throbber
JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 24, NO. 12, DECEMBER 2006
`
`4433
`
`Optical MEMS for Lightwave Communication
`
`Ming C. Wu, Fellow, IEEE, Olav Solgaard, Member, IEEE, and Joseph E. Ford
`
`Invited Paper
`
`Abstract—The intensive investment in optical microelectro-
`mechanical systems (MEMS) in the last decade has led to many
`successful components that satisfy the requirements of lightwave
`communication networks. In this paper, we review the current
`state of the art of MEMS devices and subsystems for lightwave
`communication applications. Depending on the design, these com-
`ponents can either be broadband (wavelength independent) or
`wavelength selective. Broadband devices include optical switches,
`crossconnects, optical attenuators, and data modulators, while
`wavelength-selective components encompass wavelength add/drop
`multiplexers, wavelength-selective switches and crossconnects,
`spectral equalizers, dispersion compensators, spectrometers, and
`tunable lasers. Integration of MEMS and planar lightwave cir-
`cuits, microresonators, and photonic crystals could lead to further
`reduction in size and cost.
`
`Index Terms—Microelectromechanical devices, optical fiber
`communication, optical signal processing, optical switches.
`
`I. INTRODUCTION
`
`N EARLY three decades ago, Petersen published a paper
`
`on the micromechanical spatial light modulator (SLM)
`array [1] and another on the silicon torsion mirror [2]. Thirty
`years later, this has become a thriving field known as optical
`microelectromechanical systems (MEMS), sometimes also
`called microoptoelectromechanical systems, with several con-
`ferences dedicated to the field. It is a key enabling technology
`for the “dynamic” processing of optical signals. The first mar-
`ket driver of optical MEMS was display [3], [4]. The digital
`micromirror devices developed by Texas Instruments Incorpo-
`rated are one of the most successful MEMS products. They
`are now widely used in portable projectors, large-screen TVs,
`and digital cinemas [3]. The applications of optical MEMS in
`telecommunications started in the 1990s [5], [6]. Early efforts
`
`Manuscript received July 7, 2006; revised October 2, 2006. This work was
`supported in part by the U.S. Defense Advanced Research Project Agency
`(DARPA)/Army Research Office under Grant W911NF-05-1-0359 and DARPA
`under Grant MDA972-02-1-0020.
`M. C. Wu is with the Berkeley Sensor and Actuator Center (BSAC) and
`Electrical Engineering and Computer Sciences Department, University of
`California, Berkeley, CA 94720 USA (e-mail: wu@eecs.berkeley.edu).
`O. Solgaard is with the E. L. Ginzton Laboratory, Stanford University,
`Stanford, CA 94305 USA (e-mail: solgaard@stanford.edu).
`J. E. Ford is with the Department of Electrical and Computer Engineering,
`University of California, San Diego, CA 92093-0407 USA (e-mail: jeford@
`ucsd.edu).
`Color versions of Figs. 3, 5, 10–12, 14, 15, 17, 18, 20, 22, and 25–28 are
`available online at http://ieeexplore.ieee.org.
`Digital Object Identifier 10.1109/JLT.2006.886405
`
`have focused on the development of optical MEMS devices and
`fabrication technologies [7]–[10]. The telecom boom in the late
`1990s and early 2000s has accelerated maturation of the tech-
`nology. A wide range of optical MEMS components were taken
`from laboratories to reliable products that meet Telcordia qual-
`ifications. Although not all commercialization endeavors were
`successful due to the market downturn, the technology devel-
`oped is available for new applications in communications and
`other areas [11].
`In this paper, we will review the recent developments in
`optical MEMS for communication applications. With the rapid
`expansion of the field and proliferation of literature, it is not
`possible to cover all developments in the last decade. Instead,
`we will focus on a selected set of applications and discuss the
`design tradeoffs in MEMS devices and systems. Topics selected
`in this paper include optical switches, filters, dispersion com-
`pensators, spectral equalizers, spectrometers, tunable lasers,
`and other dense-wavelength-division-multiplexing (DWDM)
`devices such as wavelength add/drop multiplexers (WADMs),
`wavelength-selective
`switches
`(WSSs),
`and wavelength-
`selective crossconnects (WSXC). Most of the practical com-
`ponents reported were based on free-space optics. There are
`increasing interests in extending the benefits of optical MEMS
`to guided-wave optics or even nanoscopic photonic structures.
