CMOS-MEMS Variable Capacitors for Reconfigurable RF Circuits Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Electrical and Computer Engineering John R. Reinke B.E.E., Electrical Engineering, University of Minnesota M.S., Electrical Engineering, Carnegie Mellon University Carnegie Mellon University Pittsburgh, PA May, 2011 © John Reinke 2011 All rights reserved. ii Acknowledgements First, I am grateful to Tamal Mukherjee and Gary Fedder for serving as my advisors during my graduate studies at Carnegie Mellon University. Their guidance has been invaluable, and their commitment to producing quality students has made the CMU MEMS lab a nurturing environment. Their sense of humor and understanding helped immensely during the inevitable frustrations of research. My colleagues from the MEMS lab have been both helpful and entertaining. Leon Wang has provided many important insights as he served as resident circuits guru. Peter Gilgunn has inspired me to work harder and always strive for excellence. Nathan Lazarus and Yu-Jen Fang were always willing to participate in random musings about science and technology. Abhishek Jajoo, Sarah Bedair, and Fernando Alfaro helped me get my research off the ground. Amy Wung and Kristen Dorsey deserve much credit for actually trying to foster a social environment within the lab. Many thanks to Suresh Santhanam and the staff of the Nanofabrication Facility for the many CMOS-MEMS chips they have fabricated. They have battled through phenomenon such as the mystery gunk and always managed to produce quality devices. Last but certainly not least, I am thankful for the support of my family back home in Minnesota. My family has been a needed refuge during holidays and vacations. I am truly lucky to have them in my life. This work was funded by the Industrial Technology Research Institute (ITRI) and by government support under and awarded by DoD, Air Force Office of Scientific Research, National Defense Science and Engineering Graduate (NDSEG) Fellowship, 32 CFR 168a. iii Abstract Microelectromechanical systems (MEMS) variable capacitors have demonstrated excellent performance in terms of tuning ratio and quality factor (Q) at radio frequencies (RF). These characteristics can enhance existing RF circuits and enable frequency-reconfigurable circuits. However, MEMS are typically not monolithically integrated with complementary metal-oxidesemiconductor (CMOS) processes. To achieve integration with CMOS, this work uses maskless post-CMOS processing to create two types of CMOS-MEMS variable capacitors, the gap-tuning capacitor and the digital capacitor. To show that these capacitors enable frequency-reconfigurable circuits, the digital capacitor was used to develop the first CMOS-MEMS phase shifter. The CMOS-MEMS gap-tuning capacitor uses lateral electrothermal actuation to control the gap between electrodes. A mechanical latch maintains the capacitance without consuming static power. This capacitor achieved series and shunt tuning ratios up to 6.9:1 and 5.7:1, respectively, with a Q of 80 at 1 GHz. Millions of actuation cycles were demonstrated without failure However, the gap-tuning capacitor had several disadvantages, specifically large device variation, a non-linear distribution of capacitance values, and low capacitance density, which inspired development of the digital capacitor. The CMOS-MEMS digital capacitor consists of many switched capacitors that are organized into capacitive bits to produce a linear distribution of capacitance values. This capacitor uses both vertical electrothermal actuation and lateral electrostatic actuation to achieve large tuning ratios without sacrificing capacitance density. The maximum capacitance state is maintained by electrostatic force to prevent static power consumption. The digital capacitor achieved series and shunt tuning ratios up to 73:1 and 19:1, respectively, with a Q of 369 at 1 GHz. Compared with that of the gap-tuning capacitor, the capacitance density was increased by 5x and the device variation was reduced by 4x. Reliability tests revealed evidence of mechanical fatigue after iv millions of actuation cycles, but no catastrophic failures occurred. The CMOS-MEMS phase shifter consists of a micromachined transmission line periodically loaded with digital capacitors. The digital capacitors were modified to achieve a shunt tuning ratio of 7.2:1 with a Q of 18.8 at 30 GHz. The phase shifter produced an average insertion loss and maximum phase shift of 2.9 dB and 337.5°, respectively, at 32 GHz. v Table of Contents Chapter 1. Introduction ............................................................................................................... 1 1.1. Motivation ........................................................................................................................ 1 1.2. MEMS Variable Capacitors ............................................................................................. 2 1.3. Thesis Statement .............................................................................................................. 9 1.4. Contributions .................................................................................................................. 10 Chapter 2. CMOS-MEMS Background.................................................................................... 12 2.1. CMOS-MEMS Process .................................................................................................. 12 2.2. Electrothermal Actuation ............................................................................................... 17 Chapter 3. Gap-tuning Capacitor .............................................................................................. 23 3.1. Motivation ...................................................................................................................... 23 3.2. Topology ........................................................................................................................ 28 3.3. Analysis .......................................................................................................................... 33 3.4. Design............................................................................................................................. 38 3.5. Characterization ............................................................................................................. 41 3.6. Discussion ...................................................................................................................... 57 Chapter 4. Digital Capacitor ..................................................................................................... 61 4.1. Motivation ...................................................................................................................... 61 4.2. Topology ........................................................................................................................ 64 4.3. Analysis .......................................................................................................................... 67 4.4. Design............................................................................................................................. 74 4.5. Characterization ............................................................................................................. 78 4.6. Discussion ...................................................................................................................... 99 Chapter 5. Phase Shifter ......................................................................................................... 102 5.1. Motivation .................................................................................................................... 102 5.2. Topology ...................................................................................................................... 104 5.3. Analysis ........................................................................................................................ 108 5.4. Design........................................................................................................................... 111 5.5. Characterization ........................................................................................................... 114 5.6. Discussion .................................................................................................................... 123 Chapter 6. Conclusions ........................................................................................................... 126 6.1. Summary ...................................................................................................................... 126 6.2. Contributions ................................................................................................................ 128 6.3. Future Work ................................................................................................................. 130 6.4. Final Thoughts.............................................................................................................. 135 References ................................................................................................................................... 137 Appendix A. Electrical Characterization ................................................................................ 142 A.1. Network Analyzer Settings .......................................................................................... 142 A.2. Open-short De-embedding ........................................................................................... 143 A.3. Extracting an Equivalent Circuit Model....................................................................... 145 A.4. Matlab Code ................................................................................................................. 157 Appendix B. List of Cadence Layouts .................................................................................... 166 vi List of Tables Table 1.1: Variable Capacitor Characteristics ................................................................................ 6 Table 3.1: Gap-Tuning Capacitor Design Parameters .................................................................. 40 Table 3.2: Electrical Characteristics of the Gap-tuning Capacitor ............................................... 45 Table 3.3: Gap-Tuning Capacitor Results..................................................................................... 58 Table 4.1: Digital Capacitor Design Parameters ........................................................................... 77 Table 4.2: Electrical Characteristics of the Initial Digital Capacitor ............................................ 83 Table 4.3: Electrical Characteristics of the Updated Digital Capacitor ........................................ 98 Table 4.4: Digital Capacitor Results ........................................................................................... 100 Table 5.1: Phase Shifter Design Parameters ............................................................................... 112 Table 5.2: Electrical Characteristics of the Phase Shifter Digital Capacitor .............................. 118 Table 5.3: Electrical Characteristics of the Phase Shifter ........................................................... 122 Table 5.4: Comparison of Phase Shifters .................................................................................... 124 Table 6.1: Comparison of Integrated Variable Capacitors.......................................................... 129 Table B.2: List of Cadence Libraries .......................................................................................... 166 Table B.1: List of Cadence Layouts ............................................................................................ 166 vii List of Figures Fig. 1.1. Diagram of capacitance tuning by changing the (a) gap (b) overlap area (c) dielectric. All capacitors have length L, width W, and gap g before part of the structure undergoes displacement d. Capacitance calculations ignore the effects of fringing fields. ............................. 2 Fig. 2.1. Cross-section of the CMOS-MEMS process flow (a) beginning with CMOS chip (b) after the vertical oxide etch (c) after the vertical silicon etch (d) after silicon release etch. ......... 13 Fig. 2.2. SEM image of (a) 20 µm squared released from the substrate (b) 60 µm square anchored to the substrate. These test structures were fabricated using the 0.18 µm BiCMOS process from TowerJazz. .................................................................................................................................... 14 Fig. 2.3. Diagram of a vertical electrothermal actuator beam. The polysilicon resistor for heating can either be lumped near the anchor point or distributed along the length of the actuator beam.20 Fig. 2.4. Diagram of a lateral electrothermal actuator beam. The offset metal layers are switched to the opposite side of the beam halfway along the length to produce guided-end motion at the tip. ................................................................................................................................................. 21 Fig. 2.5. Microscope image of lateral electrothermal actuator (a) deflected due to residual stress (b) after heating. (c) Focused-ion beam image of electrothermal actuator beam cross-section. ... 22 Fig. 3.1. Concept of the actuation method for the gap-tuning capacitor. ...................................... 24 Fig. 3.2. SEM image of the (a) 1st (b) 2nd (c) 3rd (d) 4th design of the gap-tuning capacitor. ........ 25 Fig. 3.3. Vertical misalignment of capacitive beams viewed as (a) SEM image (b) diagram. The observed vertical misalignment is not as exaggerated as in the diagram but still significant enough to affect the tuning ratio. .................................................................................................. 26 Fig. 3.4. SEM image of the (a) 1st (b) 2nd (c) 3rd design of the mechanical latch for gap-tuning capacitors. ..................................................................................................................................... 27 Fig. 3.5. Diagram of gap-tuning capacitor. Arrows denote the directions of self-assembly and electrothermal actuation displacements for both the gap and latch actuators. Resistor symbols indicate the location of embedded polysilicon resistors used for electrothermal actuation. The specific numbers of capacitive, actuator, and interconnect beams are not representative of the numbers of beams used in the actual device. ................................................................................ 28 Fig. 3.6. Vertical curling of CMOS-MEMS beams. Anchoring MEMS beams at a common location prevents vertical misalignment. ...................................................................................... 29 Fig. 3.7. SEM image of the mechanical latch for the gap-tuning capacitor. As pictures, the latch is engaged in an intermediate capacitance position, neither minimum nor maximum capacitance. ....................................................................................................................................................... 31 Fig. 3.8. SEM image of the gap actuator for the gap-tuning capacitor. The shading of the actuator beams changes half-way along their length because the offset metal layers are switched to the other side of the beams. ................................................................................................................ 31 Fig. 3.9. Thermal isolation unit viewed as (a) SEM image (b) cross-section diagram. The metal 3 and metal 4 pictured in the diagram are not floating; these metals are connected to ground along a different cross-section. .................................................................................................................. 32 Fig. 3.10. Cross-section of CMOS-MEMS beams used in the electrothermal gap-tuning capacitor. ....................................................................................................................................... 33 Fig. 3.11. Diagram of capacitive beams for displacement x. ........................................................ 34 Fig. 3.12. Electric field magnitude for cross-sections of capacitive beams in both minimum and maximum capacitance positions. The white, unmeshed areas represent the discrete metal layers of the CMOS-MEMS beams. ........................................................................................................ 35 viii Fig. 3.13. Capacitance per unit length as a function of vertical offset for different gaps. Vertical misalignment between capacitive beams has a much larger impact when the gap is small. ........ 36 Fig. 3.14. SEM image of CMOS-MEMS variable capacitor. ....................................................... 41 Fig. 3.15. Microscope image of the gap-tuning capacitor in (a) maximum capacitance position (b) minimum capacitance position. The latch appears retracted because voltage was applied to the heating resistors in the latch actuator. ........................................................................................... 42 Fig. 3.16. Microscope image of the latch at position (a) 1 (b) 2 (c) 3 (d) 4 (e) 5 (f) 6. Each position corresponds to a different capacitance value at which the variable capacitor can be maintained without consuming power. ......................................................................................... 42 Fig. 3.17. Microscope image of capacitive beams pushed into mechanical contact (a) for the bestperforming capacitor (b) for a capacitor with significant lateral curling. The amount of curling is small but still has significant effects of the minimum effective air gap. ...................................... 42 Fig. 3.18. Optical profilometer image of electrothermal gap-tuning capacitor. The capacitive beams are vertically aligned to within 1.6 µm. ............................................................................. 43 Fig. 3.19. S-parameter magnitude as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. S12 is not plotted because S12 and S21 are equivalent for passive networks. ................................................................................................... 44 Fig. 3.20. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ........................ 44 Fig. 3.21. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 20 GHz for the minimum capacitance position and decreases to 11 GHz for the maximum capacitance position. .................................................................................................... 44 Fig. 3.22. Quality factor as a function of frequency. This plot shows the Q for the maximum capacitance position. ..................................................................................................................... 47 Fig. 3.23. Plot of series capacitance versus latch position. Error bars indicate the 1-σ standard deviation between different devices based on 7 measurements. ................................................... 48 Fig. 3.24. Capacitance as a function of actuator voltage. Continuous tuning is possible with the latch retracted. Hysteresis is caused by capacitive beams sticking together as they are being pulled apart.................................................................................................................................... 48 Fig. 3.25. Capacitance as function of DC bias voltage. With the latch retracted (unlatched), capacitance beams snap together using less than 5 V. With the latch in place, up to 14 V can be applied before electrostatic pull-in occurs. ................................................................................... 49 Fig. 3.26. Electrothermal actuator frequency response. The thermal 3 dB frequency limits the speed of both the gap and latch actuators. .................................................................................... 51 Fig. 3.27. Electrothermal actuator transient response. The 90% rise time is 4.9 ms. .................. 51 Fig. 3.28. Temperature profile of gap-tuning capacitor viewed with quantum infrared microscope. Voltage is applied to resistors in (a) the gap actuator only (b) the latch actuator only (c) both the gap and latch actuators. ............................................................................................. 54 Fig. 3.29. Latched capacitance as a function of ambient temperature. While latched in position, the capacitance remains stable up to 65°C. ................................................................................... 55 Fig. 3.30. Minimum and maximum capacitance as a function of actuation cycles for two DUT‟s. Maximum capacitance decreased as the variable capacitor was cycled between latch positions. 57 Fig. 3.31. SEM images of mechanical latch after failure (a) zoomed out (b) zoomed in. Repeated impacts of the latch interface created sub-micron debris on the latch. ......................................... 57 Fig. 4.1. Concept of the actuation method for the switched capacitor. ......................................... 62 ix Fig. 4.2. Diagram of CMOS-MEMS switched capacitor viewed from above (a) with annotations (b) with circuit schematic; (c) SEM image of four CMOS-MEMS switched capacitors in parallel. ....................................................................................................................................................... 64 Fig. 4.3. Steady-state thermal circuit for the CMOS-MEMS switched capacitors. ...................... 66 Fig. 4.4. Cross-section of the switched capacitor along the rotor electrode. ................................ 67 Fig. 4.5. Diagram of CMOS-MEMS switched capacitor viewed (a) from above with nominal layout dimensions (b) from the side with the rotor electrode deflected vertically into minimum capacitance position. Line A–A’ and line B–B’ show the positions of the cross-sections along the rotor electrode and stator electrode, respectively. ......................................................................... 68 Fig. 4.6. Normalized parameters as a function of MIM capacitor length. Tuning ratio and capacitance density are normalized to their respective values for the optimum MIM capacitor length of 35 µm. ............................................................................................................................ 75 Fig. 4.7. Normalized parameters as a function of actuator beam length. Tuning ratio, capacitance density, and quality factor are normalized to their respective values for an actuator beam length of 90 µm to illustrate the tradeoffs necessary for increased tuning ratios. ................................... 76 Fig. 4.8. SEM image of 3-bit digital capacitor. Bits are separated by dummy structures to improve the temperature uniformity of the electrothermal actuators. Floating electrodes appear darker due to charging................................................................................................................... 78 Fig. 4.9. Microscope picture of CMOS-MEMS switched capacitors (a) in minimum capacitance position before actuation (rotor electrodes are deflected out of focus) (b) after electrothermal actuation (c) in maximum capacitance position after electrostatic actuation. .............................. 79 Fig. 4.10. Profilometer image of bit 3 of the digital capacitor (a) before actuation (b) after actuation. The rotor beams are not visible when deflected upwards because their angle with respect to the chip surface is too large for their profile to be captured by the Wyko unless settings are modified which sacrifice the quality of the image. ................................................................. 79 Fig. 4.11. Microscope image of the 3-bit digital capacitor in state (a) 0 (b) 1 (c) 2 (d) 3 (e) 4 (f) 5 (g) 6 (h) 7. Each capacitive bit is controlled independently to select the desired capacitance. .... 80 Fig. 4.12. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. .......................................................................................................................... 82 Fig. 4.13. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ........................ 82 Fig. 4.14. Lumped impedances as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. The behavior of |1/(Y11+Y12)| is explained by silicon remaining beneath the suspended MIM capacitor. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 11.4 GHz for the maximum capacitance position. ..................................................................................................................... 82 Fig. 4.15. Quality factor as a function of frequency. This plot shows the Q for the maximum capacitance position with port 2 connected to ground. ................................................................. 84 Fig. 4.16. Plot of series capacitance versus state. Error bars indicate the 1-σ standard deviation between different devices based on 7 measurements. ................................................................... 85 Fig. 4.17. Plot of series capacitance during an actuation cycle. The x-axis is arbitrary; all measured points shown were allowed to reach steady-state and maintained for approximately one minute. “Electrothermal” and “electrostatic” are abbreviated as “ET” and “ES”, respectively. The plotted data represents a single actuation cycle from a single device and thus may vary over time or between different devices. ........................................................................................................ 86 Fig. 4.18. Electrothermal actuator frequency response. ................................................................ 87 x Fig. 4.19. Simulated temperature profile of electrothermal actuator. Symmetry boundary conditions are applied so the simulation represents an array of many switched capacitors in parallel........................................................................................................................................... 89 Fig. 4.20. Series capacitance of a 3-bit digital capacitor as a function of ambient temperature. .. 91 Fig. 4.21. SEM images of surface damage caused by repeated temperature cycling at 250°C. The white boxes indicate the locations of the embedded heating resistors. ......................................... 92 Fig. 4.22. Vertical gap during heating as a function of the number of actuation cycles. A positive gap indicates that the tip of the rotor electrodes are deflected above the stator electrodes. ......... 93 Fig. 4.23. Diagram of capacitive bit with (a) interwoven actuator beams resulting in large parasitic capacitance (b) updated topology to reduce parasitic capacitance. ................................ 95 Fig. 4.24. SEM image of 3-bit digital capacitor updated to reduce parasitic capacitance. The chunk of debris was left in place since manual removal could damage the actuators. ................. 96 Fig. 4.25. Microscope image of the updated 3-bit digital capacitor in state (a) 0 (b) 1 (c) 2 (d) 3 (e) 4 (f) 5 (g) 6 (h) 7. Each capacitive bit is controlled independently to select the desired capacitance. ................................................................................................................................... 96 Fig. 4.26. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. .......................................................................................................................... 97 Fig. 4.27. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ........................ 97 Fig. 4.28. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 10 GHz for the maximum capacitance position. .................................................................................................... 97 Fig. 4.29. Quality factor as a function of frequency for the updated digital capacitor. This plot shows the Q for the maximum capacitance position with port 1 connected to ground. ............... 99 Fig. 5.1. Schematic of phased antenna array. Phase shifters create a variable time delay between antennas to enable electronic beam steering. .............................................................................. 103 Fig. 5.2. Schematic of distributed MEMS phase shifter (for N = 3 sections) as (a) distributed components (b) lumped component (c) lumped components with a single shunt capacitances. 104 Fig. 5.3. Diagram of the CMOS-MEMS switched capacitor viewed from above (a) with annotations (b) with circuit schematic; (c) SEM image of eight CMOS-MEMS switched capacitors in parallel. .................................................................................................................. 105 Fig. 5.4. Cross-section of the phase shifter (a) along the rotor electrode of a switched capacitor (b) along a section of transmission line without a switched capacitor........................................ 106 Fig. 5.5. Schematic of 4-bit digital capacitor used in the phase shifter. ..................................... 107 Fig. 5.6: SEM image of phase shifter digital capacitor. The rotor electrodes are vertically deflected into the minimum capacitance state. “D” indicates dummy bits which are included for temperature uniformity. .............................................................................................................. 113 Fig. 5.7: SEM image of one CMOS-MEMS phase shifter section. ............................................ 114 Fig. 5.8: SEM image of the CMOS-MEMS phase shifter composed of eight phase shifter sections. ....................................................................................................................................... 114 Fig. 5.9: Microscope image of CMOS-MEMS digital capacitor (a) before actuation (b) after actuation in state 10 (1010) with bits 4 and 2 turned on. ............................................................ 115 xi Fig. 5.10. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. ........................................................................................................................ 116 Fig. 5.11. Equivalent circuit model in terms of (a) Z-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ...................... 117 Fig. 5.12. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 10 GHz for the maximum capacitance position. .................................................................................................. 117 Fig. 5.13. Quality factor as a function of frequency for the digital capacitor used in the phase shifter. This plot shows the Q for the maximum capacitance position with port 2 left floating. 119 Fig. 5.14: Capacitance of phase shifter digital capacitor as a function of state. State 3 (0011) indicates that the two least significant bits are turned on. .......................................................... 120 Fig. 5.15: Measured phase shift as a function of frequency for all 16 states. Phase shift is negative because the time delay increases with increasing capacitance. ................................................... 121 Fig. 5.16: Measured insertion loss as a function of frequency for all 16 states. ......................... 121 Fig. 5.17: Measured reflection as a function of frequency for all 16 states. ............................... 122 Fig. 5.18. S-parameter magnitude as a function of frequency for (a) the minimum phase shift state (b) the maximum phase shift state. S12 is not plotted because S12 and S21 are equivalent for passive networks. ........................................................................................................................ 123 Fig. 5.19: Comparison of passive phase shifters. The average insertion loss is plotted versus the minimum operating frequency. ................................................................................................... 124 Fig. A.1. Layout view of (a) open test structure (b) short test structure (c) DUT. Measurements of the open and short test structures are necessary to remove the parasitic effects of the probe pads using open-short de-embedding. „G‟ and „S‟ represent „ground‟ and „signal‟, respectively. ..... 143 Fig. A.2. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ...................... 147 Fig. A.3. Lumped impedance magnitude as a function of frequency for the maximum capacitance position. The extracted model is compared with measured data. ............................................... 147 Fig. A.4. Series capacitance as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 148 Fig. A.5. Series resistance as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 148 Fig. A.6. Shunt capacitances as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 149 Fig. A.7. Quality factor as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 150 Fig. A.8. S-parameter magnitude as a function of frequency. The extracted model is compared with measured data. S12 is not plotted because S12 and S21 are equivalent for passive networks. ..................................................................................................................................................... 151 Fig. A.9. Equivalent circuit model in terms of (a) Z-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. ...................... 152 Fig. A.10. Lumped impedance magnitude as a function of frequency. The extracted model is compared with measured data. .................................................................................................... 152 Fig. A.11. Capacitance to ground as a function of frequency. The extracted model is compared with measured data. .................................................................................................................... 153 xii Fig. A.12. Shunt resistance as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 153 Fig. A.13. Series resistances as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 154 Fig. A.14. Series inductances as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 155 Fig. A.15. Quality factor as a function of frequency. The extracted model is compared with measured data.............................................................................................................................. 156 Fig. A.16. S-parameter magnitude as a function of frequency. The extracted model is compared with measured data. S12 is not plotted because S12 and S21 are equivalent for passive networks. ..................................................................................................................................................... 156 xiii Chapter 1. Introduction This chapter introduces the subject of CMOS-MEMS variable capacitors by outlining the motivation and goals of this work. Section 1.1 proposes multi-band radio as the motivation for developing MEMS variable capacitors. Section 1.2 explains why CMOS-MEMS variable capacitors are necessary when semiconductor variable capacitors are readily available. The thesis statement is presented in Section 1.3, and the contributions of this work are listed in Section 1.4. 1.1. Motivation Consumer radio systems are transmitting and receiving data over an increasing number of frequency bands. Modern cellular phones use four different frequency bands for transmitting voice and a fifth frequency band for wireless local area networking [1]. Multiple frequency bands are required to provide different functions and so that the total bandwidth is large enough to accommodate many users. Multi-band operation is currently achieved by using multiple narrowband circuits, each of which is capable of operation over a single frequency band. For example, dual- and quad-band receivers simply include two or four low-noise amplifiers, respectively [2], [3]. Similarly, dual-band and quad-band transmitters require two or four power amplifiers, respectively [4], [1]. As the trend for operation over more frequency bands continues, the current implementation of multi-band radio systems requires more circuits and thus becomes more expensive. To reduce the cost of multi-band radio, there are two possible solutions: wide-band circuits and reconfigurable circuits. Wide-band circuits with a bandwidth spanning multiple frequency bands would eliminate the need for multiple circuits, but the performance of wide-band circuits is inferior to that of narrow-band circuits. The ideal solution is reconfigurable circuits, which are Chapter 1: Introduction 1 tunable narrow-band circuits capable of reconfiguration between different frequency bands. Reconfigurable circuits have the potential to eliminate the need for multiple circuits while providing performance typical of narrow-band circuits. The performance of reconfigurable circuits has been limited by the available tunable devices. Because many devices must be integrated onto a single CMOS chip to minimize cost, the tunable devices available at RF are semiconductor variable capacitors. These variable capacitors perform poorly when designed for large tuning capabilities. To enable reconfigurable circuits for multi-band radio, this work proposes to use MEMS variable capacitors. Gap Tuning d g–d C 0 LW g d Area Tuning d C Movable Dielectric Metal L–d d 0 L d W g C L–d 0 d r L d W g Dielectric (εr > 1) All objects have width W into the page (a) (b) (c) Fig. 1.1. Diagram of capacitance tuning by changing the (a) gap (b) overlap area (c) dielectric. All capacitors have length L, width W, and gap g before part of the structure undergoes displacement d. Capacitance calculations ignore the effects of fringing fields. 1.2. MEMS Variable Capacitors MEMS variable capacitors are electrical devices whose capacitance value is controlled through mechanical actuation. Capacitance tuning is accomplished by mechanically changing the gap [5], overlap area [6], or dielectric [7] between the capacitive electrodes (see Fig. 1.1). Depending on the nature of the actuation, the tuning range can be continuous from minimum capacitance to maximum capacitance or consist of a set of achievable discrete capacitance values. Devices with continuous tuning range are often referred to as variable capacitors or varactors Chapter 1: Introduction 2 whereas devices with two states, on and off, are referred to as capacitive switches or switched capacitors. Devices that achieve many discrete capacitance values by using many independently controlled switched capacitors are referred to as digital capacitors. 1.2.1. Characteristics of MEMS Variable Capacitors The electrical performance of MEMS variable capacitors is described by three primary metrics. The tuning ratio, TR, is the ratio of maximum capacitance, Cmax, to minimum capacitance, Cmin. TR C max C min . (1.1) The quality factor, Q, is the ratio of energy stored to energy dissipated when charge is stored on a capacitor. Typically, Q is limited by the series resistance, Rs, of the metal path. Q 1 CRs , (1.2) where ω is the electrical frequency and C is the capacitance. The self-resonance frequency, ωres, is the frequency at which the variable capacitance resonates due to parasitic inductance, L. res 1 LC , (1.3) Above the self-resonance frequency, the total imaginary impedance is inductive meaning that the device no longer functions as a capacitor. All three of these metrics should be maximized to optimize a variable capacitor. Several other characteristics of MEMS variable capacitors are also important. Capacitance density is the ratio of the maximum capacitance to the device area, which is important when the variable capacitor is monolithically integrated with other devices. The actuation voltage or Chapter 1: Introduction 3 actuation power is the specified voltage or power required for mechanical actuation. Switching time is the time required to change the capacitance, which can differ depending on the state of the variable capacitance. For example, the turn-on time of switched capacitors is often different from the turn-off time. Reliability is the measure of the lifetime of a variable capacitor, often characterized by the number of actuation cycles achieved before failure. Power handling is defined as the maximum power that can be transmitted through the variable capacitors, typically assuming a 50-Ω source and load impedance. Power handling is limited by either the maximum voltage or maximum current a variable capacitor can tolerate. Linearity refers to the relationship between the voltage applied across the capacitor terminals and the stored electrical charge. Ideally, the charge increases linearly with applied voltage, but all capacitors are nonlinear to some extent. The linearity of a variable capacitor is often characterized by measuring the 3rdorder input-referred intercept power (IIP3) of a circuit using that variable capacitor. The mechanical actuation of MEMS variable capacitors can be electrostatic, electrothermal, piezoelectric, or magnetic. Actuation characteristics vary according to the particular actuator design, but several generalizations can be stated about the actuation types that hold true for most designs. Electrostatic actuation produces small displacements (< 10 µm) using high actuation voltages (> 10 V). The other three types of actuation, electrothermal actuation, piezoelectric actuation, and magnetic actuation, are at first glance superior because those types produce large displacements (> 10 µm) using low actuation voltages (< 10 V). Large displacements are required for high tuning ratios, and low actuation voltages are desirable so that the variable capacitor can be easily controlled. However, piezoelectric actuation and magnetic actuation are limited because those two types require piezoelectric and magnetic materials, respectively, which are difficult to integrate with semiconductors. In contrast, electrostatic actuation and Chapter 1: Introduction 4 electrothermal actuation can be used with any MEMS technology because those two types of actuation only require materials that are electrically conductive. Electrothermal actuation requires power for heating, typically on the order of milliwatts, and the switching time is limited by the thermal time constant, typically on the order of milliseconds. Because electrical time constants are on the order of nanoseconds, the switching time of electrostatic actuation is limited only by the mechanical response, typically on the order of microseconds. Electrostatic actuation is by far the most popular type of MEMS actuation because the choice of materials is virtually unlimited, the power consumption is negligible, and the switching time is fast. 1.2.2. Advantages of MEMS Variable Capacitors Because semiconductor variable capacitors are more mature than MEMS variable capacitors, semiconductor variable capacitors must be discussed to illustrate the advantages of MEMS variable capacitors. The three types of semiconductor variable capacitors available in a standard CMOS process are the diode varactor, the metal-oxide-semiconductor (MOS) varactor, and the switched metal-insulator-metal (MIM) capacitor. All three of these types have switching times on the order of nanoseconds and large capacitance densities (> 100 pF/mm2). The diode varactor and the MOS varactor have high Q‘s (> 100 at 1 GHz), but the tuning ratios of these varactors are small (< 5:1). Both varactors are fundamentally nonlinear devices because the varactors only have two electrical terminals such that the applied voltage directly affects the capacitance value. The switched MIM capacitor, which consists of a MIM capacitor in series with the channel of a metal-oxide-semiconductor field-effect transistor (MOSFET), has a third terminal, the gate of the MOSFET, which controls the capacitance and thus is more linear. Switched MIM capacitors can be designed for large tuning ratios (> 5:1), but Q decreases as the Chapter 1: Introduction 5 tuning ratio increases. This tradeoff occurs because, for large tuning ratios, the MOSFET must be small to minimize parasitic capacitance, but a small MOSFET has high channel resistance, which degrades the Q. None of the semiconductor variable capacitors can simultaneously achieve both large tuning ratio (> 10:1) and high Q (> 100 at 1 GHz). TABLE 1.1: VARIABLE CAPACITOR CHARACTERISTICS Parameter Switched MIM Capacitor University of University of University of Imperial California– Los California – San California – San College London Angeles [8] Diego [9] Diego [10] [11] MEMS MEMS MEMS MEMS 8.6 pF 192 fF 1.00 pF 600 fF 270 fF 63 fF 51 fF 60 fF Type CMOS Cmax 1.11 pF Cmin 111 fF Shunt Tuning 10:1 31:1 3.0:1 Ratio Q at 1 GHz* 9 9 1940 FoM 90 279 5820 (TRshunt*Q) Self-Resonance NA 2.7 GHz > 30 GHz Frequency* Capacitance 1000 pF/mm2 NA ≈ 9.5 pF/mm2 Density Switching MOSFET Electrostatic Electrostatic Method Tuning Range Digital Analog Digital Switching Time ~ 1 ns NA < 0.5 µs Actuation 3.3 V 40 V 32 V Voltage Actuation ≈ 0** ≈ 0** ≈ 0** Energy for 1 pF * The parameter corresponds to the maximum capacitance state. * *The power consumed for actuation is negligible. 19.6:1 10:1 265 352 5194 3520 > 10 GHz 17 GHz ≈ 12.8 pF/mm2 12 pF/mm2 Electrostatic Electrostatic Digital 8 µs Analog NA 40 V 30 V ≈ 0** ≈ 0** MEMS variable capacitors have demonstrated several advantages over semiconductor variable capacitors. For comparison, Table 1.1 presents the performance of several MEMS variable capacitors along with that of a switched MIM capacitor simulated for the 0.35 µm BiCMOS process from TowerJazz Semiconductor. First, MEMS variable capacitors can provide the combination of large tuning ratio (> 5:1) and high Q (> 100 at 1 GHz). Large tuning ratios are made possible by the mechanical displacements produced by MEMS actuators, and the Q‘s are high because MEMS variable capacitors are composed entirely of metal. Semiconductor devices Chapter 1: Introduction 6 are limited in Q due to their close proximity to and inclusion of silicon, which is a poor electrical conductor at RF. The figure of merit (FoM) introduced in Table 1.1 is the product of the shunt tuning ratio and the Q at 1 GHz for the maximum capacitance. FoM TRshunt * Q . (1.4) Second, MEMS variable capacitors are highly linear because mechanical motion is attenuated beyond the mechanical resonance frequency, which is typically less than 10 MHz [12]. As a result, the variable capacitance is not significantly affected by the applied RF voltage. Third, the power handling of MEMS variable capacitors is often excellent because large voltages (> 10 V) can be applied without causing electrical breakdown [13]. Due to their advantages, MEMS variable capacitors have been applied to both enhance existing RF circuits and enable new applications. MEMS capacitive switches proposed to serve as transmit-receive switches have demonstrated reduced insertion loss when compared with stateof-the-art semiconductor switches [14]. Phase shifters for phased antenna arrays have been improved using either MEMS variable capacitors or MEMS capacitive switches [15]. MEMS variable capacitors have been combined with fixed inductors or transmission lines to form highQ tunable resonators for a number of applications. The phase noise of voltage controlled oscillators has been improved by replacing the semiconductor variable capacitors with MEMS variable capacitors [16], [17]. Reconfigurable filters have been enabled by the large tuning ratios available with MEMS variable capacitors [18]. Antenna tuners have been developed to alleviate source-load mismatch between power amplifiers and antennas [19]. Using this work, narrowband RF circuits, specifically low-noise amplifiers [20] and power amplifiers [21], have been made reconfigurable by MEMS variable capacitors. However, there are drawbacks to MEMS variable capacitors. When compared with Chapter 1: Introduction 7 semiconductor variable capacitors, MEMS variable capacitors are slow, typically switching in milliseconds or microseconds as opposed to nanoseconds. Capacitance density is low (< 100 pF/mm2) due to the large dimensions of the MEMS structure, which are typically on the order of tens or hundreds of microns. For electrostatically actuated MEMS, power consumption is negligible but the required voltage is often large, on the order of tens of volts [13]. The opposite is true of electrothermally actuated MEMS; the required voltage is small, on the order of volts, but power consumption is significant, on the order of tens of milliwatts [22]. Reliability is not guaranteed since RF MEMS can fail from dielectric charging [23], mechanical creep or fatigue, or from degradation related to repeated mechanical contact [24]. Fortunately, many of these drawbacks can either be mitigated or tolerated by the target applications. Faster switching times are desired but not explicitly required for antenna tuners and many reconfigurable networks. The large area consumption of MEMS is acceptable because, to form resonators, MEMS variable capacitors are often combined with spiral inductors or transmission lines, both of which have dimensions on the order of hundreds of microns or larger. Large voltages for electrostatic actuation can be provided by on-chip CMOS charge pumps if high voltage CMOS is available [25]. Mechanical latches or electrostatic force [26] can be used to eliminate the static power consumption of electrothermal actuation [27]. Research on the reliability of RF MEMS has yielded methods to mitigate dielectric charging [28] and ultimately achieve billions of actuation cycles without failure [9]. 1.2.3. Monolithic Integration Despite their excellent performance, MEMS variable capacitors are not widely used in RF circuits because most MEMS variable capacitors are not monolithically integrated with CMOS. Chapter 1: Introduction 8 Monolithic integration is required because the inclusion of MEMS variable capacitors into RF circuits as discrete components is simply too expensive to warrant their use. This requirement is challenging because MEMS fabrication is complicated by the need for integration. The structural and sacrificial layers necessary for MEMS require additional processing beyond that of a standard CMOS process, especially when integration with CMOS is required. Low-temperature micromachining can be used to fabricate MEMS devices directly on top of an existing CMOS process [29]. Alternatively, MEMS can be fabricated on a separate substrate and flip-chip bonded onto a CMOS chip [30], [31]. Both methods of integration grant the MEMS designer the freedom to select the materials and layer thicknesses, but this freedom comes at the cost of additional lithography steps, which ultimately increases the cost of any system using such integrated MEMS. To solve this problem, the post-CMOS micromachining process in [32], hereafter called the ―CMOS-MEMS process‖ can be used for monolithic integration of MEMS devices without using additional lithography steps. The CMOS-MEMS process uses maskless, post-CMOS etching to create MEMS devices composed of the metal-dielectric stack of a foundry CMOS process. The resulting MEMS structure can consist of multiple metal layers connected in parallel for low resistance. The air gap between the MEMS structure and the remaining silicon substrate is approximately 50 µm, which minimizes parasitic capacitance to the resistive silicon. Polysilicon resistors can be embedded within the MEMS structure to act as heaters for low-voltage electrothermal actuation [22]. These features make CMOS-MEMS a natural choice for integrating MEMS variable capacitors in a low-cost manner. 1.3. Thesis Statement MEMS variable capacitors have demonstrated significant advantages over semiconductor Chapter 1: Introduction 9 variable capacitors but are not cost-effective if included in electrical systems as discrete components. The typical methods available for monolithically integrating MEMS variable capacitors with CMOS require additional lithography steps, which increases cost. The CMOSMEMS process is an alternative method for integrating MEMS with CMOS without using additional lithography steps. Thus, using the CMOS-MEMS process, CMOS-MEMS variable capacitors can be designed with performance comparable to that of other integrated MEMS variable capacitors and enable the design of low-cost reconfigurable RF circuits. To assess the performance of CMOS-MEMS variable capacitors, devices were designed, fabricated, and characterized. The electrical characteristics were measured to calculate the tuning ratio, Q, and self-resonance frequency. Actuation methods were tested to determine the switching times along with voltage and power requirements. The effects of repeated actuation were analyzed to determine reliability. Results are compared with that of other integrated MEMS variable capacitors. To show that CMOS-MEMS variable capacitors enable the design of reconfigurable RF circuits, a CMOS-MEMS phase shifter was designed, fabricated, and characterized. The electrical characteristics were measured to calculate the phase shift and insertion loss. Results are compared with that of other MEMS phase shifters. 1.4. Contributions The first contribution of this work, the design and characterization of a CMOS-MEMS electrothermal gap-tuning capacitor, is presented in Chapter 3. The gap-tuning capacitor uses lateral electrothermal actuation to control the capacitive gap between interdigitated electrodes. A mechanical latch, also controlled by electrothermal actuation, was used to maintain the capacitive Chapter 1: Introduction 10 gap without consuming static power for heating. The gap-tuning capacitor requires 4 V and 15 mW for electrothermal actuation and can be reconfigured in 30 ms. The average measured tuning ratio is 4.9:1, and the measured Q for the maximum capacitance state is 80 at 1 GHz. Reliability testing demonstrated that the lifetime exceeds millions of actuation cycles. Chapter 4 presents the second contribution of this work, the design and characterization of CMOS-MEMS digital capacitors. The digital capacitors are composed of many independently controlled CMOS-MEMS switched capacitors, which can be toggled on or off using a combination of vertical electrothermal and lateral electrostatic actuation. Electrostatic force is used to maintain the maximum capacitance state without consuming static power. The digital capacitors can be switched in 1 ms using 4 V for electrothermal actuation and 20 V for electrostatic actuation. The average measured tuning ratio is 61:1, and the measured Q for the maximum capacitance state is 369 at 1 GHz. Reliability testing demonstrated millions of actuation cycles successfully but revealed a drift in the actuation characteristics after 10 million cycles. These digital capacitors represent the state-of-the-art for CMOS-MEMS variable capacitors. The design and characterization of a 4-bit CMOS-MEMS phase shifter for phased antenna arrays is presented in Chapter 5 as the third contribution of this work. The phase shifter is composed of a micromachined transmission line periodically loaded with CMOS-MEMS digital capacitors. The digital capacitors are modified to achieve a Q of 18.8 at 30 GHz for the maximum capacitance state and a tuning ratio of 7.2:1. Measured results demonstrate the required phase shift of 337.5° at 32 GHz with an average insertion loss of 2.9 dB. This phase shifter is the first RF MEMS phase shifter to be monolithically integrated with CMOS. Chapter 1: Introduction 11 Chapter 2. CMOS-MEMS Background This chapter presents the CMOS-MEMS process that is used to fabricate the MEMS variable capacitors presented in this work. Section 2.1 describes the process flow along with several characteristics of CMOS-MEMS structures relevant to variable capacitors. Section 2.2 discusses how the CMOS-MEMS process is used to create both lateral and vertical electrothermal actuators. 2.1. CMOS-MEMS Process The CMOS-MEMS process is used to fabricate MEMS that are monolithically integrated with standard CMOS. The process can be applied to any CMOS technology provided that the placement of metal layers can be controlled by layout. CMOS technologies with strict metal fill requirements cannot be used for CMOS-MEMS since the metal fill blocks the post-CMOS etching. The CMOS-MEMS process can be applied to CMOS from different foundries and from different technology generations, such as 0.35 µm or 0.18 µm. The CMOS-MEMS devices fabricated for this work all use the 0.35 µm BiCMOS process from TowerJazz. This CMOS process provides four aluminum metal layers, one thick top metal and three thin metals, one polysilicon layer, and tungsten vias. The minimum width and spacing for the thick top metal is 2.6 µm and 1.9 µm, respectively. The minimum width and spacing for the three thin metals are both 0.5 µm. 2.1.1. Process Flow The CMOS-MEMS process consists of maskless, post-CMOS etching to create MEMS structures from the metal-dielectric stack of a standard CMOS process (see Fig. 2.1) [32]. First, Chapter 2: CMOS-MEMS 12 the exposed oxide is anisotropically etched using a CHF3/CF4/O2 reactive ion etch (RIE). During the oxide etch, the top-most metal layer acts as a mask both to define the resulting MEMS structure and to protect devices such as transistors which require the silicon substrate. The oxide etch exposes the silicon surface in areas which contain no metal. Next, the exposed silicon is anisotropically etched using an SF6/O2 RIE. This anisotropic silicon etch is necessary to create a large vertical air gap (≈ 50 µm) between the resulting MEMS structure and the bottom of the silicon etch pit. Finally, the silicon is etched isotropically using SF6 plasma to release the resulting MEMS structures from the substrate. The isotropic silicon etch is timed such that narrow structures are undercut while wide structures remain anchored to the substrate (see Fig. 2.2). Key: Metal 1 (a) (b) (c) (d) Metal 2 Metal 3 Metal 4 Oxide Via Silicon Fig. 2.1. Cross-section of the CMOS-MEMS process flow (a) beginning with CMOS chip (b) after the vertical oxide etch (c) after the vertical silicon etch (d) after silicon release etch. Chapter 2: CMOS-MEMS 13 50 µm 20 µm (a) (b) Fig. 2.2. SEM image of (a) 20 µm squared released from the substrate (b) 60 µm square anchored to the substrate. These test structures were fabricated using the 0.18 µm BiCMOS process from TowerJazz. 2.1.2. CMOS-MEMS Characteristics The CMOS-MEMS process is very flexible because a wide variety of beams can be created. CMOS-MEMS beams can consist of a single metal layer, such as a metal-1 beam, or multiple metal layers, such as a metal-321 beam. Beams consisting of multiple metal layers can carry multiple independent signals or be connected in parallel to serve as low-resistance wires. Both vias and polysilicon can be embedded within the CMOS-MEMS beams. Embedded polysilicon is encased in oxide and thus does not etch during the isotropic silicon etch. Fig. 2.1(d) shows three examples of released CMOS-MEMS beams. The left beam is a metal-321 beam with metal 1 and 2 offset layers. The middle beam is a metal-4321 beam with vias and polysilicon. The right beam is a metal-1 beam. Several characteristics of the CMOS-MEMS process are particularly beneficial for RF MEMS. First, the large vertical air gap (≈ 50 µm) below the MEMS structure minimizes parasitic capacitance to the silicon substrate. Since silicon is a poor conductor at RF, coupling to the silicon must be avoided to maximize the Q of CMOS-MEMS variable capacitors. Second, Chapter 2: CMOS-MEMS 14 CMOS processes often include a thick metal layer to minimize the series resistance of on-chip spiral inductors. This thick metal layer is useful for minimizing the series resistance of variable capacitors. Third, polysilicon resistors can be embedded in the CMOS-MEMS beams to enable electrothermal actuation. These resistors allow the temperature of electrothermal actuator beams to be controlled using an applied voltage. During the vertical oxide etch, a passivation layer is deposited on the side-walls of the CMOS-MEMS beams. This side-wall polymer is required for the oxide etch to be anisotropic. The polymer layer also prevents the titanium-tungsten adhesion layers, which are deposited above and below the aluminum metal layers as part of the CMOS process, from etching during the silicon release etch. Without sufficient side-wall polymer, the aluminum metal layers can delaminate and peel away from the oxide. The side-wall polymer is a dielectric material of unknown relative permittivity. Since observations of the thickness in SEM images approximately match the thickness calculated from capacitance measurements, the relative permittivity is expected to be near unity. The thickness of the side-wall polymer determines the effective air gap between CMOS-MEMS beams pushed into mechanical contact (≈ 0.2 µm). The air gap is referred to as effective because there is no actual air gap, simply a gap containing side-wall polymer whose permittivity is assumed to be equal to that of air. This effective air gap limits the maximum capacitance per unit length between two CMOS-MEMS beams. Note that CMOSMEMS beams pushed into mechanical contact do not short circuit. Shorting would be unexpected because aluminum has a native oxide that prevents electrical contact. The side-wall polymer is believed to further help prevent electrical breakdown. Two CMOS-MEMS beams in mechanical contact can withstand greater than 30 V before electrical breakdown occurs. Electrical breakdown typically occurs when more than 50 V has been applied. Breakdown Chapter 2: CMOS-MEMS 15 destroys the MEMS structures unless the current is limited by a resistor. After the silicon release etch, all CMOS-MEMS beams curl vertically with a radius of curvature specific to their metal-dielectric stack. Because beams composed of multiple metal and dielectric layers are thicker, these beams curl much less than those composed of a single layer of metal and dielectric. Metal-1 beams can be used to achieve large vertical deflections whereas metal-4321 beams can be used when vertical curling is unwanted. Most beams curl up away from the silicon, but the direction depends on the particular CMOS process. Because metal and dielectric have different temperature coefficients of expansion (TCE), the stress difference between the two materials, and hence the radius of curvature, is temperature dependent. CMOS-MEMS beams are prone to another type of curling, referred to as unintended lateral curling. Unintended lateral curling is fundamentally different from the vertical curling associated with residual stress because this curling occurs even for beams with symmetric cross-sections. For example, a metal-1 beam has an asymmetric cross-section for the purpose of vertical curling because the beam consists of metal on top and dielectric on bottom. Thus vertical curling is expected. The metal-1 beam has a symmetric cross-section for the purpose of lateral curling because the beam cross-section is uniform in the lateral direction. Thus lateral curling is not expected. Yet the fact remains that all CMOS-MEMS beams curl laterally, even when such curling is not intended. Unintended lateral curling has been observed in many CMOS-MEMS devices and occurs to a greater extent for beams which are long and narrow. The cause of this curling is unknown because the cross-sections of the affected beams are symmetrical such that there is no apparent reason for curling in one direction or the other. Several theories exist to explain this curling. Because beams with the same local geometry on the same die curl the same direction, the Chapter 2: CMOS-MEMS 16 problem may be caused by the CMOS-MEMS process. Further optimization of the process, such as a reduction in the thickness of the sidewall polymer, has been theorized to reduce the curling but has not yet been demonstrated. Misalignment of the metal layers may contribute to unintended lateral curling but cannot be solely responsible since beams composed of only a single metal layer still curl. 2.2. Electrothermal Actuation Electrothermal actuation produces mechanical displacements using the thermal expansion of materials. The actuation is controlled by a voltage applied to a resistor, which causes electrical power to be dissipated as heat. That heat causes a temperature change resulting in thermal expansion. The electrothermal actuators used in this work are thermal bimorphs or, if the actuators include more than two layers, thermal multi-morphs. Thermal bimorphs are beams that consist of two materials with different temperature coefficients of expansion (TCE). Neglecting the thin (< 100 nm) titanium-tungsten adhesion layers, the materials used in this work are aluminum, which has a TCE of 23.1 µK-1, and oxide, which has a TCE of 0.5 µK-1 [33]. Because the TCE are different, any change in temperature causes a difference in mechanical stress between the two materials. This stress difference results in a temperature-dependent bending moment applied along the length of a beam, which causes the beam to curl in a circular arc. As discussed in Section 2.1.2, all CMOS-MEMS beams, thermal bimorphs included, deflect at room temperature according to the particular metal-dielectric stack. This vertical deflection occurs immediately after the beams are released from the substrate and is thus referred to as the self-assembly displacement. This displacement is caused by residual stress in the CMOS materials after fabrication, so the magnitude and direction of the displacement depends on the Chapter 2: CMOS-MEMS 17 CMOS process. Heating the beams above room temperature typically produces displacement in the opposite direction of self-assembly. The temperature to which the beams must be heated to reduce the total displacement to zero is defined as the zero-stress temperature. For the 0.35 µm BiCMOS process used in this work, the zero-stress temperature is approximately 180°C. The analysis necessary to calculate the temperature-dependant deflection of thermal multimorphs is presented in [22]. Consider a simple thermal bimorph with length L and width w that consists of one layer of metal with thickness hm on top one layer of oxide with thickness ho. Assuming that the position of the neutral axis is near the metal-oxide interface, [22] calculates the vertical displacement at the tip, z(L), according to the dimensions of the bimorph. z L wL2 4EI eff o ho 2 m hm 2 , (2.1) where (EI)eff is the effective flexural rigidity and σo and σm represent the mechanical stress in the oxide and metal, respectively. The mechanical stress results from thermal expansion and thus depends on the change in temperature, ∆T. ET , (2.2) where α is the TCE and E is the Young‘s modulus of the material. The change in temperature is relative to the zero-stress temperature, T0. E T0 Ta , (2.3) where Ta is the temperature of the actuator beam. Substituting (2.3) into (2.1) yields the vertical displacement as a function of the actuator beam temperature. zL wL2 4EI eff o E o ho 2 m E m hm 2 T0 Ta , (2.4) The dimensions and material characteristics of the actuator beam cross-section can be lumped into a single parameter defined as the sensitivity of the actuator beam cross-section, S. Chapter 2: CMOS-MEMS 18 S w 4EI eff o E o ho 2 m E m hm 2 . (2.5) This sensitivity simplifies the expression for the vertical displacement of the actuator beam. zL SL2 T0 Ta , (2.6) Note that once the sensitivity is determined for a particular actuator beam cross-section, the length of the actuator beam can be scaled to provide the necessary displacement. The radius of curvature, ρ, is calculated using the length and tip displacement of the actuator beam. L2 2 z L , (2.7) Experimental results from [22] provided both the zero-stress temperature and the actuator cross-section sensitivity for a variety of different cross-sections. For the CMOS-MEMS variable capacitors presented in this work, the actuator beam cross-sections with the largest sensitivity were chosen to maximize displacement. The length of the actuator beams were determined by (2.6) or (2.7) to achieve the required displacement or radius of curvature, respectively. 2.2.1. Vertical Electrothermal Actuators The vertical electrothermal actuators used in this work are metal-1 beams, which closely resemble ideal thermal bimorphs (see Fig. 2.3). The heating resistor can be distributed along the length of the actuator beam or lumped at the anchor points. For the 0.35 µm BiCMOS process used in this work, the direction of the self-assembly displacement is up, away from the silicon, but this direction can vary according to the particular CMOS process. Measurements of test structures showed that a 35 µm long actuator beam without polysilicon deflected by 1.30 µm at room temperature (25°C). Using (2.7) the radius of curvature resulting from the self-assembly displacement was calculated to be 471 µm. This calculated radius of curvature is used in the Chapter 2: CMOS-MEMS 19 design of the digital capacitor presented in Section 4.4. Assuming that the zero-stress temperature for this actuator is 180°C, the sensitivity of the actuator beam cross-section was calculated to be 6.84 m-1K-1 using (2.6). Side View Cross-sections Self-assembly displacement due to residual stress No displacement after heating to the zero-stress temperature Key: Metal 1 Actuator beam without polysilicon resistor Actuator beam with polysilicon resistor Oxide Silicon Fig. 2.3. Diagram of a vertical electrothermal actuator beam. The polysilicon resistor for heating can either be lumped near the anchor point or distributed along the length of the actuator beam. 2.2.2. Lateral Electrothermal Actuators The lateral electrothermal actuators used in this work are metal-321 beams with metal-2 and metal-1 offset to one side of the beam (see Fig. 2.4). The lateral electrothermal actuator beams are far from ideal thermal bimorphs since both the top metal and most of the oxide must span the width of the actuator beam. These thermal multi-morphs operate according to the same principles as a thermal bimorph except that only the offset metal layers contribute a temperature-dependent stress difference that causes displacement. If the offset metal layers are confined to one side of the actuator beams, the lateral actuator beams will curl in a circular arc similar to the vertical electrothermal actuator beams. This type of curling is undesirable for many applications since the non-zero angle at the tip prevents many actuator beams from being connected in parallel to drive Chapter 2: CMOS-MEMS 20 a single structure. To achieve a zero angle at the tip, the offset metal layers are switched to the opposite side of the beam halfway along the length to produce guided-end motion. This guidedend configuration is equivalent to two half actuator beams connected in series. Since the displacement is proportional to the length squared, the guided-end actuator produces half the displacement of a non-guided-end actuator of the same length. Top View Cross-sections Self-assembly displacement due to residual stress A A’ B B’ B A’ B’ No displacement after heating to the zero-stress temperature Key: A Metal 1 Metal 2 Metal 3 Oxide Fig. 2.4. Diagram of a lateral electrothermal actuator beam. The offset metal layers are switched to the opposite side of the beam halfway along the length to produce guided-end motion at the tip. For the 0.35 µm BiCMOS process used in this work, the actuator beam cross-section that maximizes displacement for a given length consists of 1 µm wide metal-3 with 0.5 µm wide metal-2 and metal-1 offset layers. The actuator beam width is too narrow to accommodate polysilicon, so the heating resistors are lumped at anchor points. For this cross-section, a 100 µm long actuator beam in guided-end configuration deflected by 2.42 µm at room temperature (25°C). The measured zero-stress temperature for that actuator was 175.7°C. Using, (2.6) the sensitivity of the actuator beam cross-section was calculated to be 1.61 m-1K-1. This calculated sensitivity is used in the design of the gap-tuning capacitor presented in Section 3.4. Using (2.7) Chapter 2: CMOS-MEMS 21 for a lateral actuator beam not in guided-end configuration, the radius of curvature resulting from the self-assembly displacement is 1.0 mm. Fig. 2.5 shows microscope images of 14 lateral electrothermal actuator connected in parallel and a focused-ion beam image of cross-section of a single actuator beam. (a) (b) (c) Fig. 2.5. Microscope image of lateral electrothermal actuator (a) deflected due to residual stress (b) after heating. (c) Focused-ion beam image of electrothermal actuator beam cross-section. Chapter 2: CMOS-MEMS 22 Chapter 3. Gap-tuning Capacitor The CMOS-MEMS variable capacitors presented in this section use lateral electrothermal actuation to control the gap between interdigitated capacitive beams. Mechanical latches are used to maintain the gap without consuming static power for heating. CMOS-MEMS gap-tuning capacitors had been developed prior to this work, so Section 3.1 explains the problems with previously developed CMOS-MEMS gap-tuning capacitors that serve as motivation. Section 3.2 presents the improved topology of the gap-tuning capacitor. Section 3.3 presents an analysis of the topology linking the degrees of freedom to performance metrics. Section 3.4 describes how that analysis was used to define the geometry of the gap-tuning capacitor. Section 3.5 presents characterization results of the fabricated design. Section 3.6 discusses the various advantages and disadvantages of the resulting gap-tuning capacitor. 