired grating
`order can be approximated by [15]
`
`(3)
`
`where Ap is the width of the ftyback region and A is the
`period of the programmed grating. The energy that is not
`directed into the desired grating order is distributed among
`numerous other grating orders, causing a loss in efficiency
`for the primary order. The overall steering efficiency is
`given by the product of (2) and (3). Depending on the
`grating period (which affects both the number of steps in
`the blaze profile as well as the relative size of the ftyback),
`
`MCMANAMON et al.: OPTICAL PHASED ARRAY TECHNOLOGY
`
`271
`
`

`

`operated modulo 21r, such adaptive optic elements would
`be dispersion limited to narrow band applications.
`
`III. BEAM STEERING APPROACHES· USING
`LIMITED MECHANICAL MOTION
`An optical phased array is not the only approach to
`realizing rapid beam steering without the use of con(cid:173)
`ventional mechanical systems. Some of the more viable
`alternate options are briefly reviewed here. All of the
`options discussed here potentially allow the redirection
`of the field-of-view of an optical sensor without the use
`of complex, costly, mechanical mechanisms. Unlike the
`optical phased array, most of these alternate options do
`not eliminate mechanical motion, but instead minimize the
`degree of mechanical motion required. To date, none. of
`these alternate approaches have demonstrated the scope of
`performance characteristics desired for laser radar and most
`other optical sensors.
`One such option is the use of cascaded microlens arrays
`[19], [20] an example of which is shown in Fig. 5. Each
`microlens array consists of a (generatly) close packed,
`periodic array of miniature lenses which can be fabricated
`in either diffractive or refractive forms. Beam steering is
`effected by translating one microlens array with respect
`to the other. The concept can be understood by first
`considering a single microlens pair from a set of aligned
`afocal arrays. A collimated input beam is focused to the
`back focal point of the first microlens, which is also the
`front focal point of the second microlens, resulting in
`an unsteered, collimated output beam. However, if the
`second microlens is offset, then the back focal point of the
`first microlens appears as an off-axis point to the second
`microlens. The point remains in the front focal plane of the
`second microlens, so the second microlens still recollimates
`the light, but the beam is redirected to a nonzero field
`angle. A paraxial ray trace shows that the tangent of this
`field angle is equal to the amount of •offset divided by
`the focal length of the second microlens. Maximum useful
`steering occurs with an offset equal to the radius of a
`microlens. It may be noted that it does not matter if the
`second microlens has a positive or negative focal length
`so long as the condition of overlapping focal planes is
`met. If the individual microlenses of the arrays are aligned,
`periodically spaced, and designed to fill the aperture, the
`output beam replicates the input beam. If the offset is small,
`the steered beam approximates a simple· redirection of the
`input beam.
`However, if the offset is large, significant fractions of the
`input beam are coupled into other grating modes. This can
`be appreciated by noting that phased arrays and microlens
`arrays both approximate blazed gratings [21]. If the periodic
`quadratic phase profiles of two offset • micro lens arrays
`are superimposed, the result is a (generally asymmetric)
`triangular waveform, which approximates a blazed grating.
`If the composite phase profile were a sawtooth, the approx(cid:173)
`imation would be exact. Motion of the lenses alters the
`slope(s) of the phase profile, thereby changing the blaze
`
`Duty Cycle, A
`
`'-y-'
`Fly
`Back
`Region
`AF
`
`Fig. 4. Flyback. When one design wavelength of optical path
`difference is subtracted it requires finite spatial extent. This region
`is referred to as the flyback region. The steering efficiency into a
`given order is influenced by the relative size of the flyback region
`with respect to the grating period.
`
`either (2) or (3) may dominate the overall steering efficiency
`of an optical phased array beam steerer.
`For a normally incident input beam the steered angle is
`given by [16]
`
`(4)
`
`.
`Ao
`()
`sm =A
`where A0 is the design wavelength for the beam steerer,
`A = qd is the period of the staircase ramp, q is the number
`of phase shifters between resets, and d is the center-to(cid:173)
`center spacing between phase shifters, which is assumed
`to equal the width of the phase shifter as well. Large
`steering angles correspond to high spatial frequencies (small
`periods) and vice versa. From (3) and (4) it can be seen
`that the steering efficiency decreases monotonically with
`steering angle, for fixed flyback.
`Two-dimensional beam steering can be achieved using
`two orthogonally oriented 1-D liquid crystal phase gratings.
`In addition, any optical distortion that is separable in
`Cartesian coordinates can be fully compensated, modulo
`27r. Spherical aberrations can be fully compensated with a
`crossed grating system. For a full adaptive optics capability,
`a third layer, with a 2-D array of phase shifters, would be
`required. This would add the ability to clean up an arbitrar(cid:173)
`ily aberrated beam and adapt for atmospheric turbulence.
`Such a liquid crystal adaptive optics layer has recently
`been discussed [17]. The spacing of elements on such an
`adaptive optics layer would be orders of magnitude courser
`than the spacing required for large angle beam steering.
`Current adaptive optics mirror systems have on the order
`of 50-400 elements correcting for turbulence while using
`apertures up to a few meters [18]. However, the adaptive
`optic element is usually used prior to final beam expansion
`and is much smaller in aperture. Pixelated phase shifters
`of about 1 mm square would probably suffice for most
`applications and could be readily fabricated with the current
`technology. Thus liquid crystal phase shifter technology
`could replace the current piezoelectrically driven adaptive
`optic components, resulting in a single three-layer compo(cid:173)
`nent that both deflects and phase compensates a beam. If
`
`272
`
`PROCEEDINGS OF THE IEEE, VOL. 84, NO. 2, FEBRUARY 1996
`
`