`This new trend will be discussed at the end of this paper.
`Various types of optical switches are needed in telecommuni-
`cation networks. Small 1 × N and N × N switches are useful
`for protection, while optical crossconnect (OXC) offers fast
`provisioning and network management at the wavelength level.
`Nodes in ring networks employ WADMs. As the networks
`evolve toward mesh configuration, WSSs and WSXC become
`important. Dispersion compensators and spectral equalizers are
`essential for improving the link performance as the data rates
`approach 40 Gb/s. Spectral filters and tunable lasers increase
`the flexibility of DWDM nodes.
`This paper is organized as follows: Section II discusses
`broadband (wavelength-independent) devices, including data
`modulators, variable optical attenuators (VOAs), and two-
`dimensional (2-D) and three-dimensional (3-D) MEMS optical
`switches. Section III describes wavelength-selective MEMS,
`including spectral equalizers, WADMs, WSSs, WSXCs, filters,
`dispersion compensators, transform spectrometers, and tunable
`lasers. Section IV focuses on the integration of MEMS and
`planar lightwave circuits (PLC). Section V introduces new de-
`vice concepts based on MEMS-actuated microresonators and
`photonic crystals, and Section VI concludes this paper.
`Capella 2025
`Fujitsu v. Capella
`IPR2015-00727
`
`0733-8724/$20.00 © 2006 IEEE
`
`

`
`4434
`
`JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 24, NO. 12, DECEMBER 2006
`
`Fig. 1. MEMS etalon modulator used for digital data modulation at over
`1 Mb/s. The circular optical aperture is 22 μm in diameter.
`
`Fig. 2. Package configuration for a MEMS data transceiver.
`
`II. WAVELENGTH-INDEPENDENT MEMS
`
`A. Data Modulators
`
`The first practical application of MEMS devices in fiber
`communications was as an optical data modulator, originally
`intended for a low-cost fiber-to-the-home network. A modulator
`is essentially a 1 × 1 switch, operated in either transmission
`(two fibers) or reflection (single fiber). The optical power is
`provided by a constant-intensity remote source, and the mod-
`ulator imprints a data signal by opening and closing in response
`to an applied voltage. Signaling in DWDM fiber networks
`usually requires an expensive wavelength-controlled laser at
`each remote terminal. Passive data modulators offered a poten-
`tially inexpensive solution, but waveguide modulators were too
`expensive and too narrow in optical spectral bandwidth to be
`practical. MEMS offered a new and practical solution.
`The mechanical antireflection switch (MARS) modulator
`is a variable air-gap etalon operated in reflection. The basic
`structure is a quarter-wave dielectric antireflection (AR) coating
`suspended above a silicon substrate [5]. The quarter-wave layer
`is made of silicon nitride with 1/4λ optical path (index times
`thickness), which is roughly 0.2 μm for the 1550-nm telecom
`wavelength. The mechanically active silicon nitride layer is
`suspended over an air gap created by a 3/4λ-thick phospho-
`silicate glass sacrificial layer (0.6 μm). Without deformation,
`the device acts as a dielectric mirror with about 70% (−1.5-dB)
`reflectivity. Voltage applied to electrodes on top of the mem-
`brane creates an electrostatic force and pulls the membrane
`closer to the substrate, while membrane tension provides a
`linear restoring force. When the membrane gap is reduced
`to λ/2, the layer becomes an AR coating with close to zero
`reflectivity. A switching contrast ratio of 10 dB or more was
`readily achieved over a wide (30-nm) spectral bandwidth.
`The initial MARS device shown in Fig. 1 consisted of a
`22-μm optical window supported by X-shaped arms and had
`a resonant frequency of 1.1 MHz. Later devices used a higher-
`yield structure with a symmetric “drum head” geometry [12],
`[13]. These devices were capable of relatively high-speed
`operation: by optimizing the size and spacing of the etch,
`access holes provide critical mechanical damping, and digital
`modulation above 16 Mb/s was demonstrated [14]. While such
`data rates are no longer relevant for telecom, even for fiber-to-
`
`the-home, related modulators are useful for low-power dissi-
`pation telemetry from remote sensors using free-space optical
`communications.
`These early devices provided a proving ground for the reli-
`ability and packaging of optical MEMS telecom components.
`Initial skepticism from conservative telecom engineers was
`combated by the parallel testing of device array operated for
`months to provide trillions of operating cycles. The packaging
`of optical MEMS devices provided new challenges for MEMS
`engineers, but the simple end-coupled configuration was rela-
`tively straightforward to implement. Fig. 2 shows the config-
`uration for a duplex modulator incorporating a MEMS etalon,
`where data can be received by a photodiode and transmitted by
`modulating the etalon reflectivity [15].
`
`B. Variable Attenuators
`
`Data modulators are operated with digital signals, but the
`fundamental response of an etalon modulator is analog. Elec-
`trically controlled VOAs at that time were constructed with
`bulk optical components with electromechanical actuation, with
`10–100-ms response. Erbium fiber amplifiers can use VOA to
`suppress transient power surges, but the time scale required
`was 10 μs, much slower than the data modulation rate. MEMS
`provided an attractive replacement for optomechanical VOAs,
`and this turned out to be the first volume application for MEMS
`devices in telecom networks.
`The first MEMS VOA was fabricated by scaling the opti-
`cal aperture of a MARS modulator from 25 to 300 μm so
`that it could be illuminated with a collimated beam. The re-
`flected signal was focused into a separate output fiber, avoiding
`the need for external splitters or circulators to separate the
`output signal [16]. The first such VOA device is shown in
`Fig. 3. The wavelength dependence of a simple etalon was
`reduced using a more complex three-layer dielectric stack as the
`mechanically active structure, where the original 1/4λ silicon
`nitride layer is sandwiched between conductive polysilicon top
`(1/2λ thickness) and bottom (1/4λ thickness) layers. This at-
`tenuator provided fast (3 μs) response with 30-dB controllable
`attenuation over the 40-nm operating bandwidth, with 0.06-dB
`polarization-dependent loss, and also supported the 100-mW
`power level present in amplifiers. However, the 3-dB insertion
`loss was excessive.
`
`

`
`WU et al.: OPTICAL MEMS FOR LIGHTWAVE COMMUNICATION
`
`4435
`
`Fig. 3. MEMS etalon variable attenuator using a 0.5-mm diameter drumhead
`geometry. The lighter area covers an air gap between the silicon substrate. The
`hexagonally distributed spots are etch access holes.
`
`Fig. 5. Schematic of 2-D MEMS optical switches.
`
`C. Two-Dimensional MEMS Switches
`Protection switches are made of 1 × N or small N × N
`switches. This can be realized by a 2-D array of vertical micro-
`mirrors commonly known as a 2-D MEMS switch. Fig. 5 shows
`the generic schematic of such a switch. The optical beams
`are collimated to reduce diffraction loss. The micromirrors are
`“digital”: They either direct the optical beams to the orthogonal
`output ports or pass them to the drop ports. Generally, only one
`micromirror in a column or row is in the reflection position
`during operation.
`The first MEMS 2-D switch (2 × 2) was reported in [22]
`and quickly followed by related work [23], [24]. For 2 × 2
`switches, low insertion loss (0.6 dB) can be achieved without
`using collimators, especially when the micromirror is immersed
`in index-matching fluid [25]. Latchable 2 × 2 switches incor-
`porating MEMS bistable structures were later commercialized
`[26], [27]. Larger switches require optical collimators to reduce
`diffraction loss. Switches with 8 × 8 and 16 × 16 ports were
`demonstrated [28], [29]. There are two basic approaches for the
`actuation of the micromirror. The first is based on the rotation
`of the micromirror [22], [28], [30], [31]. The mirror is initially
`parallel to the substrate (OFF position). When actuated, it is ro-
`tated to the vertical position (ON). The second approach moves
`the vertical micromirrors in and out of the optical paths without
`changing the mirror angle [23]–[25], [29], [32], [33]. The
`2-D switches have been realized by both bulk-micromachining
`[22]–[25] and surface-micromachining [28]–[30], [32] technol-
`ogies. Electrostatic actuation is most commonly used [22]–[29],
`[32]. Magnetic actuation has also been demonstrated [23], with
`some in conjunction with electrostatic clamping [30].
