:1... .3 i. i 2.7.x 9. a. x. « . 2.: 73 31:... 1. In: .2 .s . .cyirzf. . .. c: A”): . .; . . $555.5»: . . .a .1535} ~ 5 i5..;.? .,: 9 “1.3.: t 1.. .Vol..! 13.33.... c. 1} v.‘ In , .. , .. . .3.“ arm” 9 .. I 1 .3: I :33. $519”) D ' it» {é : u G 3 ‘I 5 1. a. A- 1,. . I... 3.1. ”1:31 x 2...... .5 H 1 .s I .t. (33.! 13.: 3.... .e . .24. 43.7.5. 314 .11...A~QI.SI.: 3.... . . . T5 :ilefl .15.? 37:35 15“.. V .PIL..;.».J.. . «Hr-7.. 1”. ...| Q. . . ,. l| wt... . has: LIBRAFéY 2 Michigan tate W _ University This is to certify that the dissertation entitled Polycrystalline Diamond RF MEMS Resonator Technology and Characterization presented by Nelson Sepulveda-Alancastro has been accepted towards fulfillment of the requirements for the Doctoral degree in Electrical and Computer EngineerinL 9/? flan»; Major Professor’s Signature December 8' 2005 Date MSU is an Affirmative Action/Equal Opportunity Institution n----4-v-----.—-----.--.-c----;—-c PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/CIRC/DateDue.indd-p.1 POLYCRYSTALLINE DIAMOND RF MEMS RESONATOR TECHNOLOGY AND CHARACTERIZATION By Nelson Sepulveda-Alancastro A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Electrical Engineering 2005 ABSTRACT POLYCRYSTALLINE DIAMOND RF MEMS RESONATOR TECHNOLOGY AND CHARACTERIZATION By Nelson Sept'tlveda-Alancastro Due to material limitations of polycrystalline silicon resonators, polycrystalline diamond has been explored as a new RF MEMS resonator material. This work presents the development of polycrystalline diamond micro and nano resonators with quality factor (Q) values as high as 116,000. Polycrystalline diamond resonator structures were tested using electrostatic and piezoelectric actuation. Similar resonator structures were tested using both testing methods, and their performance showed a difference in resonant frequency of about 3%, while the measured Q values differed by approximately a factor of 10. The resonant frequency shifts due to different testing temperatures was quantified by the temperature coefficient (TCf) value, which ranged from -1.59x10'5 / °C to -2.56 x10’5 / °C for the different polycrystalline diamond structures. The Q values were not limited by clamping losses, phonon-phonon interaction or thennoelastic dissipation and they were measured as a function of temperature. The results showed an apparent thermally activated relaxation process, with an activation energy of 1.9 eV responsible for limiting the highest achievable Q value in the tested polycrystalline diamond resonators. The fabrication technology and the performance of polycrystalline diamond resonators with dimensions in the nano scale are also presented. Polycrystalline diamond resonator cantilevers and torsional resonators with dimensions in the nanometer range (as small as 100 nm) have been fabricated and tested. The performance of these structures shows resonant frequencies and Q values in the range of 23 KHZ — 805 KHZ and 2,800 — 103,600 respectively. T 0 all those who defend peace and happiness at the same time iv ACKNOWLEDGMENTS First, I would like to thank my big and great family. My dad (Don Agustin Sept'tlveda), who always motivated me to give the best of me. My mom (Dona Matilde Alancastro) who has been part of the “big events” in my life (even watching me strike out in a little league baseball game). My brothers (Isaias, Agustin, Ricardo, Jorge, “Cacho”, and Javier) and sisters (Nelly and Tere), which I have always felt close even though most of the time we have been physically distant. During my graduate school I married Laura, the most beautiful and wonderful girl in the planet. She has been very patient and also motivating during my career. Thanks also to “las malas compafiias" from “El Colegio”. They showed me that friendship is a blessing. This thesis would not have ever been written if it was not for those who inspired and motivated me to begin this journey after my BS. degree. Thanks to Yuxing, Xiangwei, and Yang who helped me in my research and with which I spend happy and painful moments in the RCE. Thanks also to Dr. John Sullivan for his knowledge and friendship, and to The Bill and Melinda Gates Foundation for their financial support. TABLE OF CONTENTS LIST OF TABLES x LIST OF FIGURES xi CHAPTER I: Research Motivation and Goals ................................................ 1 1.1 Introduction ............................................................................... 1 1.2 Objective of this Work .................................................................. 2 1.3 Overview of this Thesis ................................................................. 3 CHAPTER I]: Background ....................................................................... 5 2.1 Introduction .............................................................................. 5 2.2 Resonator Theory and Operation ...................................................... 5 2.2.1 Resonance Frequency ....................................................... 5 2.2.2 Quality Factor ................................................................ 9 2.2.2.1 Lorentzian Fit .................................................... 11 2.2.3 Actuation and Detection Methods ....................................... 12 2.2.3.1 Electrostatic Actuation .......................................... 13 2.2.3.2 Piezoelectric actuation and Optical Detection ............... 19 2.2.3.3 Magnetomotive Actuation ...................................... 22 2.3 Energy Dissipation Mechanisms .................................................... 22 2.3.1 Extrinsic Mechanism ....................................................... 23 2.3.2 Intrinsic Mechanism ........................................... ‘ ............. 24 2.4 Micromechanical Resonator Structures and Limitations ......................... 28 2.4.1 Micromechanical Resonators ............................................. 28 vi 2.4.2 Resonator Geometry ........................................................ 28 2.4.3 Resonator Materials ........................................................ 33 2.4.4 Resistance Rx ............................................................... 35 2.5 Polycrystalline Diamond Material ................................................... 36 2.5.1 Why Diamond? ................................................................................. 36 2.5.2 Chemical Vapor Deposition (CVD) of Diamond Fihns ............... 38 2.5.3 Seeding ....................................................................... 40 2.5.4 Doping ........................................................................ 40 2.5.5 Patterning .................................................................... 41 CHAPTER III: Polycrystalline Diamond Film Technology ............................... 43 3.1 Introduction ............................................................................. 43 3.2 Seeding .................................................................................. 43 3.3 Film Growth Process ................................................................... 44 3.4 Doping ................................................................................... 46 3.5 Patterning ............................................................................... 48 3.6 Surface Roughness ..................................................................... 55 3.7 Microresonator Fabrication Process ................................................. 57 3.7.1 Fabrication Technology .................................................... 57 3.7.1.1 Samples for Electrostatic Testing ............................. 58 3.7.1.2 Samples for Piezoelectric/Optical Testing ................... 61 3.7.2 Post-Processing ............................................................. 63 CHAPTER IV: Resonator Testing Methods .................................................. 65 4.1 Introduction ............................................................................. 65 vii 4.2 Electrostatic Testing ................................................................... 65 4.2.1 Electrostatic Testing Equipment Set-Up ................................. 65 4.2.2 Electrostatic Testing Measurement Results ............................. 66 4.3 Piezoelectric Actuation with Optical Detection .................................... 69 4.3.1 Piezoelectric Actuation and Optical Detection Results ............... 69 4.4 Comparison of Results ................................................................ 73 CHAPTER V: Study of Q and Frequency Shifts in Polycrystalline Diamond Resonators ....................................................................... 79 5.1 Introduction ............................................................................. 79 5.2 Quality Factor Limitations ............................................................ 79 5.3 Quality Factors in Polycrystalline Diamond Resonators ......................... 81 5.4 Polycrystalline Diamond Fihn Microstructure .................................... 85 5.4.1 Seeding Layer and Film Layer ............................................ 86 5.5 Temperature Dependence. .................................................................. .92 5.5.1 Young’s Modulus and Frequency Temperature Dependence. . . . . ....92 5.5.2 Quality factor temperature dependence.......................................94 CHAPTER VI: Polycrystalline Diamond Nanoresonators......... ....101 6.11ntroduction........... .................................................................................... 101 6.2 Fabrication of Nanoresonators. . . .. ............................................................... 101 6.3 Torsional Resonators. . . .. ........................................................................... 102 6.3.1 Testing of Torsional Resonators. ................................................ 104 6.4 Nanocantilevers. . . . . . .. ............................................................................... .105 6.4.1 Testing of Nanocantilevers.. ...................................................... 109 CHAPTER VII: Conclusions and Future Research... .............. 116 viii 7.1 Summary and Conclusions. ..................................................................... 116 7.2 Future Research. ................................................................................... 117 BIBLIOGRAPHY .............. 1 1 9 ix 2.1 2.2 2.3 2.4 3.1 4.1 4.2 5.1 5.2 6.1 6.2 LIST OF TABLES Constant Kn for the first five flexural modes of a bridge structure .................... 7 Resonator structures fabricated from polycrystalline diamond. The torsional resonator shown in the second row was patterned at Sandia National Laboratories. The other structures were completely fabricated at Michigan State University ................................................ 31 Properties of materials that can be used to fabricate micromechanical resonators .................................................................................... 34 Most recent results for diamond resonators before the work reported in this thesis ................................................................................. 35 Growth parameters for the three polycrystalline diamond samples studied and dry etching parameters (superscript denotes the sample on which the value was used) .......................................................................................... 46 Results for piezoelectically-actuated cantilever beams (first vibration mode) madefromsample.............. ............................................................................. 70 Polycrystalline diamond growth parameters for the sample tested electrostatically and piezoelectrically ................................................... 73 Polycrystalline diamond and polycrystalline silicon properties used for plotting the dissipation curves in figure 5.1 ...................................................... 81 Ratio of seeding and film thickness to total thickness for each sample and the average of the 15 higher measured Q values for each sample (Qtot )... . . .. ...88 Measurement results for single and double torsional resonators ................... 105 Results of polycrystalline diamond nanocantilevers fabricated from sample 1. The resonant frequencies were calculated using the width as the resonator thickness ( a), and the film thickness as the resonator thickness (b). M“ represents the number of measurements taken from each structure (the values shown in the table correspond to the measurement with the highest measured Q) ............................................................................... 114 1.1 2.1 2.2 2.3 2.4 2.5 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 LIST OF FIGURES Thesis overview .............................................................................. 4 Bridge resonator diagram to be tested using electrostatic actuation [8]. . . .....6 Example of resonant peak obtained using electrostatic actuation. The resonant frequency and Q are obtained directly from the measured data shown in the spectrum ..................................................................................... 1 1 Schematic diagram for piezoelectric actuation and detection. The reflected laser beam from the vibrating structure is sent to a photodetector. A network analyzer uses the output of the photodetector to plot the resonance spectrum. . . . . . . . .....21 Schematic diagram for the magnetomotive actuation and detection ................ 22 a) HF CVD and b) MPCVD polycrystalline diamond deposition .................... 39 Raman spectra for the three samples studied ........................................... 48 Grass eflect diagram. The sputtering of aluminum during the dry etching of diamond prevents the plasma from etching the diamond under the sputtered aluminum .................................................................................... 51 Evidence of the grass eflect found on (a) the dry etching of the polycrystalline diamond resonators reported in this thesis and (b) the DRIE of silicon-on- insulator wafers [78] ....................................................................... 52 Released polycrystalline diamond structures (comb-drives and bridges) patterned at Michigan State University. No grass eflect can be observed ..................... 53 a) Fabricated and patterned polycrystalline diamond resonators from [31] and b) Michigan State University (reported in this thesis) ................................... 54 AF M images showing the polycrystalline diamond surface roughness of Sample 3 (a) and Sample 2 (b). The difference in the film roughness is about 5 nm ........................................................................................... 56 Starting substrate for the fabrication of polycrystalline diamond resonators. . .....57 Fabrication process flow for electrostatically tested resonators ..................... 59 Fabrication process flow for resonators to be tested using piezoelectric actuation and optical detection ....................................................................... 63 xi 4.1 4.2 4.3 4.4 4.5 5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 Connections to the resonator structure for electrostatic testing. In order to measure only the change in current across the resonator, a bias tee is used to isolate the dc- bias from the IN-port of the network analyzer ..................... 66 Electrostatic measurements on polycrystalline diamond resonators: a) bridge structure, b) comb-drive. The resonant frequency (mwas calculated by finding the frequency at maximum magnitude and the Q was calculated using thist value and the 3dB bandwith (Af) ......................................................... 68 Young’s modulus of sample 1 was calculated from a linear fit done to the measured data. .............................................................................. 72 Raman spectra for samples measured using different actuation methods: a) piezoelectric b) electrostatic .............................................................. 74 Testing results on two similar polycrystalline diamond cantilevers using different actuation and detection methods ......................................................... 76 Energy loss mechanism curves for a 1 pm thick poly-crystalline diamond and polycrystalline silicon cantilever beam .................................................. 80 Q value limiting curves for a 0.6 pm polycrystalline diamond cantilever beam and measured data for the three samples ..................................................... 82 Measured Q values as a function of frequency for the three studied samples. Resonant peaks show the highest Q value for each sample ........................... 84 Cross section TEM images of the three studied polycrystalline diamond samples ...................................................................................... 86 Plot of l/Qm, as a function of ts/ttot. The values for Qf and Q, are obtained fiom the slope and intercept of the linear fit .................................................. 