High-resolution Micromachined Interferometric Accelerometer by Nin C. Loh Submitted to the Department of Mechanical Engineering in partial fulfillment of the requirements for the degree of Master of Science at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2001 C 2001 Massachusetts Institute of Technology All rights reserved. MASSACHUSETTS INSTITUTE OF TECHNOLOGY JUL 16 2001 LIBRARIES BRKER Authored by ................ ........... Department of Mechanical Engineering May 23, 2001 Certified by ...... ...... .. ...................... Scott R. Manalis ssistant Professor, Media Arts & Sciences and Bioengineering -I R ead by .......... Accepted by ........... Thesis Supervisor ......... ...... ............................................ George Barbastathis Assistant Professor, Mechanical Engineering Thesis Reader .... ............................................ . Ain Sonin Chairman, Departmental Committee on Graduate Students 1 2 High-resolution Micromachined Interferometric Accelerometer by Nin C. Loh Submitted to the Department of Mechanical Engineering On May 23, 2001, in partial fulfillment of the requirements for the degree of Master of Science ABSTRACT A miniature high-resolution accelerometer with a bulk-micromachined silicon proof mass and an interferometric position sensor was developed. The interferometer consists of interdigitated fingers that are alternately attached to the proof mass and support substrate. Illuminating the fingers with coherent light generates a series of diffracted beams. The intensity of a given beam depends on the out-of-plane separation between the proof mass fingers and support fingers. Displacements of the proof mass can be detected with a resolution of 10-3 A/rt Hz over a range of 600 A. Structures with a mechanical resonance ranging from 80 Hz to 8.2 kHz were fabricated with a two mask process involving two deep reactive ion etches, an oxide etch stop, and a polyimide protective layer. The structures were packaged with a laser diode and photodiode into 8.6 cm 3 acrylic housings. The 1 kHz resonant structure detected 200 ng/rt Hz at 400 Hz with a dynamic range 7x10 5 . Although the acceleration resolution of the 80 Hz resonant structure is currently limited by the background seismic noise, we speculate that the ultimate limit is the thermomechanical noise of 6.8 ng/rt Hz. The advantage of the interferometric sensor over tunneling accelerometers is its simple fabrication process and large open-loop dynamic range. Thesis Supervisor: Scott Manalis Title: Assistant Professor, Media Arts & Sciences and Bioengineering 3 4 ACKNOWLEDGEMENTS I wish to thank my adviser, Professor Scott Manalis, for his guidance and for leading a group that fosters creativity, teamwork, enthusiasm, and hands-on research. My two years at MIT have been most enjoyable, thanks to the members of the our group, Emily Cooper, Jirgen Fritz, Cagri Savran, and Andrew Sparks. I also want to thank our administrative assistants, Rosanne Kariadakis, Martha Lugo, and Alicia Peyrano for making my research more efficient. Our UROP Peter Russo has helped with the packaging of the devices. The interferometric accelerometer spawned from two project classes co-taught by Scott and Professor Martin Schmidt. The class members, Emily Cooper, Saul Griffith, Jeremy Hui, Jeremy Levitan, Ole Nielsen, Oluwamuyiwa Olubuyide, Rehmi Post, Cecily Ryan, and Jbrg Scholvin, performed groundbreaking work on this sensor. I want to especially thank Emily for documenting and transferring the technology from the first class, Professor Schmidt for his technical insight, and J6rg for helping me to develop the fabrication process. The fabrication of our accelerometer went smoother thanks to the expertise and help of the MIT Microsystems Technology Laboratory staff and students, especially Vicky Diadiuk, Bill Teynor, Tom Takacs, Kurt Borderick, Andy Fan, and Isaac Lauer. I also want to thank Saul Griffith for showing me the marvels of the lasercutter, a tool that was essential to the packaging. Funding for this research was provided by the MIT Media Laboratory's Things That Think (TTT) consortium. During my last two semesters I was supported by a graduate research fellowship from Motorola. Finally, I am deeply grateful to my parents and older siblings for their support and sacrifice in my academic journey. 5 CONTENTS 1 INTRODUCTION 1.1 Motivation: current accelerometers................................................. 1.2 High-resolution tunneling accelerometers......................................... 1.3 Optical interference accelerometers................................................. 1.4 Thesis overview ........................................................................ 11 11 13 14 15 2 THEORY 2.1 Interdigital accelerometer............................................................. 2.2 Interdigital position sensing.......................................................... 2.3 N oise analysis.......................................................................... 16 16 17 20 3 DESIGN AND FABRICATION 3.1 Proof mass wafer...................................................................... 3.1.1 D esign...................................................................... 3.1.2 Fabrication process........................................................ 3.1.3 Fabrication results......................................................... 3.2 Packaging............................................................................. 3.2.1 First generation design.................................................... 3.2.2 Second generation design................................................ 3.3 Photodiode wafer..................................................................... 3.3.1 D esign...................................................................... 3.3.2 Fabrication process....................................................... 3.3.3 Electrical properties...................................................... 24 24 24 28 31 33 33 36 38 38 39 41 4 RESULTS 4.1 Signal conditioning.................................................................. 4.2 Mechanical drive stage............................................................. 4.3 Package assembly................................................................... 4.4 Sensor performance................................................................. 4.4.1 Sensitivity................................................................. 4.4.2 Noise spectrum............................................................ 4.4.3 Linearity................................................................... 4.4.4 Cross-axis sensitivity and drift......................................... 42 42 42 43 46 46 49 51 54 5 CONCLUSION 57 6 APPENDICES A MECHANICAL MODELS A. 1 Folded pinwheel analytical model.................................................. A.2 Cantilever proof mass deflection................................................... 58 58 60 B MASK DESIGN B.1 Proof mass wafer...................................................................... B.2 Photodiode wafer...................................................................... 61 61 62 C FABRICATION DETAILS C.1 Proof mass wafer..................................................................... C.2 Photodiode wafer..................................................................... C.3 Packaging............................................................................. 65 65 68 70 D FIRST GENERATION PACKAGE RESULTS 72 7 LIST OF FIGURES 1.1 1.2 1.3 2.1 2.2 2.3 2.4 2.5 2.6 2.7 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 3.10 3.11 Mechanical model of a typical accelerometer and useful parameters.................. Trade-off between volume and resolution at 100 Hz of commercial single-axis accelerom eters................................................................................. Simplified cross-section of typical micromachined tunneling accelerometer......... Schematic of micromachined accelerometer using interdigitated fingers as a . position sensor................................................................................ Diagram of diffraction off (a) parallel and (b) vertically offset interdigitated ..... fingers.................................................................................... wavelength about eight intensity of sine squared intensity in linearizing Error offset............................................................................................ Position noise of interdigital cantilever for atomic force microscopes. (Reprinted with permission from APL paper, [12])..................................................... Thermomechanical noise as a function of resonant frequency for a 30 mg proof mass w ith Q of 100 at 300 K ................................................................. Interdigital noise as a function of operating frequency for various resonant frequencies..................................................................................... Seismic noise in 4th floor laboratory....................................................... Schematic of folded-pinwheel as an interdigital accelerometer proof mass........... Layout and dimensions of folded pinwheel proof masses............................... Folded pinwheel proof mass resonant frequencies versus spring length L............ Fabrication process sequence for interferometric accelerometer proof mass wafer. a.) Thermal oxidation of SOI wafer. b.) Topside lithography and DRIE etch to define fingers, springs, and mass. c.) Spin-on and cure polyimide. d.) Bottomside lithography and DRIE etch to define mass. e.) BOE etch of box oxide. f.) Ash polyim ide in oxygen plasm a................................................................. 5x micrograph of 10 kHz proof mass corner............................................... 1Ox backlit micrograph of 10 kHz interdigitated fingers showing release............. Cross-section of first generation package.................................................. Actual size schematic of the first generation package pieces. (A) is the bottom piece on which the proof mass rests. (B) is the bottom shim. (C) is the top shim with rectangular holes for the wirebonds and electrical leads. (D) is the top piece with a circular hole for the laser diode, a circular recess (red) for the plastic lens, and a rectangular recess (yellow) for the bonding plates. The blue lines represent shallow grooves to align the proof mass die (A) and photodiode die (D)............. Normalized signal-to-noise ratio versus incident power for 8.2 kHz sensor driven at 4 kH z ......................................................................................... Cross-section of second generation package............................................... Laser diode-lens system for generating a small beam.................................... 8 11 12 14 16 17 19 20 21 22 23 24 26 27 30 31 31 33 34 35 36 37 3.12 3.13 3.14 3.15 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9 4.10 A. 1 B. 1 B.2 B.3 C. 1 D.1 D.2 D.3 Actual size schematic of the second generation package pieces. (A) is the bottom piece on which the proof mass rests. (B) is the bottom shim. (C) is the top shim with rectangular holes for the wirebonds and electrical leads. (D) is the top piece with a semicircular cut-out for the laser dio de and a rectangular recess (yellow) for the bonding plates. The blue lines represent shallow grooves to align the proof mass die (A) and photodiode die (D)................................................. Layout of photodiode die.................................................................... Fabrication process sequence for photodiode wafer. a.) Thermal oxidation. b.) Topside lithography to open field oxide. c.) Implant dopants. d.) Dopant drivein/activation and reoxidation. e.) BOE etch to open contact holes. f.) Deposit aluminum. g.) Pattern aluminum. h.) DRIE etch through-hole...................... Plot of custom photodiode output versus incident power............................... Schematic of piezoshaker used to actuate the sensors................................... Diagram of tabletop and packaging experimental set-ups.............................. Comparison of resolution of 8.2 kHz sensor in tabletop and package set-ups....... Sensitivity of (a) 1020, (b) 432, and (c) 80 Hz packaged sensors...................... Noise spectrum of (a) 1020 and (b) 432 Hz packaged sensors.......................... Noise spectrum of 80 Hz packaged sensor................................................ Output versus acceleration of (a) 1020, (b) 432, and (c) 80 Hz packaged sensors... Illustration of cross-axes...................................................................... Maximum cross-axis sensitivities of 432 Hz packaged sensor compared to z-axis sensitivity ....................................................................................... Thirty-minute drift of 432 Hz packaged sensor........................................... Diagram and definition of parameters of folded-pinwheel structure with springs parallel to mass................................................................................ 37 39 40 41 43 44 45 48 50 51 53 54 55 56 58 Frontside (a) and backside (b) mask for proof mass wafer.............................. Mask set for photodiode wafer. a.) Mask 1 defined photodiode active regions. b.) Mask 2 opened contact holes. c.) Mask 3 pattern interconnects. d.) Mask 4 defined D RIE through-hole.................................................................. Sample dies from (a) Mask 1, (b) Mask 2, (c) Mask 3, and (d) Mask 4............... 63 64 Lasercutter drawings for (A) top piece, (B) top shim, (C) bottom shim, and (D) bottom piece of first and second generation packages. Black lines represent through cuts, yellow areas are 1 mm deep rasters, red areas are 4 mm deep, orange lines are 2 mm clean-up lines, and cyan lines are shallow (<500 pim) alignment lin es............................................................................................. 71 Sensitivity of 428 Hz first generation package............................................ Noise spectrum of 428 Hz first generation package...................................... Linearity of 428 Hz first generation package............................................. 72 73 73 9 61 LIST OF TABLES 3.1 3.2 4.1 4.2 4.3 C. 1 Gravitational deflection, higher resonant frequencies, and load at failure values calculated by Mechanica for each proof mass design.................................... Design and measured resonant frequencies............................................... 28 32 Noise level of each package during the packaging process............................. Parameters of fitted sensitivity transfer functions........................................ Acceleration and deflection linearity and dynamic range of packaged interdigital accelerom eters................................................................................. 46 49 Lasercutter settings for fabricating packages............................................. 71 10 52 1 INTRODUCTION 1.1 Motivation: current accelerometers reson ant frequen cy b k 2~ m quality factor Q b Figure (1.1). Mechanical model of a typical accelerometer and useful parameters. Accelerometers are a vital part of inertial guidance, airbag deployment, and earthquake detection systems. Figure (1.1) is a simple mechanical model of an accelerometer, in which a inertial, or proof, mass of mass m is suspended by a spring of spring constant k and a dashpot of damping constant b. This class of sensors determines acceleration by measuring either the position z of the proof mass relative to the rigid support (passive) or the force required to keep the proof mass stationary (force-feedback). Accelerometers can vary greatly in the way they detect this lag, as well as in size and minimum detectable acceleration. Before the advent of silicon micromachining, accelerometers were large electromechanical structures, sometimes Starting at the end of the 1970's, bulk- involving several discrete components [1]. micromachining (through-wafer etching) was used to fabricate silicon accelerometers with piezoresistive, capacitive, or piezoelectric position sensing [2]. Although bulk-micromachining offered a way to make smaller and more reliable sensors, its cost prevented such devices from proliferating into low-cost markets such as the automotive industry. In the early 1990's, capacitive accelerometers batch-fabricated from thin films on silicon (surface micromachining) provided a low cost alternative to traditional sensors in airbag deployment systems. Today, virtually all airbags are controlled by such sensors because of their size (DIP chip package measuring less than 0.5 cm 3), cost (<$5), and accuracy (1%). While adequate for detecting 50 g (fifty times the acceleration of gravity (g = 9.81 m/s 2 )) collisions, surface micromachined capacitive sensors can only resolve disturbances around a milli-g. Accelerometers that can detect accelerations less than a millionth that of gravity, or micro-g, are necessary for measuring seismic disturbances as well as gravitational waves [3, 4]. 11 Present-day sensors with such low noise are typically larger than the size of an egg and cost several hundreds of dollars. Figure (1.2) is a plot of the volume of various single-axis commercial accelerometers as a function of their resolution at 100 Hz. The trade-off between size and resolution is apparent. Several research groups are currently developing microfabricated sensors with sub-micro-g resolution and a package volume of less than 10 cm 3, The goal of this thesis work is to develop a new approach for fabricating and packaging an accelerometer with these characteristics. 1000 Piezoelectric sensing Capacitive sensing 100 E E 10 SA Sx 0 SX xe 1 0 .1 1 1 1 1111111 I 0.001 0.01 1111111 1111111 0.1 1 1 11111 1 10 1 111111 100 111 1111 1000 Resolution (ug/rt Hz) Figure (1.2). Trade-off between volume and resolution at 100 Hz of commercial single-axis accelerometers. 12 Microfabricating sensors that resolve accelerations in the 10 to 100 nano-g range is a major challenge [5]. The mass of the proof mass often decreases with sensor size, increasing the resonant frequency f.. This reduces the displacement z of the proof mass resulting from small accelerations because the static deflection of the proof mass (ignoring damping) is inversely proportional to the square of its resonant frequency. kz = ma => z = a a - O)g 4 2 (1.1) f2 where a), is the angular resonant frequency. For example, a bulk-micromachined capacitive accelerometer with a 126 Hz resonant proof mass of 14.7 mg can resolved 0.1 micro-g at 1 Hz [6]. This noise is over a thousand times less than that of surface-micromachined capacitive accelerometers whose masses and resonant frequencies are typically 1 pg and 10 kHz, respectively. In addition to large, compliant proof masses, high resolution accelerometers also require highly sensitive displacement detection. However, some position detection schemes do not scale well with size. For instance, the sensitivity of capacitive sensors decreases as the area of the sensing capacitor is reduced. As the sensitivity goes down, parasitic capacitances can become significant if the readout circuitry is not close enough to the sensor. 1.2 High-resolution tunneling accelerometers Over the last decade, a major effort in microfabricating nano-g accelerometers has used electron tunneling as the position detector of a bulk-micromachined proof mass. Tunneling transducers, which measure the tunneling current between proof mass and a reference electrode, are ideal for microfabrication because of their insensitivity to miniaturization. The tunneling current depends predominantly on the two closest atoms between the mass and reference. Furthermore, because the tunneling current is exponentially dependent on the tunneling gap, electron tunneling transducers can resolve displacements down to 10-4 A/rt Hz. However, to achieve a significant tunneling current, the proof mass must be positioned several angstroms away from the tip. This strict tolerance increases the likelihood of the proof mass crashing into the tip during large accelerations. For this reason, tunneling accelerometers typically operate in 13 force-feedback where the tunneling current is used to control a deflection voltage that keeps the proof mass stationary. Deflection electrode Tunneling tip Silicon Metal Figure (1.3). Simplified cross-section of typical micromachined tunneling accelerometer. Most high resolution tunneling accelerometers consist of a large proof mass (10 to 40 mg) with a conductive side that is brought within 10 angstroms of the reference tip as shown in Figure (1.3). To keep a constant tunneling gap, deflection electrodes are placed around the proof mass to limit its movement to about 1 A. Sometimes the tunneling tip is also actuated, electrostatically [7] or piezoelectrically [8], to achieve higher bandwidth. Recently, researchers have microfabricated accelerometers based on electron tunneling transducers that resolve 30 ng/rt Hz between 7 Hz and 1.1 kHz [9]. The tunneling accelerometer has several drawbacks. Although feedback can extend the bandwidth of a low resonant frequency, high Q (low damping) proof mass, such a mechanical structure is difficult to constrain to within 1 A of movement. The microfabrication process is complex and can involve up to seven masks and two or more bonded wafers. Furthermore, process variations complicate the design of the controller. The performance of the tunneling accelerometer is also highly sensitive to the cleanliness of the electrodes [10]. 1.3 Optical interference accelerometers Optical interference transducers are capable of achieving a position resolution equivalent to electron tunneling transducers. Interferometric displacement sensors can take many forms, including a Michelson interferometer [11] or a Fabry-Pdrot cavity [3]. In the mid-1990's, an interferometer was developed to increase the resolution and decrease the alignment sensitivity of atomic force microscope cantilevers [12]. This interferometer measured the intensity of a single 14 mode from the diffraction pattern reflected off interdigitated fingers on the cantilever. The offset between two sets of fingers could be resolved to 5xl04A at 200 Hz and was linear up to a tenth of the wavelength of the light source, or 600 A of motion. An accelerometer based on such a position transducer can achieve a dynamic range of over 10 4 without the use of any force- feedback. Moreover, such an optical interferometer is a good candidate for micromachining because the dimensions of the interdigitated fingers must be on the order of the wavelength of the light for significant separation of the diffraction modes. In the fall of 1999, the first accelerometer to use interdigitated fingers as a position detector was fabricated [13]. Although the sensor resolved micro-g accelerations, the fabrication had low reproducibility and yield (<10%), and the device lacked a package. 