`The port count of 2-D switches is determined by several
`factors, including mirror angle, size, fill factor (mirror width
`divided by unit cell width), and curvature. The expandabil-
`ity of the 2-D switch has been studied in [34] and [35].
`To minimize optical diffraction loss, a confocal geometry
`is used with the average optical path length equal to the
`Rayleigh range, which is proportional to the square of the
`optical beam waist. Larger mirrors are therefore required to
`support longer Rayleigh length in higher port-count switches.
`In an N × N switch, the mirror size scales as N, whereas
`
`Fig. 4. Lightconnect’s diffractive MEMS VOA.
`
`The most direct possible approach to attenuation is to use
`a MEMS actuator to insert an optical block between the input
`and output fiber. This was implemented with a surface micro-
`machining (MUMPS process) [17] and with a comb-driven
`silicon-on-insulator (SOI) device [18]. Such VOAs offered
`excellent dynamic range (measurement limited at 90 dB), but
`the polarization-dependent loss could be large ((cid:2) 1 dB) at high
`attenuations.
`Further improvement was needed and was made. Combin-
`ing the collimated beam geometry with a first-surface torsion
`mirror reflector provided a low-insertion-loss structure with
`excellent spectral and polarization performance. For example,
`the device demonstrated by Isamoto et al. [19] achieved 40-dB
`attenuation with a 600-μm mirror driven with 5 V to tilt up
`◦
`. Similar configurations were commercialized, although
`to 0.3
`the specific designs have not been published.
`Another commercial MEMS VOA is based on a diffractive
`MEMS device [4] also used with a collimated beam. This
`device provides excellent optical performance as well as high
`speed: stable operation with 30-dB contrast and less than
`40-μs response time using an 8-V drive. A novel structure with
`circularly symmetric features, shown in Fig. 4, was used to
`suppress the polarization-dependent loss to under 0.2 dB [20].
`This device was one of the first Telcordia-qualified MEMS
`components, with 40 000 units reportedly shipped by 2005 [21].
`
`

`
`4436
`
`JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 24, NO. 12, DECEMBER 2006
`
`Fig. 7. Schematic of a 3-D MEMS switch.
`
`D. Three-Dimensional MEMS Switches
`
`A transparent optical crossconnect (OXC) with large port
`count can be realized by 3-D MEMS switches illustrated in
`Fig. 7. The input and output fibers are arranged in 2-D arrays.
`The optical beams are steered in three dimensions by two stages
`of dual-axis micromirrors, directing it toward the desired output
`port. The 3-D MEMS switch has a favorable scaling law with
`respect to port count: Assuming the maximum scan angle of the
`mirror is fixed, the optical path length is proportional to N in
`an N × N switch. To maintain confocal configuration for min-
`√
`imum loss, the beam waist, and therefore the mirror size, needs
`N · √
`√
`N. As a result, the linear dimension of the mirror
`to scale as
`N = N [39]–[41]. In addition, it has low
`chip scales as
`and uniform insertion loss. The 3-D MEMS OXC is a subject
`of intense interest during the telecom boom around the turn
`of the century [42]–[46]. Early efforts (before 2002) focused
`on OXCs with port count ∼1000 × 1000 [47], [48], driven by
`the explosion of Internet data transport. Recently, interest has
`shifted to applications in metropolitan area networks, including
`metro access and metro core networks, which requires OXC
`with medium port count (∼100 × 100), with emphasis on low
`cost, low-power consumption, and small footprint [44], [49].
`Our discussion here will focus on this trend.