89 The approximation for the minimum space of solutions is determined by the rectangle, which touches all the curves at least in one single point. . .. . . ....91 Young’s modulus temperature dependence for the three samples. The three samples have very similar slopes and the sample 1 showed the highest Young’s modulus at room temperature ......................................... 93 Frequency shift as a function of testing temperature for sample l............ . . . . ...95 Frequency shift as a function of testing temperature for sample 2 .................. 96 xii 5.10 5.11 5.12 6.1 6.2 6.3 6.4 6.5 6.6 6.7 6.8 Frequency shifi as a function of testing temperature for sample 3 .................. 97 Measured Q values vs. temperature a) Three samples plotted in the entire testing temperature range; b) Sample 2 in the range of temperature where a minimum in Q was observed ................................................... 99 Measured Q values vs. activation energy for the three samples. A minimum in Q was observed for two of the three samples at an activation energy around 1.9 eV ..................................................................... 100 SEM image of polycrystalline diamond single torsional resonator ................ 103 Excitation of torsional resonator. The offset of the paddle support from the center of the paddle causes a torsional vibration mode .............................. 104 Smooth sidewalls of polycrystalline diamond nanocantilevers fabricated using e-beam lithography ................................................................ 107 Use of piezoelectric actuation and optical detection for a cantilever with vibration perpendicular to the substrate (a) and parallel to the substrate (b). The vibration will be in the direction parallel to the smallest dimension of the cantilever .............................................................. 108 100 nm wide polycrystalline diamond cantilever. The top view shows that the width of the cantilever is in not uniform along the beam length. . . . . . . . . . 109 SEM image and performance of the nanocantilever with the highest measured Q ............................................................................................. 1 10 SEM image and performance of a 100 nm wide polycrystalline diamond nanocantilever ............................................................................. 1 1 1 Plot of measured Q values on cantilever beams made of silicon nitride [26], single crystal silicon [26] and polycrystalline diamond (reported in this work). All the cantilevers had a thickness between 170 nm and 200 nm ........................... 115 xiii Chapter 1 Research Motivation and Goals 1.1 Introduction Microsystems and Broadband/wireless communication systems need high performance, low-cost, low-power and small-size components. Such components can lead to single-chip transceivers and microsystems. However, the majority of the high-Q bandpass filters commonly used in the radio frequency (RF) and intermediate frequency (IF) stages of transceivers are realized using off-chip, mechanically resonant components, such as crystal filters and surface acoustic wave (SAW) devices. These off-chip resonator components can contribute to the substantial percentage (often up to 80%) of portable transceiver area taken up by board-level passive components. Recently, silicon-based low-power resonant micromechanical structures have been fabricated. Microelectromechanical systems (MEMS) technologies that make it possible to fabricate high quality factor (Q) on-chip micromechanical resonators [1,2] now suggest a method for miniatruizing and integrating highly selective filters alongside transistors leading to miniaturized transceivers. Among the RF MEMS components that could replace conventional components in cellular and cordless applications are image reject and IF filters with center frequencies in the ranges of 0.8 - 2.5 GHz and 0.455 - 254 MHz, respectively [1]. Some of the specific devices fabricated are tunable micromachined capacitors, integrated high-Q inductors, low-loss micromechanical switches, and micromechanical resonators with quality factors (Qs) above 10,000. Microfabricated resonant structures for sensing pressure, acceleration, and vapor concentration have been demonstrated [3]. Recent advances in IC-compatible MEMS technologies suggest methods for board-less integration of wireless transceiver components [4]. In fact, given the existence already of technologies that merge micromechanics with transistor circuits onto single silicon chips [4-7], single-chip RF MEMS transceivers are not a far goal anymore. However, before one can take firll advantage of RF MEMS, a number of issues need to be addressed including higher frequency, selectivity, vacuum encapsulation and reliability. In order to solve these issues, researchers have tried to use different resonator structures, designs and materials. 1.2 Objective of this Work Polycrystalline silicon has been the material of choice for RF MEMS micromechanical resonators. It is, however, found that reducing the resonator dimensions (to achieve higher frequencies) close to or below 1 um for polycrystalline silicon resonators is expected to lead to size related limitations due primarily to enhanced adsorption properties of water-related species to the silicon surface. As diamond surfaces are known to be chemically inert to such adsorption and diamond has the highest Young’s Modulus (which allows higher resonant frequencies) among all the materials, polycrystalline diamond micromechanical resonators are expected to be superior to silicon based resonators. The goal of this work is to study the polycrystalline diamond material for its use as a structural material in RF MEMS micromechanical resonators. The major accomplishments and/or contributions to the scientific community reported in this thesis include the following: 1) 2) 3) 4) 5) 6) 1.3 The development of the fabrication technology for polycrystalline diamond RF MEMS resonators. The testing of polycrystalline diamond resonator using different techniques showed similar resonant frequencies but differences in Q values by a factor of 10. Achievement of the highest Q of 116,000 measured for any polycrystalline diamond resonators structure or cantilever beams made of any polycrystalline material. Study of the energy dissipation mechanisms (which influence the Q) in polycrystalline diamond micromechanical resonators and their relation to the film microstructure. Study of the temperature dependence of the resonant frequency and Q in polycrystalline diamond resonators. Fabrication and testing of polycrystalline diamond nanocantilevers with widths as small as 100 nm. Overview of this Thesis This thesis presents the development and characterization of a polycrystalline diamond microresonator fabrication technology capable of achieving high Q and nano- sized polycrystalline diamond resonators. Chapter 2 explains the theory of the resonator dynamic behavior and the different actuation and detection methods that can be used for the testing of micromechanical resonators. It also discusses the energy dissipation mechanisms, the micromechanical resonators limitations and the polycrystalline diamond material. The third chapter on this thesis talks about the polycrystalline diamond microresonator technology used for the fabricated and tested devices. The details for polycrystalline diamond film growth and patterning are presented with some of the issues faced and ways to avoid/solve them. The fourth chapter describes the measurement techniques used and the equipment set up. It discusses the measurement results on the two actuation and detection techniques used: 1) electrostatic actuation and detection and 2) piezoelectric actuation with optical detection. Then, in chapter five, the energy dissipation mechanisms present in polycrystalline diamond micromechanical resonators and the thermal stability of such devices are discussed. The last chapter shows the technology for the fabrication of nano sized cantilever beams and the results obtained from such structures. The following figure (figure 1.1) shows an overview of this thesis. Polycrystalline Diamond RFMEMS Resonator Technology and Characterization Figure 1.1: Thesis Overview Chapter 2 Background 2.1 Introduction This chapter presents the theory necessary to understand the dynamic behavior of micromechanical resonators and discusses the energy dissipation mechanisms, which are responsible for limiting the Q. It also mentions the resonator geometries and materials commonly used, their limitations and a summary of the latest results obtained for RF MEMS resonators. Finally, the justification for using polycrystalline diamond and some of the common techniques for the fabrication and processing of polycrystalline diamond films are discussed. 2.2 Resonator Theory and Operation For the understanding of the operation of an RF MEMS micromechanical resonator there are important concepts and parameters such as resonance frequency and Q that need to be clear. Some related concepts are the actuation and detection methods, which help to understand the testing of such devices. The following section describes the concept of resonant frequency assuming electrostatic actuation and detection is used as the testing technique. 2.2.1 Resonance frequency Figure 2.2 shows a typical clamped-clamped (bridge) structure [8]. It is clamped at both ends, and a capacitance is formed between the beam and the underlying electrode. Resonator Beam Figure 2.1 Bridge resonator diagram to be tested using electrostatic actuation [8]. An infinite number of resonance modes are possible for the structure presented in figure 2.1. The frequency of each mode is determined by the structural material properties and resonator geometry. The resonance frequency fm of a given mode I: for the bridge in figure 2.1 is given by the expression [9] 1 k fm =7 _r (2-1) 7: m, where k, and m, are the stiffness and mass of the beam respectively. The stiffness is given by [9]: k, = EW, — (2.2) where E is the Young's modulus of the structural material; Lr, wt and h are the beam length, width and thickness specified in figure 2.1 respectively. By substituting equation (2.2) in equation (2.1) the following expression for the resonance frequency of a bridge resonator is obtained: 1&2 fm where p is the density of the structural material. Different modes are distinguished by a constant Kn. If this constant is taken into consideration a more general expression for the resonant frequency of a bridge or cantilever beam can be obtained [10]: E h fm =Kn ‘——2 (2-4) P L, The constant Kn will depend on the vibration mode the beam is operated and is tabulated for the first five modes of a bridge structure in table 2.1. Table 2.1 Constant Kn for the first five flexural modes of a bridge structure Mode n Nodal Points Kn fnflo Fundamental (fo) 1 2 1-03 1 15‘ Hmonic (f2) 2 3 2.83 2.76 nd . 2 Harmomc (f3) 3 4 5.55 5.4 3rd Hmonic (f4) 4 5 9.18 8.93 th . Now, equation (2.4) is the mechanical resonance fiequency of the resonator if there were no electrodes or applied voltages. The presence of electrostatic forces on the vibrating beam seems to alter the stiffness of the beam [11]. Mathematically, it may be appropriate to re-define the spring constant k,. If k,- is the spring constant in the absence of electrostatic forces, then k; =k, —ke (2.5) where Ice is the electrostatic spring stiffness, which is subtracted fi’orn the spring constant k, to give the “modified” spring constant k} . Thus equation (2.1) can be modified to: _k_r_ _ z-kr ke __ £r_[1-£e_]y2 (2.6) m, m, 27: k, A good approximation for Ice, can be made to get [1 1]: (MI/,3) d3 (27) e: where A is the electrode overlap area, a is the permittivity of free space, V p is the dc bias applied to the resonator and d is the electrode to resonator gap as shown in figure 2.1. After substituting equation (2.2) and equation (2.7) in equation (2.6) and taking in account the constant Kn discussed earlier the following equation is obtained: 2 3 E h aAV L 1 fm=Kn -——2[1— —” 1/2 3 3 (2.3) pL, d EWh It is important to remember that equation (2.8) adds the effect of electrostatic actuation testing, and therefore it can only be used when this testing method is used. When testing methods that do not affect the spring constant (k, ) of the resonator structure are used, the resonant frequency can be calculated using equation 2.4. Due to the fact that the polycrystalline diamond Young’s Modulus is about 6 times larger than that of is about 6 times smaller for a 2 3 k EAV L polycrystalline silicon, the ratio <—£> = ___3 P 3 r d E Wh polycrystalline diamond beam when compared to a polycrystalline silicon. For a typical bridge structure (100 um long, 10 um wide, 2 um thick) with a gap spacing of 0.2 pm, an k electrode beam overlap of 20 um and a DC bias (Vp) of 2V, the ratio <_e> is 0.06 for a r polycrystalline silicon beam and 0.01 for a polycrystalline diamond beam. 2.2.2 Quality factor Another parameter that is important to study in order to characterize the performance of a resonator is the Q. The Q is defined as the ratio between the energy stored in the system to the energy lost per cycle. It describes the frequency selectivity of the device (i.e. the device response to a specific frequency). Processes or mechanisms, which increase the energy dissipated in the resonator are sources of Q degradation. Section 2.3 explains in more detail the energy dissipation mechanisms for a micromechanical resonator. High Q resonators are necessary for sensor and RF MEMS applications. In sensor applications, if a resonator is exposed to a chemical leading to a change in its mass, 3 shift in its resonant frequency could be used to detect the chemical. The shift in resonant frequency needed for detection or sensing of a mass change is determined more precisely if the resonator has a high Q. Thus, the sensitivity of a microresonator based sensor increases with Q. Also, a high Q value is desirable for oscillators and filter applications, where the frequency of operation needs to be carefully controlled. The Q of a resonator can be obtained in different ways. When electrostatic actuation and detection is used for testing the resonator, the measured output is the magnitude of the 821 scattering parameter (y-axis) as a function of frequency (x-axis). The scattering parameters are commonly used in RF circuits to measure the ratio of the output power signal to the input power signal. When the resonator beam is vibrating, the changes in the output current (which are discussed in section 2.2.3.1) are reflected on the output power signal used to determine the scattering parameters. In most cases, the magnitude of the 821 parameter is converted to decibels by using [521' dB = 20xlog(]521|). The frequency at maximum magnitude is used as the resonant fiequency f0. The half power points (1/J2 max Iszrl) [12] are determined on either side of the resonant frequency and the difference of those frequency positions is the bandwidth Af If the |821| value is in decibel units (ISzlde), then the half power points are located at the two points of the curve which are 3dB apart from the maximum magnitude (20 x log (1/«/2 max ISZII) = lszlldB — 3dB)- Figure 2.2 shows an example of a resonant peak obtained using electrostatic testing and the procedure for calculating the Q. This method relies solely on the measured 10 data. However, there is a more accurate method for calculating the resonator Q, which uses a curve fit. Af3dB ISZI| (dB) FREQUENCY (Hz) Figure 2.2 Example of resonant peak obtained using electrostatic actuation. The resonant frequency and Q are obtained directly from the measured data shown in the spectrum. 2.2.2.1 Lorentzian fit The equation for a Lorentzian curve is given by [12]: y = Ymax 2 (29) J1+4Q2[i 4] 1‘0 where x0 is the x value that corresponds to ymax. It has been found that the most accurate fit for the resonant curve of micromechanical resonators is a Lorentzian curve using nonlinear least-squares fit using frequency in the x-axis and a variable directly proportional to the resonator amplitude of motion (such as [321' or the voltage read from 11 a photodetector) in the y-axis [12]. Therefore, the data from which the resonant peak is plotted can be fit with a Lorentzian curve, and the frequencny and Q parameters can be extracted from the fit. 2.2.3 Actuation and detection methods Different resonator actuation and detection methods have been described and used [13,14,15]. The two techniques most commonly used for micromechanical resonator measurement are the capacitive (electrostatic actuation and detection) and the combination of mechanical actuation (piezoelectric actuation) with optical detection. The electrostatic method measures the change in the scattering parameter S21 (caused by a change in the output current) for the detection of motion and the mechanical method measures the change in an optical (laser) signal reflected from the resonator surface for motion detection. The change in output current for the electrostatic actuation is caused by the change in charge across the vibrating resonator which behaves as a time varying capacitor. However, as the structure gets smaller and its geometry more complicated, the difficulty to use these two techniques for movement detection increases significantly. The problem with optical detection is the minimum spot size of the laser. If the laser spot size is larger than 1 pm, it would difficult to measure the reflected signal from the surface of a resonator with dimensions smaller than 1 pm. Also, the laser induces noise such as heating. The capacitive method suffers from various drawbacks such as parasitic capacitances (which are difficult to account for) [16]. Restrictions on the device design and changes to the structural material (e. g. the stiffness of the material) are introduced due to the current flow, which limits the material capabilities [17]. 