3.1. Motivation There are two problems that MEMS variable capacitors must solve to enable frequencyreconfigurable circuits for multi-band radio. First, the variable capacitors must have large tuning ratios so that the circuits can be reconfigured for operation over a wide range of frequencies. Second, the variable capacitors must be controlled using on-chip voltages. Generating control voltages off-chip requires additional discrete components, which defeats the purpose of monolithic integration with CMOS. Ideally, the required control voltages are below the supply voltage of the CMOS process to avoid the need for CMOS charge pumps, which consume both power and area. Electrothermal actuation is the solution to both problems since electrothermal actuators produce the large displacements (> 10 µm) necessary for large tuning ratios and require only low voltages (< 5 V). However, electrothermal actuation introduces a third problem since Chapter 3: Gap-tuning Capacitor 23 electrothermal actuators require power for heating. To use electrothermal actuation without consuming static power for heating, the CMOS-MEMS gap-tuning capacitor includes a mechanical latch. 3.1.1. Concept The gap-tuning capacitor uses lateral electrothermal actuation to control both the gap between the capacitive beams and the position of the mechanical latch (see Fig. 3.1). After the MEMS structure is released from the substrate by the CMOS-MEMS process, the electrothermal actuators self-assemble such that the capacitive beams are pushed into mechanical contact, resulting in maximum capacitance. To reconfigure the gap-tuning capacitor, both actuators are heated to simultaneously disengage the latch and change the gap between the capacitive beams. Then, the actuator controlling the latch is cooled to cause the latch to engage and maintain the new position of the capacitive beams. Finally, the actuator controlling the gap is cooled so that no static power is consumed for heating. Maximum capacitance is maintained by self-assembly Both actuators are heated to reverse the self-assembly Minimum capacitance is maintained by the latch actuation self-assembly self-assembly self-assembly Key: actuation Latch interface Actuator Beams self-assembly Capacitive beams Fig. 3.1. Concept of the actuation method for the gap-tuning capacitor. Chapter 3: Gap-tuning Capacitor 24 The combination of lateral electrothermal actuation and mechanical latching is not limited to gap-tuning capacitors. The same actuation method could be applied to create an area-tuning capacitor. Gap-tuning was pursued over area-tuning because gap-tuning has the potential to create a large maximum capacitance when the capacitive beams are pushed into mechanical contact. Area-tuning capacitors typically maintain an air gap, which results in lower maximum capacitances. (a) (b) (c) (d) Fig. 3.2. SEM image of the (a) 1st (b) 2nd (c) 3rd (d) 4th design of the gap-tuning capacitor. 3.1.2. Related Work Prior to this work, four generations of electrothermal gap-tuning capacitors had been designed and characterized (see Fig. 3.2). The common goal in every design was to use electrothermal actuation to control the lateral gap between capacitive beams. Most designs Chapter 3: Gap-tuning Capacitor 25 included a mechanical latch designed to maintain a given capacitive gap without consuming power for heating. While performance metrics such as Q and capacitance density were considered, achieving a large tuning ratio was the primary focus. The first two designs (see Fig. 3.2(a) and Fig. 3.2(b)) included methods for both lateral and vertical actuation but did not achieve tuning ratios greater than 1.5:1 [34]. The vertical curling of CMOS-MEMS beams was not anticipated, so, as a result, the capacitive beams were vertically misaligned (see Fig. 3.3), which limited the maximum. The third design (see Fig. 3.2(c)) achieved the first significant tuning ratio of 3.5:1 by focusing purely on lateral actuation to control the capacitance gap [35]. The fourth design (see Fig. 3.2(d)) was a refinement of the third design but still suffered from vertical misalignment between the capacitance beams, which limited the tuning ratio to approximately 3:1. The common trend in these previous designs was a failure to compensate for the vertical curling of CMOS-MEMS beams. Side View Misaligned beams (b) (a) Fig. 3.3. Vertical misalignment of capacitive beams viewed as (a) SEM image (b) diagram. The observed vertical misalignment is not as exaggerated as in the diagram but still significant enough to affect the tuning ratio. The previous designs had two other problems, both associated with routing the RF signal through the electrothermal actuator. First, since the effective RF circuit and thermal circuit shared the same metal path, the transient power consumption of the electrothermal actuators could not be decreased without sacrificing Q. Increasing the length of the metal path to increase thermal isolation would result in increased electrical resistance and decreased Q. Second, the Chapter 3: Gap-tuning Capacitor 26 actuators contained separate routing for actuation control signals, which are low frequency voltages that act as high frequency ground. The close proximity between the RF routing and the actuation control routing caused large parasitic capacitances to ground (on the order of 100 fF) in the equivalent circuit. Mechanical latches were included in the previous designs to hold the variable capacitance constant without consuming power for heating (see Fig. 3.4). All latch designs used the selfassembly displacement of electrothermal actuators to close a gap at room temperature. Initial designs were intended to hold part of the capacitor in place with static friction [34] (see Fig. 3.4(a) and Fig. 3.4(b)). All such latches failed because friction was not strong enough compared with mechanical forces applied by the lateral actuators. The first successful latch pushed a peg into a slot connected to a moveable set of capacitive beams [27] (see Fig. 3.4(c)). This peg-andslot latch could accommodate multiple positions using multiple slots, but the number of slots was limited by the width of the gap required to accommodate a minimum width peg. (b) (a) st nd (c) rd Fig. 3.4. SEM image of the (a) 1 (b) 2 (c) 3 design of the mechanical latch for gap-tuning capacitors. The aforementioned problems associated with the previous designs of CMOS-MEMS gaptuning capacitors served as motivation for the topology presented in this work. The gap-tuning capacitor topology was updated to ensure vertical alignment between the capacitive beams, which was expected to increase the tuning ratio. The RF routing was separated from the actuation Chapter 3: Gap-tuning Capacitor 27 control routing to decrease parasitic capacitance and enable thermal isolation for lower transient power consumption. The latch was modified to provide a greater number of positions for a given capacitive gap. Latch actuator Latch actuation self-assembly RF interconnect y z La x Gap actuator actuation self-assembly Port 2 Port 1 Capacitive beams Thermal isolation / Open circuit Fig. 3.5. Diagram of gap-tuning capacitor. Arrows denote the directions of self-assembly and electrothermal actuation displacements for both the gap and latch actuators. Resistor symbols indicate the location of embedded polysilicon resistors used for electrothermal actuation. The specific numbers of capacitive, actuator, and interconnect beams are not representative of the numbers of beams used in the actual device. 3.2. Topology Fig. 3.5 shows the topology of the CMOS-MEMS gap-tuning capacitor. The majority of the area consists of interdigitated capacitive beams, whose lateral gap determines the series capacitance between Port 1 and Port 2. One set of capacitive beams is fixed and connected directly to a pad (Port 1). The other set of capacitive beams is movable and electrically connected to a second pad (Port 2) through the RF interconnect. The position of the movable capacitive beams is controlled by the displacement of the gap actuator. The latch actuator controls the position of the latch, which can be engaged to hold the movable capacitive beams in place without consuming static power for heating. Polysilicon resistors are embedded in the actuators Chapter 3: Gap-tuning Capacitor 28 to control the temperature and resulting displacement. The gap-tuning capacitor is operated by applying voltages to those resistors. Thermal isolation units are included to thermally isolate the actuators from both their anchors and the capacitive beams. The structure of the CMOS-MEMS gap-tuning capacitor was designed to ensure vertical alignment between the capacitive beams. Because all CMOS-MEMS beams curl vertically at room temperature, the capacitive beams will be vertically misaligned if those beams are anchored at opposite ends of the etch pit (see Fig. 3.6). Any vertical misalignment reduces the maximum capacitance, which negatively impacts both the tuning ratio and capacitance density. To avoid misalignment, the gap actuator and RF interconnect are ―U‖-shaped so that both the stator and rotor capacitive beams share a common anchor location. The ―U‖-shaped structures still curl vertically due to residual stress, but one section of the ―U‖-shape curls upwards whereas the other section curls downwards, such that the two tips of the ―U‖-shape are vertically aligned. Top View Side View Anchored at opposite ends Misaligned beams Common anchor location Aligned beams Fig. 3.6. Vertical curling of CMOS-MEMS beams. Anchoring MEMS beams at a common location prevents vertical misalignment. The RF interconnect is necessary to electrically isolate the RF signal from the actuation control signals, which results in low parasitic capacitance to ground. This interconnect Chapter 3: Gap-tuning Capacitor 29 contributes to the series resistance of the gap-tuning capacitor and acts as a mechanical load for the gap actuator. The ideal RF interconnect would have low series resistance for high Q and a low spring constant to enable large displacements. However, low series resistance requires short and wide structures whereas long and narrow structures are required for a low spring constant. To compromise between these two extremes, the RF interconnect consists of many narrow beams connected in parallel to provide a low-resistance yet mechanically flexible electrical connection. These beams appear similar to those of the electrothermal actuator but have different crosssections. The RF interconnect beams are not as wide and do not contain offset metal layers. The electrothermal actuators provide a self-assembly displacement which causes the latch to remain engaged without heating the actuators (see Fig. 3.5). The offset metal layers in each of the actuators are oriented such that the gap actuator self-assembles in the positive x-direction and the latch actuator self-assembles in the negative y-direction. These self-assembly displacements cause the latch to act as a limit-stop which prevents the gap actuator from pushing the capacitive beams into mechanical contact. Heating the actuators produces electrothermal actuation in the opposite direction of self-assembly, which is necessary to reconfigure the gap-tuning capacitor. To change the capacitance, the latch actuator is first heated to disengage the latch, which allows the gap actuator to move freely. Then, the gap actuator is heated to select the desired capacitive gap. Finally, removing power from the latch and gap actuators, in turn, allows the latch to close such that the capacitor is held at a new position without consuming static power for heating. The latch consists of interlocking stair-step structures with a 2 µm pitch to provide 6 positions across the 10 µm air gap (see Fig. 3.7). The latch actuator is composed of two actuators side-by-side so that latch position is not significantly affected by x-direction forces applied by the gap actuator. The choice to include two identical stair-step structures was arbitrary, and in Chapter 3: Gap-tuning Capacitor 30 hindsight, seems unnecessary. 5 µm Fig. 3.7. SEM image of the mechanical latch for the gap-tuning capacitor. As pictures, the latch is engaged in an intermediate capacitance position, neither minimum nor maximum capacitance. Each electrothermal actuator consists of many actuator beams with three heating resistors for temperature control (see Fig. 3.8). Multiple actuator beams are connected in parallel to supply the relatively large force necessary to push the RF interconnect. Each resistor was initially wired to a separate control voltage so that the thermal circuit could be experimentally characterized by using the resistors simultaneously as both heating elements and temperature sensors. 50 µm Rb Actuator beams Heating resistors Rc Ra Thermal isolation Fig. 3.8. SEM image of the gap actuator for the gap-tuning capacitor. The shading of the actuator beams changes half-way along their length because the offset metal layers are switched to the other side of the beams. Thermal isolation units are placed at actuator anchor points to lower the power consumption of the actuators (see Fig. 3.9). The actuator for gap control is also thermally isolated from the Chapter 3: Gap-tuning Capacitor 31 capacitive beams to prevent those beams from heating significantly. Because the thermal conductivity of aluminum (250 W/m/K) is much greater than that of oxide (≈ 1 W/m/K), the length and width of the metal path determines the thermal resistance of a CMOS-MEMS structure. A CMOS-MEMS structure without a continuous metal path would maximize thermal resistance. However, a metal path is required since the heating resistors must be electrically connected. The thermal isolation units, similar to those used in other thermal CMOS-MEMS devices [22], increase the thermal resistance by lengthening the metal path available for heat conduction. To create a long metal path while minimizing area, the metal is meandered between metal 1 and metal 2 and passed through a series of vias. The vias help to increase thermal resistance since the thermal conductivity of tungsten (170 W/m/K) is lower than that of aluminum. B Resistor A 5 µm A Key: Metal 4 B Metal 1 Metal 2 Metal 3 Oxide Via Silicon (a) (b) Fig. 3.9. Thermal isolation unit viewed as (a) SEM image (b) cross-section diagram. The metal 3 and metal 4 pictured in the diagram are not floating; these metals are connected to ground along a different cross-section. The gap-tuning capacitor was designed in a 4-metal 0.35 µm BiCMOS process from TowerJazz Semiconductor. Fig. 3.10 shows the resulting cross-sections of several different beams used in the gap-tuning capacitor. All four metals are aluminum. The capacitive beams are metal-4321 beams with the maximum density of vias to both maximize the capacitive overlap Chapter 3: Gap-tuning Capacitor 32 area and minimize series resistance. Since metal-4 is a thick top metal requiring larger widths, the interconnect beams and actuator beams are metal-321 beams. The smaller width and spacing of metal-321 beams are necessary for achieving large mechanical displacements. Metal-2 and metal-1 are offset within the actuator beams to form lateral electrothermal actuators as described in Section 2.2.2. A metal-4 ground plane serves as the perimeter for the MEMS structure. Interconnect Beams Capacitive Beams Actuator Beams ≈ 10 µm ≈ 50 µm Key: Metal 1 Metal 2 Metal 3 Oxide Via Silicon Metal 4 Fig. 3.10. Cross-section of CMOS-MEMS beams used in the electrothermal gap-tuning capacitor. 3.3. Analysis Assuming a parallel plate capacitance, the series capacitance of the gap-tuning capacitor, Cs, is defined by the maximum gap, gmax, and the displacement of the movable set of capacitive beams, x. 1 1 , C s N c 0 Lc h g max x g max x (3.1) where Nc, Lc, and h are the number, overlap length, and height of the capacitive beams, respectively. The minimum capacitance occurs for zero displacement. C s,min 2 N c 0 Lc h , g max Chapter 3: Gap-tuning Capacitor (3.2) 33 gmax+x gmax-x Capacitive beams have height h out of the page Lc Nc=2 for this example Fig. 3.11. Diagram of capacitive beams for displacement x. The maximum capacitance occurs when the movable capacitive beams are pushed into mechanical contact with the fixed capacitive beams. The maximum capacitance per unit length for two CMOS-MEMS beams pushed into mechanical contact is determined by the thickness of the side-wall polymer deposited during the CMOS-MEMS process. Capacitive test structures were used to calculate an effective minimum air gap, which can be used in conjunction with the height of the metal-dielectric stack to calculate the maximum capacitance. This effective air gap is a non-physical parameter which accounts for the thickness of the side-wall polymer, the effects of fringing fields, and the possibility that the permittivity of the side-wall polymer is greater than the permittivity of free space. The maximum capacitance can be calculated using the effective minimum air gap, gmin. 1 1 Cs N c 0 Lc h 2g max g min g min . (3.3) The expression for maximum capacitance can be simplified assuming that the minimum gap is much less than the maximum gap. Chapter 3: Gap-tuning Capacitor 34 Cs,max N c 0 Lc h . g min (3.4) The tuning ratio is calculated using the (3.2) and (3.4). C s,max C s,min g max 2g min (3.5) 4 µm 4 µm Minimum Capacitance Min . Maximum Capacitance Electric Field Magnitude Max Fig. 3.12. Electric field magnitude for cross-sections of capacitive beams in both minimum and maximum capacitance positions. The white, unmeshed areas represent the discrete metal layers of the CMOS-MEMS beams. The capacitance calculations above can be refined using finite element analysis (FEA) to account for fringing fields. A 2-D electrostatic analysis was used to calculate the capacitance per unit length between two CMOS-MEMS beams as the gap was varied (see Fig. 3.12). The permittivity of the side-wall polymer was assumed to be equal to the permittivity of free space. For large gaps, the capacitance is increased by fringing fields above and below the capacitive beams. For small gaps, the capacitance is decreased because electric fields become concentrated between the discrete metal layers that form only part of the capacitive beams. Given a 10 µm maximum gap and a 0.25 µm minimum gap, the tuning ratio calculated with FEA is approximately half as large as that calculated with (3.5). Chapter 3: Gap-tuning Capacitor 35 The 2-D electrostatic analysis was used to vary the vertical misalignment of the capacitive beams. Capacitance per unit length was simulated for several different gaps as the difference in vertical position between the fixed and movable capacitance beams was increased (see Fig. 3.13). The capacitance for small gaps is much more sensitive to vertical misalignment than that of large gaps, which underscores the need to minimize this misalignment. Capacitance / Length (aF/µm) 250 10 µm gap 2 µm gap 200 1 µm gap 0.3 µm gap 150 100 50 0 0 2 4 6 8 10 Vertical Misalignment (µm) Fig. 3.13. Capacitance per unit length as a function of vertical offset for different gaps. Vertical misalignment between capacitive beams has a much larger impact when the gap is small. The electrothermal actuator provides a displacement according to its geometry and the temperature of the actuator beams. The position, or tip displacement, of an unloaded actuator, xunloaded, can be determined by the length of the actuator beams. x unloaded 2S a La 2 T0 Ta , (3.6) where Sa is the sensitivity of the actuator beam cross-section in m-1K-1 and La, Ta, and T0 are the length, actuator beam temperature, and zero-stress temperature (≈ 180°C for this work) defined by the fabrication process, respectively. The unloaded displacement evaluates to zero when the actuator beam temperature is equal to the zero-stress temperature. A factor of two is included because the actuator consists of two sets of guided-end beams folded in series (see Fig. 3.5). The Chapter 3: Gap-tuning Capacitor 36 sensitivity can be calculated from the geometry of the actuator beam cross-section using the equations presented in [22]. The actuator must push the beam flexure comprising the RF interconnect, so the loaded displacement, x, includes a term based on the relative spring constants of the actuator and interconnect, ka and ki, respectively. Furthermore, the temperature of the actuator depends on both the ambient temperature, Tr, and the power dissipated as heat in the polysilicon resistors, P. Incorporating these terms into (3.6) yields an expression for the loaded displacement. x ka ka ki 2S a La 2 T0 Tr K th P , (3.7) where Kth is the effective thermal resistance observed as heat is dissipated in the three resistor locations. Note that the loaded displacement is greater than zero unless the actuator is heated to a temperature greater than the zero-stress temperature. By setting the displacement equal to the maximum gap and the actuator power to zero, (3.7) is used to calculate the maximum ambient temperature, Tr,max, at which the actuator pushes the capacitive beams into mechanical contact. Tr,max T0 k a k i g max k a 2 S a La 2 , (3.8) In addition to a requirement on ambient temperature, the voltage across the capacitive beams must be sufficiently small such that the beams are not clamped together by electrostatic force. The voltage V across the capacitor beams creates an electrostatic force, Fe, pulling the beams together. Fe V 2 dC s , 2 dx (3.9) Using the approximation in (3.4) and evaluating x at gmin results in a simple expression for the electrostatic force. Chapter 3: Gap-tuning Capacitor 37 Fe V 2 C s,max 2 g min . (3.10) The mechanical restoring force of the gap actuator under normal operation, Fm, is the force applied to move the capacitive beams from maximum to minimum capacitance position. Fm k a k i g max g min , (3.11) The voltage at which the gap actuator can no longer separate the capacitive beams occurs when the electrostatic force equally opposes the mechanical restoring force. Combining (3.10) and (3.11) yields the maximum applied voltage, Vmax. Vmax 2 g min k a k i g max g min . C s,max (3.12) If the frequency of the applied voltage is much greater than the mechanical resonance frequency of the gap actuator, then the electrostatic force is determined by the root-mean-squared (RMS) value of the voltage. This voltage is what ultimately limits the power handling of the gap-tuning capacitor. However, an explicit number representing the maximum power cannot be calculated without knowing how the gap-tuning capacitor is to be used in a circuit. The quality factor of the series capacitance depends on the series resistance, Rs, which is primarily determined by the resistance of the RF interconnect. Q 1 , 2fR s C s (3.13) where f is the frequency of operation. The series resistance can be calculated using the sheet resistances of the metal layers in combination with the geometry of the design. 3.4. Design Measurements from several test structures were used during design of the electrothermal Chapter 3: Gap-tuning Capacitor 38 gap-tuning capacitor. The optimum actuator beam cross-section was determined using measurements of the self-assembly displacement of electrothermal actuators. An actuator beam consisting of a 1-µm metal-3 top metal with 0.5-µm metal-2 and metal-1 offset layers produces the maximum displacement for a given actuator beam length, La. The corresponding actuator beam cross-section sensitivity, Sa, and the zero-stress temperature, T0, were derived by measuring the displacement with the actuator heated to a known temperature. Based on capacitance measurements of previous CMOS-MEMS gap-tuning capacitors, the minimum effective air gap, gmin, was estimated to be 0.3 µm. The maximum gap, gmax, was chosen to be 10 µm to increase the tuning ratio beyond that of previous electrothermal gap-tuning capacitors. Choosing the maximum gap determined the tuning ratio according to (3.5) in conjunction with FEA results. The number and dimensions of the capacitive beams were selected to create a maximum capacitance of approximately 400 fF. Since capacitive beams that are too long and too narrow can curl laterally resulting in decreased maximum capacitance, the beam length and width were set to 220 µm and 4 µm, respectively. The lateral curling resulting from these dimensions was expected to be minimal based on qualitative observations of test structures. To achieve the desired maximum capacitance, the number of capacitive beams, Nc, was set to 8 according to (3.4). These parameters define the structure of the capacitive beams. The gap actuator was designed in tandem with the RF interconnect to provide the displacement necessary to traverse the maximum gap. The width of the interconnect beams, wi, was set to 0.5 µm, which is the minimum metal width for the 0.35 µm BiCMOS process, so that the spring constant of the actuator beams, ka, would be larger than that of the interconnect beams, ki. The length of the interconnect beams, Li, was set equal to that of the actuator beams, La, so Chapter 3: Gap-tuning Capacitor 39 that the resulting device would fit in a rectangular area. Equal amounts of area were devoted to the actuator beams and interconnect beams, which, in conjunction with the beam cross-sections, determined the ratio of the actuator beams spring constant to that of the interconnect beams. The actuator beam length was scaled to 210 µm to increase the displacement according to (3.7). Using (3.8), the resulting maximum ambient temperature, Tr,max, was calculated to be 86°C. That maximum temperature was deemed sufficient since many consumer electronics must operate at ambient temperatures up to 70°C. Table 3.1: Gap-Tuning Capacitor Design Parameters Geometry Value Constants Value gmax 10 µm gmin* 0.3 µm Lc 220 µm Sa** 1.61 m-1K-1 wc 4 µm T0** 176°C Nc 8 La 210 µm Calculation Value wa 1 µm Cs,min 52.9 fF wo 0.5 µm Cs,max 399 fF Na 14 Tuning Ratio 7.5:1 Li 210 µm ka 0.50 N/m wi 0.5 µm ki 0.14 N/m Ni 18 Tr,max 86°C Vmax 3.1 V Q at 1 GHz 210 * Estimation based on measured data ** Parameter extracted from measured data The number of actuator beams, Na, and number of interconnect beams, Ni, were scaled to 14 and 18, respectively, to increase the Q. Increasing the number of interconnect beams increases the width of metal available for conduction which lowers the series resistance, Rs. Based on the sheet resistances provided by the 0.35 µm BiCMOS design manual, the series resistance was calculated to be 1.9 Ω. Using (3.13), the minimum Q for the maximum capacitance, Cs,max, was calculated to be 210 at 1 GHz. Multiple design iterations were necessary to reach an area-efficient design with sufficient Q Chapter 3: Gap-tuning Capacitor 40 and tuning ratio. The air gaps between the MEMS structure and the metal-4 ground plane were set to 20 µm to minimize parasitic capacitance to ground. Table 3.1 lists all the parameters used in the design of the CMOS-MEMS electrothermal gap-tuning capacitor presented in this work. Fig. 3.14 shows an SEM image of the fabricated device. RF interconnect Latch actuator Port 2 Port 1 Latch Capacitive beams 200 µm Gap actuator Fig. 3.14. SEM image of CMOS-MEMS variable capacitor. 3.5. Characterization All fabricated electrothermal gap-tuning capacitors correctly self-assembled to close the 10 µm capacitive gap at room temperature. The gap-tuning capacitors were actuated using no more than 3 V and 15 mW (see Fig. 3.15). The mechanical latch was successful at maintaining each of the six latch positions without consuming power (see Fig. 3.16). While all the capacitors successfully demonstrated basic functionality, effective minimum air gap varied from device to device due to unintended lateral curling (see Fig. 3.17). Chapter 3: Gap-tuning Capacitor 41 (a) (b) Fig. 3.15. Microscope image of the gap-tuning capacitor in (a) maximum capacitance position (b) minimum capacitance position. The latch appears retracted because voltage was applied to the heating resistors in the latch actuator. (a) (b) (c) (d) (e) (f) Fig. 3.16. Microscope image of the latch at position (a) 1 (b) 2 (c) 3 (d) 4 (e) 5 (f) 6. Each position corresponds to a different capacitance value at which the variable capacitor can be maintained without consuming power. (a) (b) Fig. 3.17. Microscope image of capacitive beams pushed into mechanical contact (a) for the best-performing capacitor (b) for a capacitor with significant lateral curling. The amount of curling is small but still has significant effects of the minimum effective air gap. Chapter 3: Gap-tuning Capacitor 42 The vertical profile of the variable capacitor was measured with a Veeco optical profilometer (see Fig. 3.18). The interdigitated capacitive beams curl out-of-plane but maintain vertical alignment. The worst-case misalignment, 1.6 µm, occurs at the tips of the movable capacitive beams, which, according to 2D cross-section FEA simulations (see Fig. 3.13), results in a 13% reduction in tuning ratio if that misalignment is constant over the length of the beams. However, misalignment varies along the length of the capacitive beams, so the tuning ratio is reduced by less than 13% due to vertical misalignment. 13.2 5.0 0.0 -5.0 Vertical Deflection (µm) 10.0 -8.1 Fig. 3.18. Optical profilometer image of electrothermal gap-tuning capacitor. The capacitive beams are vertically aligned to within 1.6 µm. 3.5.1. Electrical Performance The digital capacitors were electrically characterized using the method described in Appendix A for all six latch positions. Fig. 3.19 shows the S-parameters for the minimum capacitance and maximum capacitance positions. Fig. 3.20 shows the equivalent circuit model in terms of Y-parameters and lumped circuit elements. Fig. 3.21 shows the lumped impedances for Chapter 3: Gap-tuning Capacitor 43 the minimum capacitance and maximum capacitance positions. The equations listed below were 10 10 0 0 S-parameters (dB) S-parameters (dB) used to calculate the equivalent circuit element values (see Fig. 3.20(b)). S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 -10 -20 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -30 -40 -40 0.1 1 0.1 10 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 3.19. S-parameter magnitude as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. S12 is not plotted because S12 and S21 are equivalent for passive networks. Port 1 Port 2 -Y21 Y11+Y12 Port 1 Rs Cs Ls Port 2 C2g Y22+Y21 1E+5 1E+5 1E+4 1E+4 Impedance (Ω) Impedance (Ω) (b) (a) Fig. 3.20. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 0.1 1 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 10 0.1 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 3.21. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 20 GHz for the minimum capacitance position and decreases to 11 GHz for the maximum capacitance position. Chapter 3: Gap-tuning Capacitor 44 2 1 res . Cs imag 1 Y21 Ls 1 resCs 2 (3.14) . (3.15) Rs real 1 Y21 . C1g C 2g 1 * imag 1 Y11 Y21 1 * imag 1 Y22 Y21 Q imag 1 Y11 real1 Y11 (3.16) . (3.17) . (3.18) . Table 3.2: Electrical Characteristics of the Gap-tuning Capacitor Latch Cs (fF) Ls (nH) Rs (Ω) C1g (fF) C2g (fF) Position ± 1.1% ± 6.4% ± 20% ± 9.2% ± 5.7% 1 54.5 NA* 13.4 13.1 30.3 2 56.1 NA* 13.3 13.1 29.7 3 61.2 NA* 12.2 13.0 28.6 4 73.4 NA* 10.8 12.9 27.4 5 111 569 8.15 12.6 26.4 6 373 536 4.56 10.9 25.1 * Inductance value could not be extracted since the self-resonance frequency exceeded 20 GHz (3.19) Q (1 GHz) ± 22% 71 71 71 72 80 80 The electrical characteristics of the best-performing electrothermal gap-tuning capacitor are presented in Table 3.2. The latch position 1 is the minimum capacitance position and latch position 6 is the maximum capacitance position. Measurement uncertainties were derived from multiple measurements of the same latch position and thus do not include systematic offsets based on the particular network analyzer calibration or probe contact resistance. As designed, the Chapter 3: Gap-tuning Capacitor 45 series capacitance, Cs, increases with latch position as the capacitive gap decreases. The shunt capacitances, C1g and C2g, decrease slightly with latch position due to fringing electric fields originating from the bottom of the capacitive beams. As the capacitive gap decreases, more of the fringing electric field lines will terminate on the capacitive electrodes rather than the silicon substrate resulting in decreased capacitance to ground. Note that the measured series resistance reported in Table 3.2 changes according to the latch position. This change in resistance was unexpected since the metal path that determines the series resistance is not altered by the latch position. Even after considering the skin effect, the proximity effect, parasitic capacitance to the silicon substrate, and the possibility of piezoresistance, such a significant change in resistance could not be explained. Only after subsequent measurements of the digital capacitor, which is presented in Chapter 4, was the cause determined to be the noise floor of the VNA. The noise floor prevented accurate measurement of a small real impedance in series with a large imaginary impedance. Accurate measurements of series resistance were achieved after lowering the IF bandwidth of the VNA from the default setting of 35 kHz to 500 Hz. Since the noise floor is directly proportional to the IF bandwidth, this modification decreased the noise floor by nearly two orders of magnitude. However, since many of the gap-tuning capacitors were damaged during reliability testing, updated measurements are not available. The series resistance presumably remains constant regardless of the latch position although this has not been verified through measurement. Since the measured Q is calculated using the measured series resistance, the Q was also limited by the noise floor of the VNA and thus does not agree with calculated predictions. Assuming a constant series resistance, the Q should decrease as the capacitance increases. However, the Q reported in Table 3.2, which is the Q with port 2 connected to ground, remains Chapter 3: Gap-tuning Capacitor 46 approximately constant for different latch positions. The measured Q for the maximum capacitance is 80, which is significantly lower than the calculated value of 210. The measured Q as a function of frequency still fits well with the trend according to (3.19) (see Fig. 3.22) but Quality Factor should be considered only a rough approximation. 100 90 80 70 60 50 40 30 20 10 0 Quality Factor (De-embedded) Extracted Model 1 2 3 4 5 6 Frequency (GHz) Fig. 3.22. Quality factor as a function of frequency. This plot shows the Q for the maximum capacitance position. The series capacitance of the electrothermal gap-tuning capacitor was measured at each of the 6 latch positions for 7 different devices (see Fig. 3.23). The distribution of capacitance values is nonlinear because the latch positions are evenly spaced over the 10 µm capacitive gap. The average gap-tuning capacitor tuned from 54.9 fF to 269 fF, an average tuning ratio of 4.9:1. The best-performing gap-tuning capacitor tuned from 54.5 fF to 373 fF, a tuning ratio of 6.8:1. The variation between devices, measured by standard deviation, differs greatly depending on the latch position. The minimum capacitance ranged from 54.0 fF to 55.3 fF while maximum capacitance ranged from 214 fF to 373 fF. The standard deviation for the minimum capacitance is only 1.0%, but the standard deviation from the maximum capacitance is 18%. The average standard deviation for all states is 4.9%. Chapter 3: Gap-tuning Capacitor 47 Series Capacitance (fF) 400 350 300 250 200 150 100 50 0 1 2 3 4 5 6 Latch Position Fig. 3.23. Plot of series capacitance versus latch position. Error bars indicate the 1-σ standard deviation between different devices based on 7 measurements. 400 Data Points Latch Positions Capacitance (fF) 350 300 250 200 150 100 50 0 0 1 2 3 4 Actuator Power (mW) Fig. 3.24. Capacitance as a function of actuator voltage. Continuous tuning is possible with the latch retracted. Hysteresis is caused by capacitive beams sticking together as they are being pulled apart. Series capacitance was measured as the power applied to the heating resistors was varied while the latch was retracted. The resulting data shows a smooth transition from maximum capacitance to minimum capacitance as the input power is decreased, or as the gap is closed (see Fig. 3.24). However, hysteresis occurs when increasing the actuator power which suggests there is some adhesion force between the capacitive beams, possibly due to humidity in the laboratory Chapter 3: Gap-tuning Capacitor 48 environment. This adhesion force was always overcome by the actuator, so the capacitive beams did not become permanently stuck. A bias-T was included to measure the series capacitance as a function of the DC voltage applied across the capacitive beams. While the capacitor was unlatched, electrostatic pull-in occurred with less than 5 V applied (see Fig. 3.25). At maximum capacitance, electrostatic force flattened the capacitive beams causing the capacitance to increase by up to 14%. While the capacitor was latched with a 2 µm air gap, which is the latch position most vulnerable to electrostatic pull-in, up to 14 V could be applied before pull-in was observed. All other latched capacitance positions were maintained for DC voltages up to 20 V, which is expected due to their larger air gaps. 500 Capacitance (fF) 450 400 Position 4 350 Position 5 300 Position 6 250 Unlatched 200 150 100 50 0 0 5 10 15 20 DC Bias (V) Fig. 3.25. Capacitance as function of DC bias voltage. With the latch retracted (unlatched), capacitance beams snap together using less than 5 V. With the latch in place, up to 14 V can be applied before electrostatic pull-in occurs. The electrothermal gap-tuning capacitor was tested over a range of RF power to determine the maximum root-mean-squared (RMS) voltage at which the actuator could overcome the electrostatic force between the capacitive beams. Port 1 of the capacitor was connected to an Agilent E8241 signal generator while port 2 was connected to an Agilent E4440 spectrum Chapter 3: Gap-tuning Capacitor 49 analyzer. The latch was retracted with the capacitor placed in maximum capacitance position. While testing with a frequency of 4 GHz, the limit at which the gap actuator would no longer retract occurred when 16 dBm was delivered to the load, which amounts to 4.2 VRMS applied across the capacitive beams. This measured voltage is greater than the calculated maximum voltage of 3.1 VRMS, likely because the maximum capacitance was reduced to due unintended lateral curling. 3.5.2. Switching Speed A computer microvision system was used to measure the frequency response of the electrothermal actuation. A sinusoidal actuation voltage was applied to the heating resistors within the electrothermal actuators in conjunction with a DC bias necessary to prevent the capacitive gap from closing entirely. Because mechanical displacement is proportional to the square of the voltage, the raw displacement data includes a double-frequency component. However, this double-frequency component is filtered such that the resulting response represents the frequency-dependant conversion from electrical power to mechanical displacement (see Fig. 3.26). The thermal 3 dB frequency occurs at 60 ± 5 Hz while mechanical resonance occurs at 5.1 ± 0.1 kHz and 13.7 ± 0.1 kHz for the gap and latch actuators, respectively. Thus the switching speed is limited by the thermal time constant (≈ 2.7 ms) rather than the mechanical response. The microvision system was also used to measure the transient response of the electrothermal actuation (see Fig. 3.25). The frequency of actuation was set to 50 Hz, and the displacement was recorded at each of 100 phase points. The measured 90% rise time was 4.9 ± 0.1 ms. Note that displacement is not observed until approximately 1 ms after power has been applied to the actuator. This delay occurs because, without applied power, the capacitive Chapter 3: Gap-tuning Capacitor 50 beams are pushed into mechanical contact, and the contact force must be relieved by a change in Displacement Amplitude(µm) temperature before displacement can occur. 10 Gap Actuator Latch Actuator 1 0.1 0.01 0.001 Thermal 3dB frequency ≈ 60 Hz Mechanical resonance frequency ≈ 5.1 kHz and ≈ 13.7 kHz 0.01 0.1 1 10 100 Frequency (kHz) Fig. 3.26. Electrothermal actuator frequency response. The thermal 3 dB frequency limits the speed of both the gap and latch actuators. 16 6 Position Power 5 12 4 10 8 3 6 2 4 Power (mW) Displacement (µm) 14 1 2 0 0 0 5 10 15 20 Time (ms) Fig. 3.27. Electrothermal actuator transient response. The 90% rise time is 4.9 ms. Reconfiguration of the gap-tuning capacitor requires many thermal time constants because three successive temperature changes are needed to change the latched capacitance. First, the latch and gap actuators must be heated to disengage the latch and set the capacitive gap, respectively. Second, the gap actuator must cool to engage the latch. Third, the latch actuator Chapter 3: Gap-tuning Capacitor 51 must cool to settle into the latch position. Each of these temperature changes can take multiple thermal time constants depending on how much power is used to heat the actuators. Reliable operation of the variable capacitor has been demonstrated using 30 ms for reconfiguration, a relatively conservative actuation method which provides 3.7 thermal time constants for each temperature change. The times required for heating could be decreased by using a different voltage pulse or by using a closed-loop control system. However, the times required for cooling cannot be decreased and will always be limited by the thermal time constant. 3.5.3. Thermal Aspects The electrothermal actuators controlling the gap-tuning capacitors contain embedded polysilicon resistors at three locations (see Fig. 3.5). In a prototype of the variable capacitor, each of these resistors was wired to a separate control voltage, such that the resistors could act as both heating elements and as temperature sensors to characterize the thermal circuit of the actuator. First, power was dissipated in Ra while observing the change in resistance for all three resistors, Ra, Rb, and Rc. Using the temperature coefficient of resistivity, the changes in each resistance were used to calculate the temperature change at each resistor location. Based on the temperature changes, three thermal resistances were calculated using the power dissipated in Ra. Repeating the process with power dissipated in Rb and then Rc yielded the thermal resistance matrix, Ta K aa T K b ba Tc K ca K ab K bb K cb K ac Pa K bc Pb , K cc Pc (3.20) ΔTx and Px represent the temperature change and power dissipated at location x, respectively. Kxy is the temperature change at location x when unit power is dissipated at location y. A uniform temperature at all three resistor locations was necessary to remain consistent with Chapter 3: Gap-tuning Capacitor 52 previous modeling and characterization of electrothermal actuators and to maximize displacement for a given maximum temperature. The thermal resistance matrix was used to calculate resistor values such that a single control voltage, Vh, applied to the resistors in parallel would create a uniform temperature. 1 Ra 1 R 1 b 2 1 Rc Vh K aa K ba K ca K ab K bb K cb 1 K ac K bc K cc Ta T . b Tc (3.21) Since the electrothermal actuators retract completely at approximately 180°C, all three ΔT‘s were set to 155°C, sufficient temperature change given an ambient temperature of 25°C. Voltage Vc was set to 3 V, slightly below the specified 3.3 V supply of the 0.35 µm BiCMOS process. For the gap actuator, (3.21) computes Ra, Rb, and Rc, to be 3.4 kΩ, 3.7 kΩ, and 4.8 kΩ, respectively. 1 3.4k 1 3.7k 1 2 1 4.8k 3V 46.9KW -1 9.1KW -1 5.2KW -1 -1 45.2KW -1 11.6KW -1 8.5KW 5.0KW -1 11.7KW -1 60.6KW -1 1 155K 155K . 155K (3.22) For the latch actuator, (3.21) computes Ra, Rb, and Rc, to be 1.8 kΩ, 1.9 kΩ, and 2.7 kΩ, respectively. 1 1.8k 1 1.9k 1 2 1 2.7k 3V 25.5KW -1 -1 3.0KW 2.1KW -1 3.4KW -1 24.8KW -1 5.6KW -1 1 2.5KW -1 155K 5.8KW -1 155K . 35.0KW -1 155K (3.23) Both the electrical resistances and thermal resistances for the latch actuator are approximately half of those of the gap actuator. This difference is expected because the latch actuator consists of two actuators with a total of six heating resistors rather than three. Note that, for both the gap and latch actuator, Ra is the smallest resistance and Rc is the largest resistance. The relative sizes of the resistances was expected since Ra is near the anchor, which requires the most power to heat, and Rc is near the capacitive beams, which requires the least power to heat (see Fig. 3.5). Chapter 3: Gap-tuning Capacitor 53 The electrothermal gap-tuning capacitor was later updated to include the resistor values calculated using (3.21). A Quantum Focus infrared microscope was used to visualize the temperature profile while the actuators were heated (see Fig. 3.28). During this measurement, the substrate was heated to 30°C to provide a known reference temperature. The actuators were heated to low temperatures (< 50°C) to avoid mechanical displacements which would invalidate the reference measurement recorded before heating the actuators. The infrared microscope measures the emitted infrared energy and calculates the temperature assuming a constant emissivity, which is derived from the reference measurement. The calculated temperature should be considered only a rough estimate because the emissivity of aluminum varies with temperature. Nonetheless, the measured temperature profiles do show that the actuators are effectively thermally isolated from their anchor points and the capacitive beams. The actuator temperatures are higher when both the gap and latch actuators are heated, which suggests thermal cross-talk between the actuators. This thermal cross-talk could be further analyzed using the resistors as temperature sensors. 50oC 45oC 40oC 35oC 30oC (a) (b) (c) Fig. 3.28. Temperature profile of gap-tuning capacitor viewed with quantum infrared microscope. Voltage is applied to resistors in (a) the gap actuator only (b) the latch actuator only (c) both the gap and latch actuators. The electrothermal gap-tuning capacitor was tested over a range of temperatures to verify the maximum ambient temperature at which the actuators would maintain the desired latch position. Capacitance was measured while the probe station chuck temperature was varied from Chapter 3: Gap-tuning Capacitor 54 25°C to 100°C with the capacitor latched in a particular position (see Fig. 3.29). Up to 65°C, the capacitor maintains its capacitance to within 0.5% of the nominal (25°C) value. Beyond 65°C, two modes of failure were observed. At higher capacitance positions, the gap actuator retracts as the temperature increases causing a decrease in capacitance. At lower capacitance positions, the latch actuator retracts as temperature increases causing the device to slip into a higher capacitance position. Although this latch failure appears to occur at 95°C, the latch actuator is incapable of closing the latch for temperatures greater than 65°C. The fact that both actuators fail 400 120 350 110 300 Position 6 Position 5 100 250 Position 4 Position 3 90 Position 2 Position 1 200 80 150 70 100 60 50 50 25 50 75 Temperature ( oC) Capacitance (fF) Capacitance (fF) at the same temperature is not coincidental since both actuators are of similar design. 100 Fig. 3.29. Latched capacitance as a function of ambient temperature. While latched in position, the capacitance remains stable up to 65°C. 3.5.4. Reliability Tests Each capacitance position of the electrothermal gap-tuning capacitor was tested to verify that capacitance values were repeatable. The variable capacitor was un-latched, actuated to the specified gap, and re-latched ten times for each latch position for two different devices. Measured capacitance values varied less than 0.3% from the average capacitance for each particular latch Chapter 3: Gap-tuning Capacitor 55 position. A shaker table was used to test whether large accelerations would cause the electrothermal gap-tuning capacitor to slip into a different latch position. Measurements of the latched capacitance were recorded prior to testing. The shaker table was used to drive the CMOS-MEMS chip through sinusoidal motion at 50 Hz. The frequency was chosen to be significantly less than the mechanical resonant frequency so that any displacement due to the acceleration would reach steady-state equilibrium. The peak acceleration was ramped up to 40 g and maintained for approximately 10 seconds. The capacitance was then measured to see if a change had occurred. This test was repeated for both maximum and minimum capacitance positions with the applied acceleration in three orthogonal directions. As a result of 40 g of acceleration, the minimum and maximum capacitance varied by less than 0.1% and 1%, respectively. To assess reliability over time, two electrothermal gap-tuning capacitors were cycled at 15 Hz between the minimum and maximum capacitance latch positions over several days. During this time, the chuck temperature was held constant at 25°C with the DUT open to the general laboratory atmosphere. The full duration amounts to millions of impacts between the capacitive beams. Testing was stopped intermittently so that the capacitance could be measured. For both capacitors, a decrease in maximum capacitance was observed with increasing cycles (see Fig. 3.30). The first capacitor continued to operate beyond 5 million cycles with no catastrophic failures. The latch of the second capacitor failed after approximately 5 million cycles. While shifting to maximum capacitance position, the latch began sliding beneath the movable capacitive beams rather than making contact with the corresponding interface. An inspection of the defective latch revealed damage to the surface of the metal covering the heating resistors and sub-micron debris scattered around the latch interface (see Fig. 3.31). Whether the Chapter 3: Gap-tuning Capacitor 56 latch failure was caused by repetitive mechanical impacts or by prolonged heating is unclear. Capacitance (fF) 500 400 300 Cs,max Cs,max (DUT (DUT#1) #1) Cs,min (DUT Cs,min (DUT#1) #1) Cs,max (DUT Cs,max (DUT#2) #2) C (DUT #2) Cs,min s,min (DUT #2) 200 100 0 1000 103 10000 104 100000 105 1000000 106 10000000 107 Cycles Fig. 3.30. Minimum and maximum capacitance as a function of actuation cycles for two DUT‘s. Maximum capacitance decreased as the variable capacitor was cycled between latch positions. 5 µm 20 µm (a) (b) Fig. 3.31. SEM images of mechanical latch after failure (a) zoomed out (b) zoomed in. Repeated impacts of the latch interface created sub-micron debris on the latch. 3.6. Discussion The CMOS-MEMS electrothermal gap-tuning capacitor presented in this work has demonstrated significant advantages when compared with gap-tuning capacitors developed in Chapter 3: Gap-tuning Capacitor 57 previous work (see Table 3.3). Increasing the maximum gap from 5 µm to 10 µm increased the series tuning ratio from 2.5:1 to 6.9:1. The inclusion of the RF interconnect to isolate the RF signal from the actuation control signals resulted in significantly less parasitic capacitance to ground. The two parasitic capacitances, C1g and C2g, were decreased from 71 fF and 425 fF to 12.6 fF and 27.9 fF, respectively. The Q for the maximum capacitance at 1 GHz was increased from 54 to 80. The improved mechanical latch increased the number of latch positions from two to six. Compared with previous work, the only drawback of this work is a slight decrease in capacitance density, which is unavoidable since a larger maximum gap is required for larger tuning ratios. Table 3.3: Gap-Tuning Capacitor Results Name Previous design This work Tuning ratio (series) 2.5:1 6.9:1 Tuning ratio (shunt) 2.0:1 5.7:1 Q (1 GHz) for Cs,max* 54 80 Self-resonance frequency for Cs,max 11.3 GHz 11.3 GHz Cs,max 380 fF 373 fF Cs,min 150 fF 54.5 fF C1g** 71 fF 12.6 fF C2g** 425 fF 27.9 fF Rs for Cs,max 7.2 Ω 4.5 Ω Ls for Cs,max 522 pH 536 pH Area 0.19 mm2 0.26 mm2 Capacitance Density 2.0 pF/ mm2 1.4 pF/ mm2 Actuation Method electrothermal electrothermal Control 2 latch positions 6 latch positions Voltage 3V 3V Power NA**** 15 / 0 mW*** Switching Time NA**** 30 ms Cs,max Standard Deviation NA**** 18% Integration maskless post-CMOS etching maskless post-CMOS etching * Quality factor is calculated from a fit to measured data ranging from 1 GHz to 4 GHz ** Shunt capacitances are averaged over all capacitor states *** Power is only consumed during switching **** Value is not reported but is expected to be similar to that of this work The tests on reliability and repeatability show that an individual gap-tuning capacitor is Chapter 3: Gap-tuning Capacitor 58 reliable at producing repeatable capacitance values. However, a comparison between gap-tuning capacitors from different chips revealed significant device-to-device variation in the maximum capacitance (see Fig. 3.23). This variation is caused by lateral curling of the interdigitated capacitance beams, which is an undesirable aspect of the CMOS-MEMS process that is not fully understood. The long, narrow capacitance beams are prone to curling on the order of tenths of microns, which causes uneven mechanical contact when the beams are pushed together in maximum capacitance position. For the seven devices measured, the minimum capacitance ranged from 54.0 fF to 55.3 fF while the maximum capacitance ranged from 214 fF to 372 fF. This variation in maximum capacitance is problematic when implementing the gap-tuning capacitor in reconfigurable circuits with multiple LC tanks requiring the same resonance frequency. While the number of latch positions has been increased, the distribution of latched capacitance values is not practical. Because the latch positions are distributed linearly every 2 µm, the series capacitance varies in a nonlinear manner with respect to the latch position (see Fig. 3.23). A different distribution of latch positions could yield more useful capacitance values. For example, the spacing of the latch positions could be redesigned to create a linear variation in capacitance, or even a linear variation in the resonance frequency of an LC tank. However, the minimum size of the latch positions is limited by ion milling which occurs during the vertical oxide etch of the CMOS-MEMS process. This ion milling creates rounded corners which can prevent the latch from correctly engaging. Latches designed with a 1 µm step size slipped out of position, so the practical limit for the size of latch steps is somewhere between 1 µm and 2 µm. Thus, decreasing the number of latch positions would be necessary to modify the distribution of capacitance values. Chapter 3: Gap-tuning Capacitor 59 Several other issues limit the practical use of the gap-tuning capacitor. The capacitance density is very low compared with that of semiconductor variable capacitors and still low compared with that of other integrated MEMS variable capacitors. Increasing the capacitance density is possible but requires a smaller maximum gap resulting in reduced tuning ratio. The self resonance frequency is limited to approximately 10 GHz due to the need to loop the current through the RF interconnect to achieve vertical alignment between the capacitance beams. These problems could not be resolved using the gap-tuning capacitor topology and ultimately lead to the development of the CMOS-MEMS digital capacitor, which is presented in Chapter 4. Chapter 3: Gap-tuning Capacitor 60 Chapter 4. Digital Capacitor The CMOS-MEMS variable capacitors presented in this section are ―digital‖ because they consist of multiple independently controlled capacitive bits. The capacitive bits can be toggled on and off through a combination of vertical electrothermal and lateral electrostatic actuation. Each capacitive bit consists of many unit cells, which are referred to as CMOS-MEMS switched capacitors. Section 4.1 explains how the digital capacitor was developed to overcome the disadvantages of the gap-tuning capacitor. Section 4.2 presents the switched capacitor topology. Section 4.3 presents an analysis of the topology linking the degrees of freedom to performance metrics. Section 4.4 describes how that analysis was used to define the geometry of the switched capacitor. Section 4.5 presents characterization results of the fabricated design. Section 4.6 discusses the results and presents several ideas for further improvement. 4.1. Motivation The CMOS-MEMS digital capacitor was designed specifically to address the shortcomings of the gap-tuning capacitor. The primary goals of the design were to improve the tuning ratio, capacitance density, distribution of capacitance values, self-resonant frequency, and device-todevice variation. Secondary goals such as simplifying the design for subsequent modification by a circuit designer, increasing the switching speed, decreasing the power consumption, and increasing the quality factor were considered but were not the primary factors driving the design. To achieve these goals, a unique method of actuation was developed. 4.1.1. Concept The digital capacitor uses a combination of vertical electrothermal and lateral electrostatic Chapter 4: Digital Capacitor 61 actuation (see Fig. 4.1) to toggle the capacitance of a capacitive bit. The minimum capacitance is limited by the large vertical gap created by the self-assembly displacement of the electrothermal actuator. After vertical electrothermal actuation, lateral electrostatic actuation is used to snap the capacitive electrodes into mechanical contact, resulting in a large maximum capacitance. The resulting electrostatic force is used to hold the electrodes together without consuming power for electrothermal actuation. Side View Top View After release etch After vertical electrothermal actuation After lateral electrostatic actuation Electrothermal actuator Capacitive electrodes Fig. 4.1. Concept of the actuation method for the switched capacitor. Many of the benefits of this actuation method, which are experimentally quantified in Section 4.5, can be realized from a qualitative discussion. Vertical electrothermal actuation alleviates the need for large lateral air gaps, so large tuning ratios can be achieved simultaneously with large capacitance densities. Lateral electrostatic actuation applies large forces between the capacitive electrodes and thus does not require the CMOS-MEMS beams to be absolutely straight. Unintended lateral curling can be tolerated as long as that curling is not so large as to interfere with the vertical electrothermal actuation, so device-to-device variation is small. The unit cell necessary to enable actuation is small and can be repeated in parallel to form capacitive Chapter 4: Digital Capacitor 62 bits. The capacitive bits can be sized in a binary fashion (1C, 2C, 4C, etc.) to produce a linear distribution of capacitance values. Furthermore, because current flows straight through the structure, parasitic inductance should be minimized. 4.1.2. Related Work MEMS variable capacitors have been previously developed which use actuation schemes similar to that presented in this work. The MEMS variable capacitors in [8] and [37] use different methods of vertical electrostatic actuation to lower rotor electrodes into plane with stator electrodes, but no further actuation is used to close the remaining lateral gap. Similar to this work, the MEMS switch in [26] uses a combination of vertical electrothermal and vertical electrostatic actuation to achieve low-voltage actuation without static power consumption. However, the 3-bit digital capacitors presented in this work are the first MEMS capacitive devices to control both the gap and overlap area between capacitive electrodes using vertical and lateral actuation. Many MEMS digital capacitors, or devices which provide multiple discrete capacitance values, have also been developed. The choice to provide discrete tuning is often mandated by the bi-stable nature of the typical electrostatic gap-tuning capacitor [29]. Most digital capacitors are combinations of identical switched capacitors, often adapted from capacitive switches, which require multiple independent control signals [38], [39]. Alternatively, some digital capacitors consist of switched capacitors with varying electrostatic pull-in voltages such that only a single control signal is needed to vary the capacitance [40]. The digital capacitor presented in this work is similar to those adapted from capacitive switches in that an array of identical capacitors is controlled using multiple actuation signals. Chapter 4: Digital Capacitor 63 Key: Metal 1 Thermal Isolation Metal 2 Electrothermal Actuator Beam Metal 3 Metal 4 Rotor Electrodes Stator Electrodes Signal Beams (a) RDC VES VET RHeat Port 1 CMIM (b) Port 2 (c) Fig. 4.2. Diagram of CMOS-MEMS switched capacitor viewed from above (a) with annotations (b) with circuit schematic; (c) SEM image of four CMOS-MEMS switched capacitors in parallel. 4.2. Topology The CMOS-MEMS switched capacitor (see Fig. 4.2(a)) is a symmetrical unit cell which can be repeated in parallel to provide the specified amount of variable capacitance. The relative position of the stator and rotor electrodes defines the series capacitance between the two electrical terminals. The rotor electrodes are connected to an electrothermal actuator beam which, due to its design for a large stress gradient, is initially deflected upward such that the vertical gap limits the minimum capacitance. The actuator beam contains an embedded polysilicon resistor which is heated to lower the rotor beams into plane with the stator beams. Then, a voltage is Chapter 4: Digital Capacitor 64 applied between the capacitive electrodes to cause lateral electrostatic pull-in, resulting in a maximum capacitance limited by the thickness of the sidewall polymer. The switched capacitor is electrically connected according to the overlaid schematic shown in Fig. 4.2(b). The electrostatic actuation voltage, VES, is routed to the stator electrodes through an on-chip resistor made sufficiently large (1 MΩ) as to not significantly affect the RF electrical characteristics. The stator electrodes are connected to the bottom plate of a MIM capacitor that is included to block the large electrostatic actuation voltage (20 V), which could damage other devices connected to Port 2. The electrothermal actuation voltage, VET, is connected to the heating resistor, RHeat, though a thermal isolation structure designed to confine heat within the actuator. Thermal isolation is achieved by lengthening the metal path using a series of vias (see Section 3.2), which also increases electrical resistance. To avoid routing the RF signal through the resistive thermal isolation structure, a pair of signal beams is included to connect Port 1 to the rotor electrodes. Fig. 4.2(c) shows an SEM image which illustrates how multiple switched capacitors are combined in parallel within a capacitive bit. Fig. 4.3 shows an approximate thermal circuit for the switched capacitor, which is only valid for steady-state analysis since the thermal capacitances have not been included. Tactuator, Tbeam, and Tanchor are the temperatures of the electrothermal actuator beam, signal beams, and anchor, respectively. Pheat is the power dissipated in the heating resistor. Risolation is the thermal resistance of the thermal isolation structure. Rair is the thermal resistance of the lateral air gap between the actuator beam and the signal beams, and Roxide is the thermal resistance of the oxide connecting the end of the actuator beam to the end of the signal beams. Because there is no metal path connecting the end of the actuator beam to the rotor electrodes, an explicit thermal isolation structure is not necessary at that location. Rbeam is the thermal resistance of the signal beams. This Chapter 4: Digital Capacitor 65 thermal circuit shows that the temperature of the signal beams will increase when the electrothermal actuator beam is heated. However, this circuit should be considered only a rough approximation since the actual temperature distributions along the beams are not uniform. The heating resistor embedded in the actuator beam is lumped near the thermal isolation structure, which results in a temperature gradient along the actuator beam. The thermal resistance of the lateral air gap is distributed, which results in another temperature gradient along those beams. Finite element analysis is required to accurately simulate the temperature distribution. Tactuator Roxide Pheat Rair Risolation Tbeam Rbeam Tanchor Fig. 4.3. Steady-state thermal circuit for the CMOS-MEMS switched capacitors. The CMOS-MEMS switched capacitor was designed in a 4-metal, 0.35 µm BiCMOS process from TowerJazz Semiconductor. All four metals are aluminum. Fig. 4.4 shows the crosssection of a switched capacitor after the CMOS-MEMS process. Because metal-4 is a thick top metal requiring large line widths and gaps, the use of metal-4 is limited to routing between multiple switched capacitors. Both the stator and rotor capacitive beams are metal-321 beams to maximize the capacitive area without sacrificing the small line widths associated with the thin metal layers. The electrothermal actuator is a metal-1 beam which, due to the small thickness of the resulting metal-dielectric stack, provides the smallest radius of curvature resulting in large vertical displacements. The signal beams are also metal-1 beams to minimize the mechanical load placed on the actuators. Chapter 4: Digital Capacitor 66 Capacitive Actuator Electrode Beam ≈ 10 µm MIM Capacitor ≈ 50 µm Key: Metal 1 Metal 2 Metal 3 Oxide Via Silicon Metal 4 Fig. 4.4. Cross-section of the switched capacitor along the rotor electrode. 4.3. Analysis Analysis of the CMOS-MEMS switched capacitor is difficult due to the complex distribution of electric fields and temperature gradients. Finite element analysis (FEA) or results from measured test structures are required for a high degree of precision. The analysis that follows involves many approximations to enable design by hand calculations. After the MEMS structure is released from the substrate, the CMOS-MEMS beams curl vertically according to their temperature-dependant radii of curvature. The vertical deflection, zr, of the rotor electrode is a function of distance along its length, x, and depends on the length of the electrothermal actuator, La (see Fig. 4.5). Assuming small-angle displacements, z r x La 2 2 a La x a x2 2 r , (4.1) where ρa and ρr are the radii of curvature for the actuator beam and rotor electrode, respectively. Since the metal dielectric stack for the rotor electrode is thicker and composed of more alternating metal and dielectric layers than the actuator beam, ρa is much smaller than ρr. Neglecting the curvature of the capacitive electrodes, the vertical gap between electrodes, gz, is Chapter 4: Digital Capacitor 67 expressed as the vertical deflection minus the height of the electrodes, hc. g z x La 2 2 a wb hc wc wa La x (4.2) a . Lc,o B 4(wc+gy) B’ A gy A’ -La 0 (a) Thick metal stack for routing between capacitive bits x Lc,m hc Radius of curvature ρa MIM capacitor gz(x) zr(x) hc A -La 0B Lc,m, A’ B’ x (b) Key: Metal 1 Metal 2 Metal 3 Metal 4 Oxide Via Fig. 4.5. Diagram of CMOS-MEMS switched capacitor viewed (a) from above with nominal layout dimensions (b) from the side with the rotor electrode deflected vertically into minimum capacitance position. Line A–A’ and line B– B’ show the positions of the cross-sections along the rotor electrode and stator electrode, respectively. The vertical gap between the capacitive electrodes limits the minimum MEMS capacitance, CMEMS,min, but, for a precise calculation, finite element analysis is required to account for the lateral fringing fields, the dielectric layers, and the angle of the rotor electrode with respect to the stator electrode. However, there are two approximations which result in a simplified expression of the minimum MEMS capacitance. An array of many switched capacitors in parallel results in a Chapter 4: Digital Capacitor 68 row of rotor electrodes deflected above a row of stator electrodes. Provided that the lateral spacing from electrode to electrode is small compared with the vertical gap between the stator and rotor electrodes, the rows of electrodes can be approximated as continuous sheets of metal. Then, because small-angle displacements result in approximately vertical electric fields, the capacitance is calculated as an integral over the overlap length of the electrodes, Lc,o. C MEMS, min Lc,m Lc,m Lc,o 0 4wc g y dx g x (4.3) . where ε0 is electrical permittivity of free space and wc and gy are the widths of the capacitive electrodes and lateral capacitive gap, respectively. Evaluating this expression results in C MEMS,min 0 4wc g y a La Lc,o ln 1 L L La hc a c,m c,o 2 La . (4.4) The maximum MEMS capacitance is process dependant and, thus, must be experimentally characterized and scaled according to the length of the electrodes. The capacitance created by two electrodes pulled into mechanical contact by electrostatic actuation depends on the nature of the mechanical contact which occurs between the beams. This capacitance also depends upon the thickness of a passivation layer deposited on the sides of CMOS-MEMS beams during the vertical oxide etch. Measured data from capacitance test structures is used to extract an effective lateral air gap, gmin,eff, which is a constant in the expression for maximum MEMS capacitance, CMEMS,max. C MEMS,max 2 0 Lc,o hc g min,eff , (4.5) where hc is the height of the metal-dielectric stack for the rotor and stator electrodes. A factor of two is included because each CMOS-MEMS switched capacitor contains two capacitive gaps. Chapter 4: Digital Capacitor 69 The MEMS capacitance density, C'MEMS, is calculated by dividing the maximum capacitance by the MEMS area. C ' MEMS 2 0 Lc,o hc 4wc g y La 2 Lc,m Lc,o g min,eff . (4.6) The MEMS tuning ratio, TRMEMS, is calculated using (4.4) and (4.5). TR MEMS 2 La Lc,o hc Lc,o a g min,eff 4wc g y ln 1 La hc a Lc,m Lc,o 2 La . (4.7) This tuning ratio represents the maximum achievable tuning ratio without including the MIM capacitance. The MIM capacitance, CMIM, is calculated using the dimensions and capacitance density of the MIM capacitor, C'MIM. C MIM k f C ' MIM LMIM 4wc g y , (4.8) where kf is the fill factor and LMIM is length of the MIM capacitor. The fill factor accounts for holes in the MIM capacitor which are necessary to ensure the structure releases during the CMOS-MEMS process. Both the minimum series capacitance, Cs,min, and maximum series capacitance, Cs,max, are calculated as the series combination of the MEMS capacitance and the MIM capacitance. Cs,min Cs,max C MEMS,min C MIM C MEMS,min C MIM . C MEMS,max C MIM C MEMS,max C MIM . Chapter 4: Digital Capacitor (4.9) (4.10) 70 The series tuning ratio, TRs, is calculated using (4.9) and (4.10). TRs Cs,max . Cs,min (4.11) The area, A, is calculated using the dimensions. A 4wc g y La 2 Lc,m Lc,o LMIM . (4.12) The series capacitance density, C's, is defined the maximum series capacitance divided by the area. C 's Cs,max A . (4.13) The series resistance, Rs, is determined using the sheet resistances of the metal-dielectric stacks and the geometry of the CMOS-MEMS switched capacitor. Given that metal 1, metal 2, and metal 3 have the same sheet resistance, Rsheet, the series resistance expression is simplified because the signal beams have a sheet resistance which is three times that of the capacitive electrodes. L Lc,m Rs Rsheet a 2w b 6 wc , (4.14) where wb is the width of the signal beams. The Q for the maximum capacitance is calculated using (4.5) and (4.14). Q 1 Rs C s,max , (4.15) where ω is the frequency in radians per second. The relative spring constants of the actuator beam, ka, and the signal beams, kb, determine the temperature to which the actuator beam must be heated to lower the rotor electrodes into Chapter 4: Digital Capacitor 71 plane with the stator electrodes. For full vertical actuation, CMOS-MEMS beams must be uniformly heated to a zero-stress temperature, T0, which depends on the thermal coefficients of expansion and residual stresses of the CMOS layers. If the signal beams remain near ambient temperature, Troom, due to thermal isolation of the electrothermal actuator, then the actuator beam must be heated to an average maximum temperature, Tmax, above the zero-stress temperature. 