`

`profile of the equivalent grating, and shifting the light to
`different grating orders. To the extent that the composite
`phase profile approximates a true sawtooth, light is steered
`to a single direction. However, the offset of two lens arrays
`inherently causes each input lenslet to illuminate adjacent
`output lenslets, resulting in the multiple-slope profile of
`the triangular wave, and steering to multiple directions. To
`mitigate this effect, designs using a third microlens array
`as a field lens have been put forth, but demonstrations have
`not yet been reported.
`The agile steering of a beam using the microlens array
`concept requires the agile motion of one microlens array
`with respect to the other. Microlens arrays inherently have
`small focal lengths (typically on the order of a few lenslet
`diameters, usually a millimeter or less); consequently, the
`amount of mechanical offset required to achieve a desired
`steering angle can also be quite small. Compared to steering
`via a displaced bulk lens having the same aperture as
`the microlens array, the reduction of motion required to
`steer to a given angle is proportional to the ratio of the
`individual microlens diameter to the array diameter. Due
`to its essentially planar structure, a microlens array can be
`made much lighter than a bulk lens of equivalent aperture.
`The combination of low mass and small motion allows agile
`positioning (and agile beam steering) to be accomplished
`with more simplified mechanical drivers than would be
`required for macroscopic lenses. The microlens arrays can
`be designed to effect substantial steering with mechanical
`motions that can be achieved with piezoelectric transducers.
`However, small errors in mechanical positioning are ampli(cid:173)
`fied by the same optical leverage that makes possible the
`reduction in mechanical motion. This means the amount of
`energy at the desired steering angle will be influenced by
`a small amount of mechanical motion. Thus fine angular
`beam steering with this approach generally requires very
`precise motion.
`Microlens arrays can be programmed onto the liquid(cid:173)
`crystal based optical phased arrays reported here, thereby
`making possible an electronic translation of one lens array
`with respect to the other, and complete elimination of
`all mechanical motion. Since only one microlens array
`must move to achieve beam steering, only a single array
`would have to be programmable. Owing to the precise
`displacement control available with an optical phased array,
`this option may be preferable to piezoelectrically driven
`motion for applications requiring precision pointing.
`Flexure beam micromirror technology is another ap(cid:173)
`proach with large numbers of small apertures arranged in
`regular arrays [22]. The individual apertures are lithographi(cid:173)
`cally fabricated mirror "pixels" on hinges with micromotion
`effected by an electrostatic field. The field attracts the ele(cid:173)
`ment and moves it rapidly, on the order of a microsecond.
`This can create a piston phase shift for the individual
`aperture. These devices have demonstrated 211' phase shifts
`at 633 nm wavelength with a 60-75% fill factor. Much of
`the same phased array theory discussed later in this paper
`applies to these array structures, although the physical im(cid:173)
`plementation is significantly different. The implementation
`
`Fig. 5. Decentered micro/ens array beam steering. The figure
`shows two decentered microlens arrays, and their influence in
`steering an incoming beam. If one array is moved with respect
`to the other array it causes the beam to steer.
`
`of this approach makes individual apertures with sizes on
`the order of visible or near infrared wavelengths imprac(cid:173)
`tical, thus limiting maximum steering angles. A second
`limitation is the nonunity fill factor. Hidden hinge concepts
`are being considered to address the fill factor issue [23].
`Different approaches are represented by electro-optical
`(EO) and acoustic-optical (AO) beam deflection [24]. There
`are no moving parts with either approach. EO deflec(cid:173)
`tion can occur in nanoseconds, while AO beam deflection
`is generally effected on the order of microseconds, the
`time for an acoustic wave to propagate across the crystal
`aperture. AO beam steering requires high drive power
`at longer wavelengths, has an aperture size limited by
`available crystal dimensions, and shifts the frequency of the
`transmitted, or receive, beam. EO beam deflectors require
`high voltages for large apertures, typically of the order
`of 10 kV/cm. Beam deflection angles for both EO and
`AO deflectors are typically limited by practical issues to a
`few milliradians; consequently, these devices are generally
`restricted in applications to fast, fine beam steering of small
`b

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket