throbber
Surface-Micromachined Microoptical
`Elements and Systems
`
`RICHARD S. MULLER, LIFE FELLOW, IEEE AND KAM Y. LAU, FELLOW, IEEE
`
`Invited Paper
`
`Optical systems are ubiquitous in the present-day societal fabric,
`from sophisticated fiber-optic telecommunication infrastructure to
`visual information display, down to mundane chores such as bar-
`code reading at the supermarket. Most of these existing systems
`are built from bulk optical components, as they have been for
`many years. Just as miniaturization and batch-process production
`have revolutionized electronics, similar advances in optics will
`certain greatly expand its applications and markets. Production
`techniques for optical systems that employ the emerging micro-
`electromechanical systems (MEMS) technologies give promise of
`achieving this success. Simple micromechanical fabrication tech-
`niques are already employed in fiber-optic components to produce
`what is generally described as silicon-optical-bench systems. New
`developments, especially those permitting the use of microactuated
`structures, make substantial increases in system sophistication pos-
`sible. Surface micromachining, in which microoptical systems are
`batch-fabricated and placed on top of a silicon wafer, has become
`a promising approach to this progression. With the demonstration
`that surface-micromachined elements can be “folded” out from
`the plane in which they are constructed, an important new degree
`of design freedom has emerged. This paper examines some of the
`results obtained and attempts to project possibilities for surface
`micromachining in future optical systems.
`Keywords—Actuated micromirrors, bar-code reader, fiber-optic
`components, gratings, laser-fiber coupler, optical scanners.
`
`I.
`
`INTRODUCTION
`The continuing spectacular growth in optical system
`applications, as exemplified by optical-fiber telecommuni-
`cation infrastructure and display technologies, has stim-
`ulated great interest in producing miniaturized, reliable,
`inexpensive photonic devices for light-beam manipulation.
`There is strong desire for such devices to function without
`
`Manuscript received February 25, 1998; revised April 21, 1998. This
`work was supported in part by a grant from the Hewlett-Packard Science
`Centers program, in part by the Defense Advanced Research Project
`Agency, in part by the National Science Foundation, and in part by the
`Berkeley Sensor & Actuator Center membership.
`The authors are with Berkeley Sensor & Actuator Center, Department
`of Electrical Engineering and Computer Science, University of California,
`Berkeley, CA 94720-1770 USA.
`Publisher Item Identifier S 0018-9219(98)05094-4.
`
`any mechanical moving parts, out of consideration that
`these traditionally contribute to high cost and unreliability.
`For several decades, the only available control options have
`been those based on electro- or magnetooptic effects. Such
`devices, however, generally suffer from high costs and
`typically operate at low efficiencies. In most cases, optical-
`beam manipulation can be more effectively carried out with
`movable mechanical elements such as mirrors and shutters,
`if only these mechanical parts can be built reliably and
`inexpensively.
`Recent developments in the rapidly emerging discipline
`of microelectromechanical systems (MEMS) show special
`promise for providing microoptical systems to perform the
`functions described above. In particular, the technique of
`surface micromachining is interesting in that it is a planar
`process that is capable of producing large-area, high-quality
`layered structures along the plane of the substrate, which
`can then be rotated out of the surface to form large optical
`surfaces angled to the surface of the substrate. Such large
`elevated surfaces (up to millimeters in height) with high
`optical quality, and which are movable with high precision,
`are very difficult to fabricate with other micromachining
`technologies.
`At the Berkeley Sensor & Actuator Center (BSAC), a
`background in the development of surface micromachining
`has provided impetus for the development of actuated
`micromirrors for fiber-optic systems. The basis of our
`approach in microoptical systems rests on the following
`developments: methods for making pin joints and moving-
`element mechanisms [1], comb drives for oscillating sys-
`tems [2], microvibromotors as actuation elements [3], and
`folding of surface structures out of the surface plane to form
`truly three-dimensional structures [4]. Other groups are also
`active in applying surface micromachining to photonics,
`demonstrating innovative ways to use this technology [5],
`[6]. Rather than attempting to capture the full extent of the
`expanding efforts by all groups working in this field, we
`focus in this paper on examples drawn from our work at
`BSAC.
`
`PROCEEDINGS OF THE IEEE, VOL. 86, NO. 8, AUGUST 1998
`
`1705
`
`0018–9219/98$10.00 ª
`
`1998 IEEE
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-1
`
`

`
`micromirrors for optical-alignment purposes will need both
`precision and range—involving mechanisms reminiscent of
`stepping motors. Here, we describe a specific type of motor
`that has been successfully employed to drive micromirrors
`with submicrometer precision and with a range of hundreds
`of micrometers.
`As shown in Fig. 2, each of the two sliders is actuated
`with an integrated microvibromotor [6], which consists
`of four electrostatic-comb resonators with attached impact
`arms driving a slider through oblique impact. (The mi-
`cromirror shown in Fig. 2 is similar to that shown in Fig. 1
`except that the front sliding plate is folded under the back
`support to save space.) Fig. 3 shows a detailed top view of
`the vibromotor.
`
`C. Fabrication
`The actuated microreflector was fabricated on a silicon
`substrate using silicon-surface-micromachining technology.
`The fabrication process is described in detail in [12]. An n
`polysilicon layer defines a ground plane. Three additional
`polysilicon structural layers are used to define the comb
`drives, sliders, hinges, and mirror beams. Phosphorous-
`doped silicon dioxide is used for sacrificial spacer layers
`between polysilicon layers. A special prerelease etch in
`5 : 1 hydrofluoric acid (HF) followed by a vigorous rinse
`is used to eliminate stringers. Then the structure is released
`for 10 min in concentrated HF to dissolve the oxide and
`dried using a critical-point CO drier [13] to avoid stiction.
`A 400 ˚A gold layer is evaporated onto the mirror surface
`to increase reflectivity.
`
`D. Actuated Micromirror Characterization
`To balance the forces, two opposing impacters are used
`for each direction of travel. The resonator is a capacitively
`driven mass anchored to the substrate through a folded-
`beam flexure. The spring constant of the flexure determines
`the resonant frequency and travel range of the resonator.
`The force exerted by the comb drive is proportional to the
`square of the applied voltage
`
`(1)
`
`typically has both dc and ac components. This
`where
`quadratic response produces a primary frequency driving
`term proportional to the product of the dc and ac voltages,
`effectively linearizing the resonator and increasing the im-
`pact force. The comb structures are driven at their resonance
`frequency (roughly 7.5–8.5 kHz),
`thereby achieving an
`amplification of the electrostatic force by the resonator
`quality factor (typically 30–100 in air [14], [15]). Since
`energy is transferred to the slider during impact (typically
`lasting only a few microseconds), the impacters can deliver
`short-duration forces that are large enough to overcome
`static friction in the sliders and hinges. Due to the damping
`(primarily due to air resistance [12]) in the comb structure,
`the resonators require a few initial cycles to build up
`sufficient amplitude and momentum for impact. Therefore,
`in air, slider motion is observed only after three or more
`
`Fig. 1. Basic micromirror structure for precision alignment of
`optical components. The size of the mirror measures approximately
`200  250 m.
`
`II. OPTICAL MEMS FOR PRECISION ALIGNMENT
`
`A. Folded-Micromirror Structures
`Optical MEMS designed for precision alignment are
`intended to facilitate automated packaging of optoelec-
`tronic devices and subsystems, hence substantially lowering
`the overall production cost of the modules. Examples of
`these modules include fiber-pigtailed laser transmitters and
`external-cavity continuously tunable laser diodes. These
`modules are expensive because they require submicrometer
`alignment
`tolerances that place tight constraints on the
`positional accuracy of such optoelectronics components as
`lasers, lenses, gratings, and fibers. Silicon-optical-bench
`(SOB) technology, commonly used to align optical sys-
`tems on a silicon substrate using etched v-grooves, solder
`bumps, and other integrated circuit (IC)-process-derived
`techniques to achieve
`1 m transverse alignment [7], [8],
`is insufficient for high-performance modules. Instead, we
`promote a paradigm, in which SOB technology—a hands-
`off automated process—is used for initial placement of the
`various optical components in the module, followed by an
`automated active alignment procedure using micromirrors
`that are prefabricated on the silicon substrate using MEMS
`technology [9], [10].
`Details of a movable micromirror is shown in Fig. 1. It
`consists of four polysilicon plates, interconnected by three
`sets of microhinges [11]. The two end plates can slide
`linearly and independently on the silicon-substrate surface,
`confined by hubs on the sides. This arrangement provides
`the mirror with rotational and translational freedom of
`motion and a high vertical aspect ratio in its operating
`position.
`
`B. Vibromotor Actuation
`On-chip actuation is necessary in order for the mi-
`cromachined components to function in a self-contained
`optical module. There exist numerous ways to achieve on-
`chip actuation of microstructures; however, actuation of
`
`1706
`
`PROCEEDINGS OF THE IEEE, VOL. 86, NO. 8, AUGUST 1998
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-2
`
`

`
`Fig. 2. Complete self-actuated micromirror. Two sets of vibromotors, comprising four comb drives
`each, drive the front and rear sliders for actuation of the mirror.
`
`m [12]. To ascertain the performance of the micromirror
`assembly driven by vibromotors, optical characterization
`was performed on the integrated microreflector system. A
`HeNe laser is reflected from the micromirror surface onto
`a charge-coupled device (CCD) camera. As the front and
`rear sliders are actuated, the beam position along each of
`the two axes is measured on the CCD and extrapolated to
`a location 200 m in front of the mirror (where a fiber
`would typically be positioned for a laser-to-fiber coupling
`application). The standard deviation of the vertical beam
`position data from a linear response depends on the selected
`slider step size. For the front slider, this deviation is roughly
`equal to the step size, while for the rear slider, the deviation
`exceeds the step size by about 50–60%. The selected
`average step sizes of 0.35 m for the front slider and
`0.42 m for the rear slider produce standard deviations of
`0.32 and 0.60 m, respectively [Fig. 4(a)]. These deviations
`are due primarily to the “play” in the hinges and the
`wobble in the slider structure. The greater length of the
`rear slider results in increased wobble and leads to a
`greater standard deviation. The horizontal beam deviation
`is 0.05 m [Fig. 4(b)] and is comparable to the 0.07
`m deviation measured in externally actuated structures.
`This precision is sufficient for laser to single-mode fiber-
`coupling applications where, due to lens magnification,
`1 m
`the beam only needs to be within approximately
`for high coupling. However, for other advanced optical
`systems such as tunable external cavity lasers, a higher
`angular precision is necessary. Since in earlier experiments
`a microreflector with no on-chip actuators and an alternate
`hub design has shown a vertical precision of 0.17 m [12],
`improving the design of the actuated slider should greatly
`improve the mirror precision.
`
`Fig. 3. Top view of the linear vibromotor.
`
`voltage cycles. The number of initiation cycles depends on
`the ambient atmosphere and decreases at lower pressures.
`When driven with a free-running resonant oscillation, the
`slider reaches a maximum velocity of over 1 mm/s. Slider
`velocity can be controlled by driving the comb drives
`with gated bursts of four to five cycles of the resonant
`waveform. Once the slider is in position, it is kept in place
`by static friction until further actuation. Similar structures
`were previously subjected to shock and vibration tests and
`showed no detectable slider motion at forces up to 500 G’s
`[12].
`Characterization of the vibromotor alone has shown the
`slider motion to have a step resolution of less than 0.3
`
`MULLER AND LAU: MICROOPTICAL ELEMENTS AND SYSTEMS
`
`1707
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-3
`
`

`
`(a)
`
`(b)
`
`Fig. 4.
`(a) Vertical and (b) horizontal microreflector precision. The vertical position data show
`standard deviations of 0.32 and 0.6 m (for the front and rear sliders, respectively) from a linear
`response. The horizontal beam position has an average 0.05 m of in-plane wobble.
`
`To observe the actuation dynamics of the micromirror,
`a reflected HeNe beam was imaged on a position-sensitive
`detector while the angular position of the mirror was swept
`in real time. The rear vibromotor was biased at 40 V dc
`and driven with 20 V (p-p) resonant (8.2 kHz) square-
`wave pulses. The position of the beam and the drive voltage
`were then monitored on a digital oscilloscope. The resulting
`trace is shown in Fig. 5. As predicted and observed [12],
`the comb structures require a few cycles (three in this
`
`case) to build up sufficient energy for impact. The first
`significant impact occurs during the third cycle, at which
`point the mirror angle begins to change. The mirror is
`further deflected during the fourth cycle; however, sufficient
`energy is lost by the resonator to keep the next impact
`from producing significant displacement. Finally, the sixth
`cycle provides the greatest impact, resulting in a total mirror
`rotation of 0.3 . The small back-and-forth motion apparent
`in the response is most likely due to slight deformation
`
`1708
`
`PROCEEDINGS OF THE IEEE, VOL. 86, NO. 8, AUGUST 1998
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-4
`
`

`
`Fig. 5. Time-domain response of the microreflector to a square-wave drive voltage. An impact is
`first achieved during the third cycle, at which point the mirror angle begins to change. The motion
`concludes shortly after the last cycle, resulting in a total angular displacement of about 0.3.