1.4 Thesis overview We present the design, fabrication, packaging, and characterization of a miniature highresolution accelerometer based on an interferometric transducer. In Chapter 2, we discuss the theory behind the sensor as well as analyze sources of noise. Chapter 3 describes the design and fabrication of our sensor, which is composed of the proof mass, plastic housing, and photodiode. In Section 3.1.1, we focused on improving the performance and robustness of the proof mass from its previous design. For example, the mechanical structure was designed for high linearity of motion in the sensitive axis and low sensitivity to external strains. Section 3.1.2 discusses the microfabrication of the proof mass, including the steps taken to release low resonant frequency structures with 100% yield and high reproducibility. The fabrication of the package and the process of assembling an interdigital accelerometer are covered in Section 3.2. Chapter 4 describes the testing apparatus and the measured performance of the sensor, including sensitivity, linearity, and noise. Section 4.3 discussed how we ensured that resolution was not lost when transferring the sensor from a tabletop set-up with well separated diffraction modes to a small package with closely packed modes. Finally, Chapter 5 summarizes the work and discusses future directions of research in and applications of interdigital accelerometers. 15 2 THEORY 2.1 Interdigital accelerometer Laser Source Light Detector Di ction Aearn. Support Substrate Proof Mass Interdigitated Figer Figure (2.1). Schematic of micromachined accelerometer using interdigitated fingers as a position sensor. A schematic of the first micromachined interdigital accelerometer is shown in Figure (2.1). Like the tunneling accelerometer in Figure (1.3), there is a proof mass suspended by a spring. However, instead of a tunneling tip and deflection electrode, the interferometric sensor has flat fingers extending from the proof mass that interleave with fingers extending from the support substrate. When the interdigited fingers are illuminated with coherent light, a diffraction pattern is reflected. The movement of the proof mass is determined by measuring the intensity of one of the diffraction beams which is a sensitive function of the offset between the two sets of fingers. The theory behind this position sensing will be detailed in the next section. 16 2.2 Interdigital position sensing Incident Beam -2nd Mode +2nd Mode Oh Mode d a.) Incident Beam +1st Mode -1st Mode -- z=W/4 b.) Figure (2.2). Diagram of diffraction off (a) parallel and (b) vertically offset interdigitated fingers. 17 Imagine a row of reflective fingers all in the same plane as shown in Figure (2.1 a). If the fingers are illuminated with coherent light, a series of diffracted beams, or modes, will be reflected. The spacing between the modes depends on the illumination wavelength, A, and the finger pitch d. The angle 0,n between the nth mode beam and the specular (0th order) mode beam off fingers with pitch d is O, = sin- (2.1) 2d When all the fingers are in the same plane, all the even modes will be bright while all the odd modes will be dark. If the distance h from the fingers to an observation plane is large compared to distance of the nth beam from specular beam at the same plane, the latter is approximately xn = nhX 2d (2.2) for the lower order modes. If one were to gradually displace every other finger in the vertical direction, the intensity of each mode will change. As the offset approaches A/4, the even modes will vanish while the odd modes will appear. The intensity of each mode is a sine squared function of this offset z. For example, the intensity of the first mode is proportional to I1MODE sin (2.3) Therefore, by measuring the intensity of any of the modes, the offset between the fingers can be determined. Because only the intensity and not the position of these modes changes, the photodiode alignment is not critical. Moreover, the relationship between the offset and intensity can be well approximated as linear relationship. Figure (2.2) shows that the relative error between a linearization of a sine squared intensity about a finger offset of an eighth wavelength 18 and the actual intensity is less than 5% within a deflection range of a tenth wavelength. When red light is used, this linear range is over 600 A. 0.1 0.05 CU 0| Zi a: -0.05 I I -U. 1~ I 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.2 Finger Offset per Wavelength Figure (2.3). Error in linearizing sine squared intensity about eight wavelength offset The use of interdigitated fingers as a position detector was first applied to measuring the deflection of cantilevers for the atomic force microscope [12]. One set of the fingers was attached to the base of the cantilever while the other was attached to the tip. The noise of this position detector was found to be 0.02 A in a 10 Hz to 1 kHz bandwidth. Figure (2.3) is the position noise spectrum obtained by Manalis et al. This noise, dominated by the wavelength and phase noise of the laser at frequencies up to 1 kHz [14], is comparable to what is achievable with electron tunneling position detectors. 19 10'2 z 10*4. 10 100 1000 Frequency (Hz) Figure (2.4). Position noise of interdigital cantilever for atomic force microscopes. (Reprinted with permission from APL paper, 1996 [12]) 2.3 Noise analysis The resolution of all accelerometers is ultimately limited by thermomechanical noise, a white noise given by a _ aTM 8lkBfO =3.2lX10'J" 2 mQQ at room temperature (T (2.10) 300 K) where kB is Boltzman's constant and Q is the quality factor. Figure (2.4) shows the thermomechanical acceleration noise for a 30 mg proof mass with a 100 (typical of ambient conditions) at room temperature as a function of resonant frequency. 20 Q of 0.1 C 0.0 8 CU 0.01 - a> 10 1000 100 104 Resonant Frequency (Hz) Figure (2.5). Thermomechanical noise as a function of resonant frequency for a 3 0 mg proof mass with Q of 100 at 300 K. As mentioned Section 2.4, the dominant noise at frequencies under I kHz of the interdigital AFM cantilever is most likely caused by wavelength and phase fluctuations of the laser. This noise changes the intensity of the diffraction modes in the same way as changing the finger offset. The translation of this laser fluctuation to position noise increases linearly with the finger offset [11, 15]. For this reason, the wavelength/phase noise can be minimized by using lower order modes as well as operating with the fingers near the zero offset bias. Figure (2.5) is a plot of the equivalent acceleration of the interdigital cantilever position noise as a function of operating frequency for structures with various resonant frequencies. The equivalent acceleration was found by multiplying the position noise values in Figure (2.3) with the square of the proof mass angular resonant frequency. Recall from Equation (1. 1) that the acceleration of a proof mass is the product of its deflection and the square of its angular resonant frequency. Therefore, the equivalent acceleration of the interdigital position noise decreases with the resonant frequency of the proof mass. For example, a 30 mg, 100 Hz proof mass should have a wavelength/phase noise of 4 ng at 50 Hz which is lower than the thermomechanical noise of 6 21 ........ . . . ...... ng. Building a sensor that is limited purely by the thermomechanical noise would be extraordinary because the noise of most accelerometers is dominated by other sources such as the measurement circuit [16, 17]. I I I I I I I I I I I I I I1 100 1-. 10 C) CU) 1 Resonant Frequency 0.1 100 Hz 500 Hz 1 kHz 10 kHz CU 03 CT 0.01 I I I I 11111 , , , 1000 100 10 , , I 1 1 1 104 Operating Frequency (Hz) Figure (2.6). Interdigital noise as a function of operating frequency for various resonant frequencies. Above 1 kHz, the noise of the interdigital cantilever noise is dominated by shot noise. Shot noise is caused by a fluctuating current due to the random arrival of electrons across the pn junction in the photodiode. It is a white noise proportional to the root of the DC current iDC across the diode. iSHOT = 2 qioc (2.12) [A/rt Hz] where q is an electron charge. There are other sources of noise present in the interdigital accelerometer that are usually lower than the laser wavelength/phase and shot noise. These sources include the intensity, 22 pointing, and flicker noise in the laser noise and the flicker, Johnson noise, and input noise in the amplifier [15]. A crucial noise source not inherent to the sensor is the background seismic noise. This low frequency noise ranges between 1 pg to 0.1 ng depending on the geographic location [18]. Figure (2.6) shows the background noise of a floated optics table measured in our 4th floor laboratory in Cambridge, MA. 10 I I 1 C 0 .10 L. Ca) 0.1 U CI) 0.01 100 10 Frequency (Hz) Figure (2.7). Seismic noise in 4th floor laboratory. 23 3 DESIGN AND FABRICATION 3.1 Proof mass wafer 3.1.1 Design Laser Source Figure (3.1). Schematic of folded-pinwheel as an interdigital accelerometer proof mass. We choose to suspend the proof mass with diagonal, "folded pinwheel" springs as shown in Figure (3.1). A variation of this structure, in which the springs run parallel to the mass, has been previously studied and modeled [19] and is shown in Figure (A. 1). For motion normal to the plane of the device, the folded pinwheel is over two orders of magnitude more linear than straight tethers. Because the folded springs allow the mass to rotate slightly, it is less sensitive to 24 external strains. This makes it more robust than a straight tether design. Furthermore, the folded springs allow the fabrication of long, highly compliant springs without significantly increasing the size of the device. We choose to move the springs to 45 degrees to allow an area to attach the fingers as well as to reduce the nonlinearity of motion in the sensitive axis. The dimensions of the proof mass and springs were determined by the desired resonant frequency. As mentioned in Section 2.3, to obtain nano-g acceleration resolution that is limited only by thermomechanical noise, it was necessary to have a 30 mg, 100 Hz resonant frequency proof mass. Because these low resonant structures have narrow operating bandwidth and a slower response, we also designed for 500 Hz, 1 kHz, and 10 kHz resonant proof masses. For reasons that will be explained in Section 3.1.2, we chose to fabricate our devices with backetched silicon on insulator wafers (BESOI). The springs and fingers would be composed of only the thin device layer while the mass would be composed of all three layers. Through a generous donation from the Schmidt Group at MIT, we obtained BESOI substrates with a 20 ptm device layer, a 1 ptm buried oxide layer, and a 381.5 pm handle layer. Given the density of silicon, we calculated that we needed a 5.6 mm square to obtain a 30 mg proof mass. We then chose sensible values for the open space around the proof mass (140 pm) and springs (200 pm) and the width of the springs (140 pm). Figure (3.2) shows the layout of the proof mass dimensions. We entered the dimensions into the analytical model and ProEngineer to determine the length of the springs that would give us the desired resonant frequencies. Figure (3.3) is a plot of the length of the springs versus the resonant frequency obtained by Mechanica finite-element modeling (FEM) and the analytical model of the folded pinwheel with springs running parallel to the mass. The simulation and model agree well up to about 1 kHz resonant frequency. Near this resonant frequency, the extra length in the springs due to its 45-degree slant becomes significant, and the spring is actually less stiff than the model. For this reason, we choose to use the FEM lengths, given in Figure (3.2) for our design. The equations for the analytical model can be found in Appendix A. 25 Figure (3.2). Layout and dimensions of folded pinwheel proof masses. It is clear from the plot that achieving a resonant frequency above 4 kHz would be difficult given the size of our proof mass. For this reason we shrunk the proof mass to a 1.4 mm square and shortened the spring widths and gaps for the 10 kHz device. We then used FEM to obtain the required spring length given in Figure (3.2). 