Performance Limits of RF Power CMOS by Usha Gogineni B.Tech. National Institute of Technology, Warangal (1996) M.S. Auburn University (1999) Submitted to the Department of Electrical Engineering and Computer Science in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy ARCHiVES MASSACHUSETTS INSTITUTE OF TECHVOLO. N MAR 10 2011 LIBRARIES at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY February 2011 © Massachusetts Institute of Technology 2011. All rights reserved. Author Department of Electrical Engineering and Computer Science October 28, 2010 Certified by Jesus A. del Alamo Professor of Electrical Engineering Thesis Supervisor Accepted by. Terry P. Orlando Chairman, Department Committee on Graduate Thesis Performance Limits of RF Power CMOS by Usha Gogineni Submitted to the Department of Electrical Engineering and Computer Science in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in Electrical Engineering Abstract Wireless and mobile communication systems have become ubiquitous in our daily life. The need for higher bandwidth and thus higher speed and data rates in wireless communications has prompted the exploration of millimeter-wave frequencies. Some of the applications in this regime include high-speed wireless local area networks and high data rate personal area networks at 60 GHz, automotive collision avoidance radar at 77 GHz and millimeter-wave imaging at 94 GHz. Most of these applications are cost sensitive and require high levels of integration to reduce system size. The tremendous improvement in the frequency response of state-of-the-art deeply scaled CMOS technologies has made them an ideal candidate for millimeter-wave applications. A few research groups have already demonstrated single chip CMOS radios at 60 GHz. However, the design of power amplifiers in CMOS still remains a significant challenge because of the low breakdown voltage of deep submicron CMOS technologies. Power levels from 60 GHz power amplifiers have been limited to around 15 dBm with power-added efficiencies in the 10-20% range, despite the use of multiple gain stages and power combining techniques. In this work, we have studied the RF power potential of commercial 65 nm and 45 nm CMOS technologies. We have mapped the frequency, power and efficiency limitations of these technologies and identified the physical mechanisms responsible for these limitations. We also present a simple analytical model that allows circuit designers to estimate the maximum power obtainable from their designs for a given efficiency. The model uses only the DC bias point and on-resistance of the device as inputs and contains no adjustable parameters. We have demonstrated a record output power density of 210 mW/mm and power-added efficiency in excess of 75% at VDs = 1.1 V and f = 2 GHz on 45 nm CMOS devices. This record power performance was made possible through careful device layout for minimized parasitic resistances and capacitances. Total output power approaching 70 mW was measured on 45 nm CMOS devices by increasing the device width to 640 gm. However, we find that the output power scales non-ideally with device width because of an increase in normalized on-resistance in the wide devices. PAE also decreases with increasing device width because of degradation in f. in the wide devices. Additionally PAE decreases as the measurement frequency increases, though the output power remains constant with increasing frequency. Small-signal equivalent circuit extractions on these devices suggest that the main reason for the degradation in the normalized output power and PAE with increasing device width is the non-ideal scaling of parasitic gate and drain resistances in the wide devices. Thesis Supervisor: Jesus A. del Alamo Title: Donner Professor 4 Acknowledgements This work would not have been possible without the support and guidance of numerous people. First, I would like to thank my thesis supervisor, Prof. Jesd's del Alamo. It was a privilege to work with him and learn from him over the past few years. He is a great teacher and an exceptional researcher. I am sure that the knowledge and skills I learnt from him will serve me well in my future career. I would also like to thank my thesis readers, Prof. Dimitri Antoniadis and Prof. Joel Dawson for their valuable guidance and suggestions that helped me improve the work presented in this thesis. I should thank my funding sponsors: MIT Presidential fellowship program, Intel PhD fellowship program and Semiconductor Research Corporation. The hardware for this thesis was generously provided by IBM Microelectronics. I need to thank David Greenberg, Alberto Valdes Garcia and Christopher Putnam for their help and guidance in designing test structures, coordinating their inclusion on the IBM test vehicles, and helping me with all the necessary approvals from IBM. I have spent two summers working at IBM in Burlington VT, and Fishkill NY on projects related to this thesis. Among the many great people I met and worked with during these summers, I should specially thank Hongmei Li for her help with understanding CMOS device modeling and small-signal extractions and Susan Sweeney for her expertise on Sparameter and Power measurements. My stay at MIT has been most pleasant mainly due to the friends and colleagues I met here. I would like to thank all the former and current members of the del Alamo research group: Jorg Scholvin, Joyce Wu, Niamh Waldron, Anita Villanueva, Yoshihiro Ikura, Dae-Hyun Kim, Sefa Demirtas, Jungwoo Joh, Ling Xia, Donghyun Jin, Tae Woo Kim, Jianqiang Lin, Alex Guo, Xin Zhao and Elizabeth Kubicki. Special thanks are due to Jungwoo Joh for the many long discussions on RF power devices and amplifiers that have directly contributed to the RF power model in this thesis. Finally, I would like to thank my parents, Dr. Ravindranath Tagore and Ramalakshmi and my sisters Sirisha, and Manisha for their support and motivation. Thanks mom and dad for encouraging me to pursue my dreams and pulling me through the hardest patches. To you, I dedicate this thesis. 