2k k a Tmax T0 Troom b k a T . room (4.16) Since the actuator beam and signal beams have the same length and cross-section, the average maximum temperature is expressed in terms of the relative widths of the actuator beam, wa, and the signal beams, wb. 2w w a Tmax T0 Troom b wa T . room (4.17) The electrothermal actuator voltage, VET, necessary to achieve the maximum temperature depends on the thermal conductance, G, the resistance of the heater resistor, RHeat, and the change in temperature. VET 2 RHeat GTmax Troom . (4.18) The dimensions of the capacitive electrodes determine the electrostatic actuation voltage, VES, necessary to pull the electrodes into mechanical contact. The pull-in voltage for a suspended parallel-plate capacitor is well known but is not directly applicable because the capacitive electrodes consist of two cantilevers anchored at opposite ends. Nonetheless, the required voltage still scales in the same manner according to the gap, capacitance, and spring constant. Finite element analysis was used to determine a constant (1.81 ± 0.03) with which to modify the typical pull-in voltage expression. The uncertainty in this constant was determined by the range of Chapter 4: Digital Capacitor 72 voltages between which pull-in occurred. VES 1.81 8k y g y 3 , 27 0 Lc,o hc (4.19) where ky is the spring constant of a capacitive electrode for a point force applied at the tip of the beam. The vertical mechanical restoring force, Fm, which pulls the capacitive electrodes apart depends on the vertical spring constant, kz, and the vertical displacement at the tip of the electrothermal actuator, zr(0), calculated using (4.1). Fm k z z r 0 . (4.20) The electrostatic holding force, FES, which opposes the mechanical restoring force, is calculated based on the maximum MEMS capacitance. FES 2 dz VES 0 Lc,o 2 2 VES dC g min,eff . (4.21) The normal force exerted between the capacitive electrodes, Fn, by the applied voltage determines the force due to friction, Ff, which also opposes the mechanical restoring force. Fn 2 dx VES 0 Lc,o hc 2 2 V ES dC g min,eff 2 Ff u f Fn , . (4.22) (4.23) where uf is the coefficient of static friction for side-wall polymer in contact with side-wall polymer. Note that the signs of FES and Ff are opposite that of Fm. The combination of the force due to friction and the electrostatic holding force must be greater than the mechanical restoring force so that the capacitive electrodes can be held in maximum capacitance position without consuming power for electrothermal actuation. Chapter 4: Digital Capacitor 73 Fm FES Ff 0 . (4.24) 4.4. Design The CMOS-MEMS switched capacitor was designed to enable a 3-bit digital capacitor with large tuning ratio, high capacitance density, and sufficient Q for operation at multi-GHz frequencies. These performance metrics cannot be simultaneously optimized and are valued differently depending on the particular circuit application. The design that follows is intended as a general purpose switched capacitor which achieves a tuning ratio of approximately 50:1 while maximizing capacitance density. Design began with the geometry of the capacitive electrodes. Equations (4.6) and (4.7) show that decreasing the y-direction pitch increases the capacitance density and tuning ratio, so wc and gy were chosen to be 1 µm and 1.5 µm, respectively, near the minimum achievable by the CMOS-MEMS process. The electrode length, Lc,m, is limited for a given electrode width, wc, because CMOS-MEMS beams which are long and narrow tend to curl laterally, which interferes with vertical actuation if the amount of curling exceeds the lateral gap. Thus Lc,m was chosen to be 95 µm, short enough to avoid lateral curling greater than gy and long enough to provide a relatively small actuation voltage according to (4.19). Since alternate versions of the switched capacitor with 135 µm long electrodes failed consistently due to excessive lateral curling, 95 µm is near the maximum electrode length for this design. The electrode overlap length, Lc,o, was set to 90 µm. These decisions defined the structure and spacing of the capacitive electrodes, which sets the maximum MEMS capacitance. The length of the MIM capacitor, LMIM, was chosen to maximize the series capacitance density, C's. The MIM capacitance should be large to maximize the series capacitance but not so Chapter 4: Digital Capacitor 74 large that the MIM capacitor area dominates the area of the MEMS structure. Equations (4.5) (4.13) were used to plot the tuning ratio and capacitance density as a function of MIM capacitor length. Fig. 4.6 shows that, based on the MEMS and MIM capacitance densities, C'MEMS and C'MIM, the series capacitance density is optimized for a MIM capacitor length of 35 µm, which results in a MIM capacitance 5.2x that of the maximum MEMS capacitance. To account for the possibility of an increased MEMS capacitance, the MIM capacitor length was set to 45 µm, which lowers the capacitance density by 7% but increases the series tuning ratio by 4%. Normalized Parameters 1.4 Tuning Ratio 1.2 Capacitance Density 1 0.8 0.6 optimum capacitance density at ≈ 35 µm 0.4 0.2 0 0 20 40 60 80 100 MIM Capacitor Length (µm) Fig. 4.6. Normalized parameters as a function of MIM capacitor length. Tuning ratio and capacitance density are normalized to their respective values for the optimum MIM capacitor length of 35 µm. Design of the electrothermal actuator is all that remains to define the geometry of the switched capacitor. The actuator beam length, La, determines the tradeoff between MEMS capacitance density, MEMS tuning ratio, and series resistance according to (4.6), (4.7), and (4.14), respectively. To produce a tuning ratio in excess of 40:1, La was chosen to be 90 µm. By increasing La, larger tuning ratios can be achieved at the expense of both capacitance density and Q (see Fig. 4.7). For example, increasing the actuator beam length to 120 µm increases the tuning ratio by 87% but decreases the capacitance density and quality factor by 11% and 20%, Chapter 4: Digital Capacitor 75 respectively. Parameters wa and wb were chosen to be 3.5 µm and 1 µm, respectively, so that the actuator beam has a larger spring constant than that of the two signal beams. These widths resulted in a calculated maximum temperature of 267°C according to (4.16). Normalized Parameters 2.5 Tuning Ratio Capacitance Density 2 Quality Factor 1.5 1 0.5 0 75 90 105 120 135 Actuator Beam Length (µm) Fig. 4.7. Normalized parameters as a function of actuator beam length. Tuning ratio, capacitance density, and quality factor are normalized to their respective values for an actuator beam length of 90 µm to illustrate the tradeoffs necessary for increased tuning ratios. The resistance of the heating resistor, RHeat, was chosen to be 3 kΩ according to (4.18) to achieve the calculated maximum temperature using an electrothermal actuation voltage, VET, of 3.5 V. The thermal conductance, G, was calculated based on heat conduction through the thermal isolation structure and through the air gap separating the signal beams from the actuator beam. Both the signal beams and the metal 4 beams, which are mechanically anchored but undercut by the isotropic silicon etch, were assumed to be thermally grounded to provide a worst-case estimate of the electrothermal actuator power consumption. Equations (4.20) through (4.23) were used to compare the mechanical restoring force, Fm, with the electrostatic holding force, FES, and the force due to friction, Ff. The mechanical restoring force was 3.0 µN, the electrostatic holding force was 1.3 µN, and the normal force, Fn, was 33.0 µN. According to (4.24), the capacitive electrodes will be held in the maximum Chapter 4: Digital Capacitor 76 capacitance position for coefficients of static friction as low as 0.03, which encompasses the vast majority of materials, using only the minimum electrostatic actuation voltage necessary for pullin. Table 4.1 lists all the parameters used in the design of the CMOS-MEMS switched capacitor presented in this work. Table 4.1: Digital Capacitor Design Parameters Geometry Value Calculation Value La 90 µm CMEMS,min 0.70 fF Lc,m 95 µm CMEMS,max 38.9 fF Lc,o 90 µm C'MEMS 16.5 pF/mm2 wa 3.5 µm TRMEMS 56:1 wb 1 µm CMIM 261 fF wc 1 µm Cs,min 0.70 fF gy 1.5 µm Cs,max 33.8 fF LMIM 45 µm TRs 48:1 A 1900 µm2 Constants Value C' 14.4 pF/mm2 h c* 4.68 µm Rs 4.0 Ω ρa * 471 µm Q (1 GHz) 1170 gmin,eff* 0.19 µm Tmax 267°C kf 0.58 ky 0.20 N/m C'MIM 1 fF/µm2 G 16.8 µW/K T 0* 178°C RHeat 3 kΩ VET 3.4 V VES 12.7 V kz 0.76 N/m Fm 3.0 µN FES 1.3 µN Fn 33.0 µN * Parameter extracted from measured data The 3-bit digital capacitor was designed using an array of 35 switched capacitors in parallel (see Fig. 4.8) to achieve a maximum capacitance of at least 1 pF. Larger or smaller maximum capacitances can be achieved by scaling the number of parallel switched capacitors. Such scaling does not affect the Q since any increase in maximum capacitance is accompanied by a corresponding decrease in series resistance. The 3 bits consist of 5, 10, and 20 switched capacitors, respectively, sized in a binary fashion to create a linear range of eight different Chapter 4: Digital Capacitor 77 capacitance values. The digital capacitor was limited to 3 bits for testing purposes; the same topology could be used to produce a digital capacitor with higher capacitance resolution. Bit 1 Port 2 Bit 2 Bit 3 MIM Capacitor Port 1 Actuators Electrodes 100 µm Fig. 4.8. SEM image of 3-bit digital capacitor. Bits are separated by dummy structures to improve the temperature uniformity of the electrothermal actuators. Floating electrodes appear darker due to charging. 4.5. Characterization The CMOS-MEMS switched capacitors were successfully toggled between minimum and maximum capacitance positions using vertical electrothermal actuation and lateral electrostatic pull-in (see Fig. 4.9 and Fig. 4.10). Each capacitive bit was controlled independently to select any of the eight capacitor states (see Fig. 4.11). Electrothermal actuation required a maximum of 4 V and 70 mW to control the 35 switched capacitors while electrostatic actuation required 20 V. Electrostatic force was capable of maintaining the maximum capacitance position when power to the heating resistors was removed. Except when otherwise specified, all measurements were recorded after electrostatic pull-in with the electrothermal actuation turned off. All characterization was performed with the capacitors exposed to the general laboratory atmosphere. Chapter 4: Digital Capacitor 78 (a) (b) (c) Fig. 4.9. Microscope picture of CMOS-MEMS switched capacitors (a) in minimum capacitance position before actuation (rotor electrodes are deflected out of focus) (b) after electrothermal actuation (c) in maximum capacitance position after electrostatic actuation. 10.1 9.0 3.0 1.0 -1.0 -3.0 8.0 6.0 4.0 2.0 0.0 -2.0 -4.5.0 -5.1 (a) Vertical Deflection (µm) 5.0 Vertical Deflection (µm) 7.0 11.2 10.0 (b) Fig. 4.10. Profilometer image of bit 3 of the digital capacitor (a) before actuation (b) after actuation. The rotor beams are not visible when deflected upwards because their angle with respect to the chip surface is too large for their profile to be captured by the Wyko unless settings are modified which sacrifice the quality of the image. Out of a total of 10 variable capacitors, each on a separate CMOS die, 7 were functional and the remaining 3 failed to actuate into maximum capacitance position. Failure was caused by lateral curling of the rotor and stator electrodes to an extent such that the bottom of the rotor electrodes collided with the top of the stator electrodes during electrothermal actuation. This collision prevented the rotor electrodes from being lowered into the plane with the stator Chapter 4: Digital Capacitor 79 electrodes and from being subsequently snapped into maximum capacitance position. The functional devices exhibited varying amounts of the same lateral curling, but not enough to prevent vertical actuation. To improving the yield, the length of the capacitive electrodes should be reduced, which decreases tuning ratio and capacitance density. (a) (b) (c) (d) (e) (f) (g) (h) Fig. 4.11. Microscope image of the 3-bit digital capacitor in state (a) 0 (b) 1 (c) 2 (d) 3 (e) 4 (f) 5 (g) 6 (h) 7. Each capacitive bit is controlled independently to select the desired capacitance. The capacitive electrodes also curl vertically, which is potentially problematic because vertical misalignment between the capacitive electrodes can reduce the capacitive overlap area. Similar to that of the electrothermal actuator beams, the vertical curling of the capacitive electrodes is temperature dependant and caused by residual stress. However, the magnitude of this vertical curling is smaller because the capacitive electrodes are composed of a thicker metaldielectric stack using multiple metal layers. Measurements from a Wyko optical profilometer show that the capacitive electrodes deflect vertically by 0.7 ± 0.05 µm relative to their anchor position. This deflection is undesirable but tolerable because the measured height of the capacitive electrodes is 4.68 µm, larger than the vertical deflection. Chapter 4: Digital Capacitor 80 4.5.1. Electrical Performance The digital capacitors were electrically characterized using the method described in Appendix A for all eight capacitor states. Fig. 4.12 shows the S-parameters for the minimum capacitance and maximum capacitance states. Fig. 4.13 shows the equivalent circuit model in terms of Y-parameters and lumped circuit elements. Fig. 4.14 shows the lumped impedances for the minimum capacitance and maximum capacitance positions states. The equations listed below were used to calculate the equivalent circuit element values (see Fig. 4.13(b)). 2 1 res . Cs imag 1 Y21 Ls 1 resCs 2 (4.25) . (4.26) Rs real 1 Y21 . C1g C 2g 1 * imag 1 Y11 Y21 1 * imag 1 Y22 Y21 Q imag 1 Y11 real1 Y11 . Chapter 4: Digital Capacitor (4.27) . (4.28) . (4.29) (4.30) 81 10 0 0 S-parameters (dB) S-parameters (dB) 10 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 -10 -20 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -30 -40 -40 0.1 1 0.1 10 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 4.12. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. Port 1 Port 2 -Y21 Y11+Y12 Port 1 Rs Cs Ls Port 2 C2g Y22+Y21 1E+5 1E+5 1E+4 1E+4 Impedance (Ω) Impedance (Ω) (b) (a) Fig. 4.13. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 0.1 1 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 10 0.1 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 4.14. Lumped impedances as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. The behavior of |1/(Y11+Y12)| is explained by silicon remaining beneath the suspended MIM capacitor. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 11.4 GHz for the maximum capacitance position. Chapter 4: Digital Capacitor 82 The electrical characteristics of a typical 3-bit digital capacitor are presented in Table 4.2. The state indicates which of the capacitive bits were active (for example, state 6 (110) means that the two most significant bits are active). Measurement uncertainties were derived from multiple measurements of the same capacitor state and thus do not include systematic offsets based on the particular network analyzer calibration or probe contact resistance. As designed, the series capacitance, Cs, increases linearly with state. The shunt capacitances, C1g and C2g, change slightly with state primarily due to varying amounts of capacitance to the silicon substrate, which lies approximately 50 µm below the MEMS structure. For minimum capacitance, the rotor electrodes are deflected vertically and shielded from the silicon substrate by the stator electrodes. As the state increases, more rotor electrodes are lowered into plane with the stator electrodes, so the associated shunt capacitance, C2g, increases slightly. Conversely, the shunt capacitance associated with the stator electrodes, C1g, decreases as the capacitance to the silicon substrate is shared with the rotor electrodes. Table 4.2: Electrical Characteristics of the Initial Digital Capacitor Cs (fF) Ls (nH) Rs (Ω) C1g (fF) C2g (fF) State ± 0.4% ± 0.2% ± 0.8% ± 0.2% ± 0.2% 0 (000) 23.0 NA* NA** 75.0 478.5 1 (001) 243.5 0.214 2.34 72.8 480.9 2 (010) 435.5 0.209 1.18 72.1 481.3 3 (011) 640.2 0.182 1.11 70.0 484.1 4 (100) 736.4 0.177 1.02 68.6 485.5 5 (101) 1015 0.149 0.66 66.2 489.1 6 (110) 1178 0.153 0.48 65.3 488.8 7 (111) 1397 0.140 0.37 62.1 492.5 * Inductance value could not be extracted since the self-resonance frequency exceeded 40 GHz ** Resistance could not reliably be extracted from the small capacitance Q (1 GHz) ± 0.4% NA** 35 55 68 75 94 110 125 Note that the data presented in Table 4.2 corresponds to a digital capacitor for which the isotropic silicon etch was incomplete. Parts of the MIM capacitor which should have been suspended were proven to be anchored to the substrate by probing the structure. The unreleased Chapter 4: Digital Capacitor 83 sections contribute a parasitic capacitance to the silicon substrate, which is a complex RC network whose effective resistance and capacitance vary with frequency. This lossy parasitic capacitance is responsible for the behavior of |1/(Y11+Y12)| shown in Fig. 4.14 and also increases the C1g and decreases the Q presented in Table 4.2. The measured Q increases with state since Cs, the high quality series capacitance, increases with state relative to C1g, the lossy shunt capacitance. Fig. 4.15 shows the measured Q for the maximum capacitance (state 7) along with a fit to the data according to (4.30). The disparity between measured data and the extracted model is similarly caused by the parasitic capacitance to the silicon substrate. 200 Quality Factor (De-embedded) Quality Factor 180 Extracted Model 160 140 120 100 80 60 40 20 0 1 2 3 4 5 6 Frequency (GHz) Fig. 4.15. Quality factor as a function of frequency. This plot shows the Q for the maximum capacitance position with port 2 connected to ground. The series capacitance of the digital capacitor was measured at each of the eight states for seven different devices (see Fig. 4.16). Measured data shows the capacitor was successful in producing a distribution of capacitance values which matches a linear trend with an R2 of 0.9997. The average digital capacitor tunes from 22.6 fF to 1374 fF, an average tuning ratio of 61:1. The best-performing digital capacitor tuned from 19.8 fF to 1360 fF, a tuning ratio of 69:1. The variation between devices, measured by standard deviation, differs slightly depending on the Chapter 4: Digital Capacitor 84 capacitor state. The standard deviation for the maximum capacitance is 4.6% while the standard deviation averaged over all states is 7.2%. The standard deviation for state 6 appears large but is only 6.4%, less than the average. The largest standard deviation, which is not apparent from Fig. 4.16, is that of state 1 at 13.3%, but that standard deviation is primarily due to poor performance of the least significant capacitance bit for a single device. The variation between devices could be due to variation in the thickness of the side-wall polymer deposited during the vertical oxide etch Series Capacitance (fF) of the CMOS-MEMS process. 1400 1200 1000 800 600 400 200 0 0 1 2 3 4 5 6 7 State Fig. 4.16. Plot of series capacitance versus state. Error bars indicate the 1-σ standard deviation between different devices based on 7 measurements. Series capacitance was measured as the 3-bit digital capacitor was slowly stepped through one actuation cycle (see Fig. 4.17). First, as predicted by theory, the measured capacitance increased gradually as the voltage applied to the electrothermal actuator resistors (VET) was increased. After electrothermal actuation, measured results temporarily diverged from the predicted theory. Once the rotor electrodes were lowered into the plane with the stator electrodes but before an electrostatic actuation voltage was applied, the tips of several electrodes displaced laterally into mechanical contact with their corresponding electrodes (see Fig. 4.9(b)). This lateral Chapter 4: Digital Capacitor 85 deflection explains the unpredicted jump in series capacitance. Since only the electrode tips rather than the full length of the electrodes moved into mechanical contact, the displacement appeared to be partial electrostatic pull-in. After the electrostatic actuation voltage (VES) reached a certain threshold, typically around 6 V – 10 V, the electrode tips that partially pulled-in were seen to retract, which caused the measured capacitance to resume agreement with theory. The observed partial pull-in does not occur on the first actuation cycle but does occur for all subsequent cycles. The fact that electrostatic force between conductors decreased with increasing voltage is evidence of dielectric charging, enough trapped charge such that an applied voltage of approximately 10 V is necessary to counteract this effect. 1200 Measurement Theory 1000 800 600 400 200 VET (V) 0 4 after heating, partial pull-in from charging electrostatic pull-in applied voltage counteracts charging effect 20 3 15 2 10 1 5 0 Turn on ET Turn on ES Turn off ET Turn off ES actuation actuation actuation actuation VES (V) Series Capacitance (fF) 1400 0 Fig. 4.17. Plot of series capacitance during an actuation cycle. The x-axis is arbitrary; all measured points shown were allowed to reach steady-state and maintained for approximately one minute. ―Electrothermal‖ and ―electrostatic‖ are abbreviated as ―ET‖ and ―ES‖, respectively. The plotted data represents a single actuation cycle from a single device and thus may vary over time or between different devices. After electrostatic pull-in, represented by the large increase in capacitance from Chapter 4: Digital Capacitor 86 approximately 200 fF to greater than 1100 fF, the capacitance continues to increase gradually as power applied to the heating resistors is ramped down. Presumably, this occurs because some rotor electrodes are lowered too far and, thus, settle into a higher capacitance position as the temperature is decreased. As the electrostatic actuation voltage is ramped down, capacitance decreases gradually until one or more capacitance bits suddenly disengage. 4.5.2. Switching Speed A computer microvision system was used to measure the frequency response of the vertical electrothermal actuation similar to the method presented in Section 3.5.2. Mechanical motion was recorded as a sinusoidal voltage with a DC bias was applied to the heating resistors within the electrothermal actuators. The thermal 3 dB frequency occurs at 503 ± 39 Hz while mechanical resonance occurs at 37.9 ± 2.9 kHz (see Fig. 4.18). Thus, the switching speed is limited by the thermal time constant (≈ 0.32 ms) rather than the mechanical response. Z-displacement (µm) 10 1 Thermal 3dB Frequency ≈ 500 Hz 0.1 Mechanical Resonance Frequency ≈ 38 kHz 0.01 10 100 1000 10000 100000 Frequency (Hz) Fig. 4.18. Electrothermal actuator frequency response. The switched capacitors were typically switched on by applying voltage to the electrothermal actuator resistors for 1 ms, or about 3 thermal time constants. During the last Chapter 4: Digital Capacitor 87 100 µs of electrothermal actuation, the electrostatic actuation voltage was increased from 0 V to 20 V and then maintained as necessary, often for up to several minutes. Faster electrothermal actuation can be achieved by using higher voltages to heat the electrothermal actuator at a faster rate. However, to avoid applying excessively high voltages which could damage the limited number of devices, the absolute minimum switching time was not determined. The shortest demonstrated switching time was 0.6 ms, or about two thermal time constants. Note that this switching time is the time necessary for electrothermal actuation and subsequent electrostatic pull-in. Since Fig. 4.17 shows that the maximum capacitance continues to increase after electrostatic pull-in while the temperature decreases, several thermal time constants may be required after the switching time for the capacitance to settle into its final value. Because the switched capacitors cannot be switched off until the after the actuators have cooled, switching faster than the thermal time constant would not be practical. The switched capacitors were typically switched off by lowering the electrostatic actuation voltage from 20 V to 0 V for 1 ms without applying any voltage the electrothermal actuator resistors. This method of switching off the capacitors should cause the rotor electrodes to vibrate at the mechanical resonance frequency. Based on the measured resonance frequency and mechanical Q, any vibration of the rotor electrodes should settle to within 2% of its initial displacement in 0.41 ms. Since switching between two digital capacitor states often involves turning on some bits while turning off other bits, the switching time is limited by the slower transition, which is the time required to switch on a capacitance bit. 4.5.3. Thermal Aspects The heating resistors embedded in the electrothermal actuator beams were used as Chapter 4: Digital Capacitor 88 temperature sensors to characterize the temperature necessary for actuation. The heater resistance was monitored as the power dissipated in the resistors was increased. Based on the measured temperature coefficient of resistivity, the maximum operating temperature necessary to lower the rotor electrodes into plane with the stator electrodes was calculated to be 249.5 ± 5.4°C. Because the electrothermal actuators are thermally isolated from the bulk of the substrate, this temperature is applied only to the MEMS structure and thus does not significantly affect other on-chip CMOS devices. 250 225 200 150 125 100 Temperature ( C) 175 75 50 25 Fig. 4.19. Simulated temperature profile of electrothermal actuator. Symmetry boundary conditions are applied so the simulation represents an array of many switched capacitors in parallel. The temperature profile of the electrothermal actuator beam when heated was simulated using Comsol finite element analysis (see Fig. 4.19). Measurements of the temperature profile were not possible since the quantum infrared microscope lacks the resolution required for the small widths of the actuator beams and signal beams. The temperature of the actuator beams when heated is not uniform because the heating resistor is lumped near the base of the actuator Chapter 4: Digital Capacitor 89 beam. Heat conduction, primarily through the air, causes the temperature to decrease along the length of the actuator beams. Some of that heat is transferred to the signal beams running parallel to the actuator beams, which causes the temperature of the signal beams to increase along their length. The capacitance of the digital capacitor was measured over a range of ambient temperatures to assess the temperature sensitivity inherent to electrothermal actuation. The maximum capacitance position is not significantly affected by temperature change since the capacitive electrodes are locked together by electrostatic force. However, for the minimum capacitance position, increasing the ambient temperature causes the electrothermal actuators to lower the rotor electrodes closer to the stator electrodes, which increases the minimum capacitance. While varying the ambient temperature from 25°C to 100°C, the minimum series capacitance of the digital capacitor increased from 20.0 fF to 35.1 fF (see Fig. 4.20). Over the same temperature range, the maximum capacitance varied between 1443 fF and 1485 fF, a variation less than 3% of the nominal (25°C) capacitance value. The theoretical maximum ambient temperature at which the digital capacitor could still function is approximately 180°C, which is the zero-stress temperature for the electrothermal actuator beams. However, operation at that temperature would not be practical since the capacitive electrodes would be in the same plane without requiring electrothermal actuation and the tuning ratio would decrease to less than 10:1. A more practical maximum ambient temperature of 85°C would result in a minimum tuning ratio still greater than 40:1. Chapter 4: Digital Capacitor 90 2000 1750 50 1500 40 1250 30 1000 750 20 500 10 250 0 Max. Capacitance (fF) Min. Capacitance (fF) 60 0 25 40 55 70 85 100 Temperature ( C) o Fig. 4.20. Series capacitance of a 3-bit digital capacitor as a function of ambient temperature. 4.5.4. Reliability Tests To assess the short-term reliability of the digital capacitor, the series capacitance was measured after each of twenty actuation cycles for all eight capacitance positions. The capacitance for each position was repeatable with a worst-case standard deviation less than 1.4% of the average capacitance value. The average standard deviation for any given position was 0.5% of the average capacitance value. To assess the long-term reliability of the digital capacitor, automated measurements were setup to measure the series capacitance as the capacitor was cycled between minimum and maximum capacitance. However, these tests failed much sooner than expected, lasting only hours rather than days. The digital capacitor operated correctly until electrostatic pull-in failed to occur. Adjusting the voltage applied to the heating resistors caused electrostatic pull-in to temporarily resume until the problem occurred again. Repeated actuation was causing a shift in the vertical position of the rotor electrodes when power was applied to the heating resistors. Chapter 4: Digital Capacitor 91 An inspection of the digital capacitor after repeated actuation revealed damage on the surface of the electrothermal actuators, which is peculiar for several reasons (see Fig. 4.21). While the damage is greatest at the location of the heating resistors, damage is not confined to that maximum temperature location and often appears farther down the length of the beam where the temperature is reduced. Also, the damaged regions are not continuous along the length of the actuator beam. Damage only occurred on the electrothermal actuator beams and never on the signal beams which experienced the same mechanical motion. Since the signal beams are significantly cooler than the actuator beams, these observations suggest that the damage must be caused or at least accelerated by high temperature. However, damage only occurs from temperature cycling; heating the actuators to a constant temperature of 250°C for more than 12 hours did not produce visible surface damage. Thus this damage is classified as thermallyinduced mechanical fatigue. Key = resistor Key 20 µm = resistor 5 µm (a) (b) Fig. 4.21. SEM images of surface damage caused by repeated temperature cycling at 250°C. The white boxes indicate the locations of the embedded heating resistors. To measure the mechanical fatigue, two digital capacitors were cycled between room temperature (23 ± 2°C) and a specified maximum temperature for more than 40 million cycles. The first capacitor was cycled with a maximum temperature of 250°C, the nominal temperature Chapter 4: Digital Capacitor 92 necessary to operate the digital capacitor. The second capacitor was cycled with a more conservative maximum temperature of 180°C, which is insufficient to enable subsequent electrostatic actuation for the current design. The frequency of actuation was set to 100 Hz to allow ample time (> 15 thermal time constants) for the actuator beams to reach their maximum temperatures. Testing was stopped intermittently while a Wyko optical profilometer was used to measure the vertical gap between the rotor and stator electrodes. The voltage and current applied to the heating resistors was also monitored to verify that the heater resistance remained constant Vertical Gap w/ Heating (µm) to within 1% of the nominal value. 9 7 5 Cycling at Low Temp. (≈ 180oC) 3 Cycling at High Temp. (≈ 250oC) 1 -1 -3 -5 1.E+03 103 1.E+04 104 1.E+05 105 1.E+06 106 1.E+07 107 1.E+08 108 # Cycles Fig. 4.22. Vertical gap during heating as a function of the number of actuation cycles. A positive gap indicates that the tip of the rotor electrodes are deflected above the stator electrodes. As a result of repeated temperature cycling, the vertical gap between the rotor and stator electrodes during heating decreased from 7.8 µm to 5.8 µm and from 0.3 µm to -4.6 µm for maximum temperatures of 180°C and 250°C, respectively (see Fig. 4.22). A vertical gap of zero indicates that the rotor and stator electrodes are perfectly aligned for electrostatic actuation. For both temperatures, repeated actuation caused the rotor electrode to move to a progressively lower position while the total actuation stroke (nominally 28 – 29 µm) remained constant to within 8%. Chapter 4: Digital Capacitor 93 Typically, electrostatic pull-in still occurs for vertical gaps up to approximately 1 – 2 µm but fails for larger gaps. Thus the current digital capacitor designed to operate at 250°C will not exceed 10 million cycles unless the actuation voltage is dynamically adjusted to account for the fatigue. Digital capacitors designed to operate at 180°C will exceed 10 million cycles, longer than capacitors operating at higher temperatures. Several qualitative observations are worth noting due to their potential impact on reliability. These observations are atypical results that are not always repeatable and do not occur for every digital capacitor. Occasionally, a capacitive bit that has been actuated into maximum capacitance position would not return to minimum capacitance position after the electrostatic actuation voltage was lowered to zero. The stuck capacitive bit could usually be unstuck using repeated electrothermal actuation. Rarely, after actuating a capacitance bit, DC current caused by the electrostatic actuation voltage would flow through the capacitive electrodes. Monitoring the current while varying the electrostatic actuation voltage revealed a diode-like current-voltage relationship which turned on at 10 V. Presumably, this diode-like behavior is electrical breakdown occurring between the capacitive electrodes. On one occasion, the digital capacitor was left in maximum capacitance position after all voltages had been disconnected and remained in that position over night. The capacitor could not be returned to minimum capacitance position using electrothermal actuation. The measured DC resistance between the electrostatic actuation voltage and port 2 was approximately 1 MΩ, the resistance of the bias resistor. The fact that the resistance was measurable at DC suggests that one or more capacitive electrodes became fused together. Chapter 4: Digital Capacitor 94 4.5.5. Updated Topology For the initial 3-bit digital capacitor presented above, the shunt capacitance connected to port 2, C2g, was large, comparable with the maximum series capacitance, Cs, (see Table 4.2). If the digital capacitor is used as a shunt capacitor, then port 2 can be connected to ground so that C2g does not affect electrical performance. However, if the digital capacitor is used as a series capacitor, then C2g is a large parasitic capacitance to ground. This parasitic capacitance is large because the signal beams carrying the RF signal were interwoven with the grounded electrothermal actuators (see Fig. 4.23(a)). To reduce this parasitic capacitance, the topology was updated such that the actuator beams are lumped together, separate from the signal beams carrying the RF signal (see Fig. 4.23(b)). This modification was expected to reduce C2g by an order of magnitude without significantly affecting other electrical and mechanical characteristics. (a) (b) Fig. 4.23. Diagram of capacitive bit with (a) interwoven actuator beams resulting in large parasitic capacitance (b) updated topology to reduce parasitic capacitance. The updated 3-bit digital capacitor (see Fig. 4.24) was successfully actuated into all eight capacitor states (see Fig. 4.25), each of which was electrically characterized using the method described in Appendix A. Fig. 4.26 shows the S-parameters for the minimum capacitance and maximum capacitance states. Fig. 4.27 shows the equivalent circuit model in terms of Yparameters and lumped circuit elements. Fig. 4.14 shows the lumped impedances for the minimum capacitance and maximum capacitance states. Equations (4.25) through (4.29) were Chapter 4: Digital Capacitor 95 used to calculate the equivalent circuit element values (see Fig. 4.27(b)). 