`
`maximum estimated scan rate of 10.2 rad/s or a beam speed
`of 2 mm/s on a plane 200 m in front of the reflector.
`
`E. A Single-Mode-Fiber Laser Transmitter Module
`The actuated microreflector system is used in a laser-
`to-fiber coupling module, as illustrated schematically in
`Fig. 7(a). A picture of the assembled coupling module is
`shown in Fig. 7(b). An aspheric microlens was used to
`image the output of a laser diode into a single-mode fiber. A
`polysilicon micromirror with two degrees of freedom (linear
`displacement and angular position) reflects the laser beam
`at a 45 angle and provides the fine alignment. Lens, laser,
`and fiber are positioned passively on the substrate using
`etched alignment grooves and photolithographically defined
`alignment aids and held in place by low-viscosity epoxy.
`The height of the fiber core is controlled by mounting it
`in a silicon subcarrier that is attached to the foundation
`substrate. Using these techniques, the axial displacement
`and tilt of the optical components can be minimized, and
`the reduction in coupling efficiency resulting from these
`sources of misalignment is negligible. Transverse misalign-
`ment of the laser and lens, however, has a significant effect
`on the coupling, particularly because of the magnification of
`the lens system. As a consequence of our simple passive-
`alignment method, the transverse offsets of the laser and
`lens are on the order of 5–10 m, which, magnified by
`the lens, lead to roughly a 20–40 m displacement of the
`beam on the fiber plane. Therefore, the micromirror must
`has sufficient travel range to compensate for these offsets.
`This coupling module is used to couple light from a stan-
`dard telecommunications-grade 1.3 m distributed feed-
`back laser into a 9 m core single-mode fiber. The actuated
`microreflector was positioned between the lens and the fiber
`
`Fig. 6. Beam scanning with the actuated microreflector. In (a), a
`40-V dc offset was used, resulting in 0.7–1.1 m step size on the
`plane 200 m from the micromirror, while in (b), a 38-V dc offset
`produces 0.3–0.6 m steps.
`
`of the hinge joints during impact as the square pin is
`forced against the staple. The remaining “roughness” of
`the response is due to noise in the detector.
`To demonstrate the use of the actuated microreflector in
`scanning and beam-positioning applications, the mirror was
`used to scan a laser beam across the detector continuously.
`The rear vibromotor was driven with bursts of four 20 V
`(p-p) resonant (8.2 kHz) square-wave cycles spaced 10 ms
`apart, with a 40- or 38-V dc offset. The resulting output
`(Fig. 6) clearly shows the step-wise nature of the mirror
`motion. The average speed of the sweep can be changed
`by varying the spacing of the bursts. Fig. 6(a) and (b)
`also demonstrates that the size of the step itself depends
`on the applied voltage and can be controlled. With an
`average angular step size of 5 mrad, the microreflector has a
`
`MULLER AND LAU: MICROOPTICAL ELEMENTS AND SYSTEMS
`
`1709
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-5
`
`

`
`(a)
`
`(b)
`
`Fig. 7.
`(a) Schematic of a laser-to-fiber coupling module showing
`two-dimensional (2-D) optical alignment with two one-dimensional
`(1-D) translational degrees of freedom in the micromirror. (b)
`Scanning electron microscope (SEM) photograph of the coupling
`chip.
`
`to provide fine alignment. Fig. 8 shows the dependence of
`the coupling efficiency on the number of rear vibromotor
`steps. The step size can be varied by changing the drive
`voltage to provide various degrees of control. The data
`show a good fit
`to a Gaussian curve, with deviations
`being due to slider wobble and angular misalignment of
`the components. The coupling efficiency was limited to
`32% in this arrangement due to the “soft-focusing” optical
`arrangement; a tighter focus at the fiber plane will produce
`a higher coupling efficiency but is less forgiving in beam
`misalignment. As seen from Fig. 8, the step size in the
`actuated micromirror is sufficiently small that it can actually
`handle a much tighter focus than what is being shown. Thus,
`a much higher coupling efficiency, in the 60–80% range,
`can readily be achieved with a more aggressive optical
`design.