26 4000 I I I I I I I I I I I I I I I I I II I I I I I I 3500 - I I mecnanic FENv Analytical Model 3000 -c 2500 2000 U- CU C 1500 0 1000 500 0 0 500 1000 1500 2000 2500 3000 3500 Spring Length L (urn) Figure (3.3). Calculated folded pinwheel proof mass resonant frequencies versus spring length L Our FEM also allowed us to predict the maximum deflection due to gravity, higher resonant frequencies, and maximum load before failure. The second and third resonant modes compose of rotations about the diagonal of the proof mass. In theory, they occur at a factor of N- from the first resonant mode [19]. In Section 3.1.3, we will present the actual resonant frequencies. The failure of the proof mass was assumed to occur when the maximum stress (located where the tethers join the support substrate) reaches the yield stress of silicon, 2 GPa. Table (3.1) lists these three values for each of the proof masses. In the center of one side of the proof mass are fifty 175 pm x 6 pm x 20 pm interdigitated fingers as shown in Figure (3.2). Although the location of the fingers guarantees that their outof-plane separation will always change during rotation about a diagonal, it also ensures that the maximum separation change will be less than if the fingers were located near a set of springs orthogonal to the diagonal axis of rotation. The fingers are twice as wide and half as long as the ones on the original interdigital accelerometer to eliminate breakage and stiction. The fingers on the proof mass overlap with the fingers on the support substrate for 125 pm, giving an area of 27 450 pm x 125 pm in which to focus the laser. This area was an ample target for the 75 pm focused laser diode spot sizes we could achieve. 1st Resonant Max Deflection in 2nd & 3rd Resonant Frequency Gravity (um) Frequency 100 31 200Hz 27 500 1.2 900 Hz 67 1.0 kHz 0.37 1.7 kHz 100 10 kHz 0.25 16 kHz 227 Load at Failure (g) Table (3.1). Gravitational deflection, higher resonant frequencies, and load at failure values calculated by Mechanica for each proof mass design. Another advantage of the folded-pinwheel design over the cantilever structure in Figure (2.1) is the reduced deflection and tilt of the fingers due to gravity. This deflection causes premature breakage, increases the effect of laser wavelength/phase noise, and limits the angle of tilt of the sensor. In Appendix A.2, we calculate that the static deflection of a 100 Hz cantilever proof mass is 46 um, or about 33% more than a folded pinwheel proof mass of equal resonant frequency. More importantly, the tilt of the cantilever proof mass with respect to the substrate would be almost half a degree. Any angle between one set of fingers and the other degrades the diffraction pattern off the array. On the other hand, the folded-pinwheel proof mass will always be parallel to the substrate in the absence of higher-order rotations about the diagonal. 3.1.2 Fabrication Process The first interdigital accelerometer proof mass was fabricated in a two mask, CMOS compatible process that had low reproducibility and yield. It consisted of two timed deep reactive ion etches (DRIE) in a plain silicon wafer and an acetone release. We used the same process with two improvements. First, a buried oxide layer in the substrate was used to provide a consistent etch stop for the DRIE, which is 100 times more selective for silicon than for oxide. This decreases the sensitivity of the final spring dimension to variations in etch rate across the wafer and on different wafers. The second improvement was a polyimide film to serve as a support for the delicate springs and fingers during the subsequent DRIE etch through the handle 28 silicon to define the proof mass. The low pressure in the DRIE chamber can create a destructive pressure differential across the thin springs. The polyimide also allows the structures to be released dry, without the breakage and stiction problems accompanying the previous acetone release. Figure (3.4) shows the process cross-sections. See Appendix B.1 for mask design details. The proof mass wafer process started with a 100 mm, single-side polished BESOI wafer with a 20 ± 1 pim device layer, a 1 + 0.05 jim buried oxide layer, and a 381.5 ± 0.5 im Si handle layer. First, 0.56 pm of thermal oxide was grown as shown in Figure (3.4a). This oxide was then patterned with Mask 1. After developing the resist, the exposed oxide was plasma etched to clear all the oxide features to the silicon. The resist was then removed. Next the exposed device layer silicon was then attached to a 150 mm quartz carrier wafer because the DRIE etcher only accepts 150 mm wafers. The exposed device layer silicon then was deep reactive ion etched all the way to the buried oxide layer as shown in Figure (3.4b). After the DRIE etch, we released the wafer from the carrier in acetone and stripped the resist in piranha. We then coated 20 pm of polyimide on the device layer side as shown in Figure (3.4c). Next we patterned the backside of the wafer with Mask 2 and attached (with the backside exposed) the wafer to a 150 mm quartz handle wafer. The handle silicon layer was then deep reactive ion etched until the only the buried oxide layer was exposed under all the features. At this point, shown in Figure (3.4d), the springs and fingers on the other side are visible through the oxide. The wafer was then separated from the carrier in acetone. We then removed the I pm of buried oxide in an 18-minute buffered oxide etch (BOE) as shown in Figure (3.4e). The wafer was then mounted with the polyimide side up on a 100 mm silicon carrier using a 5 mm ring of 10 jim-thick resist on the wafer edge. Finally, we etched away the polyimide in an oxygen plasma. This last step, shown in Figure (3.4f), releases the proof mass and allows the dies to be removed by only breaking the twelve 200 jm x 100 pm x 20 pm tabs. Appendix C. 1 lists the proof mass process steps in greater detail. 29 d.) a.) a.) d.) Li ___M b.) e.) C.) f.) El Li Silicon Silicon Dioxide Cross-section on proof mass Polyimide Figure (3.4). Fabrication process sequence for interferometric accelerometer proof mass wafer. a.) Thermal oxidation of SOI wafer. b.) Topside lithography and DRIE etch to define fingers, springs, and mass. c.) Spin-on and cure polyimide. d.) Bottomside lithography and DRIE etch to define mass. e.) BOE etch of box oxide. f.) Ash polyimide in oxygen plasma. 30 3.1.3 Fabrication results Figure (3.5). 5x micrograph of 10 kHz proof mass corner. Figure (3.6). 1 Ox backlit micrograph of interdigitated fingers showing release. 31 Using the process outlined above, we released proof masses with 100% yield. Figure (3.5) shows a 5x micrograph of the corner of a 10 kHz proof mass. Figure (3.6) is a backlit lOx magnification of the 6 um interdigitated fingers showing all fifty fingers perfectly intact and released. In Section 3.1, we used FEM to predict the resonant frequencies of the proof masses. We measured the resonances by looking at the peaks in the frequency spectrum of the first diffraction mode intensity in the tabletop set-up. See Section 4.3 for details on tabletop experimental set-up. Table (3.2) compares the measured resonant frequencies to their design values. The measured frequencies deviated from the design values due to deviations between the actual and expected mask and wafer dimensions. Process variations had less effect since the variation in resonant frequencies from wafer to wafer was about 5%. As predicted by theory, the rotational resonances occurred near a factor of V[ from the natural frequency. More importantly, these higher order resonances occurred well beyond our frequency range of interest for each proof mass. Design 1st Resonant Actual 1st Resonant Design 2nd & 3rd Actual 2nd & 3rd Frequency Frequency Resonant Frequency Resonant Frequency 100 Hz 80 ±4 Hz 200 Hz 150 Hz 500 Hz 430 + 10 Hz 900 Hz 780 Hz 1.0 kHz 1020 1 50 Hz 1.7 kHz 1.6 kHz 10 kHz 8.2+0.1 kHz 16 kHz 14 kHz Table (3.2). Design and measured resonant frequencies 32 3.2 Packaging 3.2.1 First generation package Laser diode Figure (3.7). Cross-section of first generation package We designed a plastic housing to hold the interdigital accelerometer with a laser diode and photodetector. Figure (3.7) shows a cross-section of this package. The first generation package consisted of four acrylic pieces cut with a CO 2 lasercutter as shown in Figure (3.8). With the exception of the top piece, which is 1/4-inch thick, all the pieces are cut from 1/8-inch thick cast black (for light exclusion) acrylic. The bottom piece holds the device die and has a hole under the proof mass to allow it to oscillate. The top piece has a hole to hold a plastic collimating lens (Thorlabs CAY033) and a 670 nm, 5mW, laser diode (Sanyo DL3149-056) whose source is placed at the focal length of the lens. The red laser was chosen to maximize diffracted mode spacing (See Equation (2.2)) and allow visual alignment. Attached to this piece is a silicon die with pn junction diodes located around a 125-pm diameter DRIE through-hole. This photodiode wafer is discussed in Section 3.3. The diodes are wire-bonded to two alumina plates (Thermomicroscope 30-600-0110) nearby. Electrical leads carrying the output signal are 33 soldered to these plates. The middle two pieces form a 1/4-inch (6.7 mm) shim to allow the diffraction modes to spread and be differentiated from each other. There is one shim attached to both the top and bottom pieces. The separation of 6.7 mm corresponds to a mode of spacing of 250 pim. A C B D Figure (3.8). Actual size schematic of the first generation package pieces. (A) is the bottom piece on which the proof mass rests. (B) is the bottom shim. (C) is the top shim with rectangular holes for the wirebonds and electrical leads. (D) is the top piece with a circular hole for the laser diode, a circular recess (red) for the plastic lens, and a rectangular recess (yellow) for the bonding plates. The blue lines represent shallow grooves to align the proof mass die (A) and photodiode die (D). The assembly of this package involves three alignments after the necessary components have been fixed onto their corresponding piece. First, the two halves are aligned parallel and as close together as possible without touching (<0.5 mm). Second, the laser diode is positioned into the hole in the top piece such that its reflection off the accelerometer is directed back into the hole in the photodiode die. Lastly, the bottom piece is moved until the laser spot is illuminating the interdigitated fingers, and the diffraction modes are incident on the correct photodiodes. Then the two halves are cemented together. The completed package is 2.33 cm x 1.84 cm x 1.67 cm (7.2 cm 3 ) and weighs about 8 g. It has five necessary leads: three for the laser diode and two for the photodiode. After testing the first package, we found that the noise was two to three orders of magnitude greater than the interdigital cantilever noise in Figure (2.3). We discovered that there was not enough light getting through the hole in the photodiode wafer. Figure (3.9) shows how the sensor signal-to-noise ratio of an 8 kHz proof mass driven at constant, 4 kHz acceleration diminishes with incident power due to the low sensitivity. 