100 µm MIM Capacitor Port 1 Bit 3 Bit 2 Electrodes Actuators Actuators Port 2 Bit 1 Fig. 4.24. SEM image of 3-bit digital capacitor updated to reduce parasitic capacitance. The chunk of debris was left in place since manual removal could damage the actuators. (a) (b) (c) (d) (e) (f) (g) (h) Fig. 4.25. Microscope image of the updated 3-bit digital capacitor in state (a) 0 (b) 1 (c) 2 (d) 3 (e) 4 (f) 5 (g) 6 (h) 7. Each capacitive bit is controlled independently to select the desired capacitance. Chapter 4: Digital Capacitor 96 10 0 0 S-parameters (dB) S-parameters (dB) 10 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 -10 -20 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -30 -40 -40 0.1 1 0.1 10 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 4.26. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. Port 1 Port 2 -Y21 Y11+Y12 Port 1 Rs Cs Ls Port 2 C2g Y22+Y21 1E+5 1E+5 1E+4 1E+4 Impedance (Ω) Impedance (Ω) (b) (a) Fig. 4.27. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 0.1 1 1E+3 1E+2 1E+1 |1/-Y21| (Measured) |1/(Y11+Y12)| (Measured) |1/(Y22+Y21)| (Measured) Extracted Model 1E+0 1E-1 10 0.1 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 4.28. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 10 GHz for the maximum capacitance position. Chapter 4: Digital Capacitor 97 Table 4.3: Electrical Characteristics of the Updated Digital Capacitor Cs (fF) Ls (nH) Rs (Ω) C1g (fF) C2g (fF) Q1 (1 GHz) Q2 (1 GHz) State ± 0.4% ± 0.2% ± 0.8% ± 0.2% ± 0.2% ± 0.4% ± 0.4% 0 (000) 18.3 NA* NA** 33.5 55.4 NA** NA** 1 (001) 234.1 0.293 2.52 33.0 55.6 170 341 2 (010) 444.7 0.256 1.28 33.1 56.0 193 329 3 (011) 648.3 0.239 0.99 31.4 57.9 190 289 4 (100) 795.4 0.226 0.76 30.9 58.7 199 293 5 (101) 960.8 0.198 0.56 29.5 60.0 223 336 6 (110) 1130 0.200 0.50 28.4 61.5 222 317 7 (111) 1335 0.191 0.42 27.9 61.9 240 369 * Inductance value could not be extracted since the self-resonance frequency exceeded 40 GHz ** Resistance could not reliably be extracted from the small capacitance The electrical characteristics of the updated 3-bit digital capacitor are presented in Table 4.3. As designed, the series capacitance, Cs, series inductance, Ls, and series resistance, Rs, are similar to those of the initial digital capacitor presented in Table 4.2. The shunt capacitance corresponding to port 1, C1g, was reduced from an average of 69.0 fF to an average of 31.0 fF, primarily because the silicon beneath the digital capacitor had been completely etched away. The shunt capacitance corresponding to port 2, C2g, was reduced from an average of 485.1 fF to an average of 58.4 fF, which was the goal of the updated design. Table 4.3 includes two different quality factors, Q1 and Q2. Q1 is the quality factor seen looking into port 1 with port 2 connected to ground. Q1 imag 1 Y11 real1 Y11 . (4.31) Q2 is the quality factor seen looking into port 2 with port 1 connected to ground. Q2 imag 1 Y22 real1 Y22 . (4.32) Only a single quality factor was reported for the initial digital capacitor in Table 4.2 because the shunt capacitance, C2g was large compared with the series capacitance, Cs. Using the initial digital capacitor as a shunt capacitor with port 1 connected to ground would result in a tuning Chapter 4: Digital Capacitor 98 ratio of only 3.8:1, an unacceptable sacrifice for moderately increased Q. However, since C2g has been reduced in the updated digital capacitor, two configurations with a large shunt tuning ratio are possible. Looking into port 1 with port 2 connected to ground, the tuning ratio is 26:1 with a Q of 240 at 1 GHz for the maximum capacitance position. Looking into port 2 with port 1 connected to ground, the tuning ratio is 19:1 with a Q of 369 at 1 GHz for the maximum capacitance position. Fig. 4.29 shows the measured Q2 for the maximum capacitance along with Quality Factor a fit to the data according to (4.32). 500 450 400 350 300 250 200 150 100 50 0 Quality Factor (De-embedded) Extracted Model 1 2 3 4 5 6 Frequency (GHz) Fig. 4.29. Quality factor as a function of frequency for the updated digital capacitor. This plot shows the Q for the maximum capacitance position with port 1 connected to ground. 4.6. Discussion Compared with the electrothermal gap-tuning capacitor presented in Chapter 3, the 3-bit digital capacitor has demonstrated improvements in tuning ratio and capacitance density by 10x and 5x, respectively. Variation between devices, represented by the standard deviation of the maximum capacitance, has decreased from 18% to 4.4%. While the transient actuation power has increased, the energy per actuation (normalized to 1 pF of maximum capacitance) has decreased Chapter 4: Digital Capacitor 99 from 670 µJ to 50 µJ because the switching time has improved from 30 ms to 1 ms. The series inductance has been reduced below 200 pH, which maintains the self-resonance frequency near 10 GHz since the maximum capacitance has increased. Table 4.4: Digital Capacitor Results Name Initial Digital Capacitor Updated Digital Capacitor Tuning ratio (series) 61:1 73:1 Tuning ratio (shunt) 16:1 19:1 Q (1 GHz) for Cs,max* 125 369 Self-resonance frequency for Cs,max 11.4 GHz 10.0 GHz Cs,max 1374 fF 1335 fF Cs,min 22.6 fF 18.3 fF C1g** 69.0 fF 31.0 fF C2g** 485.1 fF 58.4 fF Rs for Cs,max 0.37 Ω 0.42 Ω Ls for Cs,max 0.140 nH 0.191 nH Area 0.16 mm2 0.17 mm2 Capacitance Density 8.6 pF/ mm2 7.9 pF/ mm2 Actuation Method electrothermal / electrostatic electrothermal / electrostatic Control 3-bit digital 3-bit digital Voltage 4 V / 20 V 4 V / 20 V Power 70 / 0 mW*** 70 / 0 mW*** Switching Time 1 ms NA**** Cs,max Standard Deviation 4.4% NA**** Integration maskless post-CMOS etching maskless post-CMOS etching * Quality factor is calculated from a fit to measured data ranging from 1 GHz to 4 GHz ** Shunt capacitances are averaged over all capacitor states *** Power is only consumed during switching **** Value was not measured but is expected to be similar to those of the initial digital capacitor The digital capacitor has two primary disadvantages when compared with the electrothermal gap-tuning capacitor. The actuation scheme is more complicated and larger voltages are required for electrostatic actuation. However, these drawbacks are minor because other examples of integrated RF MEMS have shown that large voltages can be provided by CMOS charge pumps, given that high-voltage CMOS is available, and that actuation control systems are necessary for even a simple vertical electrostatic pull-in structure [38]. Thus future work should be focused on improving the reliability of the digital capacitor. Chapter 4: Digital Capacitor 100 The 3-bit digital capacitor produces repeatable capacitance values over short periods of time but suffers from mechanical fatigue due to repeated electrothermal actuation. The current lifetime, approximately 10 million cycles, limits the capacitor to applications that do not require continuous switching. The observed fatigue does not cause sudden catastrophic failure of the capacitor but mandates an adjustment of the electrothermal actuation voltage to align the capacitive electrodes for electrostatic actuation. If a control system was designed to dynamically adjust the electrothermal actuation voltage, the lifetime of the digital capacitor could be extended at the cost of increased minimum capacitance over time. Alternatively, the digital capacitor design could be modified to operate at lower temperatures, which may increase the lifetime by an order of magnitude. Other solutions may exist such as using different materials for the electrothermal actuators. Copper, which is readily available in more advanced CMOS processes, may be more resistant to thermally-induced fatigue than aluminum due to a higher melting temperature. Chapter 4: Digital Capacitor 101 Chapter 5. Phase Shifter This section presents a 4-bit CMOS-MEMS phase shifter for phased antenna arrays. The phase shifter is composed of a micromachined transmission line periodically loaded with CMOSMEMS 4-bit digital capacitors. MIM capacitors are embedded in the transmission line in series with the digital capacitor bits to trade tuning ratio for increased capacitor Q. Section 5.1 explains why the CMOS-MEMS phase shifter has been developed. Section 5.2 presents the phase shifter topology. Section 5.3 presents an analysis of the topology linking the degrees of freedom to performance metrics. Section 5.4 describes how that analysis was used to define the geometry of the phase shifter and digital capacitors. Section 5.5 presents characterization results of the fabricated design. 5.1. Motivation Phase shifters are required for phased antenna arrays, which consist of many antennas each driven by a phase shifter in series with a power amplifier (see Fig. 5.1). The phase shifters create a variable time delay between the different antennas enabling the direction of the transmitted beam to be steered electronically rather than mechanically. Beam steering is necessary to simply improve the antenna gain for a specific direction or to enable more complex radar applications which require continuous scanning [15]. For example, phased antenna arrays operating at 77 GHz are being developed for automotive collision avoidance systems [41]. Electronic beam steering alleviates the need to mechanically shift or rotate the antenna, which potentially improves both speed and reliability. Chapter 5: Phase Shifter 102 ∆φ1 ∆φ2 ∆φ3 Antennas Power Amplifiers Phase Shifters Fig. 5.1. Schematic of phased antenna array. Phase shifters create a variable time delay between antennas to enable electronic beam steering. Phased antenna arrays can be large, sometimes requiring tens or hundreds of thousands of components per array for military applications, so the cost and power consumption are primary concerns. Both of these characteristics are influenced by the insertion loss of the phase shifters. MEMS phase shifters have been developed to reduce phase shifter insertion loss by using RF MEMS switches and variable capacitors [15]. Reduced phase shifter insertion loss enables a reduction in both the number of required power amplifiers and the power consumption of phased antenna arrays. However, previously developed MEMS phase shifters are not monolithically integrated with CMOS. MEMS phase shifters which are monolithically integrated with CMOS can be combined with on-chip power amplifiers to further reduce the number of components required for phased antenna arrays. Thus the goal of this work was to demonstrate the feasibility of MEMS phase shifters which are monolithically integrated using the CMOS-MEMS process. Using this process, a distributed MEMS phase shifter was fabricated in 0.35 µm BiCMOS with four metal layers. The phase shifter is composed of a micromachined transmission line periodically loaded with 4bit CMOS-MEMS digital capacitors. The digital capacitors, originally presented in Chapter 4, Chapter 5: Phase Shifter 103 were modified for increased Q resulting in an average phase shifter insertion loss of 2.9 dB at 32.0 GHz. Phase Shifter Section T-Line Lsect Cd T-Line Cd Lsect Cl’lsect Lsect Cd Lsect Cd Cl’lsect Cd Cl’lsect Lsect Csect T-Line Cd Lsect Csect Csect Fig. 5.2. Schematic of distributed MEMS phase shifter (for N = 3 sections) as (a) distributed components (b) lumped component (c) lumped components with a single shunt capacitances. 5.2. Topology The CMOS-MEMS phase shifter is a distributed MEMS phase shifter composed of a micromachined transmission line periodically loaded with CMOS-MEMS digital capacitors (see Fig. 5.2(a)). The digital capacitors are spaced evenly along the transmission line such that the phase shifter is composed of a number of identical sections. The variable capacitance of the digital capacitor adds in parallel to the capacitance of the transmission line to form a distributed MEMS transmission line (see Fig. 5.2(b)) [42]. The composite circuit is a discrete transmission Chapter 5: Phase Shifter 104 line with tunable shunt capacitance (see Fig. 5.2(c)). Varying the capacitance changes the phase velocity of the line resulting in a phase shift at the output. Key: Thermal Isolation Electrothermal Actuator Beam Signal Beams Metal 1 Metal 2 Metal 3 Transmission Line Rotor Electrodes Metal 4 Rotor Electrodes Stator Electrodes Stator Electrodes Electrothermal Actuator Beam Thermal Isolation Signal Beams (a) VES RB Port 1 VET VET RHeat CMIM CMIM RHeat Port 2 (b) (c) Fig. 5.3. Diagram of the CMOS-MEMS switched capacitor viewed from above (a) with annotations (b) with circuit schematic; (c) SEM image of eight CMOS-MEMS switched capacitors in parallel. The CMOS-MEMS digital capacitors used in the phase shifter are composed of many CMOS-MEMS switched capacitors that are nearly identical to those described in Chapter 4 (see Fig. 5.3). The dimensions of the electrothermal actuators and capacitive electrodes were left unchanged to achieve the same actuation characteristics presented in Section 4.5.2. The switched capacitors are mirrored on either side of the micro-machined transmission line for symmetry. The stator electrodes are anchored to the transmission line whereas the rotor electrodes are anchored to the metal-4 ground plane. Because only a shunt capacitance is required rather than a series capacitance, the rotor electrodes are electrically connected to the ground plane. Chapter 5: Phase Shifter 105 The CMOS-MEMS phase shifter was designed in a 4-metal, 0.35 µm BiCMOS process from TowerJazz Semiconductor. All four metals are aluminum. Fig. 5.4 shows the resulting cross-sections of the phase shifter after the CMOS-MEMS process. The micromachined transmission line is composed of a metal-4321 beam with the maximum density of vias to minimize series resistance. Near the switched capacitors, the transmission line is split into two segments to ensure that the structure releases from the substrate during the CMOS-MEMS process (see Fig. 5.4(a)). Splitting the line is necessary because the lateral gaps between the topmost metal layers are smaller in the vicinity of the switched capacitors. The majority of the transmission line consists of a single segment since the lateral gaps between the transmission line and the ground plane are large enough to guarantee release (see Fig. 5.4(b)). Transmission Line Actuator Electrode Electrode Actuator ≈ 10 µm ≈ 50 µm (a) Ground Plane Ground Plane Transmission Line ≈ 10 µm ≈ 50 µm (b) Key: Metal 1 Metal 2 Metal 3 Metal 4 Oxide Via Silicon Fig. 5.4. Cross-section of the phase shifter (a) along the rotor electrode of a switched capacitor (b) along a section of transmission line without a switched capacitor. The micromachined transmission line provides high-Q series inductance such that the insertion loss of the phase shifter is primarily dependant on the Q of the digital capacitors. Characterization in Section 4.5.1 showed that the Q of the digital capacitor when used as a shunt Chapter 5: Phase Shifter 106 capacitor is 125 at 1 GHz for the maximum capacitance. The Q increases to 160 at 1 GHz after additional silicon etching is used to remove all silicon beneath the MEMS structure. Assuming series resistance is the dominant loss mechanism, the capacitor Q decreases to 5.3 at 30 GHz, which results in unacceptably large insertion losses. Note that this calculation is an approximation since other factors such as skin depth and losses associated with the silicon substrate affect the capacitor Q. To minimize insertion loss, the digital capacitors were modified for increased Q at frequencies greater than 30 GHz. Bit 1 Port 1 Bit 2 Bit 3 Bit 4 CMIM1 CMIM2 CMIM3 CMIM4 RB1 CMEMS1 RB2 CMEMS4 RB3 CMEMS3 RB4 CMEMS4 VES1 RMEMS1 VES2 RMEMS2 VES3 RMEMS3 VES4 RMEMS4 RHeat1 RHeat2 RHeat3 Port 2 RHeat4 VET Fig. 5.5. Schematic of 4-bit digital capacitor used in the phase shifter. Fig. 5.5 shows the schematic of the modified digital capacitor used in the phase shifter. CMEMS[n] and RMEMS[n] are the capacitance and resistance, respectively, of the CMOS-MEMS electrodes for bits n = 1 to 4. MIM capacitors, CMIM[n] are embedded in the transmission line and connected in series with the MEMS capacitance. For the digital capacitors presented in Chapter 4, the MIM capacitors were connected in parallel to form a DC blocking capacitor with capacitance much larger than the maximum MEMS capacitance. For the phase shifter, the MIM capacitors are sized such that the MIM capacitance is smaller than the maximum MEMS capacitance but much larger than the minimum MEMS capacitance. The net result is a decrease in maximum capacitance which decreases tuning ratio but increases the Q to enable the design of Chapter 5: Phase Shifter 107 a phase shifter with low insertion loss. The MIM capacitors still function as DC blocking capacitors to enable capacitive bits to be independently selected by their respective electrostatic actuation voltage, VES[n]. The bias resistors, RB[n], are large, on the order of MΩ‘s, and composed of polysilicon with a sheet resistance of 1 kΩ. These resistors connect the electrostatic actuation voltages to the stator electrodes. The heating resistors, RHeat[n], are necessary for electrothermal actuation and do not significantly affect RF performance. All digital capacitors are connected such that a single electrothermal actuation voltage, VET, and four electrostatic actuation voltages, one for each bit, control all sections of the phase shifter. 5.3. Analysis The capacitances of the digital capacitor bits, Cd[n], n = 1 to 4, are determined by the capacitances of the MEMS electrodes, CMEMS[n], and the MIM capacitances, CMIMS[n]. Cd[n ] C MEMS[n ]C MIM[n ] C MEMS[n ] C MIM[n ] (5.1) , The total capacitance of the digital capacitor is the sum of the capacitive bits, which varies depending on which bits are activated. Cd Cd[1] Cd[ 2] Cd[ 3] Cd[ 4] , (5.2) Assuming series resistance is the dominant loss mechanism, the Q can be calculated with the effective series resistance, Rs. Qmax 1 C d,min Rs , Chapter 5: Phase Shifter (5.3) 108 Qmin 1 C d,max Rs , (5.4) where ω is the electrical frequency. The maximum Q corresponds to the minimum capacitance while the minimum Q corresponds to the maximum capacitance. The capacitance of the digital capacitors, Cd, adds in parallel with the capacitance of the transmission line resulting in the capacitance of one phase shifter section, Csect. Csect, min Cl ' lsect Cd, min , (5.5) Csect, max C l ' lsect C d, max , (5.6) where Cl‘ and lsect are the transmission line capacitance per unit length and section length, respectively. Assuming that the digital capacitor contributes negligible series inductance, the inductance of one phase shifter section, Lsect, is determined by the transmission line inductance per unit length, Ll‘, and section length. Lsect Ll 'lsect , (5.7) Provided that the frequency of operation is much less than the Bragg frequency, fBragg, the phase shifter performance can be approximated assuming that the variable capacitance is distributed. f Bragg, min 1 Lsect C sect, max , (5.8) Alternatively, the impedance of the transmission line sections can be combined with that of the digital capacitors in ABCD matrices. The resulting matrices can be cascaded to calculate phase shifter performance without approximation [43]. However, for simplicity, the remainder of this analysis assumes the variable capacitance is distributed. The loaded transmission line is also assumed to have low loss. Chapter 5: Phase Shifter 109 Rl ' Ll ' , (5.9) Gl ' Cl ' , (5.10) where Rl‘ and Gl‘ are the series resistance and shunt conductance per unit length of the transmission line, respectively. Circuit simulation is used for the final calculation of phase shift and insertion loss. The characteristic impedance of the loaded transmission line, Z0, varies according to the digital capacitance. Z 0,max Z 0,min Lsect (5.11) C sect,min , Lsect , C sect,max (5.12) The characteristic impedance determines the maximum reflected power, S11,max, which occurs when the total length of the phase shifter is such that the loaded transmission acts as a quarter wavelength transformer. S11,max 20 log 10 Z0 2 ZL 1 2 Z0 2 ZL 1 2 , (5.13) where ZL is the load impedance, which is 50 Ω for this work. Assuming that the reflected power is negligible, the insertion loss, IL, can be calculated based on the number of phase shifter sections, N, the digital capacitor Q, and transmission line characteristics [43]. C d,min Z L lsectZ i ILmin N 2Q ZL max Chapter 5: Phase Shifter , (5.14) 110 C d,max Z L l sectZ i ILmax N 2Q ZL min , (5.15) where α and Zi are the loss per unit length and characteristic impedance, respectively, both corresponding to the unloaded transmission line. The phase constant, β, of the loaded transmission line varies according to the capacitance of the phase shifter sections. min max LsectCsect,min lsect LsectCsect,max lsect , (5.16) , (5.17) The maximum phase shift, ∆φmax, is determined by the difference in achievable phase constants and is proportional to the number of phase shifter sections. max Nlsect max min . (5.18) Using (5.16) and (5.17) to expand (5.18) shows the maximum phase shift in terms of the capacitance and inductance of each phase shifter section. max N Lsect Csect,min Csect,max . (5.19) Note that the phase shift does not scale linearly with increasing capacitance, but rather as difference of two square roots. The phase shift scales linearly with frequency, which indicates a constant time delay rather than a constant phase shift over a given frequency range. 5.4. Design The micromachined transmission line was designed to maximize transmission line quality factor, QTL, which minimizes loss for a given phase shift. The unloaded characteristic impedance Chapter 5: Phase Shifter 111 was chosen to be approximately 100 Ω, because the addition of digital capacitors reduces the loaded characteristic impedance according to (5.11) and (5.12). The transmission line dimensions were determined through iterative simulations with HFSS. The signal line was made 20 µm wide with 20 µm air gaps to the ground plane. The line consists of all four metal layers with the maximum density of vias. Mechanical supports are placed periodically along the line to minimize deflection caused by the vertical force exerted by the electrothermal actuators. The simulated Sparameters of the micromachined transmission line were used to calculate the transmission line parameters at 30 GHz (see Table 5.1). The unloaded characteristic impedance is 98.8 Ω, the loss per unit length is 0.07 dB/mm, and the transmission line QTL is 43 at 30 GHz. Table 5.1: Phase Shifter Design Parameters Parameters Value Calculation Value CMEMS[1],min* 0.85 fF Cd[1],min 1.44 fF CMEMS[2],min* 0.85 fF Cd[2],min 1.57 fF CMEMS[3],min* 1.71 fF Cd[3],min 3.13 fF CMEMS[4],min* 3.41 fF Cd[4],min 6.26 fF CMEMS[1],max* 4.42 fF Cd[1],max 1.44 fF CMEMS[2],max* 4.42 fF Cd[2],max 1.57 fF CMEMS[3],max* 8.84 fF Cd[3],max 3.13 fF CMEMS[4],max* 17.7 fF Cd[4],max 6.26 fF CMIM[1] 2.02 fF Cd,min** 18.6 fF CMIM[2] 6.84 fF Cd,max** 123 fF CMIM[3] 12.7 fF Qmax 123 CMIM[4] 24.7 fF Qmin 18.6 l’ 540 µm Csect,min 33.7 fF N 8 Csect,max 138 fF ω 30 GHz Lsect 189 pH ZL 50 Ω fBragg,min 62 GHz R' 1.19 Ω/mm Z0, max 75 Ω L' 350 pH/mm Z0, min 37 Ω C' 36.0 fF/mm S11,max -10 dB G' 39.1 MΩ-1/mm ILmin 0.7 dB Zi 98.8 Ω ILmax 2.4 dB α 0.069 dB/mm βmin 0.94 rad/mm β 0.689 rad/mm βmax 1.90 rad/mm QTL 43.4 ∆φmax 223.8° * Parameter extracted from measured data ** Includes parasitic capacitance from mechanical supports Chapter 5: Phase Shifter 112 Using the simulated transmission line characteristics, the digital capacitors were designed to shift the loaded characteristic impedance from 75 Ω to 37 Ω to ensure that S11 remained below −10 dB according to (5.13). The characteristic impedance, determined by (5.11) and (5.12), was controlled by sizing the MIM capacitors to create a 6.6:1 tuning ratio according to (5.1). Since the MIM capacitors are embedded in the micromachined transmission line, there is a significant oxide capacitance that adds in parallel with each of the MIM capacitances. The MIM capacitances listed in Table 5.1 include the contributions of the oxide capacitances, which were calculated using finite element analysis. Using (5.4), the capacitor Q was calculated to be 18.6 at 30 GHz for the calculated maximum capacitance of 123.3 fF. Fig. 5.6 shows one half of the fabricated digital capacitors attached to the micromachined transmission line. The other half of the digital capacitor is mirrored on the other side of the transmission line. Bit 1 D Bit 2 Bit 3 Bit 4 D 100 µm MIM Capacitors Actuators Electrodes Fig. 5.6: SEM image of phase shifter digital capacitor. The rotor electrodes are vertically deflected into the minimum capacitance state. ―D‖ indicates dummy bits which are included for temperature uniformity. The length of one phase shifter section (see Fig. 5.7) was set to 540 µm resulting in a minimum Bragg frequency of 62 GHz according to (5.8). The number of sections was limited to eight by the size of the CMOS die. Using (5.18), (5.14), and (5.15), the maximum phase shift and Chapter 5: Phase Shifter 113 average insertion loss were calculated to be 223.8° and 1.5 dB at 30 GHz. Note that the calculated phase shift is less than 337.5°, which is the minimum phase shift required for sixteen steps distributed linearly between 0° and 360°. However, the phase shift can be scaled by increasing the number of sections, so a calculated phase shift of 223.8° was acceptable for characterization. The fabricated phase shifter measures 4320 µm by 560 µm, a total area of 2.42 mm2 (see Fig. 5.8). Table 5.1 list all the parameters used for design of the phase shifter. Digital Capacitors Transmission Line Ground Plane Mechanical Supports 100 µm Fig. 5.7: SEM image of one CMOS-MEMS phase shifter section. Phase Shifter Section G S G 540 µm 560 µm G S G 1 mm Fig. 5.8: SEM image of the CMOS-MEMS phase shifter composed of eight phase shifter sections. 5.5. Characterization To verify the phase shifter design, both the phase shifter and a single copy of the digital capacitor used in the phase shifter were characterized. All digital capacitors were controlled Chapter 5: Phase Shifter 114 using the actuation methods presented in Chapter 4 except that capacitive bits were independently selected using multiple electrostatic actuation voltages rather than multiple electrothermal actuation voltages (see Fig. 5.9). Probe pads were included to access the heater resistors for the dummy bits independently from those of the capacitive bits in case the dummy bits required more power to reach a uniform temperature. In hindsight, multiple electrothermal actuation voltages were unnecessary because a single electrothermal actuation voltage applied to both the capacitive bits and dummy bits proved sufficient for actuation. The digital capacitors required 4.8 V for electrothermal actuation and 20 V for electrostatic actuation. During electrothermal actuation, a single digital capacitor consumed 40 mW while the phase shifter consumed 320 mW. Each of the eight digital capacitors in the phase shifter operated identically, which qualitatively demonstrates the uniformity of the electrothermal actuators. (a) (b) Fig. 5.9: Microscope image of CMOS-MEMS digital capacitor (a) before actuation (b) after actuation in state 10 (1010) with bits 4 and 2 turned on. 5.5.1. Digital Capacitor Measurements The digital capacitor used in the phase shifter was electrically characterized using the method described in Appendix A for all sixteen capacitor states. Fig. 5.10 shows the Sparameters for the minimum capacitance and maximum capacitance states. Fig. 5.11 shows the equivalent circuit model in terms of Z-parameters and lumped circuit elements. Fig. 5.11 shows Chapter 5: Phase Shifter 115 the lumped impedances for the minimum capacitance and maximum capacitance states. The equations listed below were used to calculate the equivalent circuit element values (see Fig. 5.11(b)). Cg 1 imag Z 21 . (5.20) Rg realZ 21 . Ls1 imag Z 11 Z 12 (5.21) . (5.22) Rs1 realZ11 Z12 . Ls2 imag Z 22 Z 21 (5.23) . (5.24) Rs2 realZ 22 Z 21 . imag Z 11 realZ 11 10 10 0 0 S-parameters (dB) S-parameters (dB) Q S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 (5.25) . (5.26) S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 -40 -40 0.1 1 10 0.1 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 5.10. S-parameter magnitude as a function of frequency for (a) the minimum capacitance state (b) the maximum capacitance state. S12 is not plotted because S12 and S21 are equivalent for passive networks. Chapter 5: Phase Shifter 116 Port 1 Z11-Z12 Z22-Z21 Port 2 Port 1 Port 2 Rs1 Ls1 Rg Rs2 Z21 Ls2 Cg (b) (a) 1E+5 1E+5 1E+4 1E+4 Impedance (Ω) Impedance (Ω) Fig. 5.11. Equivalent circuit model in terms of (a) Z-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. |Z21| (Measured) |(Z11-Z12)| (Measured) |(Z22-Z21)| (Measured) Extracted Model 1E+3 1E+2 1E+1 1E+3 1E+1 1E+0 1E+0 1E-1 1E-1 0.1 1 10 |Z21| (Measured) |(Z11-Z12)| (Measured) |(Z22-Z21)| (Measured) Extracted Model 1E+2 0.1 1 10 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 5.12. Lumped impedances as a function of frequency for (a) the minimum capacitance position (b) the maximum capacitance position. The series impedance behaves like an RLC circuit, and the shunt impedances are capacitive below 10 GHz. The self-resonance frequency is greater than 40 GHz for the minimum capacitance position and decreases to 10 GHz for the maximum capacitance position. The electrical characteristics of the digital capacitor used in the phase shifter are presented in Table 5.2. The state indicates which of the capacitive bits were active (for example, state 12 (1100) means that the two most significant bits are active). Measurement uncertainties are not available since only a single device was measured due to lack of available CMOS-MEMS chips. As designed, the shunt capacitance, Cg, increases linearly with state. Although there must be an inductance in series with the shunt capacitance, no such inductance was measured because the self-resonance frequency of the shunt impedance was greater than 40 GHz. The series inductances, L1s and L2s, and series resistances, R1s and R2s, change slightly with state because the different capacitive bits are located at different distances along the transmission line between port Chapter 5: Phase Shifter 117 1 and port 2. The total sum of both series inductance and resistance stays approximately constant with changing state. Table 5.2: Electrical Characteristics of the Phase Shifter Digital Capacitor State Cg (fF) Rg (Ω) Ls1 (pH) R1s (Ω) Ls2 (pH) 0 (0000) 15.5 NA* 97.4 NA* 80.8 1 (0001) 24.2 NA* 81.9 NA* 96.7 2 (0010) 32.1 3.31 80.5 0.53 98.0 3 (0011) 37.9 3.08 76.1 0.52 102.5 4 (0100) 46.3 2.02 86.7 0.69 91.5 5 (0101) 54.1 1.92 81.7 0.69 96.7 6 (0110) 59.3 1.60 81.3 0.73 96.9 7 (0111) 64.6 1.55 78.7 0.73 99.6 8 (1000) 69.9 1.26 109.6 1.23 69.1 9 (1001) 77.2 1.10 104.0 1.18 75.7 10 (1010) 84.3 0.95 101.4 1.18 78.1 11 (1011) 89.3 0.91 98.6 1.17 81.3 12 (1100) 93.4 0.79 101.1 1.21 77.5 13 (1101) 102.1 0.65 97.5 1.18 81.8 14 (1110) 106.6 0.59 96.7 1.20 82.5 15 (1111) 111.6 0.54 94.5 1.20 85.1 * Resistance could not reliably be extracted from the small capacitance R2s (Ω) NA* NA* 1.53 1.58 1.44 1.49 1.48 1.51 1.10 1.22 1.26 1.31 1.28 1.37 1.38 1.43 Q (30 GHz) NA* NA* 63.3 46.2 46.6 38.8 38.6 35.2 25.4 24.5 23.3 22.0 21.0 20.3 19.4 18.8 The Q reported in Table 5.2 is the Q for the digital capacitor with port 2 unconnected, or left floating. This Q calculation differs from that of the variable capacitors in Chapter 3 and Chapter 4 because those capacitors were connected in series between the two ports, which requires one port to be grounded to produce a capacitor with a single terminal. Since the phase shifter digital capacitor already has one terminal connected to ground, one port must be left floating to produce a capacitor with a single terminal. Note that, unlike the updated digital capacitor presented in Section 4.5.5, the measured Q decreases significantly with increasing state. This change in Q occurs because the loss due to the series resistances, R1s and R2s, is comparable to the loss due to the resistance of each capacitive bit, Rg, such that the total series resistance remains roughly constant with state. Fig. 5.13 shows the measured Q for the maximum capacitance state as a function of Chapter 5: Phase Shifter 118 frequency. To calculate the extracted model, the self-resonance frequency was assumed to be 55 GHz, which provides the best fit between measurement and calculation. The measured Q is noisy even though measurements were recorded with the IF bandwidth of the VNA lowered to 500 Hz. presumably, the measurement is simply limited by the accuracy of the VNA. The noise increases at low frequencies as the series resistance becomes smaller in comparison with the Quality Factor imaginary impedance. 100 90 80 70 60 50 40 30 20 10 0 Quality Factor (De-embedded) Extracted Model 10 20 30 40 Frequency (GHz) Fig. 5.13. Quality factor as a function of frequency for the digital capacitor used in the phase shifter. This plot shows the Q for the maximum capacitance position with port 2 left floating. The series capacitance of the digital capacitor was measured at each of the sixteen states for a single device (see Fig. 5.14). The measured capacitance as a function of state matches a linear trend with an R2 of 0.9963. The digital capacitor tunes from 15.5 fF to 111.6 fF, a tuning ratio of 7.2:1. Device variation in terms of standard deviation is not available because only a single device was measured. Chapter 5: Phase Shifter 119 120 Capacitance (fF) 100 80 60 40 20 0 0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 State Fig. 5.14: Capacitance of phase shifter digital capacitor as a function of state. State 3 (0011) indicates that the two least significant bits are turned on. 5.5.2. Phase Shifter Measurements The phase shifter was electrically characterized using the method described in Appendix A. S-parameters were measured for each of the sixteen states. The phase shift for a given state was calculated by subtracting the phase of S21 for that state from the phase of S21 for State 0. state[x] phaseS 21, state[0] phaseS 21, state[x] . (5.27) At 32.0 GHz, the phase shifter achieves a maximum measured phase shift of 337.5° for state 15 (see Fig. 5.15). The insertion loss is calculated as the magnitude of S21. ILstate[x] 20 log 10 S 21,state[x] . (5.28) The insertion loss varies by state from 1.9 dB to 3.9 dB with an average of 2.9 dB at 32.0 GHz. The return loss, RL, is calculated as the magnitude of S11. RLstate[x] 20 log 10 S11,state[x] . (5.29) The return loss varies by state since changing the capacitance determines the characteristic impedance of the distributed MEMS transmission line. The maximum return loss is -5.4 dB at Chapter 5: Phase Shifter 120 30 GHz. The measured electrical characteristics of the phase shifter at 32 GHz are summarized in Table 5.3. The characteristics are presented at 32 GHz since that frequency is the minimum at which the phase shifter produces sixteen steps between 0° and 360° phase shift. Phase Shift (Degrees) 0 -45 -90 -135 -180 -225 Increasing state -270 -315 -360 0 10 20 30 Frequency (GHz) 40 Fig. 5.15: Measured phase shift as a function of frequency for all 16 states. Phase shift is negative because the time delay increases with increasing capacitance. Insertion Loss (dB) 0 -0.5 -1 -1.5 -2 -2.5 -3 Increasing state -3.5 -4 -4.5 -5 0 10 20 30 40 Frequency (GHz) Fig. 5.16: Measured insertion loss as a function of frequency for all 16 states. Chapter 5: Phase Shifter 121 0 Increasing state Return Loss (dB) -5 -10 -15 -20 -25 -30 -35 -40 0 10 20 30 40 Frequency (GHz) Fig. 5.17: Measured reflection as a function of frequency for all 16 states. TABLE 5.3: ELECTRICAL CHARACTERISTICS OF THE PHASE SHIFTER State 0 (0000) 1 (0001) 2 (0010) 3 (0011) 4 (0100) 5 (0101) 6 (0110) 7 (0111) 8 (1000) 9 (1001) 10 (1010) 11 (1011) 12 (1100) 13 (1101) 14 (1110) 15 (1111) Reflection |S11| (dB) -6.1 -13.5 -19.4 -14.9 -13.3 -14.8 -18.3 -21.0 -23.3 -28.7 -32.9 -52.2 -32.4 -23.1 -19.5 -17.9 Insertion Loss |S21| (dB) -2.7 -2.0 -1.9 -2.2 -2.3 -2.6 -2.6 -2.8 -2.8 -3.1 -3.2 -3.4 -3.4 -3.6 -3.7 -3.9 Phase Shift ∆φ (°) 0 -40.3 -75.8 -98.0 -127.2 -154.0 -173.0 -189.3 -206.1 -231.3 -254.4 -269.5 -285.7 -307.6 -322.8 -337.5 Phase Step ∆φn-∆φn-1 (°) NA -40.3 -35.8 -22.2 -29.2 -26.8 -19.0 -16.3 -16.8 -25.2 -23.1 -15.1 -16.2 -21.9 -15.2 -14.7 The de-embedded S-parameters for the phase shifter in state 0, zero phase shift, and state 15, maximum phase shift, are plotted in Fig. 5.18. Note that the measured data does not match that of the initial simulations, particularly for S11, which was designed to remain below −10 dB. Analysis of the digital capacitor revealed additional series resistance and inductance which were not included in the original model. Presumably, these parasitics occur because the digital Chapter 5: Phase Shifter 122 capacitors periodically break up the ground plane resulting in a longer return path for the current. The additional resistance increased the insertion loss, and the additional inductance increased the characteristic impedance resulting in a worse match, or larger S11, for state 0 and a better match, or smaller S11, for state 15. Re-simulation of the phase shifter accounting for these parasitics resulted in closer agreement between measurement and simulation. 