`
`F. Shock and Vibration Tests
`The micromirror structures are constructed from very thin
`polysilicon plates (typically 2 m thick) and have very
`low inertial masses. As a consequence, the structure can
`be maintained in position purely by frictional and stictional
`
`Fig. 8. Coupling efficiency from semiconductor laser to sin-
`gle-mode fiber using the actuated micromirror. A peak efficiency
`of 32% was achieved.
`
`forces alone. For illustration, a raised micromirror structure
`like the one shown in Fig. 1, without fixation by adhesives
`or any other means, was subjected to a random-vibration
`test (MIL-SPEC 883C) on all three axes, as well as to a 500
`G drop test. To the extent that can be determined visually
`(at a resolution of approximately 1 m), no movement
`of the structure can be detected after all of these tests.
`When the structure is subjected to a 1000 G drop test, the
`structure moved by several micrometers, but no breakage
`was observed on any of the specimens.
`In a separate experiment, a raised micromirror was
`mounted on a variable-frequency vibrating platform and
`was subjected to vibration from 5 to 100 kHz on three
`axes consecutively. The motion of the micromirror was
`monitored by optical
`interferometry through an optical
`fiber rigidly attached to the vibrating platform. There, the
`cleaved fiber facet and the micromirror form a Fabry–Perot
`interferometer and a relative motion between the fiber and
`the micromirror results in variations in the intensity of the
`reflected light. The vibrating platform is piezoelectrically
`driven, and the vibration amplitude is calibrated separately
`using a similar optical
`interferometry technique. The
`results are shown in Fig. 9. The frequency response of
`the micromirror structure can be seen to be rather complex
`and to consist of multiple resonances. This is expected
`from the relatively complex mechanical structure, which
`consists of multiple plates of different sizes interconnected
`by hinges. One conclusion that can be drawn from these
`measurements is that
`the resonance frequency in any
`excitation direction occurs well above 25 kHz, a frequency
`range beyond that of common mechanical or microphonic
`disturbances. These micromechanical devices are thus
`expected to be mechanically robust,
`in contrast
`to the
`strength of bulk-mechanical devices.
`On the other hand, the small inertial mass that contributes
`to the robustness of the microdevices also makes them
`vulnerable to disturbances from air currents. No quantitative
`tests have yet been performed on the effect of air currents,
`
`1710
`
`PROCEEDINGS OF THE IEEE, VOL. 86, NO. 8, AUGUST 1998
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-6
`
`

`
`other designs are possible and have been demonstrated
`elsewhere. For example, an electrostatically driven, linear-
`translating microactuator known sometimes as “scratch
`drive” has been demonstrated by Akiyama [16]. Scratch-
`drive microactuators employ polysilicon plates that are
`electrostatically attracted to an electrode on the silicon
`surface. The plates are made with vertical lips that impact
`the surface and are pushed forward when the plate flexes
`under the strain induced by the attractive force. When
`the force is released, the plate tends to “walk” forward.
`Hence, when a series of electrodes are laid out on the
`surface, sequentially applying bias to the electrodes makes
`it possible to walk the plate forward in a controlled fashion.
`These “scratch-drive actuators” have been demonstrated in
`optical applications at the University of California, Los
`Angeles, as effective movers for surface-micromachined
`foldout structures, including a frame to hold a separately
`machined microlens [17].
`
`III. MICROSCANNERS
`Optical scanners have long been used for scientific and
`industrial applications ranging from laser imaging and
`displays to laser surgical
`tools and appliances such as
`facsimile machines and printers. Perhaps their most familiar
`application is to bar-code scanning, a proliferating product
`area that contributes strongly to continuing progress in
`cost reduction. Production of a scanning system for bar-
`code reading or perhaps for a head-mounted raster-scanning
`display by MEMS techniques gives promise not only of cost
`but also of power reduction, as well as of portability and
`shrinkage in size when compared to present designs. The
`eventual integration of activation and detection circuits to
`produce an integrated scanner-chip set (eventually, perhaps,
`a fully integrated MEMS scanner chip) is a motivating
`longer term goal.
`MEMS microscanners built with rotating mirrors have
`been demonstrated [18], [19]. The speed, size, and deflec-
`tion angles of existing micromirrors are, however, severely
`limited by their planar structures and actuation mechanisms.