12 The third method for actuation and detection of a resonator is the magnetomotive technique. This method makes use of the Lorentz force [18]. A drawback of this method is that in order to prevent heating the structure has to have low resistance [19]. For a polycrystalline diamond resonator this can only be accomplished with heavy doping or a thick metal film on top of the resonator, which makes this method unsuitable if the intrinsic characteristics of diamond are of interest. These three techniques are described in more detail in the following sections. 2.2.3.1 Electrostatic actuation When electrostatic force is used, a dc-bias Vp is applied to the suspended beam, and an ac input voltage v,- = Vm sin cot (Vm is in the range of the mV) is applied to the electrode under the beam (see figure 2.1). Therefore, the resulting electrostatic potential across the beam is v=vi+Vp. (2.10) In a vibrating resonator, the capacitance formed between the bottom electrode and the beam is changing with time. Therefore, the charge stored in the capacitor is also changing with time, which implies a current flow. From equation (2.10) three different inputs can be analyzed: 1) v = Vp (vi = 0), 2) v = v,- (Vp = 0) and 3) v = vi + Vp. Each one of these cases is analyzed separately. First, let the input to the system be purely dc (i.e. v = VP). The electrostatic force will make the beam bend towards the bottom electrode, generating a change in capacitance for a short period of time. However, once the system reaches its equilibrium, l3 the beam stops moving, and the equivalent circuit can be represented by a dc voltage source connected in series with a capacitor. The expression for the current flowing through a capacitor is: ._6Cv_C6v+ ac ___ _. _ 2.11 ’0 at at ya: ( ) From (2.11) it can be seen that if neither the voltage nor the capacitance varies with time, the current will be zero. Therefore, in steady state, the current for a purely dc input would be zero. Now, let v=v,-. If the frequency of this input signal ( fi :31) is far from the 7: natural frequency of the beam ( f0 = 2% ), the beam does not vibrate, and therefore the 72' capacitance between the beam and the electrode does not change with time. The second term in (2.11) becomes zero, and the total current becomes i0 = C 6vi/6t. As the frequency of the ac signal approaches the natural frequency of the beam, it begins to vibrate, creating a time varying capacitor. Then the current is given by (2.11) with v = vi. The third input v = v,- + Vp results in a current given by 6(Vi + VP) i0 =C 6t +(Vi +Vp)%€ (2.12) Since the magnitude of the ac signal (Vm) is in the range of mV and the dc signal is in the range of Volts, the current is significantly increased when the dc signal is added to an ac signal with frequency close to the natural frequency of the resonator. l4 Knowing that the change of Vp with respect of time is zero (6Vp/6t = 0), equation (2. 12) can be simplified to: 1'0 =C%—+(vi +Vp)%€— (2.13) Now, in the second term in equation (2.13) Vp is added to v,- and since the magnitude of the ac signal is in the mV range, it can be neglected and we get the following equation: i0 =C%+Vp%€— (2.14) The derivative of capacitance with respect of time (6C / at) can be expressed as: 66-2993 __ 2.15 61 6x61 ( ) where the derivative of capacitance with respect to the displacement x can be obtained directly fi'om the plate capacitance formula: 80A C = (d - x) (2.16) where so is the permitivity of free space, A is the electrode area, d is the initial electrode gap, and x is the beam displacement (see figure 2.1). So, after taking the derivate of the capacitance with respect to displacement x we get; 6C EA ax =(d_x)2 (2.17) For a typical vibrating beam the beam displacement is very small when compared to the gap spacing (x< 350 °C), making it increasingly unsuitable for high temperature applications [30]. Materials with higher Young’s and shear moduli will provide a larger resonant frequency as the equations in the fourth column of table 2.2 show. That is why polycrystalline silicon carbide and polycrystalline diamond materials are being studied for use in RF MEMS devices. Both materials have a larger Young’s and shear moduli than polycrystalline silicon (see table 2.3). Early results on polycrystalline silicon carbide [30] and polycrystalline diamond [31,32] show that indeed devices made of these two materials show a higher resonance frequency than that of equivalently sized polycrystalline silicon versions. Polycrystalline diamond folded beam comb-drive nricromechanical resonators have now been measured with resonance frequencies 1.77 times higher than that of identical polycrystalline silicon counterparts, 1.20 times higher than achievable by polycrystalline SiC, and 1.53 times higher than a previous attempt at using CVD polycrystalline diamond as a resonator structural material [32]. Also, polycrystalline diamond cantilever beams have shown the highest Q for cantilevers made of any polycrystalline material [33]. Table 2.4 shows the most recent and outstanding results 33 (prior to the work shown in this thesis) for diamond resonators tested using different actuation methods. Table 2.3 Properties of crystalline materials that can be used to fabricate micromechanical resonators many/Mm... haggline figggggflggg sec on. Density, p 2300 3300 3500 Kg/m3 Young’s Modulus, E 150 448 1200 GPa Shear Modulus, G 52 160 478 GPa Acoustic Velocity, «1(B1p) 8075 11652 18516 m/s Normalized Acoustic Velocity l 1.44 2.29 -- 34 Table 2.4 Most recent results for diamond resonators before the work reported in this thesis Actuation Method Structure f0 Q Disk* [7] 1.51GHz 11,555 Electrostatic Bridge [31] 2.938 MHz 6,225 Comb-Drive [32] 27.35 KHz 36,460 Doubly-Clamped Paddle * [34] 640 MHz 3,000 Piezoelectric Doubly-Clamped Paddle * [14] 157 MHz 9,000 Magnetomotive 13:23:13.3?)ng 157 MHz ~9,000 * These structures (Doubly-Clamped Paddle and Disks) were made of nanocrystalline diamond 2.4.4 Resistance Rx Now that the GHz frequency range has been reached, micromechanical resonators are emerging as viable candidates for on-chip versions of the high-Q resonator used in wireless communication systems for frequency generation and filtering. Among the more important of the remaining issues that still hinder development of these devices in RF front ends is their larger-than-conventional impedance. In particular, it is their large impedance (commonly called motional resistance, Rx) that presently prevents micromechanical resonator devices in the VHF and UHF ranges from directly coupling to antennas in wireless cormnunication applications, where matching impedances in the range of 500 and 3300 are often required [35]. 35 The most direct methods for lowering the motional resistance Rx of electrostatically actuated micromechanical resonators are [36]: 1) decrease the electrode- to-resonator gap; 2) raising the dc-bias voltage; and 3) summing together the outputs of an array of identical resonators. Each of these methods comes with drawbacks. The first two methods are very effective in lowering Rx, with exponential dependences. However, they do so at the cost of linearity [37]. On the other hand the third method improves linearity while lowering Rx. Unfortrmately, the third method is difficult to implement, since resonators with identical responses are required. This is even harder for large Q devices. 2.5 Polycrystalline Diamond Material In this thesis, the polycrystalline diamond material is studied for its use as a structural material for RF MEMS resonators. The following sections show the justification for the use of polycrystalline diamond and a summary of the reported properties and techniques for the processing of this material. 2.5.1 Why diamond? The use of strategic geometries, and alternative materials, are the two methods of choice for increasing the resonance of the device. Going for smaller dimensions (sub-um range) on polycrystalline silicon resonator structures is not practical due to enhanced adsorption properties of water-related species to the silicon surface. Diamond, with the . . -1 . . highest acoustic velocrty of 18,076 ms [38] and chemrcal inertness, seems to be the 36 most superior among the new materials including polycrystalline silicon carbide. The 3 . . . . . carbon-carbon sp bonds in tetrahedral configuration, crucral for the formation of srngle crystal diamond (SCD), are responsible for a unique combination of diamond properties. However, SCD is too expensive for most applications in wireless systems. Furthermore SCD technology is not compatible with the conventional silicon technologies. Fortunately, the polycrystalline diamond, with physical properties approaching those of SCD, can be fabricated on silicon substrates using the chemical vapor deposition (CVD) techniques and its cost is comparable to that of polycrystalline silicon for film thicknesses in the range of l — 2 pm. Table 2.3 shows a comparison of properties of the three structural materials that have been studied for RF MEMS applications. Polycrystalline diamond films with the same acoustic velocity of SCD can be obtained by controlling the sp3/sp2 ratio during its growth [38]. As the polycrystalline diamond films contain sp and sp2 carbon-carbon bonds leading to non-diamond phases at the grain boundaries, their properties, such as acoustic velocity can be affected adversely if the polycrystalline diamond quality, indicated by the sp3/sp2 ratio as measured by Raman spectroscopy, is not carefully controlled. As mentioned earlier, reducing the resonator dimensions close to or below 1 pm for polycrystalline silicon resonators is expected to lead to size related limitations due primarily to enhanced adsorption properties of water-related species to the silicon surface. The diamond surfaces are known to be chemically inert to such adsorption, which makes this material more suitable and reliable for resonators with dimensions below 1 pm. 37 2.5.2 Chemical Vapor Deposition (CVD) of diamond films Therrnodynamically, graphite, not diamond, is the stable form of solid carbon at ambient pressures and temperatures. The CVD technique is based on decomposition of carbon-containing precursor molecules (typically CH4) diluted in H2 gas. This generally involves thermal (i.e. hot filament) or plasma (DC, RF, or microwave) activation, or use of a combustion flame (oxyacetylene or plasma torches). The fact that polycrystalline diamond films can be formed by CVD techniques is linked to the presence of H2 and hydrocarbon molecules. The H2 atoms are believed to play a number of crucial roles in the CVD process [39]. There are several techniques for the deposition of polycrystalline diamond including microwave plasma CVD (MPCVD), hot filament CVD, radio frequency CVD, and dc-arc jet CVD. MPCVD is the most widely used technique due to its efficacy to produce high film quality, large substrate size, less contamination and better controllability [40]. Figure 2.5 includes schematics of HFCVD and MPCVD techniques. These processes have been described in the past [41, 42, 43, 44]. All the deposition techniques rely on the ability to set up a dynamic non- equilibrium system, in which only sp3 carbon bonding can survive. This is achieved by the presence of hydrocarbon radicals, and more importantly, by large quantities of atomic hydrogen in the deposition gas. The hydrocarbon radicals provide the vapor source of carbon for diamond deposition. The atomic hydrogen is believed to play two important roles: it eliminates sp2 bonded carbon (graphite), while establishing the dangling bonds of the tetrahedrally bonded carbon sp3. By creating a plasma system where atomic hydrogen 38 predominates, any graphitic bonding is etched away. Atomic hydrogen etches both diamond and graphite out, but under typical CVD conditions, the rate of diamond growth exceeds its etch rate, while this is the opposite for other forms of carbon (graphite). a) —hydv8m — methane lPC power swirl .11.; pyrometer pump heating element \ thermocouple b) microwave 10de generator 333"“ urethane 8‘; : m... . -———substrate Wesmde: heater pimp Figure 2.5 a) HFCVD and b) MPCVD polycrystalline diamond deposition When the sample is ready for diamond growth it is loaded into the chamber and taken to a low pressure level. A mixture of gases (gases and their quantities depend on the type of polycrystalline diamond film that needs to be deposited) is then introduced into the chamber. The energy, provided by either DC. power supply or microwave generator, helps creating the conditions inside the chamber for creating the plasma (ionized gas). The sample is heated by the plasma and usually monitored with a pyrometer. 39 2.5.3 Seeding One of the most crucial steps in CVD growth of polycrystalline diamond films is generating seeds on the substrate (seeding) before the growth begins. In order to guarantee a continuous polycrystalline diamond film, a high nucleation density of diamond particles is necessary. Some of the techniques for this seeding step, are the abrasive polishing of the substrate with various grain sizes of diamond [45,46] and the ultrasonic agitation of the substrate with a diamond powder suspension [47,48,49,50]. Bias enhanced nucleation (BEN) has also been reported for effectively creating nucleation sites on silicon substrates [51] and preparing highly oriented diamond films [52,53]. An IC-compatible nucleation technique which does not cause surface damage [54] is the use of Diamond Powder Loaded Fluids (DPLF) with different carrier fluids, mean powder sizes and densities. The idea of this method is to spread diamond crystals, suspended in carrier fluids, on the substrate surface. During the diamond deposition process, the carrier fluids are evaporated at initial stages leaving behind the diamond particles which act as seeds for diamond growth. There are different ways to apply the DPLF to the substrate such as direct writing, spinning, spraying or brush painting. 2.5.4 Doping As a wide band gap semiconductor material (Eg = 5.5 eV), CVD diamond films deposited without intentional doping are usually good insulators. Since the micromechanical resonators which are meant to replace oscillating components in transceivers need to be conductive, the diamond fihns used for the fabrication of such devices, need to be doped. Because the 2000 K temperature necessary for effective 40 diffusion in diamond is too high [55], diamond doping is performed either during growth, or by subsequent ion-implantation. Boron, aluminum, phosphorous, lithium and nitrogen have been tested as dopants for diamond [56]. Polycrystalline diamond film doping (p-type) during the CVD process is usually achieved by using pure boron powder [57], boron trioxide (8203) [5 8], diborane (B2H6) [59,60], or trymethylboron [61]. IR measurements have confirmed that boron atoms occupy substitutional sites [62]. Diamond quality, grain size, growth rate, hydrogen and oxygen contents, dislocation and planar defect densities are affected by boron doping [63-66]. The changes in polycrystalline diamond resistivity as a function of doping concentration have been reported by [40]. 2.5.5 Patterning Due to diamond’s chemical inertrress, standard wet etching techniques are not possible. Two patterning techniques can be applied: selective deposition and selective dry etching. Selective deposition is achieved by selective nucleation or by masking the areas where diamond growth is not desired. SiOz was successfully used as a masking layer by Masood et al. [54], Roppel et al. [66] and Davidson et al. [67]. A nucleation technique, which consists of spinning a layer of photoresist pre-mixed with diamond powder and lithographically patterning it, was developed at Michigan State University [54]. Selective etching of CVD diamond with SiOz or Si3N4 as a mask, was performed at atmospheric pressure, in oxygen environment at 700 °C, in a rapid thermal processor [54]. 41 The dry etching of polycrystalline diamond, which uses different active gas species such as oxygen, argon, CR; and SF6 with metal or SiOz masks [69-71] seems to be an excellent choice for polycrystalline diamond patterning. Most researchers have used conventional reactive ion etching (RIE) methods where the gas species are excited by RF power [71-72]. Dry etching using ECR assisted microwave plasma at low substrate temperatures and pressures has led to very clean structures with small feature sizes and sharp edges [73-74]. 42 Chapter 3 Polycrystalline Diamond Film Technology 3.1 Introduction The work reported in this thesis focuses on the use of polycrystalline diamond as a micromechanical resonator structural material. This chapter explains the polycrystalline diamond film technology used for the fabrication of such devices as well as the fabrication process flow. 3.2 Seeding The seeding process influences the grain sizes in the polycrystalline diamond film and is the fabrication step responsible for guaranteeing a continuous film. The seeding technique used for the polycrystalline diamond fihns reported in this thesis consists of the use of a diamond powder loaded water suspension, commonly known as diamond in water (DW). This suspension is prepared by mixing of 25 carats of diamond powder, with an average particle size of 25 nm, in 1000 ml of deionized (DI) water and a suspension reagent. The DW suspension is placed in an ultrasonic bath for 30 nrinutes before it is intended to use. For the fabrication of the resonators reported in this work, the surface over which the seeding will be done is mainly SiOz. The starting substrate depends on the fabrication process that will be used, which are discussed in section 3.7. The first step is to clean the substrate, which was done using RCA cleaning [74]. Then, before the DW was poured over the substrate a quick dip of the sample in a diluted 43 HF solution (1% HF in H20) for 10-20 seconds was done to clean the SiOz layer. It was observed that this HF dip step improved the DW suspension surface coverage over the sample. After the HF dip, the sample was rinsed in DI water for 5 nrinutes, and then blown dried with nitrogen. Now the samples are ready to be seeded. The DW suspension is taken out from the ultrasonic bath and it is poured over the substrate. The sample, which is now covered with the DW suspension, was loaded in a conventional spinner, and was spun for 30 seconds at 1000 rpm. For improving the seeding density, the spinning process was repeated more than once. This seeding method is particularly suitable for clean hydrophilic surfaces [75] such as SiOz or Si3N4. A nucleation density of 1011 cm-z, and an average surface area of film grains of 0.06 pm2 has been obtained by using this seeding method [76]. 3.3 Film Growth Process All the CVD polycrystalline diamond films reported in this work were grown in a bell jar type MPCVD chamber (WavematTM MPDR 313EHP) with a 9 inch chamber diameter and a 5 inch quartz bell jar diameter. A 2.45 GHz, 5 kW microwave power TM . . supply (Sairem GMP60KSM) and a large chamber size ensured the uniformity of the plasma. The sample wafer was heated by the plasma and its temperature was monitored by a pyrometer. The gas mixture introduced to the chamber depended on the properties needed for the diamond film. Since the polycrystalline diamond film quality increases with lower methane concentrations [77], the methane concentration was never above 2% of the total gas mixture. As it will be seen in later chapters, a comparative study between different diamond films was made to identify the growth parameters that affect the diamond microstructure and consequently the Q. More precisely, the parameters varied in order to get diamond films with different microstructures were the growth temperature (600 °C - 780°C) and the gas percentage of tri-methyl-boron (TMB) in the MPCVD chamber during growth . The following samples are referred to in the rest of the thesis. 0 Sample 1: Undoped film grown 780 °C (66 measurements) 0 Sample 2: Highly doped film grown at 780 °C (69 measurements) 0 Sample 3: Undoped film grown at 600 °C (62 measurements) Each sample represents one specific type of polycrystalline diamond fihn from which different resonator structures were tested at different vibration modes and temperatures. Table 3.1 shows the polycrystalline diamond growth parameters used for each one of the films. 45 Table 3.1 Growth parameters for the three polycrystalline diamond samples studied and dry etching parameters (superscript denotes the sample on which the value was used) Polycrystalline Diamond Film Growth H2 1001,213 Gas(I;lcocv;1 Rate CH4 11,25 TMB 42; 0"3 Temperature (°C) 780”2 ; 6003 Microwave Plasma Power (kW) 2.01’2'3 MPCVD Gas Pressure (torr) 35i’2’3 Seeding Diamond Powder Size (nm) 251’2’3 Dry Etching Parameters Gas Flow Rate CE. 1’"3 (seem) 02 30"”3 Chamber pressure (mbar) 41110'2 1’2’3 DC Power (W) 400 W Figure 3.1 shows the Raman spectra for three representative polycrystalline diamond films. It can be noticed that from the three samples, sample 1 has the best -1 . looking Raman peak at ~1332 cm . It has been reported how the doping of polycrystalline diamond films and the low temperature synthesis affect adversely the sp3to spzratio in polycrystalline diamond films [77]. This can also be seen from figure 3.1, where the sharpest diamond (sp3 carbon-carbon bonding) peak corresponds to sample 1, which is the undoped film grown at the highest temperature (780°C). 3.4 Doping If the micromechanical resonator is intended to be used in oscillator or filter applications, the resonator structural material needs to be conductive. The doping technique used in this work is done in situ using TMB. The boron doping of 46 polycrystalline diamond films using TMB has been investigated in the past as a function of TMB/CH4 gas ratio, growth temperature and post growth anneal [40]. In this work, the diamond films that were heavily doped (e. g. sample 2) had a hole . 1 -3 . . . . concentration around 1 x 10 cm and a resrst1v1ty of 0.15 Q-cm. Thrs came as a result of a diamond growth gas mixture of 100:1 :4 (szCH4zTMB). 47 Raman Spectra 6000‘ 1 1 z 1 1 *i l g s s 3 3 . 1 i 5000 e ..................................................................... DZ 3; 4000 1 ................................................................................... 3‘ , 17: c 3C3 3000 . """"""" """""""""""""""""""""" "" ; Sample 2 2000 L ............. ............. ...... .............. .............. ............ .. l Sample 3 1000 ' 1 i 1 1 1000 1100 1200 1300 1400 1500 1600 Wavenumber (cm'1) Figure 3.1 Raman spectra for the three samples studied 3.5 Patterning As it was mentioned in the last chapter, due to chemical inertness of diamond, it can not be patterned using wet etching. The polycrystalline diamond patterning technique used in this work was dry etching using electron cyclotron resonance (ECR) assisted microwave plasma for the samples that were patterned at Michigan State University. For 48 the samples that were patterned at Sandia National Laboratories the dry etching was done using RF plasma. The dry etching system at Michigan State University etched polycrystalline diamond at an approximate rate of 0.05 11ml rrrinute, while the polycrystalline diamond etch rate for the system at Sandia National Laboratories was about 0.03 urn/minute. Aluminum was used as a hard mask material in both cases for the selective dry etching of polycrystalline diamond. The etch rate for aluminum in both dry etching systems is much slower than the etch rate of polycrystalline diamond. For the dry etching system at Michigan State University, the aluminum was etched around 13 times slower than polycrystalline diamond [40], whereas for the system at Sandia National Laboratories the aluminum was etched around 4 times slower [13]. Some of the recipes for dry etching of polycrystalline diamond can cause aluminum sputtering over some areas of the diamond surface. When this happens, the plasma can not attack the diamond under the sputtered aluminum. As a result, one might end up getting “spikes” around the patterned structure after the dry etching of diamond is completed in the rest of the sample. This is called the grass eflect (process shown in figure 3.2) and it has been reported in the past when deep reactive ion etching (DRIE) was used to pattern silicon-on-insulator wafers [78]. Figure 3.3-a shows SEM images of the grass effect noticed in the polycrystalline diamond structures patterned at Michigan State University with the spikes created by the grass eflect. These spikes are likely to be polycrystalline diamond. Since the fabricated polycrystalline diamond resonator structures lie on top of the $102 layer, when the structures are released, the spikes that are not attached to the structure are released as well. Figure 3.3-3 shows areas where the spikes are still attached to the structure, and some of the released spikes. In order to 49 obtain these SEM images with the released diamond spikes, the samples did not go through the cleaning process after the release step, and they were loaded into the SEM right after release. Figure 3.3-b shows an SEM image of a patterned silicon structure showing also the spikes generated by the grass eflect [78]. For the dry etching done at Sandia National Laboratories the gas mixture was a combination of oxygen plasma with CF4 (02:CF4 ; 30:1). This recipe etches A1 at a faster rate and eliminated the grass effect by etching the sputtered aluminum which caused the diamond spikes. The samples that were patterned at Michigan State University had to go through a more vigorous rinsing process in order to eliminate the released diamond spikes from the sample. Figure 3.4 shows SEM images of released polycrystalline diamond structures patterned at Michigan State University. Very clean surfaces and smooth sidewalls can be observed and the spikes discussed earlier are not noticed in these SEM images. For the patterning of these structures, the temperature inside the dry etching chamber was kept below 100 °C and the dry etching process was done for only 3 minutes. Since 3 minutes was not enough for etching the polycrystalline diamond fihn, the dry etching process was repeated until the polycrystalline diamond fihn was entirely removed. The dry etching system was not used for 30 minutes between each run. This suggests that a careful cleaning of the released samples together with a low temperature and short dry-etching intervals can also help reduce the sidewall roughness and lower the grass eflect. Figure 3.5 shows a comparison of the structures fabricated and patterned at Michigan State University and the very first set of results for polycrystalline diamond 50 resonators reported by [31]. The close up views of the structures show a comparable sidewall roughness. Aluminum Sputtering A! P-type Silicon (substrate) Cross Section View A-A’ Aluminum 1:] Diamond Silicon Dioxide Sputtered Aluminum Figure 3.2 Grass effect diagram. The sputtering of aluminum during the dry etching of diamond prevents the plasma from etching the diamond under the sputtered aluminum. 51 | Silicon Spikes] Figure 3.3 Evidence of the grass efi'ect found on (a) the dry etching of the polycrystalline diamond resonators reported in this thesis and (b) the DRIE of silicon-on-insulator wafers [78]. 52 Anchors \ . 5011111 Corjib ‘ \ Anchors\§~ 10 11111 Figure 3.4 Released polycrystalline diamond structures (comb-drives and bridges) patterned at Michigan State University. N o grass effect can be observed. 53 Figure 3.5 a) Fabricated and patterned polycrystalline diamond resonators from [31] and b) Michigan State University (reported in this thesis) 54 3.6 Surface Roughness As it was mentioned in the previous chapter, researchers have found the surface roughness of a micromechanical resonator to influence the Q [19,23,26]. Randomly oriented polycrystalline diamond films (which are the ones used for this work), show very rough surfaces when compared to polycrystalline silicon fihns [79]. It is therefore important to have some control over the surface roughness in polycrystalline diamond fihns which are going to be used for micromechanical resonators. The results obtained in this work, show similar surface roughness for undoped films grown at 600 °C, and highly doped films grown at 780 °C (sample 2 and sample 3). Figure 3.6 shows AFM images of these two samples. The difference in surface roughness was only about 5 nm. 55 / / Img Rms (Rq) 57.52 nm X scale 2p./div Z scale lit/div b) Img Rms (Rq) 62.35 nm X scale 2p./div Z scale lu/div Figure 3.6 AFM images showing the polycrystalline diamond surface roughness of Sample 3 (a) and Sample 2 (b). The difference in the film roughness is about 5 nm. 56 u. -.. 1.810. f I. . ...—..., Silicon Substrate Figure 3.7 Starting substrate for the fabrication of polycrystalline diamond resonators 3.7 Microresonator Fabrication Process The following section describes the fabrication process steps for the resonators used in this work. 3.7.1 Fabrication technology In this work, two testing methods were used: 1) the electrostatic actuation and detection and 2) the piezoelectric actuation with optical detection. Depending on the testing method that will be used, micromechanical resonators can go through different fabrication steps. A device that has been fabricated for electrostatic actuation can be tested using either electrostatic or piezoelectric actuation with optical detection. However, if the device is going to be tested using only piezoelectric actuation with optical detection, the fabrication process can be simplified significantly. Both fabrication processes, begin with the same substrate which is shown in figure 3.7. It consists of a commercially available silicon wafer with a 2 pm thick thermal SiOz (wet grown) layer on top. The doping and orientation of the silicon wafer is not critical. In this thesis, a p- type (100) silicon wafer was used. 57 3.7.1.1 Samples for electrostatic testing Figure 3.8 shows the fabrication process flow for a micromechanical resonator which is meant to be tested using electrostatic actuation and detection. The thermally grown SiOz layer that comes with the starting substrate is necessary to electrically isolate the devices from the substrate. The fabrication of these devices consists of a 3 mask fabrication process. The first step for the fabrication of a micromechanical resonator that will be tested using electrostatic actuation and detection (after the starting substrate was cleaned using standard RCA cleaning process [74]) is the deposition of a thin (~ 300 nm) layer of silicon nitride (Si3N4). The deposition of this layer prevents the etching of the underlying Si02 layer at the time of release. The justification of this Si3N4 layer is explained in more detail later on, when the releasing of the structure is discussed. Since the purpose of this layer is to simply isolate the underlying $102 layer from the future wet processing of the sample, the details about its deposition are not critical. The second step is the deposition and patterning of the layer that will serve for making the electrical connections for testing (mask #1). Highly doped polycrystalline silicon has been used for this layer in the past. However, a metal had to be deposited over it, in order to decrease the contact resistivity [80]. In the work reported in this thesis, a single layer of chromium (100 nm thick) was deposited and patterned for electrical connections. 58 q MATERIALNTEROSS SECT'OMROCESSING STEP“ VIEW Chromium SisN4 3102 * a) Metal deposition Silicon Substrate ‘ PECVD S'O ' 2 b) Metal patterning and PECVD Si02 deposition Pogprystalcljine c) PECVD Si02 \ 'amO" j ‘ Patterning and diamond deposition d) Diamond patterning e) Release (PECVD @2 removal) / Figure 3.8 Fabrication process flow for electrostatically tested resonators After this metal layer was deposited and patterned, it follows to deposit the material that will serve as a sacrificial layer. In this work, SiOz was deposited via PECVD and used as the sacrificial layer. 