0 -5 |S21| -10 -15 -20 -25 -30 -35 -40 |S11| Measured Simulated Re-simulated S-Parameters (dB) S-Parameters (dB) 0 -5 -10 -15 -20 -25 -30 -35 -40 Measured Simulated Re-simulated |S21| |S11| 0 10 20 30 40 20 30 40 Frequency (GHz) Frequency (GHz) (a) (b) Fig. 5.18. S-parameter magnitude as a function of frequency for (a) the minimum phase shift state (b) the maximum phase shift state. S12 is not plotted because S12 and S21 are equivalent for passive networks. 0 10 5.6. Discussion The CMOS-MEMS phase shifter is compared with a representative sample of other passive phase shifters, both MEMS and non-MEMS, in Table 5.4. The frequencies listed in Table 5.4 are the minimum frequencies at which the phase shifters produce an even distribution of phase shifts up to 360°. For digital phase shifters, the insertion loss is averaged over all states. For analog phase shifters, the insertion loss is recorded for a 180° phase shift. The insertion loss of the phase shifter in this work, 2.9 dB, is less than that of GaAs phase shifters but greater than that of other MEMS phase shifters (see Fig. 5.19). However, the CMOSMEMS phase shifter is the first MEMS phase shifter which can be monolithically integrated with Chapter 5: Phase Shifter 123 CMOS using only maskless post-processing. Despite the increased insertion loss, this work represents a significant achievement because only the MEMS phase shifter in this work could be combined with on-chip CMOS power amplifiers to further reduce the number of components required for phased antenna arrays. Insertion Loss (dB) Table 5.4: Comparison of Phase Shifters Frequency Insertion Reference MEMS? Type # Bits (GHz) Loss (dB) [44] yes Reflect-Line 7 4 1.2 [45] yes Switched-Line 10 4 2.2 [46] yes Switched-Line 10 4 1.1 [47] yes Switched-Line 32 4 2.3 yes [48] Distributed 14 2 1.2 yes [49] Distributed 16 4 3 yes [50] Distributed 26 3 1.7 yes This work Distributed 32 4 2.9 yes [48] Distributed 38 2 1.5 yes [51] Distributed 60 2 2.2 yes [52] Distributed 94 analog 2.5 no [50] Distributed 6 analog 4 no [53] Distributed 6 analog 5.1 no [54] Distributed 20 analog 4.2 * Resistance could not reliably be extracted from the small capacitance 8 7 6 5 4 3 2 1 0 Switched-Line Distributed This Work Substrate Si GaAs GaAs Si SiO2 Si SiO2 CMOS SiO2 SiO2 SiO2 GaAs GaAs GaAs Reflect-Line GaAs 0 10 20 30 40 50 60 70 80 90 100 Frequency (GHz) Fig. 5.19: Comparison of passive phase shifters. The average insertion loss is plotted versus the minimum operating frequency. The insertion loss of the CMOS-MEMS phase shifter was larger than expected due to the Chapter 5: Phase Shifter 124 series resistance of the digital capacitor, measured as R1s and R2s (see Table 5.2). This series resistance cannot be accounted for by the sheet resistance of the metal layers and is actually frequency dependant, which suggests another source of loss. Further investigation of this loss would require finite element analysis to account for how the digital capacitors break up the ground plane of the transmission line and force the current to return through a longer path. The bias resistors also warrant additional investigation since lumped resistors were used for simulation. The bias resistors could be more accurately represented as a series resistance with a distributed shunt capacitance to ground. This representation would add a low-Q parasitic capacitance to simulation which may be responsible for the additional insertion loss. Further optimization of the phase shifter design could also improve the insertion loss. The design presented in Section 5.4 specifies the digital capacitor design after determining the geometry of the micromachined transmission line. If the transmission line loss and phase velocity can be parameterized, then the transmission line dimensions can be chosen specifically to minimize insertion loss [43]. Such parameterization would require further simulation with HFSS of transmission lines with varying dimensions. The only reason this optimization was not performed was lack of available time. The power consumption required for electrothermal actuation is potentially problematic. Power is only consumed during the 1 ms switching time, resulting in 0.32 mJ consumed per actuation cycle. This energy consumption is beneficial if the phase shifter is not switched often but may limit this particular phase shifter to applications which do not require continuous scanning. Thus a CMOS-MEMS phase shifter based purely on electrostatic actuation could enable applications that require continuous scanning. The use of only electrostatic actuation would also provide faster switching times. Chapter 5: Phase Shifter 125 Chapter 6. Conclusions This chapter begins with Section 6.1, a summary of the three contributions of this work, the CMOS-MEMS gap-tuning capacitor, the CMOS-MEMS digital capacitor, and the CMOSMEMS phase shifter. Next, the thesis statement is revisited and evaluated in Section 6.2 by comparing the CMOS-MEMS variable capacitors of this work with other state-of-the-art variable capacitors. Section 6.3 discusses several possibilities for future work on CMOS-MEMS variable capacitors. Finally, concluding remarks on the viability of RF CMOS-MEMS are presented in Section 6.4. 6.1. Summary Prior to this work, MEMS variable capacitors had been developed performance superior to that of semiconductor variable capacitors. The increased tuning ratios and higher quality factors made possible by MEMS are desirable for enhancing existing RF circuits and enabling new frequency-reconfigurable circuits. However, MEMS variable capacitors are not widely used in RF circuits because most MEMS are not monolithically integrated with CMOS. Those MEMS which are integrated require potentially expensive fabrication processes. To achieve monolithic integration without expensive fabrication processes, this work used the CMOS-MEMS process to create CMOS-MEMS variable capacitors. The first CMOS-MEMS variable capacitor presented in this work is the gap-tuning capacitor that uses lateral electrothermal actuation to control the capacitive gap between interdigitated electrodes. A mechanical latch, also controlled by electrothermal actuation, was used to maintain the capacitive gap without consuming static power for heating. The routing for the RF signal was separated from that of the actuation signal to minimize parasitic capacitance to ground. The Chapter 6: Conclusions 126 structure was designed to achieve vertical alignment between the capacitive beams. The gaptuning capacitor achieved series and shunt tuning ratios up to 6.9:1 and 5.7:1, respectively, with a Q of 80 at 1 GHz for the maximum capacitance. The switching time was measured to be 30 ms. Millions of actuation cycles were demonstrated without failure. Because the tuning ratio and Q are greater than those of semiconductor variable capacitors, the gap-tuning capacitor was fundamentally successful. However, practical implementation of circuits using the gap-tuning capacitor was difficult due to the large variation in maximum capacitance, the non-linear distribution of latched capacitance values, and the low capacitance density. The second CMOS-MEMS variable capacitor presented in this work is the digital capacitor, which was designed to overcome the limitations of the gap-tuning capacitor. The digital capacitor consists of many CMOS-MEMS switched capacitors organized into independently controlled capacitive bits. The capacitive bits are sized in increasing powers of two to produce a linear distribution of capacitance values. The switched capacitors use a combination of vertical electrothermal actuation and lateral electrostatic actuation to achieve large tuning ratios without sacrificing capacitance density. Because the minimum capacitance state is maintained by residual stress and the maximum capacitance state is maintained by electrostatic force, power for heating is only consumed during actuation. The initial digital capacitor topology was updated to minimize parasitic capacitance to ground. The updated digital capacitor achieved series and shunt tuning ratios up to 73:1 and 19:1, respectively, with a Q of 369 at 1 GHz for the maximum capacitance. Compared with the gap-tuning capacitor, the capacitance density was increased by 5x and standard deviation in maximum capacitance was reduced from 18% to 4.4%. The switching time was decreased to 1 ms. Reliability tests revealed a drift in the actuation characteristics after 10 million cycles due to thermally-induced mechanical fatigue, but no Chapter 6: Conclusions 127 catastrophic failures occurred. These characteristics represent the state-of-the-art for CMOSMEMS variable capacitors. The CMOS-MEMS phase shifter was developed as an application of CMOS-MEMS variable capacitors. The CMOS-MEMS phase shifter consists of a micromachined transmission line periodically loaded with CMOS-MEMS digital capacitors. The digital capacitors are sufficiently distributed such that the circuit can be approximated as a transmission line with variable capacitance per unit length. The digital capacitor was modified to enable operation at frequencies beyond 30 GHz. The phase shifter digital capacitor achieved a shunt tuning ratio of 7.2:1 with a Q of 18.8 at 30 GHz for the maximum capacitance. The phase shifter produced an average insertion loss and maximum phase shift of 2.9 dB and 337.5°, respectively, at 32 GHz. This work represents the first MEMS phase shifter to be monolithically integrated with CMOS. 6.2. Contributions The first goal of this work was to show that, using the CMOS-MEMS process, CMOSMEMS variable capacitors can be designed with performance comparable to that of other integrated MEMS variable capacitors. To work towards this goal, CMOS-MEMS variable capacitors were designed and characterized. Table 6.1 compares the characteristics of the best performing variable capacitor of this work, the updated CMOS-MEMS digital capacitor presented in Section 4.5.5, with that of several other MEMS variable capacitors capable of monolithic integration with CMOS. Also included are the characteristics of one type of semiconductor variable capacitor, the switched MIM capacitor simulated using the same CMOS technology, TowerJazz 0.35 µm BiCMOS. The other two types of semiconductor variable capacitors, the varactor diode and MOS varactor, are not included because, although their Q is Chapter 6: Conclusions 128 higher than that of switched MIM capacitors, their tuning ratios are limited to less than 5:1. A figure of merit (FoM), the product of the shunt tuning ratio, TRshunt, and Q, is presented for comparison between variable capacitors with different tuning ratios. FoM TRshunt * Q . (6.1) This FoM remains constant if, by decreasing the maximum capacitance, tuning ratio is decreased to increase Q. Parameter Type Integration Table 6.1: Comparison of Integrated Variable Capacitors Switched MIM University of Raytheon This Work Capacitor Waterloo [55] Systems [29] MEMS Semiconductor MEMS MEMS Maskless postMaskless post- Built on top of CMOS CMOS etching CMOS etching silicon 1.34 pF 1.11 pF 1.64 pF 33.2 pF 18.3 fF 111 fF 290 fF 1.5 pF Cmax Cmin Shunt Tuning 19:1 10:1 5.6:1 Ratio Q at 1 GHz* 369 9 ≈ 180 FoM 7011 90 1008 (TRshunt*Q) Self-Resonance 10 GHz NA > 20 GHz Frequency* Capacitance 7.9 pF/mm2 1000 pF/mm2 ≈ 6.6 pF/mm2 Density Switching Electrothermal / MOSFET Electrostatic Method electrostatic Tuning Range Discrete Discrete Discrete Switching Time 1 ms ~ 1 ns NA Actuation 4 V / 20 V 3.3 V 70 V Voltage Actuation 50 µJ ≈ 0** ≈ 0** Energy for 1 pF * The parameter corresponds to the maximum capacitance state. ** The energy consumed for actuation is negligible. WiSpry [38] MEMS Built on top of CMOS 8 pF 720 fF 22:1 11:1 < 20 160 < 440 1760 NA > 5 GHz NA NA Electrostatic Electrostatic Discrete < 10 µs Discrete < 10 µs 55 V 30 V ≈ 0** ≈ 0** Based on the FoM presented in Table 6.1, the CMOS-MEMS digital capacitor achieves a superior combination of tuning ratio and Q compared with that of other MEMS variable capacitors capable of monolithic integration with CMOS. The self-resonance frequency is sufficiently high for use at mobile communications frequencies (1 GHz – 6 GHz). Compared to Chapter 6: Conclusions 129 MEMS variable capacitors using only electrostatic actuation, the digital capacitor has a longer switching time consumes power during actuation. However, if the digital capacitor is used in frequency-reconfigurable circuits that do not require continuous switching, the slow switching time and transient power consumption are not significant problems. Thus this work has satisfied the first goal of the thesis statement by showing that CMOS-MEMS variable capacitors can be designed with performance comparable to that of other integrated MEMS variable capacitors The second goal of this work was to show that, CMOS-MEMS variable capacitors enable the design of reconfigurable RF circuits. To prove this statement, the first CMOS-MEMS phase shifter was designed using CMOS-MEMS variable capacitors. The phase shifter operated at frequencies greater than 30 GHz, which demonstrates the potential for CMOS-MEMS to work at even higher frequencies. The performance of the phase shifter, measured by the insertion loss, is close to that of other MEMS phase shifters and better than that of semiconductor phase shifters. Therefore, this work has also satisfied the second goal of the thesis statement by showing that CMOS-MEMS variable capacitors enable the design of reconfigurable RF circuits. 6.3. Future Work CMOS-MEMS variable capacitors have come a long way since their inception, but there are still options for future work. The tuning ratio and Q have been significantly improved and are superior to those of semiconductor variable capacitors, so future work should focus on improving the characteristics which are inferior to those of semiconductor variable capacitors, specifically capacitance density, switching time, reliability, and actuation voltage. This section presents suggestions for further improvement based on process modifications, topology improvements, and applications. Chapter 6: Conclusions 130 6.3.1. CMOS-MEMS Process Modifications Because this work focused on the design of CMOS-MEMS variable capacitors, the CMOSMEMS process was not modified. Process modifications were neglected in lieu of device design because, at the time of this work, the CMOS-MEMS process was well established and optimized for general purpose CMOS-MEMS. However, there are several options for future work related to the CMOS-MEMS process which will specifically improve CMOS-MEMS variable capacitors. A simple way to significantly improve CMOS-MEMS variable capacitors is to decrease the thickness of the sidewall polymer deposited during the vertical oxide etch. Since this polymer determines the maximum capacitance per unit length for two CMOS-MEMS beams pushed into mechanical contact, any decrease in thickness would increase both the tuning ratio and capacitance density of CMOS-MEMS variable capacitors. Control over this thickness has previously been limited by the capabilities of the Plasma-Therm 790 RIE. Work is currently underway to use the STS Aspect Advanced Oxide Etcher to perform the vertical oxide etch while depositing less sidewall polymer. However, there are risks associated with depositing too little sidewall polymer because this polymer prevents the titanium-tungsten barrier metals from etching during the silicon release etch. Furthermore, if the polymer is too thin, the breakdown voltage between two CMOS-MEMS beams pushed into mechanical contact may become too low (< 20 V) for electrostatic pull-in. If these potential pitfalls can be averted, reducing the thickness of the sidewall polymer would significantly improve any CMOS-MEMS capacitive devices. Another method to improve CMOS-MEMS variable capacitors is to increase the lengths of both the vertical silicon etch and the isotropic etch. Using the current CMOS-MEMS process, the silicon etches create a substantial air gap (≈ 50 µm) between the MEMS structure and the remaining silicon substrate. As large as this gap is, there is still significant parasitic capacitance Chapter 6: Conclusions 131 (≈ 10 fF) that couples the MEMS structure to the silicon. Because the silicon is a poor conductor at RF, this parasitic capacitance decreases the Q of CMOS-MEMS variable capacitors. Capacitance to the silicon can be avoided by including grounded metal near all capacitive electrodes but, similar to adding a ground shield beneath an inductor, adding grounded metal increases the total parasitic capacitance, which causes the shunt tuning ratio to decrease. If more silicon is removed, capacitance to the silicon can be minimized without increasing the total amount of parasitic capacitance. The third suggestion related to processing is not a process modification but instead regards unintended lateral curling, which is believed to be caused by the CMOS-MEMS process. Because unintended lateral curling of CMOS-MEMS beams is problematic for nearly all CMOSMEMS devices, a better understanding of this curling is needed to help the designer compensate accordingly. Currently, information concerning this curling is based on qualitative observations. Since even CMOS-MEMS beams of a single metal layer exhibit lateral curling, the phenomenon is not entirely caused by misalignment of the metal layers. The sidewall polymer is believed to contribute to lateral curling because sidewall polymer that has delaminated curls to a greater extent than CMOS-MEMS beams. Thus unintended lateral curling may be alleviated by reducing the thickness of the sidewall polymer. If not, statistical measurements of curling from multiple CMOS-MEMS chips and multiple CMOS process runs are needed to fully understand the phenomenon. Perhaps an upper bound for the magnitude of lateral curling can then be predicted given the length and cross-section of a CMOS-MEMS beam. 6.3.2. Device Topology Improvements The CMOS-MEMS variable capacitors presented in this work have three main Chapter 6: Conclusions 132 disadvantages when compared with semiconductor variable capacitors: low capacitance density, slow switching speed, and limited reliability. Capacitance density is nearly optimized by the digital capacitor topology since the capacitive electrodes consist of a dense array of CMOSMEMS beams. Thus capacitance density should be improved through process modifications, such as a reducing the thickness of the sidewall polymer, rather than changes to the device topology. The slow switching speed and limited reliability are both a result of electrothermal actuation. Both of these characteristics could be improved by changes to the device topology. If the actuation can be accomplished using only electrostatic actuation without electrothermal actuation, the switching speed will increase and the reliability will not be limited by thermally induced mechanical fatigue. The sensitivity of the variable capacitance to changes in ambient temperature should also decrease significantly. However, achieving the same tuning ratio, Q, and capacitance density without electrothermal actuation will be difficult since electrothermal actuation enables large vertical displacements. Because the CMOS-MEMS process cannot produce a top and bottom electrode, there is no obvious method for vertical electrostatic actuation. One method for achieving vertical electrostatic actuation without a top electrode involves the use of fringing fields to pull vertically curled capacitive electrodes into the same plane [37]. Perhaps this actuation method can be modified to also include lateral electrostatic pull-in for achieving a large tuning ratio. Another technique that may be useful is a method to produce upwards vertical deflection using electrostatic force without a top electrode. By pulling one end of a MEMS beam down, the angle of the beam can be changed causing the other end to tilt upwards [8]. These two actuation techniques can potentially be combined to create a digital capacitor using vertical and lateral electrostatic actuation. If a device topology can be developed to use only electrostatic force, the second challenge Chapter 6: Conclusions 133 will be minimize the required actuation voltage. The high voltages (> 20 V) typically required for electrostatic actuation are undesirable because those voltages can cause dielectric charging, which eventually causes the capacitive electrodes to stick together even when the applied voltage is reduced to zero. Furthermore, the CMOS process must include high-voltage transistors so that those high voltages can be generated on-chip with integrated charge pumps. One idea to decrease the required actuation voltage is to use resonant electrostatic actuation [56]. Rather than applying only a DC voltage, a DC voltage can be combined with an AC voltage at the mechanical resonance frequency. Because sinusoidal motion at resonance is amplified by the mechanical Q, the necessary displacements can be achieved using lower voltages. The switching speed is slowed by the need for resonance but could still switch faster than electrothermal actuation. 6.3.3. Potential Applications In principle, any circuit that requires variable capacitors can be improved by replacing the semiconductor variable capacitors with CMOS-MEMS variable capacitors. However, the improvements to a given circuit must be sufficient to justify the large area consumption, slow switching times, and added complexity inherent to MEMS. Frequency-reconfigurable circuits, which use variable capacitors to select the frequency band of operation, are an example of a good application of CMOS-MEMS variable capacitors. Circuits that perform band selection through reconfiguration consume less area than those that perform reconfiguration by switching between multiple versions of the same circuit. Furthermore, fast switching times are necessarily required for frequency-reconfigurable circuits. A good application of CMOS-MEMS variable capacitors that has yet to be explored is the impedance tuner. Impedance tuners are circuits that maintain the impedance match between the output Chapter 6: Conclusions 134 network of a power amplifier and the antenna. Such circuits are necessary to extend battery life in mobile handsets because the impedance of an antenna is affected by the local surroundings. Impedance tuners require tunable passive components with high Q to minimize loss because these tuners are placed in series between the power amplifier and antenna. Fast switching times are not required since the antenna impedance changes slowly in response to manual manipulation of mobile handsets. Because the CMOS-MEMS process consists of maskless processing, a CMOS-MEMS impedance tuner could extend battery life with little extra fabrication cost. 6.4. Final Thoughts The field of RF MEMS is odd because there is no fundamental reason why mechanical devices are needed to provide the functions of RF MEMS. RF MEMS devices typically involve transduction of an electrical control signal into a mechanical displacement that determines another electrical quantity. This use of the mechanical domain as an intermediary between electrical signals is at first glance unnecessary since variable capacitance, switching, and tunable resonance can all be provided by semiconductor devices. In contrast, MEMS is a natural choice for inertial sensors since the purpose of sensors such as accelerometers and gyroscopes is to sense a mechanical quantity. Because the mechanical domain is not strictly required to provide the functions of RF MEMS, RF MEMS devices are in direct competition with semiconductor devices. For now, RF MEMS are competitive because MEMS can be constructed entirely from metal, without the need for resistive semiconductor materials. But commercialization of RF MEMS has been slow due to reliability problems introduced by the mechanical domain and also because RF MEMS have not been monolithically integrated with CMOS. Chapter 6: Conclusions 135 This work represents significant progress since RF CMOS-MEMS are monolithically integrated with CMOS using maskless post-CMOS processing. However, even the maskless processing required for CMOS-MEMS adds cost and complexity beyond that required for semiconductor devices. In exchange, RF CMOS MEMS provide superior performance in terms of tuning ratio and quality factor. The question that remains is whether or not there exists an application that benefits sufficiently from this superior performance to justify that additional cost and complexity. I sincerely hope so because I have thoroughly enjoyed the work. Chapter 6: Conclusions 136 References [1] ―iPhone 4 Technical Specifications,‖ internet: http://www.apple.com/iphone/specs.html, Mar. 2011. [2] S. Wu and B. Razavi, ―A 900-MHz/1.8-GHz CMOS receiver for dual-band applications,‖ IEEE J. Solid-State Circuits, vol. 33, pp. 2178–2185, Dec. 1998. [3] K. Muhammad, Y.-C. Ho, T. L. Mayhugh, C.-M. Hung, T. Jung, I. Elahi, C. Lin, I. Deng, C. Fernando, J. L. Wallberg, S. K. Vemulapalli, S. Larson, T. Murphy, D. Leipold, P. Cruise, J. Jaehnig, M.-C. Lee, R. B. Staszewski, R. Staszewski, and K. Maggio, ―The first fully integrated quad-band GSM/GPRS receiver in a 90nm digital CMOS process,‖ IEEE J. Solid-State Circuits, vol. 41, no. 8, pp. 1772–1783, Aug. 2006. [4] M. Zargari, M. Terrovitis, S. H.-M. Jen, B. J. Kaczynski, M. Lee, M. P. Mack, S. S. Mehta, S. Mendis, K. Onodera, H. Samavati, W. W. Si, K. Singh, A. Tabatabei, D. Weber, D. K. Su, and B. A. Wooley, ―A singlechip dual-band tri-mode CMOS transceiver for IEEE 802.11a/b/g wirelass LAN,‖ IEEE J. Solid-State Circuits, vol. 39, no. 12, pp. 2239–2249, Dec. 2004. [5] D. J. Young and B. E. Boser, ―A micromachined variable capacitor for monolithic low-noise VCOs,‖ in SolidState Sens. Actuators Workshop Tech. Dig., Hilton Head Island, SC, Jun. 1996, pp. 86–89. [6] J. J. Yao, S. Park, and J. DeNatale, ―High tuning-ratio MEMS-based tunable capacitor for RF communications applications,‖ in Solid-State Sens. Actuators Workshop Tech. Dig., Hilton Head Island, SC, Jun. 1998, pp. 124– 127. [7] J.-B. Yoon and C. T.-C. Nguyen, ―A high-Q tunable micromechanical capacitor with movable dielectric for RF applications,‖ in IEEE Int. Electron Devices Meeting Tech. Dig., San Francisco, CA, Dec. 2000, pp. 489 –492. [8] H. D. Nguyen, D. Hah, P. R. Patterson, R. Chao, W. Piyawattanametha, E. K. Lau, and M. C. Wu, ―Angular vertical comb-driven tunable capacitor with high-tuning capabilities,‖ J. Microelectromech. Syst., vol. 13, no. 3, June 2004. [9] B. Lakshminarayanan, D. Mercier, and G. M. Rebeiz, ―High-reliability miniature RF-MEMS switched capacitors,‖ IEEE Trans. Microw. Theory Tech., vol. 56, no. 4, pp. 971–981, Apr. 2008. [10] I. Reines, B. Pillans, and G. M. Rebeiz, ―Thin-film aluminum RF MEMS switched capacitors with stress tolerance and temperature stability,‖ J. Microelectromech. Syst., vol. 20, no. 1, pp. 193–203, Feb. 2011. [11] S. H. Pu, A. S. Holmes, E. M. Yeatman, C. Papavassiliou, and S. Lucyszyn, ―Stable zipping RF MEMS varactors,‖ J. Micromech. Microeng., vol. 20, pp. 1–8, Mar. 2010. References 137 [12] L. Dussopt and G. M. Rebeiz, ―Intermodulation distortion and power handling in RF MEMS switches, varactors, and tunable filters,‖ IEEE Trans. Microw. Theory Tech., vol. 51, no. 4, pp. 1247–1255, Apr. 2003. [13] G. M. Rebeiz and J. B. Muldavin, ―RF MEMS switches and switch circuits,‖ IEEE Microw. Mag., vol. 3, pp. 72–81, Jun. 2002. [14] C. Goldsmith, Z. Yao, S. Eshelman, and D. Denniston, ―Performance of low-loss RF MEMS capacitive switches,‖ IEEE Microw. Guided Wave Lett., vol. 8, no. 8, pp. 269–271. Aug. 1998. [15] G. M. Rebeiz, G.-L. Tan, and J. S. Hayden, ―RF MEMS phase shifters: Design and applications,‖ IEEE Microw. Mag., vol. 3, pp. 72–81, Jun. 2002. [16] D. Ramachandran, A. Oz, V. K. Saraf, G. K. Fedder, and T. Mukherjee, ―MEMS—Enabled reconfigurable VCO and RF filter,‖ in Proc. IEEE RFIC Symp., Fort Worth, TX, Jun. 2004. pp, 251–254. [17] A. Dec and K. Suyama, ―A 1.9-GHz CMOS VCO with micromachined electromechanically tunable capacitors,‖ IEEE J. Solid-State Circuits, vol. 35, no. 8, pp. 1231–1237, Aug. 2000. [18] J. Brank, J. Yao, A. Malczewski, K. Varian, and C. L. Goldsmith, ―RF MEMS-based tunable filters,‖ RF Microwave Computer-Aided Eng., vol. 11, pp. 276–284, Sep. 2001. [19] J. Papapolymerou, K. L. Lange, C. L. Goldsmith, A. Malczewski, and J. Kleber, ―Reconfigurable double-stub tuners using MEMS switches for intelligent RF front-ends,‖ IEEE Trans. Microw. Theory Tech., vol. 51, no. 1, pp. 271–278, Jan. 2003. [20] A. Jajoo, ―Frequency reconfigurable narrow-band low noise amplifiers using CMOS-MEMS passives for multiband receivers,‖ Ph.D. thesis, Carnegie Mellon University, Pittsburgh, PA, May 2009. [21] L. Wang, and T. Mukherjee, ―A 3.4 GHz to 4.3 GHz frequency-reconfigurable class E power amplifier with an integrated CMOS-MEMS LC balun,‖ in Proc. IEEE RFIC Symp., Anaheim, CA, Jun. 2010, pp. 167-170. [22] P. J. Gilgunn, J. Liu, N. Sarkar, and G. K. Fedder, ―CMOS-MEMS lateral electrothermal actuators,‖ J. Microelectromech. Syst., vol. 17, no. 1, pp. 103–114, Feb. 2008. [23] C. Goldsmith, J. Ehmke, A. Malczewski, B. Pillans, S. Eshelman, Z. Yao, J. Brank, and M. Eberly, ―Lifetime characterization of capacitive RF MEMS switches,‖ in IEEE Int. Microw. Symp. Tech. Dig., Baltimore, MD, Jun. 1998, pp. 1569–1572. [24] J. DeNatale and R. Mihailovich, ―RF MEMS reliability,‖ in Proc. Int. Conf. Transducers Solid-State Sens. Actuators Tech. Dig., Boston, MA, Jun. 2003, pp. 943–946. References 138 [25] N. Yan and H. Min, ―High efficiency all-PMOS charge pump for low-voltage operations,‖ Elect. Lett., vol. 42, no. 5, pp. 277–279, Dec. 2006. [26] P. Robert, D. Saias, C. Billard, S. Boret, N. Sillon, C. Maeder-Pachurka, P. L. Charvet, G. Bouche, P. Ancey, and P. Berruyer, ―Integrated RF-MEMS switch based on a combination of thermal and electrostatic actuation,‖ in Proc. Intl. Conf. Transducers Solid-State Sens. Actuators Tech. Dig., Boston, MA, Jun. 2003, pp. 1714-1717. [27] A. Oz and G. K. Fedder, ―CMOS/BiCMOS self-assembling and electrothermal microactuators for tunable capacitors, gap-closing structures, and latch mechanisms,‖ in Solid-State Sens. Actuators Workshop Tech. Dig., Hilton Head Island, SC, Jun. 2004, pp. 251–254. [28] Z. Peng, X. Yuan, J. C. M. Hwang, D. Forehand, and C. L. Goldsmith, ―Dielectric charging of RF MEMS capacitance switches under bipolar control-voltage waveforms,‖ in IEEE Int. Microw. Symp. Dig., Honolulu, HI, Jun. 2007, pp. 1817–1820. [29] C. L. Goldsmith, A. Malczewski, Z. J. Yao, S. Chen, J. Ehmke, and D. H. Hinzel, ―RF MEMS variable capacitors for tunable filters,‖ RF Microw. Comput.-Aided Eng.—(Special Issue), vol. 9, no. 4, pp. 362–374, Jul. 1999. [30] Z. Feng, H. Zhang, K. Gupta, W. Zhang, V. Bright, and Y. C. Lee, ―MEMS-based series and shunt variable capacitors for microwave and millimeter-wave frequencies,‖ Sens. Actuators A, Phys., vol. 91, no. 3, pp. 256 – 265, Jul. 2001. [31] A. Morris, ―Tunable RF modules for mobile applications,‖ in Proc. IMS Workshop Emerging Appl. RF MEMS, Boston, MA, Jun. 2009. [32] G. K. Fedder, S. Santhanam, M. L. Reed, S. C. Eagle, D. F. Guillou, M. S.-C Lu and L. R. Carley, ―Laminated high-aspect ratio microstructures in a conventional CMOS process,‖ Sens. Actuators A, Phys., vol. 57, no. 2, pp. 103–110, pp. 103–114, Mar. 1997. [33] CRC Materials Science and Engineering Handbook, CRC Press, Boca Raton, FL, 2001. [34] A. Oz, ―CMOS/BiCMOS self-assembling and electrothermal microactuators for tunable capacitors,‖ M.S. Thesis, Carnegie Mellon University, Pittsburgh, PA, Dec. 2003. [35] A. Oz and G. K. Fedder, ―RF CMOS-MEMS capacitor having large tuning range,‖ in Proc. of 12th International Conference on Transducers, Solid-State Sensors, Actuators, and Microsystems (Transducers ’03), Boston, MA, pp. 851-854, June 2003. References 139 [36] J. Costa, T. Ivanov, J. Hammond, J. Gering, E. Glass, J. Jorgenson, D. Dening, D. Kerr, J. Reed, S. Crist, T. Mercier, S. Kim, and P. Gorisse, ―Integrated MEMS switch technology on SOI-CMOS,‖ in Solid-State Sens. Actuators Workshop Tech. Dig., Hilton Head Island, SC, Jun. 2008, pp. 18–21. [37] X. Rottenberg, S. Bregels, P. Ekkels, P. Czarnecki, P. Nolmans, R. P. Mertens, B. Nauwelaers, R. Puers, I. De Wolf, W. De Raedt, and H. A. C. Tilmans, ―An electrostatic fringing-field actuator (EFFA): application towards a low-complexity thin-film RF-MEMS technology,‖ J. Micromech. Microeng., vol. 17, pp. S204–S210, July 2007. [38] A. Morris, ―Tunable RF modules for mobile applications,‖ presented at the IMS Workshop Emerging Appl. RF MEMS, Boston, MA, Jun. 2009. [39] L. Dussopt and G. M. Rebeiz, ―An X- to Ku- band 3-bit digital MEMS varactor,‖ IEEE Microwave and Wireless Components Letters, vol. 13, no. 9, pp. 361–363, Sept. 2003. [40] J. K. Luo, M. Lin, Y. Q. Fu, L. Wang, A. J. Flewitt, S. M. Spearing, N. A. Fleck, and W. I. Milne, ―MEMS based digital variable capacitors with a high-k dielectric insulator,‖ Sensors & Actuators, vol. A132, no. 1, pp 139-146, May 2006. [41] J. Schoebel, T. Buck, M. Reimann, M. Ulm, M. Schneider, A. Jourdain, G.J. Carchon, and H.A.C. Tilmans., ―Design considerations and technology assessment of phased-array antenna systems with RF MEMS for automotive radar applications,‖ IEEE Trans. Microw. Theory Tech., vol. 53, no. 6, pp. 1968–1975, Jun. 2005. [42] N. S. Barker and G. M. Rebeiz, ―Distributed MEMS true-time delay phase shifters and wideband switches,‖ IEEE Trans. Microwave Theory Tech., vol. 46, pp. 1881-1890, Nov. 1998. [43] A.S. Nagra and R.A. York, ―Distributed analog phase shifters with low insertion loss,‖ IEEE Trans. Microwave Theory Tech., vol. 47, pp. 1705-1711, Sep. 1999. [44] A. Malczewski, S. Eshelman, B. Pillans, J. Ehmke, and C.L. Goldsmith, ―X-Band RF MEMS phase shifters for phased array applications,‖ IEEE Microwave Wireless Comp. Lett., vol. 9, pp. 517-519, Dec. 1999. [45] M. Kim, J.B. Hacker, R.E. Mihailovich, and J.F. DeNatale, ―A DC-to-40 GHz four-bit RF MEMS true-time delay network,‖ IEEE Microwave Wireless Comp. Lett., vol. 11, pp. 56-58, Feb. 2001. [46] G.L. Tan, R.E. Mihailovich, J.B. Hacker, J.F. DeNatale, and G.M. Rebeiz, ―A very-low-loss 2-bit X-band RF MEMS phase shifter,‖ presented at the 2002 IEEE Int. Microwave Symp., Seattle, WA. References 140 [47] B. Pillans, S. Eshelman, A. Malczewski, J. Ehmke, and C. Goldsmith, ―Ka-band RF MEMS phase shifters,‖ IEEE Microwave Guided Wave Lett., vol. 9, pp. 520-522, Dec. 1999. [48] J. S. Hayden and G.M. Rebeiz, ―A low-loss Ka-band distributed MEMS 2-bit phase shifter using metal-air-metal capacitors,‖ presented at the 2002 IEEE Int. Microwave Symp., Seattle, WA. [49] J. S. Hayden, A. Malczewski, J. Kleber, C. L. Goldsmith, and G. M. Rebeiz, ―2 and 4-bit DC-18 GHz microstrip MEMS distributed phase shifters,‖ in 2001 IEEE MTT-S Int. Microwave Symp. Dig., May 2001, pp. 219-222. [50] Y. Liu, A. Borgioli, A.S. Nagra, and R.A. York, ―K-band 3-bit low-loss distributed MEMS phase shifter,‖ IEEE Microwave Guided Wave Lett., vol. 10, pp. 415-417, Oct. 2000. [51] H.