34 0.8 E C14 Ce) E v 0.6 N E 0 0.4-- z 00 iT 0.4 0 0.01 1 0.1 Incident Power (mW) Figure (3.9). Normalized signal-to-noise ratio versus incident power for 8.2 kHz sensor driven at 4 kHz The focal lengthf of the collimating lens and angular divergence of the laser diode beam 0,s, gave a collimated beam radius of r,,, = ftan,, = 3.33mm- tan80 = 468pm (3.1) This width is 56 times the area of the 125 pm hole in the photodiode. This agrees with our finding that the power on the other side of the hole was diminished by a factor of 30. For this reason we abandoned the first generation package. Results from a 428 Hz sensor packaged in this manner are shown in Appendix D. 35 FN I - _-E iR- - ,- - - - - - - 3.2.2 Second generation Dackane Output Epoxy Laser diode hotodiodesW wire Wirebond Accelerometer f ingers Side view Front view Figure (3.10). Second generation package. The second generation package used a similar four-piece housing that aligned the proof mass die parallel to the photodiode wafer. However, instead of bringing the light through the hole we directed it in from the side at a small angle (about 250) as shown in Figure (3.10). This time we used a faster (shorter focal length) glass lens (Thorlabs 350140-B) with a focal length of 1.45 mm. When attached directly to the can of the laser diode, the lens brought the light to a focus around 15 mm away from the lens. This meant that the laser source was s =x+f -WD =x+1.45mm-0.88mm = (f - =( I-- S") 1.45mm I ,= 1.6mm =>x = I.mm(32 15mm away from the tip of the can, agreeing with the laser diode manufacturer's value of 0.93 + 0.05 mm. The distances in Equation (3.2) are shown in diagram of the laser diode-lens system in Figure (3.11). More importantly, this diode-lens system gave us a pre-focus beam width of less than 220 pm. r,,,, s tan0,,, =1.55mm -tan8 0 = 218gm 36 (3.3). This meant that we could fit the entire beam in the 450 pm width of the finger region. If the beam were to exceed the width of the fingers, the specular reflection would overlap into the 1st mode, increasing the noise. With this new package design we can deliver over 3 mW to the laser diode, increasing our signal-to-noise ratio compared to the first generation package. Lens Center s"t Laser Diode r7:' source ' Lens f= focal length - f-WD S Figure (3.11). Laser diode-lens system for generating a small beam. The four acrylic package pieces are shown in Figure (3.12). Like the first generation package, all the pieces are 1/8-inch thick except for the 1/4-inch thick top piece. Because the 3 dies are offset from each other and the laser diode is outside, the total volume, 8.6 cm (2.33 cm x 2.20 cm x 1.67), is larger than its predecessor. This new package still weighs about 9 g. A C B D Figure (3.12). Actual size schematic of the second generation package pieces. (A) is the bottom piece on which the proof mass rests. (B) is the bottom shim. (C) is the top shim with rectangular holes for the wirebonds and electrical leads. (D) is the top piece with a semicircular cut-out for the laser diode and a rectangular recess (yellow) for the bonding plates. The blue lines represent shallow grooves to align the proof mass die (A) and photodiode die (D). Although the hole in the photodiode wafer was no longer needed, we kept with our custom photodiodes because the exposed active areas were easier to align to and had better 37 signal-to-noise ratios than a glass-covered commercial photodiode of equal area (United Detector Technology HRO08). The assembly of the second generation package followed the same procedure as the first generation except that the laser diode was glued to the outside of the package instead of behind the through-hole in the photodiode die. The results in Section 4 are taken from sensors in second generation packages. 3.3 Photodiode wafer 3.3.1 Design The photodiode wafer was designed around the first generation plastic housing. Figure (3.13) shows the layout of the photodiode die. We placed four pn junction diodes around a 125pm through-hole in the positions of the -1st, +2nd, -3rd, and +4th modes when placed 6.7 mm away from the fingers (giving a 250 pm mode separation). The rationale behind the multiple photodiodes was the ability to test multiple modes at once and to perform a subtraction of two adjacent photodiode signals to lower the laser intensity noise [12]. We found the first mode from specular to be the most sensitive, and hence all the results in Chapter 4 are from the measurement of this mode. Furthermore, subtraction of the signal from two adjacent modes did not increase the signal-to-noise ratio because the laser intensity noise was not the dominant noise source. The area of each photodiode was 200 pm wide and 150 pm long plus trapezoids on the top and bottom to extend the active area to cover a 540 arc around the hole, allowing for theta misalignment with the fingers. Each photodiode has an anode contact that extends along 200 pm wide traces to a 1.6 mm square bond pad. 38 __ - M IN _ _ __ - - __ - - I Figure (3.13). Layout of photodiode die 3.3.2 Fabrication Process The photodiode wafer fabrication was a four mask CMOS compatible process. See Appendix B.2 for mask design details. Figure (3.14) shows cross-sections of the wafer during various steps in the process. The process to create the pn junction diode was borrowed from the Microelectronics Processing Technology class (6.152J) at MIT. The diodes are formed by implanting boron dopants into an n-type substrate. The starting material was a 100 mm, 525 pmthick, single-side polished (100) n-type (phosphorus doped) silicon wafer with resistivity of 1.0 ohm-cm. The process began with growing 560 nm of oxide and pattering with Mask 1 to expose the diode regions (Figure (3.14a)). The wafers were then sent to Implant Sciences (Wakefield, MA) to have boron implanted (Figure (3.14b)). When they returned, we annealed the wafers to drive-in and activate the dopants (Figure (3.14c)). The oxide over the contacts were then etched away (Mask 2) leaving only the lines on which the anode interconnects would run (Figure (3.14d)). Then 1 gm of aluminum (1% Si) was deposited (Figure (3.14e)), patterned with Mask 3, and sintered (Figure (3.14f)). Finally, the through-holes were patterned on the backside with Mask 4 and etched with the DRIE (Figure (3.14g)). Appendix C.2 lists the photodiode process steps in greater detail. 39 d.) a.) e.) C.) g.) n-type Silicon Silicon Dioxide p-type Silicon Aluminum Figure (3.14). Fabrication process sequence for photodiode wafer. a.) Topside lithography on thermal oxidation to open field oxide. b.) Implant dopants. c.) Dopant drive-in/activation and reoxidation. d.) BOE etch to open contact holes. e.) Deposit aluminum. f.) Pattern aluminum. g.) DRIE etch through-hole. 40 3.3.3 Electrical properties We measured the sensitivity of our custom photodiode by focusing a 670 nm laser diode on the active area and measuring the output through a current amplifier (106 V/A gain) while varying the incident power. The intensity was varied by using neutral density filters and by adjusting the laser current and calibrated using a power meter (Coherent LM+). Figure (4.6) is a plot of the output versus the incident power. The slope gives a sensitivity of 0.33 A/W which is comparable to commercial photodiodes. 1.4 I - I I -I I.I I 1.2 - - 1 E - - 0.8 4-a 0. 0.6 - - 0.4 - 0.2 -S 0' 0 0 I 0.5 1 1.5 2 2.5 3 3.5 Incident Power (mW) Figure (3.15). Plot of custom photodiode output versus incident power 41 4 4 RESULTS 4.1. Signal conditioning The output of the photodiode used to detect changes in diffraction beam intensity was amplifier with a commercial low noise current amplifier (Keithley 428). Previously we dropped the photocurrent across a large resistor but we found that the voltage across the resistor saturated around 0.4 V. This is because resistors have a limited voltage compliance [20]. The transimpedance amplifier, on the other hand, faithfully converts the photocurrent in a signal as well as allows bandpass filters and voltage biases. We operated our photodiodes with zero bias because we did not need responsivities greater than 100 ps (10 kHz) and the negative bias would increase the dark current. The gain was set between 104 and 105 V/A, which was adequate for amplification above the noise of the devices down the line such as an FFT spectrum analyzer (Stanford Research Systems SR760) and a DSP lock-in amplifier (Stanford Research Systems SR830). Furthermore, the 10 MQ resistance of the photodiodes restricted the gain to less than 107 V/A since the propagation of voltage error across the current amplifier increases with the feedback resistance or gain. Voieu = V110gi 1+ Rfedbak Rsource) (4.1) The laser diodes were driven by an ultra low noise driver (ILX Lightwaver LDX-3620) set in constant power mode. The laser current was set around 55 mA, which gave a laser output of about 3.3 mW. 4.2 Mechanical drive stage In order to test our accelerometer, we needed a system to impart known accelerations. First we tried a permanent magnet, electro-dynamic shaker (LDS V456 driven by LDS PA1000) but we found that voltage or frequency changes in the signal generator could cause the shaker to impart large acceleration impulses that could destroy a low frequency device. For this reason, we built a piezoelectrically actuated shaker like the one in Figure (4.1) using five piezoelectric 42 stacks (Thorlabs AE0505D16). The stacks were mounted horizontally between a 2 mm-thick aluminum plate with 10-32 tapped holes and right angle plate (Thorlabs AP90). The piezostacks were connected electrically in parallel and driven by a function generator (Stanford Research Systems DS345). On the aluminum plate, we mounted our test sensor along with an Analog Devices ADXL105JQC capacitive accelerometer as a reference. This latter device measures accelerations up to 12 kHz with a 250 mV/g sensitivity and a noise density of 225 pg/rt Hz. Off to the side of the stage we mounted a Wilcoxon 731 seismic piezoelectric accelerometer to measure seismic disturbances down to 3 ng between 0.05 and 400 Hz with a sensitivity of 10 V/g. All sensors were mounted with their sensitive axes in the direction of stage actuation. Both reference signals were filtered and amplified with a low noise voltage amplifier (Stanford Research Systems SR560). Direction of actuation 10-32 tapped holes Outline of piezostack behind plate Aluminum plate Piezostack Front View Side View Figure (4.1). Schematic of piezoshaker used to actuate the accelerometers. 