`Detailed design tradeoffs and system implementations of the
`3-D MEMS OXC have been reported recently [42]–[46]. Two
`schemes have been proposed to reduce the size of the switch
`and tilt angle of the micromirror. Lucent inserted a Fourier lens
`between the two micromirror chips with the focal length equal
`to the Rayleigh range of the optical beam (Fig. 8) [50]. This
`√
`reduces the required scan angle of the mirror. In addition, the
`2 times
`mirrors can be placed at the beam waist, resulting in
`smaller optical beams. This permits the use of smaller mirrors
`and/or reduction of the crosstalk. Fujitsu used a “rooftop”
`mirror to connect two adjacent micromirror chips (photograph
`show in Fig. 9) [44]. The rooftop mirror shifts the optical beams
`laterally, reducing the tilt angle requirement. Folding of the
`optical beam also shrinks the footprint of the switch.
`In the compact switch category, Lucent’s 64 × 64 switch
`has a size of 100 × 120 × 20 mm3, which can be mounted
`on a standard circuit board [49]. The insertion loss is 1.9 dB.
`Fujitsu’s 80 × 80 switch has a packaged size of 77 × 87 ×
`53 mm3 [44]. The average insertion loss is 2.6 dB. Impressively,
`the switch continues to operate under vibration or 50G shock
`
`(a) SEM of OMM’s 16 × 16 switch (reprinted from [29] with
`Fig. 6.
`permission). (b) Photograph of the packaged switch (reprinted from [36] with
`permission).
`
`the linear dimension of the chip scales as N 2 [35]. Large
`chips are more susceptible to imperfections in mirror angles,
`which cause walkoff of optical beams at the receiving fibers.
`Ultimately, the chip size will be limited by the fabrication
`precision of the micromirrors. 16 × 16 switches have been
`realized, and 32 × 32 switches are within the capability of
`today’s technology.
`Fig. 6(a) shows a scanning electron micrograph (SEM) of
`OMM’s 2-D switch [29]. A vertical mirror is attached at the
`tip of a cantilever. The tilted cantilever can be pulled down
`◦
`during
`electrostatically. The mirror angle is maintained at 90
`switching. The switch is fabricated using a standard three-
`polysilicon-layer surface-micromachining process. The mirrors
`are assembled into vertical position with angular distribution of
`(90 ± 0.1)◦
`. The hermetic switch package is shown in Fig. 6(b)
`[36]. Maximum insertion losses of 1.7 and 3.1 dB have been
`obtained for 8 × 8 and 16 × 16 switches, respectively, and
`the crosstalk is less than −50 dB. The switching time is less
`than 7 ms. Packaging is critical to attain long-term reliabil-
`ity and satisfy Telcordia qualification for telecommunication
`applications [36].
`There were also significant efforts in nonmirror-based
`MEMS 2-D switches [37], [38]. Both Agilent’s Champaign
`switch [37] and NTT’s OLIVE switch [38] used microfluidic
`actuation to switch light between intersecting waveguides. The
`Champaign switch used thermally generated bubbles to dis-
`place index-matching fluids at waveguide intersections, causing
`the light to bend by total internal reflection (TIR). The OLIVE
`switch used thermal-capillary force to move trapped bubbles.
`One drawback of these approaches is the cumulative losses
`and crosstalks through multiple waveguide intersections. The
`maximum port counts achieved are 32 × 32 and 16 × 16 for
`the Champaign and the OLIVE switches, respectively.
`
`

`
`WU et al.: OPTICAL MEMS FOR LIGHTWAVE COMMUNICATION
`
`4437
`
`Fig. 8. Lucent’s optical system layout for OXC (reprinted from [50] with
`permission). A Fourier lens is inserted between the two MEMS chips to reduce
`the required tilt of the mirror and beam size.
`
`(a) Dynamic spectral equalizer package and (b) transmission spectra
`Fig. 10.
`showing the improvement in channel uniformity for a 36-channel DWDM
`transmission.