59 This layer will be removed entirely at the last step of the process (releasing step). The thickness of this layer will dictate the electrode to resonator gap. The gap value used in this work is 200 +/- 100nm. This range of values is very crucial for the electrostatic operation of polycrystalline resonators. As the gap increases, the electromechanical transduction factor (n) discussed in chapter 2 decreases. When this factor decreases, the output current also decreases, limiting the detection capability. If it is too thin, it can get completely etched away by the plasma in the diamond growth step, leaving a short circuit between the beam and the bottom electrode. A thickness around 200 nm has been successfully used in the past [81]. After this sacrificial layer has been deposited, it has to be patterned to create the contact areas where the resonator will be in contact with the metal layer underneath (mask #2). These areas are commonly called “anchors”. Once the anchors are patterned, the sample is ready for diamond seeding and growth (sections 3.3 and 3.3). A device that will be tested using electrostatic techniques needs to be conductive. For achieving this, in situ doping was done (section 3.4). For patterning the diamond (mask #3), dry etching was used (section 3.5). After the diamond film has been patterned, the exposed areas of the sample are SiOz and diamond. Now the devices are ready for releasing. In the release step, the purpose is to remove the entire sacrificial SiOz layer. This is done by submerging the sample in a 5:1 BHF solution (5 40% NH4F :l 49% HF). This solution typically etches the PECVD SiOz at a rate between 250 nm/min and 500 nm/min [82]. The time that it will take for the solution to remove entirely the sacrificial layer depends on how large is the area under the resonator structures. This 60 etching solution (5:1 BHF) is isotropic. Therefore, if a 10 um wide cantilever needs to be released, it will take the solution at least 20 minutes to etch 5 am from each side and completely release the structure. If a faster releasing is desired, a solution with higher HF concentration can be used or even pure HF (49% by weight, remainder water). Both etching solutions remove the aluminum that was used to mask the diamond. If the Si3N4 layer was not deposited, the SiOz etching solution (BHF or HF) will etch parts of the 2 um thick thermal SiOz that came with the starting substrate. If this happens, the corner areas of the metal interconnect layer, which in this case would be over the thermally grown SiOz layer, would be partially released as well. This could lead to metal peeling off. Also, the undercutting of this metal layer will affect the resonator Q since the rigidity of the anchor can be affected. The Si3N4 isolates the thermally grown SiOz from the sacrificial PECVD SiOz in the release step, which avoids this problem and also gives more flexibility for the releasing time. 3.7.1.2 Samples for piezoelectric/optical testing The fabrication process for a micromechanical resonator that is going to be tested using piezoelectric actuation and optical detection is much simpler. There are fewer processing steps and restrictions on the design. These include: 1) The underlying electrode required by the electrostatic method is not necessary. 2) The material does not need to be doped 3) The resonator gap is not critical. 61 Figure 3.9 shows the fabrication process flow for devices to be used using piezoelectric actuation and optical detection. Since there will be no current flow across the resonator and no time varying capacitance needs to be formed, metal electrodes are not necessary and the resonator does not need to be conductive. Since there is no capacitance formed or electromechanical transduction factor (n) to keep high, the gap between the beam and the substrate is not critical. In this work, the 2 pm thick SiOz layer that comes with the commercially available sample (figure 3.7) was used as the sacrificial layer. This would lead to a resonator to substrate gap of 2 am. The fabrication consists of a two mask fabrication process. The processing for a micromechanical resonator that will be tested using piezoelectric actuation and optical detection starts with the standard RCA cleaning of the substrate. Then mask #1 is used to pattern the thermally grown SiOz layer that comes with the starting substrate and form the anchors. Now the sample is ready for diamond seeding and growth (sections 3.2 and 3.3). The second and final mask is used to pattern the material (aluminum for purposes of this thesis) that will be used to mask and pattern the diamond (section 3.5). The releasing step is the same used for the releasing of devices for electrostatic testing. The difference is that now the sacrificial thermal SiOz (wet grown) etches at a much slower rate in the BHF solution (100 nm/min) than the PECV D SiOz. Therefore, the releasing time will be longer. It will now take at least 50 minutes to completely release a 10 um wide cantilever beam. 62 VIEW q MATERIAL PFROSS SECT'OFI‘4ROCESSING STEPI\ $102 Silicon Substrate : Polycrystalline Diamond a)Si02 patterning and k 1‘ diamond deposition b) Diamond patterning c) Release (PECVD SiO removal) U / Figure 3.9 Fabrication process flow for resonators to be tested using piezoelectric actuation and optical detection 3.7.2 Post-processing After the samples are released, some cleaning and drying steps were made. The samples that were tested at Michigan State University (tested electrostatically) could not go through standard RCA cleaning or any other cleaning process that involved hydrogen peroxide (piranha cleaning for example). This is because the metal layers present in such samples could suffer oxidation or be removed. After the sample was released, it was submerged in acetone for 5 minutes, then rinsed in a beaker with running DI water for 5 minutes and then submerged in methanol for 5 minutes. Since there might be released diamond spikes to be removed, the sample was then rinsed again with running DI water 63 for about 10 minutes. Finally, the sample was then heated in a hot plate to 100 °C and then annealed to 500 °C in a nitrogen environment for removing any moisture. The samples that were processed and tested at Sandia National Laboratories (piezoelectric actuation and optical detection) went through a different post processing. After the sample was released, it was taken into an ultraviolet ozone cleaning system to remove the organic residues, and finally it was heated to 800 °C in an argon atmosphere using rapid thermal annealing to remove any residual moisture. 64 Chapter 4 Resonator Testing Methods 4.1 Introduction As it was explained in chapter 2, there are several methods for testing micromechanical resonators. In this thesis, two different techniques were used to test similar resonator structures. The two methods used are the electrostatic testing and the piezoelectric actuation with optical detection. 4.2 Electrostatic Testing This section describes the equipment set-up for the electrostatic testing system and the measurement results of polycrystalline diamond resonator structures. The equipment for electrostatic actuation was set-up at Michigan State University for the first time. 4.2.1 Electrostatic testing equipment set-up Figure 4.1 shows a diagram of the testing system and the connections to the resonator structure. The system consists of a Low Temperature Microprobe vacuum chamber supplied by MMR Technologies capable of reaching a vacuum level of 10 mtorr. The chamber had four probes with electrical connections to the outside of the chamber by SMA (Sub-Miniature A) connectors. By using these SMA connectors, the network analyzer and the dc-voltage supply were connected to the probes inside the chamber and consequently to the resonator structure. The network analyzer was used to 65 supply the input ac signal and to constantly monitor the output current. A bias tee was used to isolate the dc-bias from the network analyzer. As soon as the beam starts vibrating, a change in the output current flow will be detected by the spectrum analyzer and it will show a maximum value at the resonant frequency. For purposes of this work a Hewlett-Packard/Agilent 8753B Network Analyzer (300KHz —- 3GHz) was used. The dc- voltage supply was used to apply the dc-bias needed in order to increase the output current and get a more noticeable peak. A more detailed explanation of the theory involved in electrostatic testing can be found in section 2.2.3. 1. Bias Tee POI-III’ Vacuum Chamber (100 mtorr) Network Analyzer Figure 4.1 Connections to the resonator structure for electrostatic testing. In order to measure only the change in current across the resonator, a bias tee is used to isolate the dc- bias from the IN-port of the network analyzer. 4.2.2 Electrostatic testing measurement results Different resonator structures were tested electrostatically. The results obtained were on bridge, comb—drive and cantilever resonator structures tested at a vacuum level 66 of 100 mtorr. Figure 4.2 shows the measurement results for a bridge and comb-drive structure tested using a dc-bias voltage of 30 V and SEM images of the tested structures. It can be noticed how the resonant peak shows a slight non-symmetry with respect to the resonant frequency point. This could be due to a non-linear effect caused by over driving the resonator structure. This problem has been found in the past by Kaaj akari [83]. 1 67 V‘ LII): 5011111 _ ‘U/I, \Q/ , f/rr. Figure 4.2 Electrostatic measurements on polycrystalline diamond resonators: a) bridge structure, b) comb-drive. The resonant frequency (f,) was calculated by finding the frequency at maximum magnitude and the Q was calculated using this f, value and the 3dB bandwith (Af) 68 4.3 Piezoelectric actuation with optical detection The piezoelectric actuation and optical detection was done using the measurement system that was already built at Sandia National Laboratories. The basic idea of the piezoelectric actuation method is to drive the resonator into vibration by using a piezoelectric transducer, which is in contact with the sample that contains the resonator structure. This piezoelectric transducer is driven into vibration at a frequency determined by the output of a network analyzer. For the optical detection, a laser is focused on the vibrating beam and the reflected signal is used to determine the frequency of vibration. A more detailed explanation is given in section 2.2.3.2. 4.3.1 Piezoelectric actuation and optical detection results The tested structures were cantilevers and torsional resonators. The cantilevers reported in this section had lengths ranging from 50 pm to 100 um, and they all had a width of 10 pm. The thickness varied from 0.45 um to 0.7 pm. The table 4.1 summarizes the set of results obtained for these cantilevers made from the sample 1 described in chapter 3. Researchers have found different values for the Young’s modulus of polycrystalline diamond at room temperature [84]. The Young’s modulus of the polycrystalline diamond resonators fabricated and tested was checked using the data shown in Table 4.1 and the following equation for the resonant frequency of a cantilever [10]. f0 =O.1615 51 (4.1) PL; 69 Table 4.1 Results for piezoelectically-actuated cantilever beams (first vibration mode) made from sample 1. Cantilever Performance Length (um) f0 (HZ) Q 100 258,900 9,970 95 276,400 7,540 90 325,500 14,470 85 346,300 10,070 80 384,900 15,260 75 443,500 8,240 70 503,800 6,390 65 570,900 3,550 60 654,300 5,390 50 856,500 6,726 where E and p are the Young's modulus and density of polycrystalline diamond respectively. The constants L,- and h are the beam length and thickness respectively. Figure 4.3 shows a plot of the data shown in table 4.1. Since the plot in figure 4.3 shows 1 the fiequency ( f0) on the y-axis and the inverse of the cantilever length squared ( 2 Lr )in 70 the x-axis, equation (46) can be fit with a linear equation by takingx ___;2’ y = f0 Lr andm = 0.1615J—E—h. This would lead to: p f0 = 0.1615 -E—h[—12-] (4-2) 7 L a p J L, W—/ m x The Young’s modulus can be calculated by using the expression for the slope m=0.l615\/§h. p The linear fit to the data shown in figure 4.3 shows a slope (m) of 0.002 Hz-umz. The value for the film thickness (h) can be approximated from the SEM images to be 0.7 pm and the value for p is commonly know to be 3520 kg-m-3. Since the variables h, p, and m are known, the Young’s modulus for the polycrystalline diamond film (E) was calculated to be 1,000 GPa. This value is close to the accepted value of the Young’s modulus for bulk single crystal diamond, which is about 1,050 GPa [85]. The inset in figure 4.3 shows the resonant peak of the 95 um long cantilever beam. 71 Data for 95 um long cantilever Lorentzian Fit f,,=276,400 Hz —-— Data Q=7.540 Linear Fit 5. 9x10 ‘ (Normalized Amplitude) 2 276,200 276.400 276.600 8x1 05- . Frequency (Hz) 7x105- 6x105- 5x105- 4x105: 3x105.- 5. 9.00x107 1.80 x108 2.70 x108 3.60 x108 1/L2 Figure 4.3 Young’s modulus of sample 1 was calculated from a linear fit done to the measured data. ~ ratio of 0.045. This ratio means that the r measured resonant frequency of this device using electrostatic actuation will be (theoretically) 95.5% of the resonant frequency that would be measured using a testing method which does not affect the resonator spring constant. The resonant frequency measured using electrostatic actuation was 315,240 Hz which is 96% the resonant frequency measured using piezoelectric actuation and optical detection (325,560 Hz). Another possible reason for the difference in frequency is a film thickness difference between the two polycrystalline diamond films from which the resonators were fabricated. 75 3) Electrostatic Actuation _.v' -'~):v...' , ~,_.. --. . J ,/ rv' / 1‘ I _,, ,.. i . .._-' i" ‘I i. ’5’/ ."J i: i . ‘ ; -. ' J l “a ‘ f“ ‘ ' ”L i 4“; "A f =315,240 Hz 8 1 0 g Q=1,902 g. - < 0.5 1 ' i g : v 0 314,900 315,200 315,500 Resonant Frequency (Hz) b) Piezoelectric Actuation N 1 ] g ‘ f0=325,560 Hz i I Q=l4,469 g 0.5 1 g 1 Z . v 0 i 325,300 325,600 325,900 Resonant Frequency (Hz) Figure 4.5 Testing results on two similar polycrystalline diamond cantilevers using different actuation and detection methods. 76 Also, although both films were fabricated under similar conditions (table 4.2), their thickness could still be slightly different due to possible fluctuations in temperature at the polycrystalline diamond grth step. The difference in Q is likely not related to the method of actuation or the properties of the polycrystalline diamond film, but rather due to the different measurement pressures used in the two methods, 100 and 0.001 mtorr for electrostatic and piezoelectric actuations, respectively. The pressure of 100 mtorr reflects a low vacuum and, thus, represents what would be readily achievable in a packaged device [87]. At 0.001 mtorr there should be negligible air damping, and the Q should reflect the true mechanical Q of the structure. The effect of air damping on Q can be readily estimated. At 100 mtorr, the Q is limited by the exchange of energy between the mechanical structure and gas molecules. This is in the molecular flow regime, and the magnitude of Q can be calculated according to the following equation [22]: _ 271nm? Q - _km P (43) where f" is the resonance frequency of the 11th vibrational mode, p is the density, P is the 1/2 gas pressure, and km = [(32M)/(92r.R7)] , where M is the molecular weight of the gas, R is the gas constant, and T is the temperature. The molecular flow regime extends down to about 10 mtorr, and below this point air damping is negligible. Therefore, decreasing the pressure item 100 mtorr to pressures less than 10 mtorr should result in an increase in Q of about a factor of 10. For the cantilever structures tested using electrostatic and piezoelectric actuation, the measured Q values were in the range of 1,000-2,000 and 3,550-15,260 respectively. The low quality 77 factors of the electrostatically structures are consistent with the Q being limited by air damping, and the quality factors of these structures are representative of what can be achieved for structures of these dimensions in a low vacuum package [87]. The piezoelectrically-actuated structures in high vacuum should not be limited by air damping. Similar resonator structures also made from polycrystalline diamond were tested electrostatically under 50 mtorr by Wang [31]. Their reported Q values range from 2,000 — 6,000. Later results, also on similar polycrystalline diamond resonators tested electrostatically under 50 utorr [32], showed Q values in the range of 6,225 — 36,460. Since the molecular flow regime extends down to about 10 mtorr, and below this point air damping is negligible, a decrease in pressure from 50 mtorr to any pressure below 10 mtorr would represent an increase of Q by a factor of 5. The results obtained in [31] and [32] show an improvement of Q close to a factor of 5 when the pressure in the chamber is lowered from 50 mtorr to 50 utorr. Based on the measured resonant frequencies and the dimensions of the structures, an acoustic velocity of 16,100 ms.1 was found for the doped fihn and 15,900 for the . . -3 . undoped one. Assuming a densrty of 3,500 kg-m , the resulting Young’s modulus for the polycrystalline diamond films are 883 GPa for the doped sample and 907 GPa for the undoped one. The slightly lower moduli found here could be due to a slightly lower density for the polycrystalline diamond films or an error in the estimation in the polycrystalline diamond film thickness. 78 Chapter 5 Study of Q and Frequency Shifts in Poly- crystalline Diamond Resonators 5.1 Introduction Two very important characteristics of micromechanical resonators are resonant frequency and Q. This chapter discusses the Q and resonant frequency shifts due to varying testing temperature for polycrystalline diamond resonators. The data obtained during this study showed the measurement of the highest quality factor (116,000) for polycrystalline diamond resonators and for cantilever beams made of any polycrystalline material [33]. 