`For optical-scanning applications, the scanned-image reso-
`lution is limited by diffraction from the smallest optical
`aperture (in many cases the mirror) in the scanner. As a
`result, mirrors having lateral dimensions measured in the
`hundreds of micrometers are required to build practical
`systems. To create high-aspect-ratio optical surfaces us-
`ing processing technologies derived from IC procedures,
`polysilicon microhinges are incorporated in the scanner
`structures. These hinges allow relatively large mirrors to
`be lifted out of the plane of the substrate after planar
`processing is completed.
`Using silicon-micromachining and SOB technologies, ex-
`tremely compact optical systems incorporating low-inertia
`scanners can be built. As an example, shown in Fig. 10 is
`a sketch of such a raster-scanner microdisplay module fab-
`ricated on a chip. The light from three semiconductor light
`emitting diodes or lasers is collimated by a microlens and
`directed onto a pair of microscanners, each consisting of a
`
`(a)
`
`(b)
`
`(c)
`
`Fig. 9. Vibration testing of the micromirror structure on three
`axes. The dotted curves shows the oscillation amplitude of the
`vibration table; the dark curves show the motion recorded at the
`mirror surface.
`
`but it is clear that some form of protective packaging,
`perhaps in partial vacuum, is necessary for field deployment
`of these devices. Packaging is currently being studied.
`
`G. Alternate Means of Actuation
`Although the linear vibromotor was employed as the
`basic workhorse for our microphotonic alignment devices,
`
`MULLER AND LAU: MICROOPTICAL ELEMENTS AND SYSTEMS
`
`1711
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-7
`
`

`
`Fig. 10. A 2-D head-mounted raster-scanning design using two orthogonal scanning micromirrors
`on a chip.
`
`micromirror and an actuator. Both mirrors stand vertically
`on the silicon substrate and are actuated by electrostatic
`comb drives. A planar image display is achieved by raster
`scanning the light beam in two orthogonal directions using
`the two mirrors. These out-of-plane mirrors interact with
`optical beams that propagate in a plane parallel to the
`substrate onto which other microoptical components can
`be integrated on the same chip. An integrated module like
`the one sketched in Fig. 10 can be packaged in vacuum or
`otherwise sealed to reduce air damping, mirror deformation,
`and particulate erosion of the mirror surface. Also, the
`mechanical Q-factors of the scanners increase significantly
`even under moderate vacuum. Furthermore, the mounting
`and partial alignment and assembling of the mirrors are
`carried out during the fabrication process, and large quanti-
`ties can be produced inexpensively because they are batch
`fabricated.
`
`A. Resonant Scanner Design
`In order that high-functionality optical systems such
`as the display module described above be realized, one
`needs to address basic issues concerning the optical and
`mechanical performances of microscanners. We consider
`resonant scanners, since they can be driven at high fre-
`quency at large scan angles, with small power consumption.
`Fig. 11(a) is an SEM micrograph showing one of the
`resonant microscanners we have fabricated [20]. The size
`of the mirror is 300
`400 m, and the rotation hinges
`are square torsion bars, each measuring 50 m in length
`and 4 m in cross section. Suspended by a frame that is
`connected to a hinged slider at its back, the micromirror is
`inclined 70 to the substrate. Electrostatic comb drives are
`used as the actuators for their low power consumption and
`high resonant frequencies. The electrostatic comb drive for
`this scanner system consists of 100 interdigitated fingers on
`
`(a)
`
`(b)
`
`Fig. 11. SEM photographs of two prototype resonant scanners.
`An electrostatic comb drive is attached to the bottom of the
`scanning micromirror through cross-weaving hinges, while torsion
`bars are used as the suspension and the rotation axis for the
`mirror. (a) A 300  400 m -scanning mirror, which deflects
`the laser beam in the out-of-plane direction. (b) A 440  300 m
`-scanning mirror that scans the beam in a direction parallel to
`the substrate.
`
`1712
`
`PROCEEDINGS OF THE IEEE, VOL. 86, NO. 8, AUGUST 1998
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-8
`
`

`
`Table 1
`
`the shuttle comb and 101 fingers on the stationary comb.
`The comb fingers are 2 m wide and 40 m long, and
`the entire comb extends 1 mm from side to side. The
`comb drive is hinged to the bottom of the mirror through a
`bar-and-folded-spring assemblage. The folded springs have
`beams that are 2 m wide and 300 m long. The maximum
`excursion of the shuttle comb is limited to
`15 m by
`the lengths of the comb fingers. Fig. 11(b) is an SEM of
`another microscanner constructed differently with a mirror
`measuring 440 by 300 m. The mirror in Fig. 11(b) stands
`upright on the wafer, and its rotational axis is normal to the
`substrate. The maximum scan angle of the microscanner
`is determined by the relative position of the rotational
`axis and the allowed range of excursion for the shuttle
`comb. Therefore, the maximum scan angle can be increased
`simply by: 1) rearranging the position of the rotational axis
`and/or 2) increasing the range of the shuttle-comb excursion
`(lengthening the comb fingers), while the rest of the scanner
`structure remains unchanged.