4.3 Package assembly Before we packaged an accelerometer, we had to verify that it was released and obtain a benchmark resolution. To do this, we tested the performance of the proof mass in a large tabletop set-up, repeating the experiments performed by Cooper et al. We attached the proof mass into the two bottom pieces of the package and mounted the half-package to the shaker, oriented such that the fingers point horizontally. On the fingers we focused a 3.3 mW laser beam with a 15.36 mm focal length aspheric lens (Thorlabs C260TM-B). The intensity of the first mode was measured from a distance of 14 cm away from the interdigitated fingers with a 3.6 mm 43 x 3.6 mm silicon photodiode (Thorlabs FDS 100). The tabletop set-up is shown in Figure (4.2). This put the modes at a well-separated distance of 5.2 mm from each other. An increased sensitivity represented by the presence of a large voltage peak near the expected resonant frequency on the spectrum analyzer indicated that the device was released. We then drove the sensor with sine waves from function generator and measured the signal-to-noise ratio at various frequencies. mass package Commercial Proof Photodiode Laser diode Tabletop set-up = XYZ positioner Photodiode package Packaging set-up Figure (4.2). Diagram of tabletop and packaging experimental set-ups. After obtaining the best case resolution of the sensor, we assembled the package. We attached a photodiode wafer to the top package piece and made the necessary gold wire bonds and solder connections. After attaching the top shim, we positioned the optics half-package less than half a millimeter away from the proof mass half-package. We then attached a focusing lens to a laser and positioned the two by the open side of the package piece as shown in Figure (4.2). The pieces were aligned until the laser illuminated the fingers and the -1st mode was incident on a photodiode. Then we repeated the resolution measurements, using the reference accelerometer 44 to ensure equal drive acceleration. At this point we could adjust the position of the laser and photodiodes to maximize the signal-to-noise ratio. While packaging the sensor, we learned how to handle the trade-off between the focus of the incident beam and of the diffraction modes. High sensor sensitivity requires high laser power incident on both the fingers and the photodiodes. Maximizing incident power on the fingers means placing them near the focus. However, a focused beam on the fingers results in a defocused beam at the photodiodes. We solved this problem by moving the fingers away from the focus toward the laser until the spot size was fully contained in the interdigitated region. Then we aligned the largest photodiode on the die to measure the defocused -1st mode. The maximum signal-to-noise ratio was achieved in this configuration. 60 Tabletop x 50 -. N Package ID Limit - 0e 40 x cn 0 30 x - 20 10 ' 20 0 300 500 400 600 700 800 Frequency (Hz) Figure (4.3). Comparison of resolution of 8.2 kHz sensor in tabletop and package set-ups. Once the resolution of the package was near that of the tabletop set-up, we gently applied quick setting epoxy around the package and allowed it to cure. Figure (4.3) shows a comparison 45 of the resolution of the 8.2 Hz sensor in both the tabletop and package set-up as well as the interdigital detection limit set by the laser wavelength/phase noise extracted from Figure (2.3). It is evident that the accelerometer did not lose much resolution going from the tabletop to our package. Furthermore, the noise of the package is only about a factor of 3 greater than the interdigital limit. This is the procedure with which we packaged a 1020 kHz, 432 Hz, and 80 Hz sensor in the second generation housing. Table (4.1) shows the resolution of each sensor in each stage of the packaging process. The resolution would often increase by a factor of 2 to 3 after the package was sealed because the laser diode, proof mass, and photodiode would be fixed with respect to each other. Resonant Drive Tabletop Noise Prepackage Postpackage Frequency (Hz) Frequency (Hz) (ptg/rt Hz) Noise (pg/rt Hz) Noise (pg/rt Hz) 1020 400 0.7 0.7 0.2 432 160 0.8 0.6 0.2 80 40 0.1 0.1 0.1 Table (4.1). Noise level of each package during the packaging process 4.4 Sensor Performance 4.4.1 Sensitivity Like most accelerometers, the interdigital sensor can be modeled by the mass suspended by a spring and damper in Figure (1.1). The second order transfer function between the acceleration A(l) and the proof mass position Z(1) is Z(f) A(f) (4.2) 1 4_ 2 2 f _f 2 )+j(ff 0 /Q)] Movement of the proof mass due to acceleration changes the intensity of the diffracted beams. The intensity change is proportional to the photocurrent from the photodetector which in turn is proportional to the voltage V(1) at the output of the current amplifier (to the frequency roll-off of the amplifier). Assuming the relationship between the proof mass position and mode intensity is 46 linear, we can use a constant, G, to account for this constant of proportionality as well as the gain of the current amplifer. Then the transfer function between output voltage and acceleration is V(f) GZ(f) A(f) A(f) G 4 2[(f 2f 2 )+ j(ff 0 /Q)] The real magnitude of the transfer function is then A(f) 44 G V(f) 47t2f4+(1/Q2 2)f2f2+f 4] Equation (4.4) is the transfer function that defines the sensitivity of the interdigital accelerometer. We measured the sensitivity of the packaged interdigital accelerometers using a LabVIEW program which controlled the function generator to drive the sensor at incremental frequencies. At each frequency, the stage is driven at two different accelerations and a lock-in amplifier extracts the component of both the test and reference accelerometer signals at that frequency. The sensitivity was calculated by taking the ratio of the change in sensor signal to the change in acceleration. We then fit a function with the form of Equation (4.4) to the sensitivity data points to obtain the Q of the device. Figure (4.4) shows the sensitivity of the 1020, 432, and 80 Hz resonant frequency packages assuming current amplifier gains of 105 V/A (we measured the highly-sensitive 80 Hz device with a gain of 104 V/A). Table (4.2) lists the parameters in the second-order transfer function. The comparable values for the gain G demonstrate that the intensity of the modes in each package is of equal magnitude. This is also evident in the fact that the low frequency sensitivities scale with the inverse of the square of the resonant frequency. This result suggests that the packaging is highly reproducible. 47 SI 1000 Measured Fit (Q=1 10) - c,) 100 % S C,) C ci) Co 10 - - 1 a -- - - 5005 w 0 V 4 0 W-19 * * >4 1000 100 (a) Frequency (Hz) Measured Fit (Q=185) _ 1000 C C - 100 - 10 1 1 * 1 (b) 400 300 200 70 80 90100 ---. 500 600 Frequency (Hz) I II I Measured Fit (Q=68) C) 104 C 1000 - 30 (c) 40 I II- 70 60 50 Frequency (Hz) 80 I- 90 100 Figure (4.4). Sensitivity of (a) 1020, (b) 432, and (c) 80 Hz packaged sensors. 48 Resonant Frequency Quality Factor Low Frequency (Hz) G (V/m) Sensitivity (V/g) 1020 110 7 1.4x10 8 432 185 30 2.4x10 8 80 68 1300 2.2x10 8 Table (4.2). Parameters of fitted sensitivity transfer functions. 4.4.2 Noise spectrum The noise spectrum of the sensor was measured by taking the FFT of the signal when the sensor was not being actuated. For this measurement, we move the shaker onto a small platform mechanically isolated by bubblewrap which has slightly lower seismic noise than the floated table alone. The voltage noise was converted to an equivalent acceleration by dividing the spectrum by the fitted sensitivity function. Figure (4.5) shows the noise equivalent acceleration of the 1020 kHz and 432 Hz sensors including the measured seismic and shot noises and the theoretical interdigital detection and thermomechanical noises. The interdigital detection noise in these sub-kHz frequencies is dominated laser wavelength/phase noise. Not shown is the measured photodiode/amplifier noise which is usually 1 to 2 orders of magnitude below the thermomechanical noise. The photodiode/amplifier noise is the noise from the sensor when the laser diode is off. The noise of the 1020 Hz sensor is limited up to 50 Hz by the seismic disturbance. At higher frequencies, this sensor appears to be limited by the theoretical laser wavelength/phase noise. This suggests that the interdigital deflection noise in our package is similar to that of the interdigital AFM cantilever. The 432 Hz accelerometer, whose laser noise has a lower equivalent acceleration, is completely limited by the seismic disturbance. To determine its true self noise, one would have to obtain two similar 432 Hz packages and determine the coherence of their noise signals when mounted next to each other. Correlated noise points to the seismic noise while uncorrelated noise corresponds to the self noise of each sensor [4]. The use of this technique is beyond the scope of this thesis. 49 W- I 0.1 I A- I I ;0- -Sensor -- Seismic I I I II --------- ID Limit -------- ThermomEechanical Shot E 0) 0,29111" 0.01 0.001 w - A - v v W 0.0001 z i 10 i I I I I I I Frequency (Hz) Sensor --Seismic --------- ID Lim it -. .-----Thermomechanical Shot N 0.01 E C 0 0.001 Cu L. a) CD) .) I 100 10 (a) i 0.0001 (b) 105 L------- ------ - ----- --M-------- -------- ----------- - 100 10 Frequency (Hz) Figure (4.5). Noise spectrum of (a) 1020 and (b) 432 Hz packaged sensors 50 =MM Figure (4.6) is the noise spectrum of the 80 Hz sensor. The noise of this sensor, is limited by seismic disturbances. Like the 432 Hz sensor, a coherence measurement on the 80 Hz sensor will also be needed to determine its self noise limit. The 80 Hz noise spectrum shows that we may be able to measure the thermomechanical noise of the proof mass because it is greater than the laser phase noise above 10 Hz. At 40 Hz, the thermomechanical noise is 6.8 ng while the interdigital noise is 2.8 ng. 0.1 N r 0) E 0.01 C 0 0.001 Sensor Seismic 0.0001 ID Limit --------- Thermomechanical Shot 10-5 _--------------------------- z 10 6 ' ' ' I ' 10 1 Frequency (Hz) Figure (4.6). Noise spectrum of 80 Hz packaged sensor. 4.4.3 Linearity The linear range of the accelerometer was tested by driving the shaker at a set frequency with increasing amplitude and measuring the output on the FFT. Recall that the intensity of each diffraction mode is a sine squared relationship with the finger offset. At a high enough drive acceleration, the output signal no longer resembles the sine wave drive signal because of this nonlinear relationship. Figure (4.7) shows the output linearity of the 1020, 432, and 80 Hz 51 packaged accelerometers while Table (4.3) summarizes their linearity in acceleration and equivalent deflection calculated using Equation (1.1). The linear deflection ranges agrees with our analysis in Section 2.2 that the interdigital sensor is only linear to about a tenth of the wavelength of the light source, or 67 nm. We can determine the open loop dynamic range of the 1020 Hz device by taking the ratio of the linear range to the resolution. At 400 Hz, the dynamic range of the 1020 Hz sensor is over 7x 05. Because we cannot determine the true self noise of 432 Hz and 80 Hz sensors with the use of coherence techniques, we can only put a lower limit on their dynamic range. The sensor dynamic ranges are also listed in Table (4.3). Device Resonant Drive Linear Acceleration Linear Deflection Dynamic Frequency (Hz) Frequency (Hz) Range (g) Range (nm) Range 1020 400 0.15 36 7x105 432 160 0.04 54 >2x10' 80 40 0.001 39 >Ix104 Table (4.3). Acceleration and deflection linearity and dynamic rnage of packaged interdigital accelerometers. 