5.2 Quality Factor Limitations As it was discussed in chapter 2, there are different mechanisms that can limit the Q. Figure 5.1 shows the effect of different energy loss mechanisms as a function of resonant frequency for polycrystalline silicon and polycrystalline diamond computed for a 1 um thick cantilever beam. The corresponding cantilever beam length for each frequency and material appears on the top axis. The parameters used for plotting the dissipation curves [25] in figure 5.1 are shown in the Table 5.1. It is interesting to note that the increase of resonant frequency achieved by reducing the length of the beam leads to a substantial decrease in Q because the anchor losses are dominant mechanisms in the whole frequency range. In other words, Q can be substantially increased by minimizing anchor losses. However, in that case, the maximum achievable Q will be determined by 79 other energy loss mechanisms as shown in figure 5.1. As most applications require a Q of at least 10,000, the anchor losses will limit the frequencies to less than 10 MHz. Thus, for applications requiring frequencies above 10 MHz, if the clamping losses can be minimized, the losses due to other mechanisms will allow Q values of 10,000 as seen in figure 5.1. Quality Factor — Diamond 1_' Silicon Resonant Frequency (Hz) Figure 5.1 Energy loss mechanism curves for a 1 pm thick poly- crystalline diamond and polycrystalline silicon cantilever beam 80 Table 5.1 Polycrystalline diamond and polycrystalline silicon properties used for plotting the dissipation curves in figure 5.1 Polycrystalline Polycrystalline Parameter Silicon Diamond Thermal Expansion Coefficient, a( x 10'6 "C’1 ) 2-67 1-1 Yorgg’s Modulus, E (GPa) 170 41050 Density, p ( kg m'3) 2329 3520 -l -1 Specific Heat Capacity, C9 ( J kg K ) 700 502 Heat Capacity per Unit Volume, C (J mn3K-1 ) L63 1.767 Thermal Conductivity, 1c( W Ill-l K.l ) 130 1500 Gruneisen’s Constant, y l 1 Velocity of longitudinal and transverse sound . . 3 -1 9; 5.5 18; 13.1 waves in the material 0| ; 0t (l x 10 ms ) 5.3 Quality Factors in Polycrystalline Diamond Resonators This section shows the results obtained for polycrystalline diamond cantilever beams with widths of 10 um, and lengths and thicknesses varying from 100 um - 500 um, and 0.45 pm - 0.8 pm respectively. Piezoelectric actuation and optical detection was used as the testing method for all the results reported in this chapter. The samples were tested at a vacuum level of 1x10" torr. The performance of the resonators (frequency and Q) was obtained from the Lorentzian fits done to the obtained data. The frequencies of the tested resonators were all in the KHz range (8 KHz — 800 KHz), and the Q values varied significantly from sample 1, where a Q of 116,000 was obtained, to the other two samples (sample 2 and 3), where the Q values did not exceed 50,100. Figure 5.2 shows some of the Q limiting curves in polycrystalline diamond resonators (solid lines) and the measured data (experimental points) on all the samples. These curves apply to a 0.7 pm thick polycrystalline diamond cantilever beam. The clamping curve does not depend on the cantilever temperature, and this is why figure 5.2 81 shows only one curve for clamping losses. The curves for erao and Qpi, will be shifted in the y-axis by the cantilever temperature. Figure 5.2 shows 2 curves for each of these temperature dependent dissipation mechanisms. They are plotted to include the lowest and highest temperature at which the samples were tested (30 °C and 400 °C). 1X108 v- ooc' _1 1 leO7 I P CTy2 1+(wrph? [ f 2 FopCp 1+[-——) 1... F0 O 6 r QTED= 3 1x10 fazm‘ a: w a . dish 1 =O.34x \ .5 3, 1x105 8 Ogegn y r 30°C :1 n n 2 gg '800 400 g 1 104 Eng ° 0 x i P o F 3%, ° [3 Samplet . 0 o O 0 . 1x103 , o 0 Samplaz 1: O Sample3 1 1x104 1x105 1x106 1x107 1x108 1x109 Frequency (Hz) Figure 5.2 Q value limiting curves for a 0.6 pm polycrystalline diamond cantilever beam and measured data for the three samples. 82 All the Q factors measured in this work are below any of the dissipation curves, suggesting that the Q values are not limited by any of these mechanisms. The most likely mechanisms for limiting Q values in polycrystalline diamond are mechanical relaxation processes at defects in the films, such as vacancy or impurity motion, grain boundary sliding, etc [13]. A study of the temperature dependence of Q in polycrystalline diamond resonators together with the study of the polycrystalline diamond microstructure will allow a better understanding of these mechanisms. Figure 5.3 shows a plot of the obtained Q values for the three samples (same experimantal points shown in figure 5.2) in a range of frequency where the data can be observed clearly. It also shows the resonant peak with the highest Q value for each sample. 83 1.6 0.5 (Normalized Amplitude) 0. fo= 318.183 Hz Q=116,400 Sample} 318.180 318.190 \. _ 318,170 g 12 k Frequency (Hz) J : l a ‘;< 10» U Sample1 a u 1: _ 0 Samplez : L 8 0 Sample3 3 II '5 5' .9 . ' - 0 _ a B (U 1.1. , 6 3‘ ~ . 'r-B 4 .- 3 O O 2* a a Frequency (Hz) Figure 5.3 Measured Q values as (g Sample 3\ (a Sample 2 r3. 1. g ...l g f fo = 89.030 Hz g fo = 120, 970 Hz E ' Q = 50,100 ”g °" 2 2 Frequency (Hz) j ”1 Q: 47, 900 0.0 120.940 120.960 120.980 121.000 Frequency (Hz) J a function of frequency for the three studied samples. Resonant peaks show the highest Q value for each sample. 84 5.4 Polycrystalline Diamond Film Microstructure In order to characterize the different polycrystalline diamond films, TEM images of the cross sections from one piece of each sample were taken. The cross sections were obtained by using the focused ion beam (FIB) technique, which uses an ion beam to raster over the surface allowing the milling of small holes in the sample at well localized sites, so that cross—sectional images of the structure can be obtained. From the TEM images, the polycrystalline diamond fihn can be studied in more detail including the nucleation layer (seeding layer) that is necessary for polycrystalline diamond growth. This nucleation layer is typically formed of small grains (~30 nm in diameter). Figure 5.4 shows how the grain sizes fiom the studied polycrystalline diamond films increases from the nucleation layer up, showing a larger percentage of the fihn composed of small grains for the sample 2. As the polycrystalline diamond fihns contain sp and sp2 carbon-carbon bonds leading to non-diamond phases at the grain boundaries, their properties can be affected adversely in this nucleation layer. According to figure 5.3 the Q values are larger for the polycrystalline diamond film with less percentage of the film composed by the seeding layer (sample 1). This preliminary result suggests a relation between Q and the nucleation layer in polycrystalline diamond resonators. The characteristics of this nucleation layer (thickness, grain sizes, nucleation density, etc.) can be controlled by the growth process. Polycrystalline diamond films can be divided into two layers: 1) The seeding layer and 2) The film layer. 85 SAMPLE 1 SAMPLE 2 Film Percentage Occupied by Nucleation Layer: ~Sample 1 ~ 10% -Sample 2 ~ 18% 08ample 3 ~ 13% »| Figure 5.4 Cross section TEM images of the three studied polycrystalline diamond samples. 5.4.1 Seeding Layer vs. Film Layer If a cantilever beam is treated as a two layer structure across its thickness [88], where the two layers are the seeding and film layer, the total dissipation can also be thought of as the combination of the dissipation in these two layers. In other words, the total dissipation in the cantilever beam would be partially due to the seeding layer, and partially due to the film layer. The total dissipation can then be expressed as [89]: .,/ tf/ 1 t tot ttot =___+— (5.1) Qtot Qs Q f 86 where ts, and tf, are the thicknesses of the seeding and film layer respectively and ttot 2 ts + t f' The values HQ, and I/Qf, are the dissipation due to the seeding and film layer respectively, and I/Qtot is the total dissipation in the cantilever structure. The purpose of this analysis is to find if there is a common dissipation among all the polycrystalline diamond films (Samples 1, 2 and 3) due to the seeding layer HQ, and another due to the film I/Qf. For this analysis, the values for ts, and tfwere obtained directly from the TEM images. For estimating the Qtot for each fihn, the arithmetic average of the 15 higher measured Q values of each sample was calculated. The 15 higher Q values for each sample were in the following ranges for each sample: 1) Sample 1: 76,560 - 116,000 2) Sample 2: 31,596 - 47,900 3) Sample 3: 23,544 - 50,100 The obtained values for ts / ttotr tf ”tot and the average of the 15 higher Q values for each sample (Qtot) are shown in table 5.2. Now that these variables are known, they can be substituted in equation (5.1) for each of the three films. The result of this substitution is shown in equations (5.2), (5.3), and (5.4). 87 Table 5.2 Ratio of seeding and film thickness to total thickness for each sample and the average of the 15 higher measured Q values for each sample (Qtot ). Sample/Parameter ts/ttot tf/ttot Qtot Sample 1 0.1 0.9 96,000 Sample 2 0.18 0.82 35,000 Sample 3 0.13 0.87 30,000 1 _ 0.1 0.9 — + =>S l 1 96,000 QS Q f amp 8* (5'2) 1 _O.18+0.82 35,000 Q3 Qf 1 _O.13+O.87 30,000 QS Qf :> Sample_ 2 (5-3) :> Sample _ 3 (5.4) Now, if equation 5.1 is rearranged in the following way 1 = 1 + t s 1 _ l (5.5) Qtot Q f ttot Qs Q f the values for the Qtot can be plotted as a function of tS/tm, and from the slope of a linear fit and the intercept in the y-axis, the values for Q5 and chan be obtained. The resulting plot is shown in figure 5.5. From the slope an intercept values, the calculated Qf and Q, values are: Qj=66,000 and QS=4,737 88 3.5‘ 1 3 _. 1 25-i /Qtot ' . (10-5) 2 - 1.5 Slope = 1.961x104 , Intercept = 1.5x10'5 1 _ I I ' I ' I ' I ' I 0.1 0.12 0.14 0.16 0.18 ty ‘tot Figure 5.5 Plot of 1/Qm as a function of tS/ttop The values for Q; and Q, are obtained from the slope and intercept of the linear fit. Another way to find the range of values for Qf and Q3, would be to plot the Qfas a function of Q, by using the functions shown in equations 5.2-5.4 and then approximate the minimum space which contains a set of values for Qfand Q5 which satisfy each of the equations in the system. One possible way and find this minimum area or space of solution is to find a rectangle that contains at least one set of values for Qfand Q5, which belong to each of the curves. This is similar to what was done by [90] and [91]. The horizontal line that would describe this rectangle (HOR) is the difference in Q, from 89 sample 1 and sample 3. The vertical line that would describe this rectangle (VER) is the difference in forom sample 1 and sample 3. Using equations (5.2) and (5.4), these lines can be taken as: _ 9,600Q f 3,900Qf - Q f — 86,400 — Q f — 26,100 (5.6) __ 86,400Qs _ 26,100QS Q3 —9,600 QS -3,900 VER (5.7) In order to minimize the area of the rectangle, it would be necessary to minimize the product of equations (5.6) and (5.7). An approximation of this minimization was obtained graphically by plotting the product of equations (5.6) and (5.7) in terms of Q3 and Qfand obtaining the values for Q3 and Qfat which the area of the rectangle is minimum. Fig 5.6 shows a plot of the obtained rectangle. The range of values for Qfand Qs that are defined by the rectangle are: Q f = 70,500 i 30,000 5.8 93 =30,250:19,750 ( ) These values suggest that the Q of a polycrystalline diamond film consisting entirely of the seeding layer would be about 30,000 +/- 20,000 whereas a polycrystalline diamond film having no seeding layer would have a Q of about 70,000 +/- 30,000. The Q of a polycrystalline diamond film containing both a seeding layer and a film layer can be estimated using equation 5.1. 90 20~-[--E-? ------------- F ------ i ------ Qs=50,000 1 g : g ; Qf=106,930 5 15—-1-4.-:r-—---—-a-- ------- i ----- Qf l a 2 . : a 4 10-- ~17? ------ E ------- i ------- i ----- \— - ----- i ----- 1‘ (1x10) i\ ‘i: (Qs=30,250 01:70.500) : : a =- a a a ‘ £1 5 e i Qs=11,223 1 104 Qf=40,000| Qs(x ) Figure 5.6 The approximation for the minimum space of solutions is determined by the rectangle, which touches all the curves at least in one single point. 91 5.5 Temperature dependence It is important to guarantee the micromechanical resonator thermal stability in order to make the device suitable for high temperature applications. Since the elastic properties of most polycrystalline materials change at different temperatures, the performance of resonators made from such films also varies with temperature. Also, it is important to be able to predict the shift in resonant frequency of a mechanical resonator when the ambient pressure changes. 5.5.1 Young’s modulus and frequency temperature dependence As it has been pointed out earlier, for most resonators the frequency depends on the Young’s modulus of the material the resonator is made of. Most polycrystalline materials have a temperature dependent Young’s modulus. Consequently, there is a frequency shift on polycrystalline resonators when they are operated at different temperatures. The Young’s Modulus of polycrystalline diamond decreases with temperature [92] and so does the frequency of resonators made of polycrystalline diamond. In this work, cantilever beams made from the three different polycrystalline diamond films (described in section 3.3) were tested at different temperatures. The temperature dependence of the polycrystalline diamond Young’s modulus was obtained from the measured frequency values at different temperatures and using a linear fit analysis similar to the one used in section 4.3.1. Figure 5.7 shows the variation of the Young’s modulus for the three polycrystalline diamond films. In the studied temperature range (30 °C — 400 °C) the samples show a Young’s modulus decrease of 2.5% (sample 1), 3.8% (sample 2) and 2.6% (sample 3). 92 Young's Modulus (GPa) 1050 — 1000 -, ____ Sample 1: Undoped, 780 °C _____ _ F A k 4 950 - ————-~— ,__ —— ~ 900 ~ —— ~ S I 2: 5 1019 '3, 780°C 850 fig ___} ampe x :n'\' 800 Sample 3- Undoped 600°C 750 -___.__.h ._ +. i J_. 700 I r I I 100 200 300 400 Tom perature (C) Figure 5.7 Young’s modulus temperature dependence for the three samples. The three samples have very similar slopes and the sample 1 showed the highest Young’s modulus at room temperature. This decrease in Young’s modulus is reflected in a resonant frequency shift of resonators made from these films. The shifts in resonant frequency due to changes in temperatures are measured by the temperature coefficient (T Cf). This temperature coefficient represents the resonant frequency shift in relation to the room temperature resonant frequency (A%0 ) by degree centi grade. Figures 5.8 - 5.10 shows the frequency shift Af f0 for 300 um and 400 um long cantilever beams made from the three different samples when operated in their first two flexural modes. The resonant frequency shift by degree centigrade was obtained by calculating the slopes of linear fits done to the measured data. The range of the obtained 93 temperature coefficients (TCf) for the three types of polycrystalline diamond structures go from -1.59x10'5 °C’1 to -2.56x10'5 °C'1 (-15.9 ppm/°C to -25.6 ppm/°C). These ranges are comparable to those obtained for poly-Si structures (-12.5 ppm/°C to -l6.7 ppm/°C ) in a shorter temperature range (30 °C — 247 °C) [11], but are far from values obtained for geometrically compensated poly-Si structures (~2.5 ppm/°C to -0.24 ppm/°C) [93,94] in a shorter temperature range (30 °C — 107 °C). 5.5.2 Quality factor temperature dependence Most of the extrinsic internal dissipation mechanisms that cause Q degradation in polycrystalline materials are associated with atomic motion or with structural reconfiguration. The motion of vacancies or substitutional impurities, interstitial motions, dislocation, grain boundary sliding cause a resonant Debye-like relaxation, which limits the Q value according to the following equation [25]: . ’1 (s 9) Qdefect = A x [Jr—)2] . l + (tor * where A is known as the “intensity” of the relaxation process, 1* is the relaxation time for the defect motion and a) is the resonator vibrational angular frequency (w=2nf0). It can be noticed from (5.9) that when a defect is limiting the Q, then its value (Qddea) will be minimum at arr * = 1. Since there is atomic motion involved, 1* is thermally activated and has the following functional form [25]: 94 0.002 -0.002 -0. 004 -0.006 Frequency Shift (A f / f0) -0.008 -0.01 I Torrie 5\3 Freq. Shift = B ..-...-...o-.~. ........... TC’ -2.5659 x 105 0 0.0013235 2 ER 0 R 0.99335 0 £300 pm 1st mode 3 5 1:1 E300 pm 281 mode —"'““\ """ """"""""" """"""" <>"'§'400'pm"1fist'mod‘e‘r x 5400 pm 2st mode O 100 200 300 Temperature (C) Figure 5.8 Frequency shift as a function of testing temperature for sample 1 95 Frequency Shift (A f / f0) 0.001 -0.001 -0.002 -0.003 -0.004 -0.005 -0.006 -0.007 Sample 2 i . ! i \ o 300 pm 151 mode ”a """" ; 3:2 """"""""" f """" film'300'firfi'23trfibdéfl _. ................ 8K ________ <.>.-...4QQ-ym-..1.s;t.mod.e.-_ = \ g x 400 pm Zst mode “' """""""""""""""" \ """""""" “ _35 ..... x ................. .............. _ \é O _ .................................................... --\. ............... E ............... 4 = - if] N o ci> __ Freq.Shift=TCf*T+B [3‘ "1:1 _______________ __ 8 0.0010015 g; g _ TC’ -1.5915x10'5 ............. 99 ............... a R 0.98262 5 9 if: i i 0 100 i 200 300 400 Temperature (C) Figure 5.9 Frequency shift as a function of testing temperature for sample 2 96 Frequency Shift (A f / f0) 0.002 -0.002 -0.004 -0.006 -0.008 -0.01 Sample 3 o 300 pm 13}: mode _ -.g ..................................... :1-..300.pm.23§t-mode._ 0 400 pm 1st mode _ .......................................... 