`
`increases the complexity of electronics [58]. Micromirrors with
`vertical comb drive actuators, first reported in [59], offer many
`advantages. They have a much larger torque, which one can use
`to reduce the operating voltage as well as increase the resonant
`frequency. In addition, they are free from the pull-in effect,
`further increasing the stable tilt angles. It should be mentioned
`that lateral pull-in between comb fingers is a potential issue but
`could be mitigated by MEMS design (such as V-shaped torsion
`beam [60] or off-centered combs [61]). Several variations of
`vertical comb drive mirrors have been reported, including self-
`aligned vertical combs [62], [63], angular vertical combs [64],
`[65], electrostatically assembled vertical combs [66], and thick
`vertical combs (100 μm) attached to mirror edges on double-
`sided SOI wafers [44], [60].
`
`III. WAVELENGTH-SELECTIVE MEMS
`
`A. Spectral Equalizers
`
`The natural extension of a single variable attenuator is to
`provide a VOA for each channel of a DWDM transmission
`system. The surface-normal geometries of the etalon mirror-
`and grating-based attenuators discussed in Section II-B were
`all compatible with a free-space imaging spectrometer. An
`input fiber is imaged through a diffraction grating so that each
`spectral channel is laterally shifted to illuminate one modulator
`in a linear array. The reflected signal, attenuated to the desired
`value, is collected into a single output fiber by a second pass
`through the imaging spectrometer. The first such MEMS spec-
`tral equalizer used a continuous etalon membrane [67]. This
`approach was later implemented in the compact package shown
`in Fig. 10, which located the MEMS device array next to a
`single input/output (I/O) fiber. A single lens is to collimate
`the multiwavelength beam onto a blazed reflective grating and
`
`Fig. 9. Photograph of Fujitsu’s 80 × 80 OXC with a rooftop reflector
`connecting the two MEMS chips (reprinted from [44] with permission). The
`packaged size is 77 × 87 × 53 mm3.
`
`without any signal degradation. The total power consumption
`of Fujitsu’s switch is only 8.5 W, thanks to the low operating
`voltage of the mirrors. NTT’s 100 × 100 switch has a size of
`80 × 60 × 35 mm3 with an insertion loss of 4 dB [43].
`The two-axis micromirror array is the key enabling device
`of the 3-D switch. Important parameters include size, tilt an-
`gle, flatness, fill factor, and resonant frequency of the mirror.
`Additionally, the stability of the mirror plays a critical role
`in the complexity of the control schemes. Early development
`focused on surface-micromachined two-axis scanners [51],
`[52]. The residue stress limits the mirror size to approximately
`1 mm, and the different thermal expansion coefficients be-
`tween the mirror and the metal coating also cause the mirror
`curvature to change with temperature. Bulk-micromachined
`single-crystalline silicon micromirrors are often used in high-
`port-count OXCs that require larger mirror size [46], [53]–[56].
`Electrostatic actuation is most commonly used because of
`its low-power consumption and ease of control. Early devices
`use parallel-plate actuators, which have high actuation voltage
`and limited scan angle due to pull-in instability [57]. Although
`the pull-in effect can be mitigated by nonlinear controllers, it
`
`

`
`4438
`
`JOURNAL OF LIGHTWAVE TECHNOLOGY, VOL. 24, NO. 12, DECEMBER 2006
`
`Fig. 11. Optical schematic for a 2 × 2 MEMS wavelength add/drop switch.
`
`refocus the spectrally separated signals with a second pass onto
`the MEMS array. A third and fourth pass through the lens
`reintegrates the signal into the I/O fiber, where it is separated
`by an external optical circulator. The use of such equalizers is
`illustrated by the before and after spectral traces at the bottom of
`Fig. 10, showing the improvement in uniformity of 36 channels
`sent through a two-stage amplifier. The equalizer setting was
`generated by an iterative algorithm running on the computer
`controller [68].
`Dynamic spectral equalization quickly went from an option
`to a practical requirement as the channel transmission rate
`increased from 2.5 to 10 and then to 40 Gb/s. The simplest
`and least expensive dynamic gain equalizers (DGEs) use a
`mid-amplifier filter that can be spectrally uniform (a VOA,
`as discussed above) or provide a constant spectral slope [69].