-T. Kim, J.-H. Park, Y.-K. Park, and Y. Kwon, ―V-band low-loss and low-voltage distributed MEMS digital phase shifter using metal-air metal capacitors,‖ presented at the 2002 IEEE Int. Microwave Symp., Seattle, WA. [52] N.S. Barker and G.M. Rebeiz, ―Optimization of distributed MEMS transmission-line phase shifters—U-band and W-band designs,‖ IEEE Trans. Microwave Theory Tech., vol. 48, pp. 1957-1966, Nov. 2000. [53] F. Ellinger, R. Vogt, and W. Bächtold, ―Ultra compact, low loss, varactor tuned phase shifter MMIC at Cband,‖ IEEE Microwave Wireless Comp Lett., vol. 11, pp. 104–105, Mar. 2001. [54] A. S. Nagra and R. A. York, ―Monolithic GaAs phase shifter with low insertion loss and continuous 0°–360° phase shift at 20 GHz,‖ IEEE Microwave Guided Wave Lett., vol. 9, pp. 31–33, Jan. 1999. [55] M. Baker-Kassem, S. Fouladi, and R. R. Mansour, ―Novel high-Q MEMS curled-plate variable capacitor fabricated in 0.35 μm CMOS technology,‖ IEEE Transactions on Microwave Theory and Techniques, vol. 56, pp. 530-541, Feb. 2008. [56] K. Vummidi, J. Hammond, J. Costa, and S. Raman, ―Resonant pull-in for a variable gap lateral contact RF MEMS switch,‖ in Proc. IEEE Microelectromechanical Syst., Wanchai, Hong Kong, pp. 767–770, Jan. 2010. References 141 Appendix A. Electrical Characterization This appendix describes how the RF electrical characteristics of the CMOS-MEMS variable capacitors were measured and used to calculate an equivalent circuit model. Section A.1 discusses the network analyzer settings that were used to measure two-port network parameters for the gap-tuning capacitor, the digital capacitor, and the phase shifter. Section A.2 presents the procedure for open-short de-embedding, which is used to remove the effects of the on-chip probe pads from the measured data. Section A.3 shows how the measured two-port network parameters were used to calculate the values of the equivalent circuit elements for both a π-model and a Tmodel. The Matlab code used to perform open-short de-embedding and to calculate the equivalent circuit element values is included in Section A.4. A.1. Network Analyzer Settings An Agilent E8364A network analyzer was used to measure the RF electrical characteristics of the CMOS-MEMS variable capacitors and the CMOS-MEMS phase shifter. Prior to measurement, short-open-load-through (SOLT) calibration was performed to eliminate the parasitic effects of the probes and cables using WinCal 3.1 software and the calibration substrate. The probes used to make electrical contact with the probe pads are Infinity probes with 100 µm pitch for ground-signal-ground (GSG) configuration. These probes were chosen because the nickel probe tips provide a low contact resistance to aluminum. The network analyzer was set to record 1601 measurements with a linear frequency range from 100 MHz to 40 GHz. The intermediate filter (IF) bandwidth was lowered from the default 35 kHz to 500 Hz to reduce the noise floor. Modification of the intermediate frequency is necessary to measure small resistances in series with large imaginary impedances. Using these settings, two-port S-parameters were Appendix A: Electrical Characterization 142 measured for the device under test (DUT) and the open-short de-embedding test structures (see Fig. A.1). G G G G G S S S S S G G G G G (a) (b) G DUT S G (c) Fig. A.1. Layout view of (a) open test structure (b) short test structure (c) DUT. Measurements of the open and short test structures are necessary to remove the parasitic effects of the probe pads using open-short de-embedding. ‗G‘ and ‗S‘ represent ‗ground‘ and ‗signal‘, respectively. A.2. Open-short De-embedding Open-short de-embedding is a procedure used to remove the parasitic effects of the on-chip probe pads from the measured data. The probe pads contribute shunt capacitance, series resistance, and series inductance to the measured two-port network parameters of the DUT. Deembedding uses measurements of open and short test structures to calculate the two-port network parameters of the DUT as if the probe pads were not present. The open test structure consists of two sets of probe pads that are not connected, or open-circuited. The short test structure consists of two sets of probe pads that are connected to ground, or short-circuited. Open-short deembedding was performed using the following steps. 1. Three sets of S-parameters, Smeas, Sopen, and Sshort, were measured for the DUT, the open test structure, and the short test structure, respectively. 2. The three sets of S-parameters were converted to Y-parameters, Ymeas, Yopen, and Yshort. using the following equations. These equations assume that both the source and load impedance are 50 Ω. Appendix A: Electrical Characterization 143 Y11 Y12 Y21 Y22 3. 1 - S11 1 S 22 S12 S 21 . 50 1 S11 1 S 22 S12 S 21 1 2S12 1 (A.1) . (A.2) . (A.3) 1 S11 1 S 22 S12 S 21 . 50 1 S11 1 S 22 S12 S 21 (A.4) 50 1 S11 1 S 22 S12 S 21 2S 21 1 50 1 S11 1 S 22 S12 S 21 1 The open Y-parameters were subtracted from both the DUT and the short Y-parameters to remove the parallel capacitance of the probe pad. 4. Ymeas ' Ymeas Yopen . (A.5) Yshort' Yshort Yopen . (A.6) The resulting Y-parameters were converted to Z-parameters, Zmeas‘ and Zshort‘, using the following equations. Z 11 Z 12 Z 21 Z 22 5. Y22 Y11Y22 Y12Y21 Y12 Y11Y22 Y12Y21 Y21 Y11Y22 Y12Y21 Y11 Y11Y22 Y12Y21 . (A.7) . (A.8) . (A.9) . (A.10) The short Z-parameters were subtracted from the DUT Z-parameters to remove the series resistance and inductance of the probe pad. Zde -embedded Z meas'Zshort ' . (A.11) The resulting Z-parameters are de-embedded two-port network parameters that represent the Appendix A: Electrical Characterization 144 electrical characteristics of the DUT as if the probe pads were not present. The de-embedded Zparameters were converted to Y- or S-parameters as needed for further calculations. The following equations were used to convert Z-parameters to Y-parameters. Y11 Y12 Y21 Z 22 Z 22 Z 11Z 22 Z 12 Z 21 Z 12 Z 11Z 22 Z 12 Z 21 Z 21 Z 11Z 22 Z 12 Z 21 . (A.12) . (A.13) . (A.14) . (A.15) Y11 Z 11Z 22 Z 12 Z 21 The following equations were used to convert Y-parameters to S-parameters. These equations assume that both the source and load impedance are 50 Ω. S11 S12 S 21 S 22 1 - 50Y11 1 50Y22 502 Y12Y21 . 1 50Y11 1 50Y22 502 Y12Y21 2 * 50Y12 (A.16) . (A.17) . (A.18) 1 50Y11 1 50Y22 502 Y12Y21 . 1 50Y11 1 50Y22 502 Y12Y21 (A.19) 1 50Y11 1 50Y22 502 Y12Y21 2 * 50Y21 1 50Y11 1 50Y22 502 Y12Y21 A.3. Extracting an Equivalent Circuit Model After open-short de-embedding, the measured two-port network parameters represent the DUT without probe pads as the DUT would be used in an integrated circuit. These network parameters are sufficient for simulating the CMOS-MEMS variable capacitors but do not provide Appendix A: Electrical Characterization 145 an intuitive understanding which would allow the variable capacitors to be parameterized and subsequently modified during circuit design. To better understand the variable capacitors, further calculations were performed to extract the element values of an equivalent circuit model. Two different equivalent circuit models, a π-model and a T-model, were used for the variable capacitors presented in this work. A.3.1. π-Model A π-model was used for the gap-tuning capacitor presented in Chapter 3 and both digital capacitor presented in Chapter 4 because those variable capacitors are connected in series between two probe pads. A π-model is better suited for series capacitors than a T-model since the π-model contains a single series impedance to represent the series capacitance. The de-embedded network parameters were transformed into Y-parameters and lumped into a π-model (see Fig. A.2(a)). A plot of the lumped impedances shows that the series impedance can be modeled as a series resistor-inductor-capacitor (RLC) circuit whereas the shunt impedances can be modeled as capacitors for frequencies below the self-resonance frequency (Fig. A.3). The equivalent circuit model was chosen to match the measured impedances (Fig. A.2(b)). The equations in this section describe how to calculate the equivalent circuit element values. For each calculation, plots comparing the measured data with the extracted model are presented to validate the proposed model. All plots presented in this section correspond to the maximum capacitance state of the updated digital capacitor presented in Section 4.5.5. Appendix A: Electrical Characterization 146 Port 1 Port 2 -Y21 Y11+Y12 Rs Port 1 C1g Y22+Y21 Cs Ls Port 2 C2g (b) (a) Fig. A.2. Equivalent circuit model in terms of (a) Y-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. Impedance Magnitude (Ω) 1E+5 1E+4 1E+3 1E+2 1E+1 |1/-Y21| |1/(Y11+Y12)| |1/(Y22+Y21)| Extracted Model 1E+0 1E-1 0.1 1 10 Frequency (GHz) Fig. A.3. Lumped impedance magnitude as a function of frequency for the maximum capacitance position. The extracted model is compared with measured data. All calculations of the equivalent circuit element values were averaged over frequencies from 1 GHz to 4 GHz. First, the self-resonance frequency, ωres, was calculated as the frequency for which the imaginary component of the series impedance was minimized. Then, the series capacitance, Cs, was calculated using the self-resonance frequency and the imaginary component of the series impedance (Fig. A.4). Cs 1 res 2 imag 1 Y21 . (A.20) The asymptote in the measured data is expected since the imaginary component of the series Appendix A: Electrical Characterization 147 impedance approaches zero at the self-resonance frequency. 2000 Cs (Measured) Capacitance (fF) 1800 Extracted Model 1600 1400 1200 1000 800 600 400 200 0 1 2 3 4 5 6 7 8 9 10 Frequency (GHz) Fig. A.4. Series capacitance as a function of frequency. The extracted model is compared with measured data. The series inductance, Ls, was calculated from the self-resonance frequency and the series capacitance. Ls 1 res 2 C s . (A.21) 5 Rs (Measured) 4.5 Resistance (Ω) 4 Extracted Model 3.5 3 2.5 2 1.5 1 0.5 0 1 2 3 4 5 6 7 8 9 10 Frequency (GHz) Fig. A.5. Series resistance as a function of frequency. The extracted model is compared with measured data. The series resistance, Rs, was calculated from the real part of the series impedance (see Fig. A.5). Appendix A: Electrical Characterization 148 Rs real 1 Y21 . (A.22) The increase in series resistance at higher frequencies could be due to the skin effect or due to the noise floor of the network analyzer. 100 C1g (Measured) C2g (Measured) Extracted Model Capacitance (fF) 90 80 70 60 50 40 30 20 10 0 1 2 3 4 5 6 7 8 9 10 Frequency (GHz) Fig. A.6. Shunt capacitances as a function of frequency. The extracted model is compared with measured data. The shunt capacitances were calculated using the imaginary component of the shunt impedances (see Fig. A.6). C1g 1 . * imag 1 Y11 Y21 (A.23) C2 g 1 . * imag 1 Y22 Y21 (A.24) The shunt capacitances vary with frequency because the shunt impedances are not purely capacitive. The shunt impedances are complex distributed networks because these impedances include capacitance to the silicon substrate. Thus the total capacitance is constant at low frequencies but appears to vary as the frequency approaches the self-resonance frequency. The Q can be calculated by two different equations depending on how the two ports are constrained. Both equations calculate the Q of the variable capacitor used as a shunt capacitor with one port connected to ground. Q1 is the Q calculated looking into port 1 with port 2 Appendix A: Electrical Characterization 149 connected to ground. Q1 Im1 Y11 Re1 Y11 . (A.25) Q2 is the Q calculated looking into port 2 with port 1 connected to ground (see Fig. A.7). Q2 Im1 Y22 Re1 Y22 . (A.26) Calculating Q as Q1 ignores the effects of C2g whereas calculating Q as Q2 ignores the effects of C1g. The Q is difficult to measure at low frequencies where the series resistance is small compared to the large imaginary impedance of the capacitance. Note that both (A.25) and (A.26) are approximations that are only valid for frequencies much less than the self-resonance Quality Factor frequency. This definition of Q equates to zero at the self-resonance frequency. 500 450 400 350 300 250 200 150 100 50 0 Quality Factor (De-embedded) Extracted Model 1 2 3 4 5 6 Frequency (GHz) Fig. A.7. Quality factor as a function of frequency. The extracted model is compared with measured data. Equations (A.20) through (A.24) were used to calculate the Y-parameters of the extracted model to provide a comparison between the measured and extracted lumped impedances (see Fig. A.3). For frequencies from 1 GHz to 4 GHz, the lumped impedances of the extracted model match the measured lumped impedances with a median error of 4.2%. The extracted YAppendix A: Electrical Characterization 150 parameters were converted to S-parameters to provide a comparison between the measured and extracted S-parameters (see Fig. A.8). For frequencies from 1 GHz to 4 GHz, the S-parameters of the extracted model match the measured S-parameters with a median error of 0.6%. 0 S-Parameters (dB) -5 -10 -15 -20 -25 S11 (De-embedded) S21 (De-embedded) S22 (De-embedded) Extracted Model -30 -35 -40 0.1 1 10 Frequency (GHz) Fig. A.8. S-parameter magnitude as a function of frequency. The extracted model is compared with measured data. S12 is not plotted because S12 and S21 are equivalent for passive networks. A.3.2. T-Model A T-model was used for the phase shifter digital capacitor presented in Chapter 5 because that variable capacitor is connected in shunt between two probe pads. A T-model is better suited for shunt capacitors than a π-model since the T-model contains a single shunt impedance to represent the shunt capacitance. The de-embedded network parameters were transformed into Zparameters and lumped into a T-model (see Fig. A.9(a)). A plot of the lumped impedances shows that the shunt impedance can be modeled as a series resistor-capacitor (RC) circuit whereas the series impedances can be modeled as series resistor-inductor (RL) circuits (Fig. A.10). The equivalent circuit model was chosen to match the measured impedances (Fig. A.9 (b)). The equations in this section describe how to calculate the equivalent circuit element values. For each calculation, plots comparing the measured data with the extracted model are presented to validate Appendix A: Electrical Characterization 151 the proposed model. All plots presented in this section correspond to the maximum capacitance state of the phase shifter digital capacitor presented in Chapter 5. Port 1 Z11-Z12 Z22-Z21 Port 2 Port 1 Port 2 Rs1 Ls1 Rg Rs2 Z21 Ls2 Cg (b) (a) Fig. A.9. Equivalent circuit model in terms of (a) Z-parameters and (b) lumped circuit elements. Circuit element values are extracted from measured 2-port network parameters. 1E+5 Impedance (Ω) 1E+4 1E+3 |Z21| (Measured) |(Z11-Z12)| (Measured) |(Z22-Z21)| (Measured) Extracted Model 1E+2 1E+1 1E+0 1E-1 0.1 1 10 Frequency (GHz) Fig. A.10. Lumped impedance magnitude as a function of frequency. The extracted model is compared with measured data. Similar to that of the π-model, the equivalent circuit element values were averaged over a range of frequencies. However, since the phase shifter digital capacitor was designed for frequencies greater than 30 GHz, results were averaged from 30 GHz to 35 GHz rather than from 1 GHz to 4 GHz. The capacitance to ground, Cg, was calculated using the imaginary component of the shunt impedance (see Fig. A.11). Cs 1 imag Z 21 . Appendix A: Electrical Characterization (A.27) 152 200 Cg (Measured) 180 Extracted Model Capacitance (fF) 160 140 120 100 80 60 40 20 0 10 20 30 40 Frequency (GHz) Fig. A.11. Capacitance to ground as a function of frequency. The extracted model is compared with measured data. The shunt impedance contains capacitance to the silicon substrate, which has a significant resistive component. Because the shunt impedance is not a simple RC circuit, both the capacitance to ground and resistance to ground vary slightly with frequency. The resistance to ground, Rg, was calculated from the real part of the shunt impedance (see Fig. A.12). R s real Z 21 . (A.28) 3 Rs (Measured) Resistance (Ω) 2.5 Extracted Model 2 1.5 1 0.5 0 10 20 30 40 Frequency (GHz) Fig. A.12. Shunt resistance as a function of frequency. The extracted model is compared with measured data. Appendix A: Electrical Characterization 153 The resistance to ground is noisy because that resistance is small compared with the large imaginary impedance of the shunt capacitance. The series resistances, Rs1 and Rs2, were calculated from the real part of the series impedances (see Fig. A.13). 3 (A.29) R s 2 real Z 22 Z 21 . (A.30) Rs1 (measured) Rs2 (measured) Extracted Model 2.5 Resistance (Ω) R s1 real Z 11 Z 12 . 2 1.5 1 0.5 0 10 20 30 40 Frequency (GHz) Fig. A.13. Series resistances as a function of frequency. The extracted model is compared with measured data. The series resistances are also noisy due to the large imaginary impedance. The series inductances, Ls1 and Ls2, were calculated from imaginary part of the series impedances (see Fig. A.14) Ls1 Ls 2 imag Z11 Z12 imag Z 22 Z 21 . (A.31) . (A.32) Appendix A: Electrical Characterization 154 200 Ls1 (measured) Ls2 (measured) Extracted Model Inductance (pH) 180 160 140 120 100 80 60 40 20 0 10 20 30 40 Frequency (GHz) Fig. A.14. Series inductances as a function of frequency. The extracted model is compared with measured data. Similar to that of the π-model, the Q can be calculated by two different equations depending on how the two ports are constrained, and both equations calculate the Q of the variable capacitor used as a shunt capacitor. Since the T-model only includes a shunt capacitance, one port must be left floating to calculate the Q. Q1 is the Q calculated looking into port 1 with port 2 left floating (see Fig. A.15). Q1 ImZ 11 ReZ 11 . (A.33) Q2 is the Q calculated looking into port 2 with port 1 left floating. Q2 ImZ 22 ReZ 22 . (A.34) Calculating Q as Q1 ignores the effects of Rs2 and Ls2 whereas calculating Q as Q2 ignores the effects of Rs1 and Ls1. The Q is difficult to measure at low frequencies where the series resistance is small compared to the large imaginary impedance of the capacitance. To calculate the extracted model in Fig. A.15, the self-resonance frequency was assumed to be 55 GHz, which provides the best fit between measurement and calculation. Note that both (A.33) and (A.34) are Appendix A: Electrical Characterization 155 approximations that are only valid for frequencies much less than the self-resonance frequency. Quality Factor This definition of Q equates to zero at the self-resonance frequency. 100 90 80 70 60 50 40 30 20 10 0 Quality Factor (De-embedded) Extracted Model 10 20 30 40 Frequency (GHz) Fig. A.15. Quality factor as a function of frequency. The extracted model is compared with measured data. S-parameters (dB) 10 0 S21 (De-embedded) S11 (De-embedded) S22 (De-embedded) Extracted Model -10 -20 -30 -40 0.1 1 10 Frequency (GHz) Fig. A.16. S-parameter magnitude as a function of frequency. The extracted model is compared with measured data. S12 is not plotted because S12 and S21 are equivalent for passive networks. Equations (A.27) through (A.32) were used to calculate the Z-parameters of the extracted model to provide a comparison between the measured and extracted lumped impedances (see Fig. A.10). For frequencies from 30 GHz to 35 GHz, the lumped impedances of the extracted Appendix A: Electrical Characterization 156 model match the measured lumped impedances with a median error of 0.6%. The extracted Zparameters were converted to S-parameters to provide a comparison between the measured and extracted S-parameters (see Fig. A.16). For frequencies from 30 GHz to 35 GHz, the Sparameters of the extracted model match the measured S-parameters with a median error of 1.2%. A.3.3. Discussion The equivalent circuit models for both the π-model and the T-model are simplified interpretations of the measured network parameters. These models capture the most important aspects of the CMOS-MEMS variable capacitors but could certainly be improved if more complexity is required. For example, the shunt impedances in the π-model could be modeled as series RC circuits rather than just capacitors to account for the resistance of the silicon substrate. The shunt impedance in the T-model could be modeled as a series RLC circuit rather than just a series RC circuit to include the parasitic inductance. Many possibilities exist for expanding the current model if a more complex model is required. A.4. Matlab Code Matlab scripts were used to perform open-short de-embedding and to calculate the equivalent circuit element values from the de-embedded two-port network parameters. To use these scripts, the directory name must be changed to the directory that contains the measured two-port S-parameters. The file names for the DUT and open and short test structures must be updated. Appendix A: Electrical Characterization 157 A.4.1. π-model Script %John Reinke %3/7/2011 %Open-short de-embedding and extraction of equivalent circuit model parameters for pi-model capacitors %Directory and file names dir='C:\Documents and Settings\jreinke\Desktop\Jazz\Jazz032\RF\092010'; file_m=[dir '\cap1c15s7r.s2p']; %File containing DUT 2-port S-parameters file_o=[dir '\gsg_open.s2p']; %File containing open 2-port S-parameters file_s=[dir '\gsg_short.s2p']; %File containing short 2-port S-parameters %Get 2-port S-parameters from capacitor file [number,unit,parameter,format,type,load]=textread(file_m,'%s %s %s %s %s %s',1,'headerlines',5); if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_m,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_m=s11_real+j*s11_imag; s21_m=s21_real+j*s21_imag; s12_m=s12_real+j*s12_imag; s22_m=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_m,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_m=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_m=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_m=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_m=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Get 2-port S-parameters from open file [number,unit,parameter,format,type,load]=textread(file_o,'%s %s %s %s %s %s',1,'headerlines',5); if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_o,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_o=s11_real+j*s11_imag; s21_o=s21_real+j*s21_imag; s12_o=s12_real+j*s12_imag; s22_o=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_o,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_o=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_o=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_o=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_o=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Get 2-port S-parameters from short file [number,unit,parameter,format,type,load]=textread(file_s,'%s %s %s %s %s %s',1,'headerlines',5); Appendix A: Electrical Characterization 158 if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_s,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_s=s11_real+j*s11_imag; s21_s=s21_real+j*s21_imag; s12_s=s12_real+j*s12_imag; s22_s=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_s,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_s=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_s=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_s=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_s=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Convert 2-port S-parameters to 2-port Y-parameters z0=50; D_m=(1+s11_m).*(1+s22_m)-s12_m.*s21_m; y11_m=1/z0*((1-s11_m).*(1+s22_m)+s12_m.*s21_m)./D_m; y12_m=-1/z0*2*s12_m./D_m; y21_m=-1/z0*2*s21_m./D_m; y22_m=1/z0*((1+s11_m).*(1-s22_m)+s12_m.*s21_m)./D_m; D_o=(1+s11_o).*(1+s22_o)-s12_o.*s21_o; y11_o=1/z0*((1-s11_o).*(1+s22_o)+s12_o.*s21_o)./D_o; y12_o=-1/z0*2*s12_o./D_o; y21_o=-1/z0*2*s21_o./D_o; y22_o=1/z0*((1+s11_o).*(1-s22_o)+s12_o.*s21_o)./D_o; D_s=(1+s11_s).*(1+s22_s)-s12_s.*s21_s; y11_s=1/z0*((1-s11_s).*(1+s22_s)+s12_s.*s21_s)./D_s; y12_s=-1/z0*2*s12_s./D_s; y21_s=-1/z0*2*s21_s./D_s; y22_s=1/z0*((1+s11_s).*(1-s22_s)+s12_s.*s21_s)./D_s; %Subtract open response in Y-parameters y11_mp=y11_m-y11_o; y12_mp=y12_m-y12_o; y21_mp=y21_m-y21_o; y22_mp=y22_m-y22_o; y11_sp=y11_s-y11_o; y12_sp=y12_s-y12_o; y21_sp=y21_s-y21_o; y22_sp=y22_s-y22_o; %Transform 2-port Y-parameters to 2-port Z-parameters D_mp=y11_mp.*y22_mp-y12_mp.*y21_mp; z11_mp=y22_mp./D_mp; z12_mp=-y12_mp./D_mp; z21_mp=-y21_mp./D_mp; z22_mp=y11_mp./D_mp; D_sp=y11_sp.*y22_sp-y12_sp.*y21_sp; z11_sp=y22_sp./D_sp; z12_sp=-y12_sp./D_sp; Appendix A: Electrical Characterization 159 z21_sp=-y21_sp./D_sp; z22_sp=y11_sp./D_sp; %Subtract short response in Z-parameters z11=z11_mp-z11_sp; z12=z12_mp-z12_sp; z21=z21_mp-z21_sp; z22=z22_mp-z22_sp; %Transform 2-port Z-parameters to 2-port Y-parameters D=z11.*z22-z12.*z21; y11=z22./D; y12=-z12./D; y21=-z21./D; y22=z11./D; %Transform 2-port Y-parameters to lumped impedances imp_se=-1./y12; %Series impedance imp_sh1=1./(y11+y12); %Shunt impedance 1 imp_sh2=1./(y22+y21); %Shunt impedance 2 %Convert frequency from Hz to rad/s w=2*pi()*f; %Find resonant frequency [value,index]=min(abs(imag(smooth(imp_se,100)))); if (value < 10) f0=f(index) w0=2*pi*f0; Cs_vs_f=(1-w.^2/w0.^2).*(-1./(w.*imag(imp_se))); Cs=mean(Cs_vs_f(max(find(f<1e9)):max(find(f<4e9)))) Ls=1/(w0.^2*Cs) else f0=40e9 Ls=0 Cs_vs_f=-1./(w.*(imag(imp_se)-w.*Ls)); Cs=mean(Cs_vs_f(max(find(f<1e9)):max(find(f<4e9)))) End %Calculate element values of equivalent circuit model Rs_vs_f=real(imp_se); Rs=mean(Rs_vs_f(max(find(f<1e9)):max(find(f<4e9)))) Q1_vs_f=-imag(1./y11)./real(1./y11); Q2_vs_f=-imag(1./y22)./real(1./y22); Q1_1GHz=mean(Q1_vs_f(max(find(f<1e9)):max(find(f<4e9))).*f(max(find(f<1e9)):max(find(f<4e9)))./(f0f(max(find(f<1e9)):max(find(f<4e9)))).*f0)/1e9*(f0-1e9)/f0 Q2_1GHz=mean(Q2_vs_f(max(find(f<1e9)):max(find(f<4e9))).*f(max(find(f<1e9)):max(find(f<4e9)))./(f0f(max(find(f<1e9)):max(find(f<4e9)))).*f0)/1e9*(f0-1e9)/f0 C1g_vs_f=-1./(w.*imag(imp_sh1)); C1g=mean(C1g_vs_f(max(find(f<1e9)):max(find(f<4e9)))) C2g_vs_f=-1./(w.*imag(imp_sh2)); C2g=mean(C2g_vs_f(max(find(f<1e9)):max(find(f<4e9)))) %Calculate de-embedded 2-port S-parameters s11=((1-z0.*y11).*(1+z0.*y22)+z0.*y12.*z0.*y21)./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); s12=-2*z0.*y12./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); Appendix A: Electrical Characterization 160 s21=-2*z0.*y21./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); s22=((1+z0.*y11).*(1-z0.*y22)+z0.*y12.*z0.*y21)./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); %Calculate modeled lumped impedances imp_sei=Rs+j*w*Ls+1./(j*w*Cs); imp_sh1i=1./(j*w*C1g); imp_sh2i=1./(j*w*C2g); %Calculate modeled 2-port Y-parameters y11i=1./imp_sh1i+1./imp_sei; y12i=-1./imp_sei; y21i=-1./imp_sei; y22i=1./imp_sh2i+1./imp_sei; %Calculate modeled 2-port S-parameters s11i=((1-z0.*y11i).*(1+z0.*y22i)+z0.*y12i.*z0.*y21i)./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s12i=-2*z0.*y12i./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s21i=-2*z0.*y21i./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s22i=((1+z0.*y11i).*(1-z0.*y22i)+z0.*y12i.*z0.*y21i)./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); Appendix A: Electrical Characterization 161 A.4.2. T-model Script %John Reinke %3/7/2011 %Open-short de-embedding and extraction of equivalent circuit model parameters for T-model capacitors %Directory and file names dir='C:\Documents and Settings\jreinke\Desktop\Jazz\Jazz032\RF\091910'; file_m=[dir '\cap1c1s15.s2p']; %File containing DUT 2-port S-parameters file_o=[dir '\gsg_open.s2p']; %File containing open 2-port S-parameters file_s=[dir '\gsg_short.s2p']; %File containing short 2-port S-parameters %Get 2-port S-parameters from capacitor file [number,unit,parameter,format,type,load]=textread(file_m,'%s %s %s %s %s %s',1,'headerlines',5); if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_m,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_m=s11_real+j*s11_imag; s21_m=s21_real+j*s21_imag; s12_m=s12_real+j*s12_imag; s22_m=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_m,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_m=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_m=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_m=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_m=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Get 2-port S-parameters from open file [number,unit,parameter,format,type,load]=textread(file_o,'%s %s %s %s %s %s',1,'headerlines',5); if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_o,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_o=s11_real+j*s11_imag; s21_o=s21_real+j*s21_imag; s12_o=s12_real+j*s12_imag; s22_o=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_o,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_o=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_o=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_o=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_o=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Get 2-port S-parameters from short file [number,unit,parameter,format,type,load]=textread(file_s,'%s %s %s %s %s %s',1,'headerlines',5); Appendix A: Electrical Characterization 162 if strcmp(format,'RI') [f,s11_real,s11_imag,s21_real,s21_imag,s12_real,s12_imag,s22_real,s22_imag]=textread(file_s,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_s=s11_real+j*s11_imag; s21_s=s21_real+j*s21_imag; s12_s=s12_real+j*s12_imag; s22_s=s22_real+j*s22_imag; elseif strcmp(format,'dB') [f,s11_mag,s11_phase,s21_mag,s21_phase,s12_mag,s12_phase,s22_mag,s22_phase]=textread(file_s,'%n %f %f %f %f %f %f %f %f',-1,'headerlines',6); s11_s=10.^(s11_mag./20).*exp(j*pi/180*s11_phase); s21_s=10.^(s21_mag./20).*exp(j*pi/180*s21_phase); s12_s=10.^(s12_mag./20).*exp(j*pi/180*s12_phase); s22_s=10.^(s22_mag./20).*exp(j*pi/180*s22_phase); else 'incompatible file type' return end %Convert 2-port S-parameters to 2-port Y-parameters z0=50; D_m=(1+s11_m).*(1+s22_m)-s12_m.*s21_m; y11_m=1/z0*((1-s11_m).*(1+s22_m)+s12_m.*s21_m)./D_m; y12_m=-1/z0*2*s12_m./D_m; y21_m=-1/z0*2*s21_m./D_m; y22_m=1/z0*((1+s11_m).*(1-s22_m)+s12_m.*s21_m)./D_m; D_o=(1+s11_o).*(1+s22_o)-s12_o.*s21_o; y11_o=1/z0*((1-s11_o).*(1+s22_o)+s12_o.*s21_o)./D_o; y12_o=-1/z0*2*s12_o./D_o; y21_o=-1/z0*2*s21_o./D_o; y22_o=1/z0*((1+s11_o).*(1-s22_o)+s12_o.*s21_o)./D_o; D_s=(1+s11_s).*(1+s22_s)-s12_s.*s21_s; y11_s=1/z0*((1-s11_s).*(1+s22_s)+s12_s.*s21_s)./D_s; y12_s=-1/z0*2*s12_s./D_s; y21_s=-1/z0*2*s21_s./D_s; y22_s=1/z0*((1+s11_s).*(1-s22_s)+s12_s.*s21_s)./D_s; %Subtract open response in Y-parameters y11_mp=y11_m-y11_o; y12_mp=y12_m-y12_o; y21_mp=y21_m-y21_o; y22_mp=y22_m-y22_o; y11_sp=y11_s-y11_o; y12_sp=y12_s-y12_o; y21_sp=y21_s-y21_o; y22_sp=y22_s-y22_o; %Transform 2-port Y-parameters to 2-port Z-parameters D_mp=y11_mp.*y22_mp-y12_mp.*y21_mp; z11_mp=y22_mp./D_mp; z12_mp=-y12_mp./D_mp; z21_mp=-y21_mp./D_mp; z22_mp=y11_mp./D_mp; D_sp=y11_sp.*y22_sp-y12_sp.*y21_sp; z11_sp=y22_sp./D_sp; z12_sp=-y12_sp./D_sp; Appendix A: Electrical Characterization 163 z21_sp=-y21_sp./D_sp; z22_sp=y11_sp./D_sp; %Subtract short response in Z-parameters z11=z11_mp-z11_sp; z12=z12_mp-z12_sp; z21=z21_mp-z21_sp; z22=z22_mp-z22_sp; %Transform 2-port Z-parameters to 2-port Y-parameters D=z11.*z22-z12.*z21; y11=z22./D; y12=-z12./D; y21=-z21./D; y22=z11./D; %Transform 2-port Y-parameters to lumped impedances imp_g=(z21+z12)/2; %Shunt impedance to ground imp_s1=z11-z12; %Series impedance 1 imp_s2=z22-z21; %Series impedance 2 %Convert frequency from Hz to rad/s w=2*pi()*f; Cg_vs_f=-1./(w.*(imag(imp_g))); Cg=mean(Cg_vs_f(max(find(f<30e9)):max(find(f<35e9)))) Rg_vs_f=real(imp_g); Rg=mean(Rg_vs_f(max(find(f<30e9)):max(find(f<35e9)))) Ls1_vs_f=imag(imp_s1)./w; Ls1=mean(Ls1_vs_f(max(find(f<30e9)):max(find(f<35e9)))) Rs1_vs_f=real(imp_s1); Rs1=mean(Rs1_vs_f(max(find(f<30e9)):max(find(f<35e9)))) Ls2_vs_f=imag(imp_s2)./w; Ls2=mean(Ls2_vs_f(max(find(f<30e9)):max(find(f<35e9)))) Rs2_vs_f=real(imp_s2); Rs2=mean(Rs2_vs_f(max(find(f<30e9)):max(find(f<35e9)))) f0=55e9; Q1_vs_f=-imag(z11)./real(z11); Q2_vs_f=-imag(z22)./real(z22); Q1_1GHz=mean(Q1_vs_f(max(find(f<30e9)):max(find(f<35e9))).*f(max(find(f<30e9)):max(find(f<35e9)))./(f0f(max(find(f<30e9)):max(find(f<35e9)))).*f0)/30e9*(f0-30e9)/f0 Q2_1GHz=mean(Q2_vs_f(max(find(f<30e9)):max(find(f<35e9))).*f(max(find(f<30e9)):max(find(f<35e9)))./(f0f(max(find(f<30e9)):max(find(f<35e9)))).*f0)/30e9*(f0-30e9)/f0 %Calculate de-embedded 2-port S-parameters s11=((1-z0.*y11).*(1+z0.*y22)+z0.*y12.*z0.*y21)./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); s12=-2*z0.*y12./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); s21=-2*z0.*y21./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); s22=((1+z0.*y11).*(1-z0.*y22)+z0.*y12.*z0.*y21)./((1+z0.*y11).*(1+z0.*y22)-z0.*y12.*z0.*y21); %Calculate modeled lumped impedances imp_gi=Rg+1./(j*w*Cg); imp_s1i=Rs1+j*w*Ls1; imp_s2i=Rs2+j*w*Ls2; %Calculate modeled 2-port Z-parameters z11i=imp_gi+imp_s1; Appendix A: Electrical Characterization 164 z12i=imp_gi; z21i=imp_gi; z22i=imp_gi+imp_s2; %Calculate modeled 2-port Y-parameters Di=z11i.*z22i-z12i.*z21i; y11i=z22i./Di; y12i=-z12i./Di; y21i=-z21i./Di; y22i=z11i./Di; %Calculate modeled 2-port S-parameters s11i=((1-z0.*y11i).*(1+z0.*y22i)+z0.*y12i.*z0.*y21i)./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s12i=-2*z0.*y12i./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s21i=-2*z0.*y21i./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); s22i=((1+z0.*y11i).*(1-z0.*y22i)+z0.*y12i.*z0.*y21i)./((1+z0.*y11i).*(1+z0.*y22i)-z0.*y12i.*z0.*y21i); Appendix A: Electrical Characterization 165 Appendix B. List of Cadence Layouts Table B.1 lists the main Cadence layouts for all the CMOS-MEMS variable capacitors and the CMOS-MEMS phase shifter presented in this work. Table B.2 lists all the Cadence libraries with layouts that were created by the author. Layout Gap-Tuning Capacitor Gap-Tuning Capacitor with Updated Heating Resistors Initial Digital Capacitor Updated Digital Capacitor Phase Shifter Digital Capacitor Phase Shifter Charge Pump Design Contest GapTuning Capacitor Table B.1: List of Cadence Layouts Description Cell Name The gap-tuning capacitor presented in cap5 Chapter 3. The gap-tuning capacitor presented in Chapter 3 with the heating resistors cap6 sized to produce a uniform temperature. The initial digital capacitor presented in switch5 Chapter 4. The updated digital capacitor with reduced parasitic capacitance presented switch7 in Section 4.5.5 The phase shifter digital capacitor switch1 presented in Chapter 5. The phase shifter presented in Chapter phase_shifter2 5. A CMOS charge pump capable of charge_pump2 producing ≈ 9 V from a 3.3 V supply. The gap-tuning capacitor used in the cap2_test2 SRC/SIAA Design Contest. Library jz60_025 Process Jazz 0.35 µm BiCMOS jz60_028 Jazz 0.35 µm BiCMOS jz60_031 Jazz 0.35 µm BiCMOS jz60_032 Jazz 0.35 µm BiCMOS jz60_032 jz60_032 jz60_031 jz18_001 Jazz 0.35 µm BiCMOS Jazz 0.35 µm BiCMOS Jazz 0.35 µm BiCMOS Jazz 0.18 µm BiCMOS Table B.2: List of Cadence Libraries Library Category Process jz60_024 John Jazz 0.35 µm BiCMOS jz60_025 John Jazz 0.35 µm BiCMOS jz60_026 John Jazz 0.35 µm BiCMOS jz60_027 John Jazz 0.35 µm BiCMOS jz60_028 John Jazz 0.35 µm BiCMOS jz60_031 John Jazz 0.35 µm BiCMOS jz60_032 John Jazz 0.35 µm BiCMOS jz60_033 John Jazz 0.35 µm BiCMOS jz18_001 John Jazz 0.18 µm BiCMOS ibm7rf_001 John IBM 0.18 µm CMOS ibm7rf_002 Everything IBM 0.18 µm CMOS ibm7rf_003 Everything IBM 0.18 µm CMOS tsmc035_001 John TSMC 0.35 µm CMOS tsmc018_001 John TSMC 0.18 µm CMOS tsmc018_004 John TSMC 0.18 µm CMOS st7rf_003 John ST 0.25 µm BiCMOS Appendix B: List of Cadence Layouts 166