8--..49980293m999L ~— ------ Fre.Shift=TCfx T.+B .......... .............. — B 0.0011124 : [5 TC" -1.8677 x 10'5 k S) ” '''' R 0.97822 """""""""""""" 5 """"""" i i 1 L 0 100 200 300 400 500 Temperature (C) Figure 5.10 Frequency shift as a function of testing temperature for sample 3 97 (5.10) . . . . . 13 where l/‘to is the characteristic atomic Vibration frequency (on the order of 10 Hz), EA is the activation energy for the relaxation process, and k3 is the Boltzmann’s constant. If a single defect type dominates, then there will be a maximum in internal dissipation (a minimum in Q) at the frequency in which on * = The Q values of a resonator can be measured as a function of temperature. Then the activation energy for the relaxation process can be identified by finding the temperature at which the Q degradation is maximum and using equation (5.10). Figure 5.11 (a) shows the Q values as a fiJnction of temperature for the three samples. It can be noticed in figure 5.11 (b) how the Q values for the sample 2 reached a minimum value at around 350 °C. This suggests the presence of a thermally activated relaxation mechanism responsible to limit the Q in highly doped polycrystalline diamond resonators or at least the existence of a higher concentration of defects having an activation energy corresponding to the measurement temperature of 350 °C. The data for the 3 samples was then plotted as a function of the activation energy by using (5.10). The activation energy axis was divided in ranges of 0.2 eV and the measured Q values for each polycrystalline diamond film that were inside each subdivision were averaged and represented by one single point. Figure 5.12 shows the results fiom this analysis, where it can be noticed — in addition to the higher Q values for sample 1 - the minimum in Q value for two of the three samples at an activation energy around 1.9 eV. 98 . Sample1 I Sample2 O Sample3 0) v 120 100 -. ......... ........... ...... ............ __________ a Quality Factor (1x103) 0 100 200 300 400 500 600 700 b) Temperatur (C) IlSample 35 30 .......... .......... i 25 _ ........... . ............ .......... - 20 .... ....... i ............ --g. ...... -.. .......... __J Quality Factor (1x1 03) 15 _ ........... ............ .......... _ 10 ... ....... .. ......... -.- ........ ...: ............ 5 1 : 1 1 1 280 300 320 340 360 380 400 420 Temperature (C) Figure 5.11 Measured Q values vs. temperature a) Three samples plotted in the entire testing temperature range; b) Sample 2 in the range of temperature where a minimum in Q was observed. 99 YTrjrfrTIfitfrfFfrlrITIIIIFTIVVT .O- ; Z 3 ' ' 0 Sample 1 D Sample 2 0 Sample 3 Quality Factor (1x104) -¥ N 0° -h (11 O) \l (D i in- 839 1111111111111lllillillllliillll (0.6 0.8 1 1.2 1.4 1.6 1.8 2 2.2 Activation Energy (eV) Figure 5.12 Measured Q values vs. activation energy for the three samples. A minimum in Q was observed for two of the three samples at an activation energy around 1.9 eV. The data suggests there are some defects or a higher concentration of defects that give rise to mechanical relaxation that have an activation energy for defect relaxation around 1.9 eV. This activation energy can be compared to the activation energies for self-interstitial or vacancy diffusion in crystalline diamond (about 1.3 eV and 2.3 eV, respectively). A comparison to diffusional activation energies was made because many defect relaxation processes, e.g. grain boundary sliding, are limited by diffusional processes. 100 Chapter 6 Polycrystalline Diamond Nanoresonators 6.1 Introduction This chapter reports on the fabrication technology and testing of polycrystalline diamond resonators with dimensions in the nanometer range (nanoresonators). These are the first fabricated and measured structures of this type reported. The fabricated and tested nanoresonators include torsional and cantilever structures. 6.2 Fabrication of Nanoresonators Electron beam (e-beam) lithography has been used in the past to fabricate nanoresonators in single crystal silicon [95]. E-beam lithography is also used in this work to pattern polycrystalline diamond nanoresonators. The fabrication process of the polycrystalline diamond nanoresonators was the same used for the fabrication of the cantilevers reported in chapter 5 up to the polycrystalline diamond deposition step. A special high resolution photoresist commonly used in e-beam patterning called “PMMA” (Poly Methyl MethAcrylate) was deposited (200 nm thick) on top of the polycrystalline diamond film. After this photoresist has been patterned, a 100 nm thick Al film was deposited and patterned via lift-off. This aluminum film was used as a hard mask to pattern the polycrystalline diamond film using dry etching to form the torsional resonators and nanocantilevers. The fabricated nanoresonator structures included cantilever structures patterned to have widths as small as 100 nm, and torsional paddle resonators with a support beam 101 width of 0.5 pm. The performance of these structures shows resonant frequencies and Q values in the range of 23 KHz - 805 KHz and 9,500 — 103,600 respectively. All the torsional resonators were fabricated from the the 3 types of polycrystalline diamond samples described in chapter 3, whereas the nanocantilevers were fabricated from the sample 1. 6.3 Torsional Resonators The torsional paddle resonator structures consisted of a proof mass (paddle) with a mass close to 6.16x10’13 kg in a rectangular shape suspended by a 0.5 pm wide, 20 pm long polycrystalline diamond beam. The paddles were supported perpendicular to the largest side of the paddle in either one side (single) or both (double). Figure 6.1 shows an SEM image of a single torsional paddle with its dimensions. The resonant fiequency of this type of resonator structure (when it is operated in its torsional vibration mode) can be calculated using the following equation [96]: 0.0522 xGx 53 f0 = 3 (6-1) pxwaxd where the variables b, w, L and d are the resonator dimensions shown in figure 6.1. The variable p and G are the polycrystalline diamond density (3520 Kg-m-3) and shear modulus (480 GPa) respectively. Equation (6.1) applies to a single torsional resonator. For a double torsional resonator, the equation differs only on the constant inside the radical (0.1 is used instead of 0.0522) [96]. In order to excite a torsional movement, the paddle support had to be ofi‘ the center of the paddle. Otherwise, the beam would vibrate in a flexural mode similar to a 102 cantilever beam. Figure 6.2 shows a diagram of this structure and the excitation mode of vibration. d=20 um 2_£Lnl Figure 6.1 SEM image of polycrystalline diamond single torsional resonator 103 Figure 6.2 Excitation of torsional resonator. The offset of the paddle support from the center of the paddle causes a torsional vibration mode. The patterning of the torsional resonators had to be done using e-bcam lithography since the thin support (0.5 pm wide) of the torsional paddle would have been hard to produce using optical lithography. The tested torsional resonators had a fixed width (w) of 5 um, and their lengths (d) varied from 50 um to 20 pm. 6.3.1 Testing of torsional resonators The measurement results from the torsional resonator structures are shown in Table 6.1. As it can be noticed from the testing results, the measured Q values did not show a dependence on the number of supports of the torsional resonator, or paddle length (d). However, sample 1 shows higher Q values than the other 2 samples. This result suggests a dependence of resonator Q on the type of polycrystalline diamond film. This 104 same observation was made on the results shown in chapter 5 of this thesis. The measured resonant frequencies for the torsional resonators ranged from 269,645 Hz — 805,415 Hz and the Q values ranged from 17,152 — 86,603. The range of obtained frequencies are lower than those obtained on torsional paddles made of nanocrystalline diamond (13 MHz — 640 MHz) [14,34], but the measured Q values are much larger (2,500 — 10,000) [14,34]. Table 6.1 Measurement results for single and double torsional resonators Torsional Paddles Single Length (um) 30 4O 50 Samnle fo(sz Q fofly Q @072) Q 1 490,446 68,498 361,632 78,106 269,645 65,132 2 506,005 17,152 364,146 17,815 295,844 19,387 3 529,678 35,311 353,149 28,388 274,199 31,553 Double Length (um) 30 40 50 Sample fo(HZ) Q fo(HZ) Q fa (HZ) Q 1 805,415 86,603 527,934 76,958 384,070 7051 2 706,756 23,796 462,139 32,209 329,973 25,382 3 752,627 22,953 461,704 46,170 327,282 27,273 6.4 Nanocantilevers For the first time, polycrystalline diamond resonators have been successfully patterned using e-beam lithography to have dimensions as small as 100 nm. The results of the patterning show sharp and smooth sidewalls as can be noticed from figure 6.3. The mask used for the e—beam patterning was designed to include cantilevers with widths and lengths ranging from 100 nm - 1 pm and 500 nm - 200 um respectively. 105 The cantilevers reported in chapter 5 had a thickness of 0.7 pm, which was much smaller than the resonator width (10 um). Consequently the cantilever spring constant for a vibration perpendicular to the substrate is smaller than that for a vibration parallel to the substrate. This is why the cantilever vibration was perpendicular to the substrate for the cantilevers reported in chapter 5. However, for the nanocantilevers reported in this chapter, the width (100 nm — 500 nm) is smaller than the fihn thickness (~ 0.7 am). In this case, the cantilever spring constant for a vibration parallel to the substrate is smaller than that for a vibration perpendicular to the film substrate. As a result, the nanocantilevers reported in this chapter vibrate parallel to the substrate. The vibration detection is done by monitoring the change in the reflected beam for both cases. Figure 6.4 shows a diagram of the vibration mode of cantilevers with a width larger (figure 6.4-a) or smaller (figure 6.4-b) than the film thickness. Figure 6.5 shows SEM images of a 100 nm wide polycrystalline diamond cantilever. 106 Patterned Diamond Grain Figure 6.3 Smooth sidewalls of polycrystalline diamond nanocantilevers fabricated using e-beam lithography. 107 8) Incident Laser ANCHOR Reso Beam b) Reflected Laser ThIckness Width Length Figure 6.4 Use of Piezoelectric actuation and optical detection for a cantilever with vibration perpendicular to the substrate (a) and parallel to the substrate (b). The vibration will be in the direction parallel to the smallest dimension of the cantilever. 108 Figure 6.5 100 nm wide polycrystalline diamond cantilever. The top view shows that the width of the cantilever is in not uniform along the beam length 6.4.1 Testing of nanocantilevers The most noticeable result was the measurement of the highest Q value (103,000) found for a nanoresonator made from any polycrystalline material. (Higher Q values (~250,000) have been found in nanoresonators made of single crystal silicon [22]). This value was found for a 500 mm wide, 50 um long cantilever. Figure 6.6 shows an SEM image of this structure and the resonant peak obtained. The tested cantilever with the smallest width was a 100 nm wide, 25 mm long cantilever beam. This structure had a resonant frequency of 475 KHz and a Q of 14,600. SEM images of this cantilever and the testing results are shown in figure 6.7. 109 Data ,3 _— LorentzianFit '5 1.01 £08 £0.61 fo=536,050 Hz 7;, Q=103,600 :1 0.4 E05 ° 1 50.0 536,020 536,040 536,060 536,080 Frequency (Hz) Figure 6.6 SEM image and performance of the nanocantilever with the highest measured 0. 110 [0:183 kHz .. Q=3,()33 g 10' H EzigntzianFit E .8 EDS] f0=183,214 Hz 8 0- . Q=3,033 g 0.4- m . E 0.2 o . 50.0 183,000 183,200 183,400 Frequency (Hz) Figure 6.7 SEM image and performance of a 100 nm wide polycrystalline diamond nanocantilever 111 Table 6.2 shows some of the measurements of nanocantilevers. The resonant frequencies for the nanocantilevers were calculated using the equation: fo=0.1615 ii (6.2) p L? where E and p are the film Young’s modulus and density respectively, h is the resonator thickness, L, is the beam length and j}, is the resonant fiequency. The values for the nanocantilever width and film thickness were used as the resonator thickness in equation (6.2) to calculate the resonant frequency of the nanocantilevers. The results are shown in table 6.2. The measured resonant frequencies are closer to the calculated values using the resonator width as the resonator thickness, which indicates that the cantilevers are moving parallel to the substrate as it was expected since the beam width was smaller than the film thickness. The difference between the measured and calculated frequency values could be due to the non-uniformity of the nanocantilevers width along the beam direction as shown in figure 6.5. Figure 6.8 shows a plot of the measured Q values on 200 nm thick cantilever beams made of single crystal silicon [26], silicon nitride [26], and polycrystalline diamond (reported in this work). The testing method for all these samples consisted of piezoelectric actuation with optical detection. A direct relationship between the measured Q values and the beam length can not be observed for any of the samples. This suggests that the dominant energy dissipation mechanism is not related to clamping losses, since the clamping losses are directly proportional to the beam length (see equation 2.32). The measured Q values for polycrystalline diamond seem to vary much more with beam length than the measured Q values for silicon nitride and single crystal silicon. One of the 112 possible reasons for the larger variations in Q for polycrystalline diamond could be related to the material itself. The single crystal silicon and silicon nitride nanocantilevers consist of only one uniform fihn. On the other hand, as it was pointed out in chapter 5, polycrystalline diamond consists of a seeding layer composed mainly by small diamond grains ( ~ 30 nm in diameter) and a film layer composed by much larger grains (~ 0.5 pm). The measured Q values on a structure made of two layers will depend on the thickness of each layer [14,89]. The seeding and film layer could vary in their thickness among different polycrystalline diamond nanocantilevers structures, leading to differences in Q. Another possible reason for this large variation could be the low yield that was obtained in the fabrication process of the polycrystalline diamond nanocantilevers. Due to this low yield, the number of measurements on a nanocantilever with specific dimensions was not larger than two. Although the yield was never mentioned in [26], it is possible that several measurements were taken on each nanocantilever, and the reported results was the one with the highest Q. This technique is completely valid since there are many processes that can decrease the Q, and the measurement with the highest Q would be the closest to the real Q of the device. It is interesting to notice how the average of the measured Q values for the polycrystalline diamond nanoresonators plotted in figure 6.8 (~ 21,000) is larger than that of single crystal silicon (~ 17,000) and silicon nitride (~ 11,500). 113 Table 6.2 Results of polycrystalline diamond nanocantilevers fabricated from sample 1. The resonant frequencies were calculated using the width as the resonator thickness (a), and the film thickness as the resonator thickness ( b). M* represents the number of measurements taken from each structure (the values shown in the table correspond to the measurement with the highest measured Q). NAN OCANTILEVERS Meas. Calc. ” Meas. Calc. " f0 (Hz) Q f0 (Hz) f0 (Hz) Q f0 (Hz) 210,023 1,003 219,640 77 97 7 114 Q Factor (1x 103) 0 50 100 150 200 250 300 Beam Length (nm) Figure 6.8 Plot of measured Q values on cantilever beams made of silicon nitride [26], single crystal silicon [26] and polycrystalline diamond (reported in this work). All the cantilevers had a thickness between 170 nm and 200 nm 115 Chapter 7 Conclusions and Future Research 7.1 Summary and Conclusions 0 Testing of polycrystalline diamond RF MEMS resonators using different methods Polycrystalline diamond resonators have been tested using different actuation methods, and the results have been compared. The results showed similar resonant fi'equencies, and a significant difference in Q (by a factor close to 10) was attributed to the resonator testing pressure. 0 Measured the highest Q values A Q of 116,000 has been measured on a cantilever beam made of an undoped polycrystalline diamond film grown at 780°C. Highly doped (5x10'9 cm'3) polycrystalline diamond films grown at 780 °C and undoped polycrystalline diamond films grown at 600 °C had about half the Q values obtained from undoped polycrystalline diamond flms grown at 780 °C. 0 Study of frequency thermal stability of polycrystalline diamond resonators. 116 The performance of polycrystalline diamond resonators at elevated temperatures showed thermal coefficient values ranging from -l.59x10'5 / °C to -2.56 x10'5 / °C. for different polycrystalline diamond films. 