`Two distinct categories of spectral equalizers emerged. DGEs
`provide a smoothly varying spectral profile used to compensate
`for the varying gain profiles in amplifiers, while dynamic chan-
`nel equalizers (DCEs) provide the discrete channel-by-channel
`power adjustment needed to compensate for nonuniform trans-
`mission source intensity or path-dependent loss. Channel equal-
`izers are preferable in general but require accurate matching
`of the equalizer passband to the transmission grid to avoid
`passband narrowing.
`Channel equalizers were implemented using discrete VOAs
`attached to waveguide spectral multiplexers [70] and using an
`oversampled array of digital tilt mirrors [71]. However, the best
`performance in channel equalizers was achieved by combining
`the type of free-space grating demultiplexer shown in Fig. 10
`with either diffractive MEMS modulators [72] or analog tilt
`mirrors [73]. The optical setup is similar to that in Fig. 11
`except without circulators. These components typically have
`40–80 channels spaced at 100 or 50 GHz with 6- and 7-dB
`insertion loss and 20- to 30-dB dynamic range. The most ad-
`vantageous characteristic of MEMS equalizers is the extremely
`flat passband transmission profile along with low chromatic
`
`dispersion at the edges. This performance was achieved after
`studying the effects of various mirror geometries [74].
`After understanding the effects of mirror profile on disper-
`sion, it became possible to use the same basic component struc-
`ture as the equalizer to provide channel-by-channel dispersion
`compensation, although this functionality has yet to be adopted
`in the deployed network [75].
`
`B. Wavelength Add/Drop Multiplexers
`
`Wavelength switching allows network operators to use opti-
`cally transparent components to pass through a network node
`without detecting and regenerating the data signal, and com-
`ponents that enable this have been the subject of intense re-
`search and development. The most basic wavelength switch is
`the dynamically reconfigurable WADM, which is essentially a
`1 × 2 or 2 × 2 optical switch operating independently on each
`wavelength channel.
`WADM was a natural extension of MEMS equalizers, and the
`first demonstration of a MEMS add/drop switch based on dig-
`ital tilt mirrors occurred almost simultaneously with the equal-
`izer [76], [77]. Add/drop requires four ports, twice as many
`as the equalizer, and so, the basic structure is slightly more
`complex (Fig. 11). The system is still based on a blazed diffrac-
`tion grating, which is now illuminated with an upper and lower
`beam path. The active device is a linear array of 16 digital tilt
`mirrors fabricated with surface micromachining in the MUMPS
`process. Each mirror defines a DWDM channel and, in switch-
`ing, directs the reflected signal back along the input direction
`or tilted into a new path. Optical circulators on the two I/O
`fibers separate the forward and reverse propagating signals. The
`mirrors in this switch tilted by ±5◦
`under a 20-V signal, switch-
`ing in 20 μs. A quarter-wave plate is used to achieve 0.2-dB
`polarization dependence on a total insertion loss of 7.5 dB.
`The DCE is closely related to the WADM, and in fact, it is
`possible to use high-contrast equalizers as 1 × 1 switches in a
`
`

`
`WU et al.: OPTICAL MEMS FOR LIGHTWAVE COMMUNICATION
`
`4439
`
`Fig. 12. Equivalent circuit of a 1 × 4 WSS. Eight wavelength channels are
`shown in this example.
`
`“broadcast and select” architecture [78]. The primary disadvan-
`tage of this architecture is that it is intrinsically lossy: Signals
`are power split, and then unwanted signals are blocked before
`combining into the output fiber. This does allow multicasting,
`i.e., duplicating signals to multiple output fibers. Broadcast and
`select was actually the first to be implemented in the network
`but is generally expected to be phased out in favor of multiport
`WSSs, which in addition to switching also provide channel
`equalization [79] with no additional cost or complexity.
`
`C. Wavelength-Selective Switches (WSSs)
`
`As optical networks evolve from a simple ring topology with
`WADM nodes to optical mesh networks, WSSs with more than
`one output port are needed to link the node to three or four
`neighboring nodes with each link carrying two-way traffic. The
`WADM concept can be extended to switches with N output
`ports, where N is larger than 2. This is called 1 × N WSS
`[80]–[82]. Fig. 12 shows the equivalent circuits of a 1 × 4 WSS.