52 700 600 - 400 Hz Drive 500 - C. M 400 - 300 0 200 100 - (a) 0.15 0.1 0.05 0 Drive Acceleration Amplitude (g) 800 . . , , 700 - 160 Hz Drive 600 600 --500 -400 o 300 200 - 100 0.02 0.01 Drive Acceleration Amplitude (g) 0 (b)0 I I I 35 0.03 I 40 Hz Drive 30 25 20 3 o 15 10 5 * 0 (c) 0 0.0008 0.0004 Drive Acceleration Amplitude (g) 0.0012 Figure (4.7). Output versus acceleration of (a) 1020, (b) 432, and (c) 80 Hz packaged sensors. 53 4.4.4 Cross-axis sensitivity and drift Figure (4.8). Illustration of cross axes. Input accelerations with components in directions other than the sensitivity z-axis can cause the finger offset to change due to proof mass rotations about its diagonal. As mentioned in Section 3.1.1, the two rotational resonant frequencies occur at a factor of V3 from the first resonant frequency. We measured the sensitivity of the accelerometer due to inputs in the axis across the finger length (transverse) and along the finger length (longitudinal) by mounting the sensor in different orientations on the shaker. The transverse and longitudinal axes are illustrated in Figure (4.8). Because we did not, use another reference accelerometer to ensure that the test sensor was not being accelerated in its sensitive axis, the actual cross-axis sensitivity may actually be lower that our result. Figure (4.9) shows the upper limit on the low frequency crossaxis sensitivity of the 432 Hz packaged accelerometer compared to the z-axis sensitivity. The plot shows that the folded pinwheel proof mass is at least 50 and 250 times more sensitivity in the z-axis than the transverse or longitudinal axes, respectively. 54 1 0 im Im A - - 10 4_-- - -- - - 10K Z-axis * x Transverse Longitudinal o 1 4-I x (~) C ci) x x (I) x x xx 00 0.1 00 70 80 90 100 Frequency (Hz) Figure (4.9). Maximum cross-axis sensitivity of 432 Hz packaged sensor compared to z-axis sensitivity. Low drift is an important characteristic of accelerometers used for inertial guidance. To measured the drift of the sensor, we recorded the zero-g bias for 30 minutes in a closed room. Figure (4.10) is a plot of the drift of the 432 Hz packaged accelerometer. This plot shows a maximum drift rate of 7 ptg/min and an overall drift of less than 30 pg. Commercial inertialgrade accelerometers have drifts on the order of + 1 mg over the span of one year. Longer tests in a temperature-controlled environment will have to be performed to fully characterize the drift of our sensor. 55 I.,RNIIIJIgN I ' ' ' ' I ' I I I II I I I I I I I I I I I 4 10~ CD .0 C 2 10~ E 0 0 +4- C: 0 -2 10~ ~ IP ' '!j F r 1 ~~ -4 10~ I 0 I I 5 I I I I i i 10 i I I 15 L - 20 25 Time (min) Figure (4.10). Thirty-minute drift of 432 Hz packaged sensor. 56 30 5 CONCLUSION The motivation for this thesis was to explore the use of an interferometric position detector in a low-volume, high-resolution accelerometer. Current efforts to fill this commercial void involve using electron tunneling to make very sensitive measurements of low resonant frequency proof masses. We designed and micromachined accelerometers based on optical interference off interdigital fingers because this transducer has the advantages of a larger open loop dynamic range and simpler fabrication process than tunneling. After developing a technique for releasing fragile structures, we released 80 Hz, 430 Hz, 1.0 kHz, 8.2 kHz interdigital proof masses with less than 5% variation in the resonant frequency due to processing. We lasercut acrylic 8.6 cm3 packages that integrate a proof mass with a laser diode and a photodiode. Packages with 80 Hz, 432 Hz, and 1020 Hz proof masses demonstrated sensitivities that scaled with the inverse square of the resonant frequency, a result that supports the reproducibility of the packaging process. The 1020 Hz package achieved a resolution that equaled the position resolution of the interdigital AFM cantilevers and an open loop dynamic range of over 105 . Analysis of the coherence of two sensors would be needed to determine the true self noise and dynamic range of the 432 Hz and 80 Hz packages, sensors which are both limited by the background seismic noise. Future work on this accelerometer would involve a more extensive study of the cross-axis sensitivity and the drift. The package volume can be further reduced by using a vertical cavity surface emitting laser (VSCEL) fabricated into a smaller photodiode die. A fabrication process that aligns and bonds an entire wafer of such dies to the proof mass wafer could reduce the cost of and time to assemble one package. In this scenario, the packaging would only involve making the electrical connections and sealing out the environment. The existing application of our accelerometer would be to quantify seismic disturbances, a measurement useful for vibration sensitive experiments. Other applications would be in triangulation, or the use of a network of three accelerometers to pinpoint a seismic disturbance such as a finger tap on a board or a person moving in a room. 57 MECHANICAL MODELS A A.1 Folded pinwheel analytical model connecting tether main tether E = modulus of elasticity L G = shear modulus g = acceleration of gravity Is WI d = displacement for 1 g z-axis input main tether moment of inertia I= when bent in z direction I' = main tether moment of inertia when bent in sideways main tether polar moment of inertia J= 12 A T - = connecting tether moment of inertia when bent in z direction IUI t 12' = connecting tether moment of inertia when bent sideways z-axis Figure (A. 1). Diagram and definition of parameters of folded-pinwheel structure with springs parallel to mass. The following analytical model can be found in detail in Mitchell Novack's Master of Science thesis [19]. Figure (A. 1) is a diagram of the folded pinwheel design with all its relevant dimensions and parameters. We used this model to calculated the resonant frequency of the oscillation in the z-direction. The resonant frequency is found by looking at the deflection due to a 1 g input in the z-direction. - k g (2irf0)2 _igiff, "2 (A.1) 58 The total displacement is the sum of three displacements that result from separate effects. The first is due to the bending of the main tethers and is equal to MgL3 24EII r 8GJ 2L+4EIs+6GJL 2GJ 2 L +EIs (I A.2) The second deflection is created by the torsion of the main tether and is equal to d2 Mgs2 L A.3) 8GJ, Lastly, the third deflection is caused by the bending of the connecting tether and is equal to ( k.4) d3 = Mgs 48EI2 For completeness, the moment of inertia formulae are given below. .tb3 1 = 12 bt3 12 il = I, +II wt3 12 = 12 . tw3 12 J2 = 12 + I2 (A.5) (A.6) Knowing the three deflections, the resonant frequency can be calculated. 27c (A.7) d,+d 2 +d 3 Using this model, one can obtain resonant frequencies that agree with a well-built simulation to better than 5%. 59 A.2 Cantilever proof mass deflection We will determine the deflection of a 100 Hz silicon cantilever proof mass like the one in Figure (2.1). The structure can be modeled as a massless cantilever with a bending moment at the end equal to the product of the weight of the mass mg and the beam length L. If the proof mass is a 1000 pm square that is 525 tm thick and weights 1.2 mg, the length of a 1000 gm-wide (W), 20 gm-thick (H) supporting cantilever would have to be EWH_ 1/3 1/3 L =EWH (160x109N/m2 2 167 f ) X3 (A .8) 1/3 -1x10-3 m -(20 x10 6 -(100/s)2 .1.2x10-6 kg where E is the modulus of elasticity of silicon, 160 GPa, and k is the spring constant of the cantilever. The static deflection at the end of the cantilever due to gravity would then be [21]. 6mgL 3 3 EWH A* 6.1.2 x 106 kg- 9.8m/s 2 _(8.8 x 10160x10 9 N/m 2 i) 3 -1x10-'m.(20x10-6 M) - 38gm (A.9) 3 Moreover, the tilt of the proof mass fingers with respect to the substrate fingers would be - 12mgL 2 EWH 3 12-1.2 x10-6 kg. 9.8m/s 2 _(8.8 x10-3 m) 160 x10 9 N/m 2 -1x10-3 m. (20 x106 m) 2 3 -OOO85rad=0.49' (A.10) Because the fingers are at the end of the deflected 1000 pm long proof mass, the total deflection of the fingers is 6 FINGERS = X + 1000gm -sin4m = 38gm + 8.5gm = 46.5gm 60 (A. 11) B MASK DESIGN B.1 Proof mass wafer breakout tabs I (b.) (a.) Figure (B. 1) Frontside (a) and backside (b) mask for proof mass wafer. 61 The proof mass wafer required two darkfield, contact masks. The first mask defined the fingers and springs as well as the top of the proof mass while the second mask defined the bottom of the proof mass as well as the openings under the springs and fingers. Both mask designs were created in Cadence and outsourced to be etched into chrome on 5" x 5" quartz using 0.5 ptm spot-size e-beam. Figure (B.1) shows the layout of each mask. The masks were designed allow the proof mass dies to be removed by breaking twelve 200 pim x 100 ptm tabs in the device layer holding it to the wafer at the end of the process. Although this would reduce the number of structures we could fit on one wafer because a frame would have to be left surrounding each die, it was absolutely necessary because the delicate structures would not survive being diced with a saw. We made each of the low frequency (100, 500, and 1 kHz) proof mass dies the same size and placed the interdigitated fingers at the same position (8 mm across and 4.5 mm down from the top left corner) so that they would require only one package size. These dies were made intentionally large, 16 mm x 16 mm, to allow handling with a vacuum pen during packaging. For the small 10 kHz proof mass, the die was shrunk to 16 mm x 8 mm to save wafer space. However, its finger position from the 16 mm edge is similar to the lower frequency proof mass dies. We made the supporting frame around each die 3.6 mm wide. On one 100 mm wafer, there are four 100 Hz, 500 Hz, and 1 kHz and two 10 kHz proof masses. B.2 Photodiode wafer The photodiode wafer required one clearfield and three darkfield contact masks. The first mask was a darkfield mask that defined the active regions of the photodiode to be implanted. The second mask was a clearfield mask that opens up the oxide for the electrical contacts. The third mask was a darkfield mask that defined the interconnects and bond pads. Lastly, the fourth mask was a darkfield mask that defined the DRIE through holes. As in the case of the proof mass wafer, all mask designs were created in Cadence and outsourced to be etched into chrome on 5" x 5" quartz using 0.5 pim spot-size e-beam. Figure (B.2) shows the layout of each mask. 62 - ----------- 4~ 44 .4 44 4* a.) b.) C.) d.) Figure (B.2). Mask set for photodiode wafer. a.) Mask 1 defined photodiode active regions. b.) Mask 2 opened contact holes. c.) Mask 3 pattern interconnects. d.) Mask 4 defined DRIE through-hole. Just as in the case of the proof mass wafer, we used the DRIE to release the dies. Each of the thirty dies, measuring 16 mm x 8 mm is held to the frame by six 200 pm x 100 ptm x 525 pm breakout tabs. Figure (B.3) shows an enlarged view of the same die on each mask. 63 a.) alignment marks b.) C.) breakout tabs d.) Figure (B.3). Sample dies from (a) Mask 1, (b) Mask 2, (c) Mask 3, and (d) Mask 4. 