0 Study of energy dissipation mechanisms in polycrystalline diamond resonators. The measured Q values on different polycrystalline diamond films were not limited by clamping losses, thermoelastic dissipation or phonon-phonon interaction. In order to identify the mechanism responsible for limiting the highest possible achievable Q, the Q values were measured as a function of temperature and an apparent thermally activated relaxation mechanism was identified with an activation energy around 1.9 eV. 0 Design, fabrication and testing of polycrystalline diamond nanoresonators Polycrystalline diamond films have been patterned to fabricate cantilever structures with widths as small as 100 nm. The measured data does not show a relationship between the Q values and the cantilever beam length. 7.2 Future Research Some areas of future research are: 117 1. Polished polycrystalline diamond films: This film can be used to produce polycrystalline diamond resonator structures with smooth surfaces. This would be very significant for the fabrication of polycrystalline diamond nanoresonators. Improvement of the yield in the fabrication of polycrystalline diamond nanoresonators: Modifications to the fabrication process of polycrystalline diamond resonators could lead to an increase in the yield of such devices. By increasing the yield, the characterization and the performance of polycrystalline diamond nanoresonators can be studied in more detail. 118 BIBLIOGRAPHY 119 BIBLIOGRAPHY: [1] C. T.-C. Nguyen, “Microelectromechanical devices for wireless communications (invited),” Proceedings, 1998 IEEE International Micro Electro Mechanical Systems Workshop, Heidelberg, Germany, Jan. 25-29, 1998, pp. 1-7. [2] C. T.C. Nguyen, “Transceiver front-end architectures using vibrating micromechanical signal processors (invited),” Dig. 0f Papers, Topical Meeting on Silicon Monolithic Integrated Circuits in RF Systems, Sept. 12-14, 2001, pp.23-32 [3] R. T. Howe and R. S. Muller, “Resonant microbridge vapor sensor,” IEEE Trans. Electron Devices, vol. ED-33, pp. 499-506, 1986. [4]N. M. Nguyen and R. G. Meyer, “Si IC-compatible inductors and LC passive filters,” IEEE J. of Solid-State Circuits, vol.SC-25, no. 4, pp. 1028-1031, Aug. 1990. [5]N. M. Nguyen and R. G. Meyer, “A 1.8-GHz monolithic LC voltage-controlled oscillator,” IEEE J. of Solid-State Circuits, vol. SC-27, no. 3, pp. 444-450. [6]S. V. Krishnaswarny, J. Rosenbaum, S. Horwitz, C. Yale, and R. A. Moore, “Compact FBAR filters offer low-loss performance,” Microwaves & RF, pp. 127-136, Sept. 1991. [7] Jing Wang, James E. Butler, Tatyana F eygelsonand Clark T.-C. Nguyen “1.51-GHz Nanocrystalline diamond micromechanical disk resonator with material-mismatched isolating support” To be published in IEEE Transducers '04. [8] Clark T.-C. Nguyen, “Vibrating RF MEMS for low power wireless communications (invited)”, Proceedings, 2000 Int. MEMS Workshop (iMEMS’Ol), Singapore, July 4-6, pp.2 1 -34 [9] L. Meirovitch, Analytical Methods in Vibrations. New York: Macmillirnan Publishing Co, Inc. 1967 [10] W. Weaver, S. P. Timoshenko, and D. H. Young, Vibration Problems in Engineering, 5th ed: Wiley Publishers, 1990 [11] Kun Wang, Ark-Chew Wong, Clark T.-C. Nguyen “VHF Free-Free Beam High-Q Micromechanical Resonators” IEEE Journal of Microelectromechanical Systems, Vol 9. No.3 (2000) pp. 347-360 [12] P. J. Petersan and S. M. Anlage, “Measurement of resonant frequency and quality factor of microwave resonators: Comparison of methods” Journal of Applied Physics, vol. 84, pp. 3392-3402, 1998 [13] D. Czaplewski, J.P. Sullivan, T.A. Friedmann, D.W. Carr, B.E.N. Keeler and J .R. Wendt; J. Appl. Phys. 97, 023517 (2005). 120 [14]A.B. Hutchinson, P.A. Truitt and KC. Schwab, L. Sekaric, J .M. Parpia, H.G. Craighead and J .E. Butler; Appl. Phys. Lett 84 (6) 2004, 972 [15] J. R. Clark, W.-T. Hsu, and C. T.-C. Nguyen, “Measurement techniques for capacitively—transduced VHF -to-UHF micromechanical resonators” Digest of Technical Papers, the 11th Int. Conf. on Solid-State Sensors & Actuators (Transducers’Ol), Munich, Germany, June 10-14, 2001, pp. 1118-1121. [16] P. Mohanty, D. A. Harrington, K. L. Ekinci, Y. T. Yang, M. J. Murphy, and M. L. Roukes, "Intrinsic dissipation in high-frequency micromechanical resonators," Physical Review B, vol. 66, pp. 085416 1-15, 2002. [17] RM. Osterberg and SD. Senturia, “M-Test: A test chip for MEMS material property measurement using electrostatically actuated test structures” J. Microelectromechanical Systems, vol. 6 pp. 107-118, 1997. [18] R. H. Blick, A. Erbe, H. Krémmer, A. Kraus, J. P. Kotthaus, Phys. E 6, 821 (2000) [19] D. W. Carr, Nanoelectomechanical Resonators, Ph.D. thesis Cornell University 2000 [20] Stephen D. Senturia Microsystem Design; Kluwer Academic Publishers 2001 [21] E. Buks, M.L. Roukes, Europhys. Lett. 54, 2220 (2001) and E. Buks, M.L. Roukes, Phys. Rev. B. 63, 033402 (2001) [22] J. Yang, T. Ono and M. Esashi, IEEE J. Microelectromech. Sys. 11(6) Dec. 2002, 775 [23] FR. Blom, S. Bowstra, M. Elwenspoek, J .H.J Fluitrnan “Dependence of the quality factor of micromachined silicon beam resonators on pressure and geometry” J. Vac. Sci. Technol. B 10(1), Jan/Feb 1992. [24] T. V. Roszhart, “The effect of thermoelastic internal friction on the Q of micromachined silicon resonators,” in Tech. Dig. Solid-State Sens. Actuator Workshop, Hilton Head, SC, 1990, pp. 13—16. [25] V. B. Braginsky, V. P. Mitrofanov, V. I. Panov, K. S. Thorpe, Systems with Small Dissipation, The University of Chicago Press, 1985 [26] K.Y. Yasumura, T.D. Stowe, E.M. Chow, T. Pfafman, T.W. Kenny, B.C. Stipe and D. Rugar, J. Microelectromech. Sys. 9(1) Mar. 2000, 117 [27]W.-T. Hsu, J. R. Clark, and C. T.-C. Nguyen, “A sub-micron capacitive gap process for multiple-metal-electrode lateral micromechanical resonators,” Technical Digest, 14th Int. IEEE Micro Electro Mechanical Systems Conference, Interlaken, Switzerland, Jan. 121 21—25, 2001, pp. 349-352. [28] RM. Osterberg and SD. Senturia, “M-Test: A test chip for MEMS material property measurement using electrostatically actuated test structures” J. Microelectromechanical Systems, vol. 6 pp. 107-118, 1997. [29] G. Pearson,W. Read, Jr., andW. F eldman, “Deformation and fracture of small Si crystals,” Acta Metall, vol. 5, p. 181, 1957. [30] Shuvo Roy, Russell G. DeAnna, Christian A. Zorman and Mehran Mehregany “Fabrication and Characterization of Polycrystalline SiC Resonators” IEEE Transactions on Electron Devices, vol. 49, No.12 December 2002. [31] Jing Wang, James E. Butler, D.S.Y. Hsu and Clark T.-C. Nguyen “CVD Polycrystalline diamond high-Q micromechanical resonators” IEEE International Conference Micro Electro Mechanical Systems, 2002. pp. 657-660 [32] J. Wang, J. E. Butler, D.S.Y. Hsu, and C. T. —C. Nguyen, “High-Q Micromechanical resonators in CH4-reactant optimized high acoustic velocity CVD polydiamond” Tech. Digest, 2002 Solid-State Sensor, Actuator and Microsystems Workshop, Hilton Head Island, South Carolina, June 2-6, 2002, pp. 61-62. [33] N. Sepulveda, D.M. Aslam, J .P. Sullivan, “Polycrystalline diamond RF MEMS resonators with the highest quality factors” to be presented at IEEE MEMS’O6. [34] L. Sekarik, J .M. Parpia, H.G. Craighead, T. Feygelson, B.H. Houston, J .E. Butler, J. Appl. Phys. 81, 23 pp.4455 (2002). [35] Mustafa U. Demirci, Mohamed A. Abdehnoneum and Clark T.-C. Nguyen “Mecanically comer-coupled square microresonator array for reduced series motional resistance”, IEEE Transducers, Solid-State Sensors, Actuators and Microsystems, 12th International Conference, vol.2 June9-12, 2003 [36] F. Bannon 111, Clark, John R., Clark T.-C. Nguyen “High Q HF Micromechanical Filters”., IEEE J SSC, pp.512-526, April 2000 [37] Navid, K; Clark J .R.; Demirci, M; Nguyen, C.T.-C. “Third-order intermodulation distortion in capacitively-driven CC-beam micromechanical resonators” IEEE International Conference on Micro Electro Mechanical Systems, 2001. pp.228-231 [38] Cohn M Flannery, Michael D Whitfield and Richard B Jackrnan “Acoustic wave properties of CVD diamond” Institute of Physics Publishing. Semicond. Sci. Technol. 18 (2003) $86-$95 [39] K. Iakoubovskii, Optical Study of Defects in Diamond, PhD thesis, Katholieke Universiteit Leuven, 2000. 122 [40]Y. Tang and D. M. Aslam “Technology of polycrystalline diamond thin films for microsystems applications” J. Vac. Sci. Technol. B 23(3), May/Jun 2005 [41] S. Matsumoto, Y. Sato, M. Kamo, and N. Setaka, Jpn. J. Appl. Phys, Part 1 21, 183 (1982) [42] R. C. Hyer, M. Green, K. K. Mishra, and S. C. Sharma, J. Mater. Sci. Lett. 10, 515 (1991) [43] A. Lettington and J. W. Steeds, Thin Film Diamond _Chapman and Hall, London, (1994) [44] J. Stiegler, Y. vonKaenel, M. Cans, and E. Blank, J. Mater. Res. 11, 716 (1996). [45] LA. Chernozatonskii, Z. Ya. Kosakovskaya, Yu. V. Gulyaev, N.I. Sinitsyn, G.V. Torgashov, and Yu. F. F. Zakharchenko, “Influence of external factors on electron field emission from thin-filrn nanofilament carbon structures” J. Vac Sci Technol. B 14(3), p. 2080, 1996. [46] T. Xie, W.A. Mackie and RR. Davis, “Field emission from ZrC films on SSi and Mo single emitters and emitter arrays”, J. Vac Sci. Technol. B 14(3), p.2090, 1996. [47] J .S. Ma, H. Kawarda, T. Yonehara, J .I. Suzuki, J. Wei, Y. Yokota and A Hiraki “Selective Nucelation and Growth of Diamond Particles by Plasma-assisted Chemical Vapor Deposition” App. Phys. Lett., vol. 55 no.11, pp.1071-1073, 1989. [48] MP. Everson and MA. Tamor :Studies of Nucleation and Growth Morphology of Boron-doped Diamond Microcrystals by Scanning Tunnelling Microscopy” J. Vac. Sci. Technol. B, vol.9 no.3 pp. 1570-1575, 1991. [49]S. Ijima, Y. Aikawa, and K. Baba, “Early Formation of Cernical Vapor Deposition Diamond Films” Appl. Phys. Lett., vol.57, no.25, pp. 2646-2648, 1990. [50] K. Hirabayashi, Y. Taniguchi, O. Takamatsu, T. Ikeda, K. lkoma and N. Iwasaki- Kurihara. “Selective Deposition of Diamond Crystals by Chemcial Vapor Deposition Using a Tungsten-filament Method” Appl. Phys. Lett., vol. 53 no.19 pp. 1815-1871, 1 988. [51] B.W. Sheldon, R. Csencstis, J. Rankin, and RE. Boekenhauer, “Bias-enhanced Nucleation of Diamond During Microwave-assisted Chemical Vapor Deposition” J. Appl. Phys, vol. 75, no.10, pp.5001-5008, 1994. [52]B.R. Stoner, G.H. Ma, S.D. Wolter, and J .T. Glass, “Characterization of Bias- enhanced Nucleation of Diamond on Silicon by in vacuo Surface Analysis and Transmission Electron Microscopy” Phys Rev. B. vol.45, pp.11067-11084, 1992. 123 [53]X. Jiang, C.P. Klages, R. Zachai, M. Hartweg and HI. Fusser, “Epitaial Diamond Thin Films on (001) Silicon Substrates” Appl. Phys Lett., vol. 62, pp.3438-3440, (1993). [54] A. Massod, M. Aslam, M.A. Tamor and T.J. Potter, “Techniques for Patterning CVD Diamodn Films on Non-diamond substrates,” J. Electrochem. Soc., vol 138. no.11pp. L67-L68, 1991. [55] RV. Spitsyn, G. Popovic and M.A. Prelas, “Problems of Diamond Film Doping” 2nd Int Conf. on the Application of Diamond F ihns and Related Materials, Ed. M. Yoshikawa, M.Murakawa, Y.Tzeng and W.A. Yarbrough, MY, Tokyo., pp. 57-64, 1993. [56] Gennady SH. Gildneblat, Stephen A. Grot, and Andrzej Badzian, “The electrical properties and Device Applications of Homoepitaxial and Polycrystalline Diamond Films” in Proc. Of the IEEE, vol. 79, pp647-667, 1991. [57] T. Takada, T. Fukunaga, K. Hayashi, and Y. Yokota, Sens. Actuators, A 82, 97 (2000) [58] J. Cifre, J. Puigdollers, M. C. Polo, and J. Esteve, Diamond Relat. Mater. 3, 628 (1994). [59] J .F. Prins, Thin Solid Films 212, 11 (1992) [60] J .F. Prins, Diamond Related Materials. 11, 612 (2002) [61] F. Zhang, B. Xie, B. Yang, Y. Cai and G. Chen, “Synthesis and Infrared Absorption Characteristics of Boron-Doped Serniconducting Diamond Thin Films”. Materials Letters, vol. 19, pp.115-118, 1994 [62] X. J iang, M. Paul, P. Willich, E. Bottger and C.-P. Klages, “Controlled Boron Doping in Chemical Vapor Deposited Diamond Films”, Diamond 96, Tours, France, Sept. 8-13, poster 11.051, 1996 [63] E. Colineau, et.al. “Minimization of Defects Concentration fiom Boron Incorporation in Polycrystalline Diamond Fihns”, Diamond 96, Tours, France, Sept. 8- 13, poster 11.0512, 1996. [64] X.Z. Liao, et al. “The influence of Boron Doping on the Structure and Characteristics of Diamond Thin Films,” Diamond 96, Tours, France, Sept. 8-13, poster 8.060, 1996 [65] K. Miyata, K. Kumagai, K. Nishimura and K.Kobashi, “Morphology of heavily B- Doped Diamond Films,” J. Mat. Res., vol.8, pp. 2845-2857, 1997. 124 [66] T. Roppel, R. Ramesham and S.Y. Lee, “Thin Film Diamond Microstructures” Thin Solid Films, vol. 212, pp.56-62, 1992. [67] J .L. Davidson, C. Ellis and R. Rarnesham, “Selective Deposition of Diamond Films,” New Diamond, vol.6, pp.29-32, 1990. [68] O. Dorsch, M. Werner, and E. Oberrneier, Diamond Relat. Mater. 4, 456 (1995). [69] M. Bernard, A. Deneuville, T. Lagarde, and E. Treboux, Diamond Relat. Mater. 11, 828 (2002). [70] C. Vivensang, L. F erlazzoManin, M. F. Ravet, and G. Turban, Diamond Relat. Mater. 5, 840 (1996). [71] R. Otterbach and U. Hilleringrnann, Diamond Relat. Mater. 11, 841(2002). [72] H. Shiomi, Jpn. J. Appl. Phys., Part 1 36, 7745 (1997). [73] F. Silva, R. S. Sussrnann, F. Benedic, and A. Gicquel, Diamond Relat. Mater. 12, 369 (2003). [74] W. Kern and D. A. Puotinen, RCA Rev. 31, 187 (1970). [75] R. R. Thomas, F. B. Kaufman, J. T. Kirleis, and R. A. Belsky, J. Electrochem. Soc. 143, 643 (1996). [76] N. Sepulveda-Alancastro and D. M. Aslam, “Polycrystalline diamond technology for RF MEMS Resonators” Microelectron. Eng. 73-74, 435 (2004). [77] H. Windischmann, Glenn F. Epps, Yue Cong and R. W. Collins, “Intrinsic stress in diamond fihns prepared my microwave plasma CVD” J. Appl. Phys. 69 (4) Feb. 1991, pp. 2231-2237. [78]P T Docker, P K Kinnell and M C L Ward, “Development of the one-step DRIE dry process for unconstrained fabrication of released MEMS devices” J. Micromech. Microeng. 14 pp. 941—944 (2004) [79] L M Phinney , G Lin, J Welhnan and A Garcia, “Surface roughness measurements of micromachined polycrystalline silicon films” J. of Micromechanics and Microengineering 14, 927-93, 2004 [80] John R. Clark, Wan-Thai Hsu, and Clark T.-C. Nguyen, “High-Q VHF Micromechanical Contour-Mode Disk Resonator” IEEE Int. Electron Devices Meeting , Dec 11-13, 2000 pp. 493-496. 125 [81] N. Sepulveda, D.M. Aslam, J .P. Sullivan, “Polycrystalline Diamond MEMS Resonator Technology for Sensor Applications” Diamond and Related Materials, Currently in press. [82]Kirt R. Williams, Kishan Gupta, Mathew Wasilik, “Etch rates for micromachining process-Part 11” IEEE J. Micro Electrom. Sys. Vol.12 no.6 Dec 2003. [83] Ville Kaajakari, Tomi Mattila, Antti Lipsanen, and Aame Oja, “Nonlinear Mechanical Effects in Silicon Longitudinal Mode Beam Resonators”, Sensors and Actuators A: Physical vol. 120(1), pp.64-70, 2005 [84] E. S. Zouboulis, M. Grimsditch, A. K. Ramdas and S. Rodriguez, “Temperature dependence of the elastic moduli of diamond: A Brillouin-scattering study” Physical Review B, vol. 57, pp. 2889-2896, 1998. [85] F. Szuecs, M. Werner, R. S. Sussrnann, C. S. J. Pickles and H. J. Fecht, “ Temperature dependence of Young’s modulus and degradation of chemical vapor deposited diamond” Jouranl of Applied Physics vo. 86, pp.6010-6017, 1999 [86] J. O. Orwa, K. W. Nugent, D. N. Jarnieson, and S. Prawer, “Raman investigation of damage caused by deep ion implantation in diamond”, Physical Review B, vol. 62, pp.5461-5472, 2000. [87] Liwei Lin, "MEMS Post-Packaging by Localized Heating and Bonding", IEEE TRANSACTIONS ON ADVANCED PACKAGING, VOL. 23, NO. 4, PP. 608-616, NOVEMBER 2000 [88] X. Liu, D.M. Photiadis, J .A. Bucaro, J .F. Vignola, B.H. Houston, H.-D. Wu, D.B. Chrisey, “Low temperature internal friction of diamond-like carbon films” Material Science and Engineering A, 5 July 2002. [89] BE. White, Jr., R.O. Pohl, Materials Research Symposium Proc. 356 (1995) 567. [90] Victor J. Milenkovic, “Rotational polygon contaimnent and minimum enclosure using only robust 2D constructions” Computational Geometry, Theory and Aplications (13)pp.3-19 1999 [91] H. Freeman and R. Shapira, “Determining the Minimum-Area Encasing Rectangle and Arbitrary Closed Curve”, Communications of the ACM, vol. 18 pp.409-413, 1975 [92] H. J. McSkimin and P. Andreatch Jr, “Elastic Moduli of Diamond as a Function of Pressure and Temperature” Journal of Applied Physics, vol. 43, pp. 2944-2948, July 1972. 126 [93] W. T. Hsu; C.T.-C. Nguyen, “Stiffiress-compensated temperature-insensitive micromechanical resonators” IEEE Int. Conference on MEMS, pp.73l — 734, Jan 20-24 2002 [94] . T. Hsu; J .R Clark,.; C.T.-C Nguyen,.; “Mechanically temperature-compensated flexural-mode micromechanical resonators” Electron Devices Meeting, International pp. 399 - 402, Dec 10-13 2000 [95] J. Yang, T. Ono, and M. Esashi, “Mechanical behavior of ultrathin micro- cantilever,” Sens. Actuators, vol. 82, pp. 102—107, 2000 [96] DA. Czaplewski, J.P. Sullivan, T.A. Friedmann and JR. Wendt, “Mechanical dissipation at elevated temperatures in tetrahedral amorphous carbon oscillators” Diamond and Related Materials, Currently in press. 127