`It consists of a WDM demultiplexer, Nλ of 1 × N space divi-
`sion switches (Nλ is the number of wavelength channels) and
`N WDM multiplexers. The WSS can be realized by a similar
`grating spectrometer configuration as the WADM, with the dig-
`ital micromirrors replace by “analog” ones. A large continuous
`scan angle is required to direct the output beam to any of the N
`output fiber collimators. High fill factor is desired to minimize
`the gaps between wavelength channels. The mirror size is
`usually several times larger than the focused optical beam to
`attain a wide and flat passband for minimal signal distortion.
`A detailed review paper on WSS was published recently
`[80]. The optical setup for Lucent’s WSS is shown in Fig. 13.
`The first subassembly maps all fiber I/Os to a common spot
`(point B), and the second subassembly (resolution lens and
`grating) separates and focuses the wavelengths onto the mi-
`cromirror array at the image plane. Tilting of the mirror changes
`the direction of the reflected beam at point B and sends the
`optical beam into a different output fiber. A refined design
`incorporates anamorphic optics in the input stage to reduce the
`physical size of the switch while maintaining the same spectral
`resolution at the expense of longer micromirrors.
`Experimentally, 1 × 4 WSSs with 128 channels spaced on
`a 50-GHz grid and with 64 channels spaced on a 100-GHz
`
`Fig. 13. Schematic optical setup of 1 × 4 WSS (reprinted from [80] with
`permission).
`
`grid have been demonstrated. The typical optical insertion loss
`ranges from 3 to 5 dB. The channel passband is directly related
`to the confinement factor, which is defined as the ratio of the
`mirror size to the Gaussian beam diameter. A confinement
`factor of > 2.7 is needed to produce a flattop spectral response
`with > 74% passband width measured at −1 dB point. JDSU
`has reported a similar 1 × 4 WSS with 3.5-dB insertion loss
`[81]. UCLA has reported a similar WSS with excellent open-
`loop stability [82].
`The analog micromirror array plays a key role in the per-
`formance of the WSS. Several types of WSS micromirror
`arrays have been reported, including electrostatic [83], [84]
`and electromagnetic [85] actuations. The key parameters are
`large continuous scan angle and high fill factor, with the mirror
`size and pitch matching those of the optical system. Lucent
`employed a fringe-field actuated SOI micromirror array [83]
`◦
`at 175 V. The
`and achieved a mechanical tilt angle of 9.2
`resonant frequency is 3.8 kHz for 80-μm-wide mirrors.
`More efficient actuation has been obtained using vertical
`comb drive actuators. Hah et al. reported a low-voltage analog
`micromirror array for WSS [84]. The schematic and the SEM
`of the micromirror are shown in Fig. 14. The mechanical struc-
`tures are completely covered by the mirror; therefore, a high
`fill factor is achieved along the array direction. The actuation
`voltage is as low as 6 V for mechanical tilt angles of ±6◦
`. High
`resonant frequency (3.4 kHz) and high fill factor (98%) are also
`achieved [86]. The excellent stability of the mirror (±0.00085◦)
`enables open-loop operation of the switch with insertion loss
`variation of < ±0.0035 dB over 3.5 h [82].
`Scaling of WSS has been analyzed in [86]. The figure
`of merit is the ratio of the port count and channel spacing
`(N/Δλch). It is proportional to the product of the effective
`aperture of the resolution lens and the grating dispersion. Most
`of the reported WSSs have a port count of four. A larger port
`count (N ≥ 8) is desirable for mesh optical networks, where it
`is necessary to provide two-way links to three or four adjacent
`neighboring nodes. The port count can be increased from N
`to N 2 by arranging the output collimator in a 2-D array. This
`is referred to as 1 × N 2 WSS [86]–[88]. Micromirror arrays
`providing two-axis beamsteering functions are needed for this
`architecture. This can be accomplished by using either a linear
`array of two-axis micromir

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