64 C FABRICATION DETAILS Microfabrication of the proof mass and photodiode wafers was performed in the MIT Microsystems Technology Laboratory (MTL). This interdepartmental laboratory includes the Integrated Circuits Laboratory (ICL) class 10 cleanroom, and the Technology Research Laboratory (TRL) class 100 cleanroom. C.1 Proof mass wafer The proof mass wafer was a two mask CMOS compatible process. The starting material was a 100 mm, single-side polished BESOI wafer with a 20 1 1 pim device layer, a 1 1 0.05 pm oxide box layer, and a 381.5 + 0.5 pm Si handle layer. Step 1 Description RCA clean (ICL) 10 min organic clean (5:1:1 H 2 0 : H2 0 2 : NH40H) Rinse 15 s 50:1 H 2 0 : HF dip Rinse 15 min inorganic clean (6:1:1 H 2 0 : H2 0 2 : HCl) Rinse, spin dry 2 3 560 nm thermal oxidation (ICL) Time (min) 10 30 20 TubeA3 recipe G224 Temp (*C) 800 1100 1100 10 10 1100 1100 70 20 20 60 1100 1100 1100 800 Measure oxide (ICL) KLA Tencor UV1280 Ellipsometer Measurement type: SiO 2 on Silicon Thickness = 560 nm 65 Gas N2 N2 N2 02 02 H2/02 02 N2 N2 4 Pattern frontside with Mask 1 (TRL) HMDS Spin cast 1 gm OGC825 positive resist (3000 rpm, 30 s) 30 min bake, 90 "C Expose 2.5 s, soft contact, 365 nm, 9 mW/cm 2 intensity Develop in OGC934 1:1 Rinse, spin dry 30 min bake, 120 C 5 Plasma etch thermal oxide on both sides (ICL) Applied Materials Precision 5000 Etcher AME5000 recipe Isabella LTO Time (s) Gas 10 10 320 02 CHF 3 CHF3 Rate RF (W) (scem) 20 15 10 100 0 350 Pressure Magnetic Field (mTorr) (Gauss) 200 200 200 50 50 50 6 Mount to 150 mm quartz carrier (TRL) Spin cast 10 pm AZ4620 positive resist rings (1500 rpm) on carrier Press on wafer with device layer up 15 min bake, 90 "C 7 DRIE etch 20 pm device layer silicon (TRL) STS Multiplex ICP etcher sts2 recipe Shallow, 12 min APC Manual 75% Base Pres = 0 mT Time (s) Overrun C4F8 Flow SF 6 Flow Cycle (s) (sccm) (sccm) Pass 11.0 0 0 35 Etch 12.5 0 140 0 8 Trip Pres = 95 mT Plate RF Coil RF (W) (W) 60 600 80 600 Dismount wafer and remove resist (TRL) 10 minute piranha clean (1:3 H2 0 2 : H 2 SO4 ) Rinse, spin dry 9 Spin cast 20 pm polyimide (TRL) Spin cast VM652 adhesion promoter (1000 rpm, 30 s) 60 s hot plate bake, 120 C Spin cast P12600 polyimide resin (500 rpm, 120 s) 5 min hot plate bake, 90 0C 30 min cure at 350 C in 40% N2 (load wafer horizontally at 150 C, 4 0C/min ramp) 66 10 Measure polyimide thickness (TRL) Nanospec Thin Film Thickness Measurement System Film Type: Thick Films (13) Index of refraction = 1.50 Thickness = 20 um. 11 Pattern backside with Mask 2 (TRL) HMDS Spin cast 1 ptm OGC825 positive resist on frontside (3000 rpm, 30 s) 30 min bake, 120 C Spin cast 10 ptm AZ4620 positive resist on backside (1500 rpm, 60 s) 60 min bake, 90 0C Expose 25 s, soft contact, 365 nm, 9 mW/cm2 intensity Develop in AZ440 MIF Rinse, spin dry 15 min bake, 90 0C 12 Mount to 150 mm quartz carrier (TRL) Spin cast 10 pm AZ4620 positive resist rings (1500 rpm) on carrier Press on wafer with device layer down 15 min bake, 90 0C 13 DRIE etch 381.5 ptm handle layer silicon (TRL) STS Multiplex ICP etcher sts2 reci e MIT 37b a, 110 min APC Manual 75% Cycle Time (s) Overrun (s) Pass 11.0 0 Etch 15.0 0.5 Base Pres = 0 mT C4F8 Flow SF 6 Flow (sccm) (sccm) 0 95 70 0 Trip Pres 95 mT Plate RF Coil RF (W) (W) 60 600 120 600 14 Dismount wafer and remove resist (TRL) 24 hour acetone dip Rinse, drip dry 15 Wet etch 1 pm buried oxide (TRL) 18 min buffered oxide etch (BOE) dip Rinse, drip dry 16 Mount to 100 mm silicon wafer (TRL) Spin cast 10 pm AZ4620 positive resist ring (1500 rpm) on carrier perimeter Press on wafer with device layer up 15 min bake, 90 C 67 17 Ash polyimide (ICL) Matrix Systems 106 Stripper Asher recipe Std, 12 min C.2 Photodiode wafer The photodiode wafer was a four mask CMOS compatible process. The starting material was a 100 mm, 525 pim-thick, single-sided polished (100) n-type (phosphorus doped) silicon wafer with resistivity of 1.0 ohm-cm. Step 1 Description RCA clean (ICL) See C.1 Step 1 2 560 nm thermal oxidation (ICL) See C.1 Step 2 3 Measure oxide (ICL) See C.1 Step 3 4 Pattern frontside Mask 1 (TRL) See C.1 Step 3 5 Plasma etch thermal oxide on both sides (ICL) See C.I Step 4 6 Ash resist (ICL) Matrix Systems 106 Stripper Asher recipe Std, 1 min 7 Boron ion implantation (outsourced) 40 keV, 5E15 cm 2 dose, minimal tilt (<8') 8 RCA clean (ICL) See C.1 Step 1 68 9 Dopant drive-in/activation and 57 nm reoxidation (ICL) ICL tubeA3 recipe G123 10 Time (min) 10 20 Temp ("C) 800 950 5 10 7 950 950 950 15 950 10 950 35 800 Measure resistivity Prometrix Omnimap 111 B Four Point Probe Sheet resistance = 40 ohms/sq. 11 Pattern frontside with Mask 2 (TRL) See C.1 Step 3 12 Wet etch thermal oxide on both sides (TRL) 8 min buffered oxide etch (BOE) dip 13 Pre-metal clean (ICL) 10 minute piranha clean (1:3 H 2 0 2 : H2 SO4 ) Rinse 10 minute piranha clean (1:3 H2 02 : H2 SO 4 ) Rinse 15 s 50:1 H 20: HF dip Rinse, spin dry 14 Deposit 1 pIm aluminum/silicon (ICL) Applied Materials Endura 5500 Endura recipe Al-Si, 1 um 15 Pattern frontside with Mask 3 (TRL) See C.I Step 3 16 Etch aluminum (TRL) 4 min PAN etch, 45 0C Rinse, spin dry 17 Sinter aluminum (TRL) Load wafers in 51% N 2 30 min N 2/H 2 sinter at 400 0C 69 Gas N2 N2 02 02 H2 /0 2 02 N2 N2 18 Pattern backside with Mask 4 (TRL) HMDS Spin cast 1 pm OGC825 positive resist on frontside (3000 rpm, 30 s) 30 min bake, 120 C Spin cast 8 pm AZ4620 positive resist on backside (3000 rpm, 60 s) 10 min bake, 90 0 C Spin cast 8 pm AZ4620 positive resist on backside (3000 rpm, 60 s) 60 min bake, 90 0 C Expose 100 s (ten 10 s intervals w/ 10 s delay), soft contact, 365 nm, 9mW/cm 2intensity Develop in AZ440 MIF Rinse, spin dry 15 min bake, 90 OC 19 Mount to 150 mm quartz carrier (TRL) Spin cast 10 pm AZ4620 positive resist rings (1500 rpm) on carrier Press on wafer with device side down 15 min bake, 90 C 20 DRIE etch through wafer (TRL) STS Multiplex ICP etcher sts2 recipe MIT_37ba 200 min See C.1 Step 11 21 Dismount wafer and remove resist (ICL) 24 hour acetone dip Rinse, drip dry C.3 Packaging We used a Universal Laser Systems X-100 lasercutter with a 100 watt CO 2 laser. The cutter takes a drawing, like a CorelDraw file, and traces the lines and rasters the solid areas like a printer. The intensity and the speed of the laser can be set to different values for each color in the drawing. The CorelDraw drawings for the plastic housings are shown in Figure (C. 1). Table (C. 1) shows the laser settings for each of the colors. The cutter works well on acrylic (Plexiglas) pieces thinner than 10 mm. Multiple passes are usually preferred because cutting more than 1 mm during a pass can cause melting of surrounding areas, leading to sloped sidewalls. The laser lens we used had a focal length of 2 inches. 70 A_ I A B C D First Generation 0 A D C B Second Generation Figure (C. 1) Lasercutter drawings for (A) top piece, (B) top shim, (C) bottom shim, and (D) bottom piece of first 1 mm deep rasters, red areas and second generation packages. Black lines represent through cuts, yellow areas are are 4 mm deep, orange lines are 2 mm clean-up lines, and cyan lines are shallow (<500 sm) alignment lines. Color Setting Power (%) Speed (%) Depth (mm) Passes Black Vector 70 10 3.2 4 Red Raster/Vector 40 20 3 3 Yellow Raster/Vector 20 40 1 3 Orange Vector 50 30 2 3 Cyan Vector 30 30 <0.5 4 Table (C. 1). Lasercutter settings for fabricating packages. 71 -M-4-1 ;6" D FIRST GENERATION PACKAGE RESULTS We packaged a 428 Hz sensor in the first generation package before realizing the low sensitivity of the housing. Figure (D.1) is the sensitivity of the 428 Hz first generation package which shows similar damping (Q=140) but a low frequency sensitivity (0.1 V/g) than is 400 times less than that of the 432 Hz second generation package. This lower sensitivity translates to a noise spectrum that is more limited by the noise of the photodiode and amplifier than by the background or laser wavelength/phase noise as shown in Figure (D.2). The shot noise, not shown, contributes about 0.7 pg/rt Hz of noise. Although the sensitivity and resolution are low, Figure (D.3) shows that the linearity of the sensor is the same as in the second generation package, as expected. 10 Measured Fit (Q=140) CD C C0 0.1 200 70 80 90100 300 400 Frequency (Hz) Figure (D. 1). Sensitivity of 428 Hz first generation package. 72 500 600 Sensor Seismic Photodiode/Amplifier -N 1 E 0.1 --------- ID Lim it --------- Thermomechanical 0 o 0.01 0.001 CO 01 100 Frequency (Hz) 10 Figure (D.2). Noise spectrum of first generation 428 Hz package. I I I I I I I I I II I I III 2.5 I I I I I I I I I I I I I I I T- 160 Hz Drive 2 t; o 1.5 1 0.5 0 - 0 ' 0.03 0.02 0.01 Drive Acceleration Amplitude (g) Figure (D.3). Linearity of first generation 428 Hz package. 73 0.04 BIBLIOGRAPHY [1] F. Goodenough, "Airbags Boom when IC Accelerometer Sees 50 g," Electronic Design, vol. 39, no. 15, pp.45-56, 1991. [2] G. Kovacs, "MicromachinedTrasnducers Sourcebook, McGraw-Hill, New York, 1998. [3] M. Stephens, "A sensitive interferometric accelerometer," Rev. Sci. Instrum., vol. 64, no. 9, pp. 2612-2614, 1993. [4] A. Barzilai, T. VanZandt, T. Kenny, "Technique for measurement for the noise of a sensor in the presence of large background signals," Rev. Sci. Instrum., vol. 69, no. 7, pp. 27672772, 1988. [5] C. H. Liu, J. K. Reynolds, A. Partridge, J.D. Grade, A. M. Barzilai, T. W. Kenny, "A highsensitivity micromachined accelerometer based on electron tunneling transducer," Proceedings of ASME winters conference - Symposium on MEMS, pp. 1 3-20, 1997. [6] F. Rudolf, A. Jomod, J. Bergqvist, H. Leuthold, "Precision Accelerometers with pg Resolution," Sensors and Actuators, A21-A23, pp. 291-302, 1990. [7] H. K. Rockstad, T. W. Kenny, J. K. Reynolds, W. J. Kaiser, T. B. Gabrielson, "A miniature high-sensitivity broad-band accelerometer based on electron tunneling transducers, Sensors and Actuators A, vol. 43, pp. 107-114, 1997. [8] T. Strobelt et al., "Electron tunneling accelerometers for high resolution," Proceedings International Conference on Solid-State Sensors and Actuators, pp. 1534-1537, 1999. [9] C. H. Liu, A. Barzilai, 0. Ajakaiyc, H. K. Rockstad, T. W. Kenny, "Performance enhancements for the micromachined tunneling accelerometer," Proceedings 1 0 th International Conference on Solid-State Sensors and Actuators, pp. 1290-1293, 1999. 1 0 th [10] C. H. Liu et al., "Characterization of a High-Sensitivity Micromachined Tunneling Accelerometer with Micro-g Resolution," JMEMS, vol. 7, no. 2, pp. 235-243, 1998. [11] D. Rugar, H. J. Mamin, P. Guethner, "Improved fiber-optic interferometric for atomic force microscopy," Appl. Phys. Lett., vol. 55, no. 25, pp. 2588-2590, 1989. [12] S. R. Manalis, S. C. Minne, A. Atalar, C. F. Quate, "Interdigital cantilevers for atomic force microscopy," Appl. Phys. Lett., vol. 69, no. 25, pp. 3944-3946, 1996 [13] E. B. Cooper et al., "High-resolution micromachined interferometric accelerometer," Appl. Phys. Lett., vol. 76, no. 22, pp. 3316-3318, 2000. 74 [14] S. C. Minne, S. R. Manalis, C. F. Quate, Bringing ScanningProbe Microscopy up to Speed, Kluwer, Boston, 1999. [15] G. G. Yaralioglu, A. Atalar, S. R. Manalis, C. F. Quate, "Analysis and design of an interdigital cantilever as a displacement sensor," J. Appl. Phys., vol. 83, no. 12, pp. 74057414, 1998. [16] T. B. Gabrielson, "Mechanical-Thermal Noise in Micromachined Acoustic and Vibration Sensors," IEEE Transactions on Electron Devices, vol. 40, no. 5, pp. 903-909, 1993. [17] S. D. Senturia, Microsystems Design, Kluwer, Boston, 2001. [18] M. J. Usher, I. W. Buckner, R. F. Burch, "A miniature wideband horizontal-component feedback seismometer, J. Phys. E: Sci. Instrum., vol. 10, pp. 1253-1260, 1977. [19] M. J. Novack, "Design and fabrication of a thin film micromachined accelerometer," Master's Thesis, Massachusetts Institute of Technology, 1992. [20] P. Horowitz, W. Hill, The Art ofElectronics, Cambridge, New York, 1998. [21] S. H. Crandall, N. C. Dahl, T. L. Lardner, An Introduction to the Mechanics of Solids, McGraw-Hill, New York, 1999. 75