`The hinges that suspend the mirror and the hinge that
`connects the mirror to the comb drive are constantly flexing
`during operation of the scanner, and the angular scanning
`precision of the mirror depends on the mechanical accuracy
`of these hinges. We have investigated pin-and-staple hinges
`similar to those described in previous sections, as well as
`torsional hinges, and found significant benefit in employing
`the latter for this application. These results will be described
`in detail below.
`
`B. Microscanners Characterization and Comparison
`We have fabricated microscanners of three different de-
`signs. Aside from the two scanning mirrors with torsion-bar
`hinges shown in Fig. 11, another scanner with a smaller
`mirror (200 by 250 m) and staple-and-pin hinges (Fig. 12)
`was also used in the experiments. The dimensions and
`relevant parameters of the three scanners are collected in
`Table 1.
`
`Fig. 12. A pin-and-staple hinge scanner. The insert shows a
`prototype design for an integrated bar-code-scanner module on a
`silicon substrate.
`
`One important issue for optical scanners is the scan-line
`repeatability from one scan to the next. To characterize
`repeatability,
`the resonant scanner was mounted in an
`optical interferometric setup, which can measure mirror
`displacement with a resolution on the order of 10 nm. The
`cleaved facet of the fiber formed a Fabry–Perot interferom-
`eter with the tip of the scanning micromirror [Fig. 13(a)].
`As the micromirror oscillates, the amount of light that is
`coupled back into the fiber decreases just as if there were a
`reduction in the reflectivity of the micromirror. The change
`in the effective reflectivity is, however, so small that it
`does not affect our measurements. We quantify the scan
`repeatability by comparing the turning points of the periodic
`interference patterns (i.e., the end points of the scan). For
`the scanner measured (scanner A), the deviation for the
`line-scan ends was measured to be on the order of 1% of
`the full scan. We believe that this inaccuracy resulted from
`the sloppiness in the journal bearings used in the design.
`To determine the dynamic accuracy of the microscanner
`
`MULLER AND LAU: MICROOPTICAL ELEMENTS AND SYSTEMS
`
`1713
`
`Petitioner Ciena Corp. et al.
`Exhibit 1048-9
`
`

`
`Fig. 13. Schematic of a Fabry–Perot interferometer formed be-
`tween the fiber front facet and the top of the scanning micromirror.
`
`over the whole scan (as opposed to only the end points),
`a position-sensing detector was used to record the position
`of the He–Ne laser beam reflected off the scanning mirrors.
`The two scanning micromirrors (scanners A and B) were
`used in the measurements. The control voltage for the comb
`drive was sinusoidal at a frequency close to mechanical
`resonance. The position of the scanning laser beam as a
`function of time is shown in Fig. 14. The dotted sinusoids
`were fitted to the measured data. For scanner A [Fig. 14(a)],
`the scan jitter, defined as the standard deviation from a
`pure sinusoid, was only 0.2 mrad (0.011 ) or
`0.06% of
`the full scan. By comparing to the results discussed in the
`previous paragraph, we conclude that the end points of the
`scan are much less repeatable than the center portion, which
`is the only part of the scan that is used in practical optical-
`scanning systems. Relative to the pixel size (see next
`section for definition), the scan jitter of scanner A was 10%.
`The scan jitter measured for scanner B [Fig. 14(b)] was 0.19
`mrad (0.01 ), or 9% of the pixel size for that mirror.
`The static-positional precision of the microscanner can
`also be measured using the same setup. By varying the
`pure dc bias voltage applied to the comb drive, we can
`statically change the position of the deflected laser beam.
`Plotted in Fig. 15 is the measured laser-beam position as
`a function of the square of the applied dc voltages (up-
`ramp and down-ramp over many voltage cycles), again
`for the two -scanning mirrors used previously. Fig. 15(a)
`shows data for scanner A, indicating an angular-positional
`accuracy of 3.3 mrad (0.19 ) in standard deviation from a
`straight line. In Fig.

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