6 Contents 1. Introduction 21 1.1 RF and Millimeter-wave Applications 21 1.2 Advantages and Challenges in using CMOS 23 1.3 Literature survey of 60 GHz power amplifiers 25 1.4 CMOS technologies for RF power applications 27 1.5 Thesis Goals 29 2. Experimental 31 2.1 Technology Description 31 2.2 Device Layout 33 2.2.1 65 nm Technologies: 33 2.2.2 45 nm Technology: 36 2.3 DC and S-Parameter Characterization 38 2.4 RF Power Characterization 41 2.5 Summary 45 3. RF Power Performance 47 3.1 65 nm Low Power Technology 3.1.1 1.2 V Standard Device 3.1.2 2.5 V 1/0 Device 47 47 52 3.2 65 nm High Performance Technology 54 3.3 45 nm Low Power Technology 60 3.3.1 Old Test Structures 60 3.3.2 New Test Structures 66 3.4 Comparison among different technologies 73 3.5 Comparison with literature 80 3.6 Summary 82 4. Model for RF Power 85 4.1 Model Description 86 4.2 Comparison of Model with Measurements 95 4.3 Discussion 102 4.4 Summary 104 5. Small-signal Circuit Extractions 105 5.1 Small-signal Circuit Extraction Methodology 105 5.2 Width Dependence of Equivalent Circuit Parameters 111 5.3 fT, fmax Sensitivity to Equivalent Circuit Parameters 113 5.4 Analytical Expressions for fT and fmax 115 5.5 Summary 118 6. Conclusions and Suggestions for Further Work 119 Appendix A. Device Library 125 A. 1. 65 nm Test Structures 125 A.2. 45 nm "Old" Test Structures 127 A.3. 45 nm "New" Test Structures 129 Bibliography 131 10 List of Figures Fig. 1. 1: Illustration of some of the applications in the millimeter-wave regime. (a) High-speed wireless local area network, (b) High data rate wireless personal area network, (c) Automatic cruise control and collision avoidance radar in automobiles, (d) Imaging for detection of hidden weapons for airport security applications. .................. 22 Fig. 1. 2: Output power, normalized to device width and supply voltage, as a function of device width. All previously published device data is shown in this figure. ......................................... 27 Fig. 1.3: Output power as a function of measurement frequency. All previously published device data is shown in this figu re ..................................................................................................................................................................... 28 Fig. 2.1: Illustration of a unit cell in 65 nm devices containing 24 fingers of 2 pm finger width. Gate is contacted on both sides of the device and is wired out at M3 level. Source and drain fingers are wired out at M4 level...........34 Fig. 2.2: Schematic of a W=8x24x2 Rm device showing 8 unit cells connected in parallel. The diffusion area of the unit cells is shown as the black outline. Gate, Source and Drain are wired out to the last metal level (LB) before connecting them to the RF pads...................................................................................................................................35 Fig 2.3 Layout of a unit cell in 45 nm technology containing 20 fingers of 2 gm finger width. Gate is connected on both sides of the device to reduce gate resistance. Gate, source and drain are wired out on alternate metal levels to reduce coupling capacitance between terminals ..................................................................................................... 36 Fig 2.4 Schematic of a W=4x20x2 pm device in 45 nm technology containing 4 unit cells connected in parallel. A ground plane on M1 and M2 levels is placed under the gate and drain signal lines to reduce substrate coupling. Source and substrate are shorted together and connected to the M1+M2 ground plane........................................ Fig. 2.5: Output characteristics (VGS and VDS = = 0.4 V to 1.1 V in steps of 0.1 V) and transfer characteristics (VDS = 38 0.05 V 1.1 V ) on 45 nm devices (W = 80 [im ).......................................................................................................39 Fig 2.6 Smith chart showing the S-parameters on a 65 nm standard device (W=96 gm) measured from 0.5 GHz to 40 GHz at VDD=1 V, ID=2 0 0 mA/mm. The blue solid lines show the S-parameters before de-embedding and the red dashed lines show the de-embedded S-parameters ................................................................................................. 40 Fig 2.7 Current gain and unilateral power gain as a function of frequency for a 96 gm wide device. Blue solid lines show data before de-embedding, while red dashed lines show the data after open and short de-embedding. f1 and fma improve w ith de-em bedding........................................................................................................................................4 1 Fig. 2.8 Schematic diagram of Maury Microwave Load pull system at MIT..........................................................42 Fig. 2.9 Sample RF power sweep at 2 GHz showing power gain, PAE and average drain current as a function of output power................................................................................................................................................................44 Fig. 3.1. Output characteristics (VGS = 0.2 V to 1.2 V in steps of 0.2 V) and transfer characteristics (VDS = 0.05 V and VDS = 1.2 V) on the standard 65 nm LP device (W = 96 pm)...............................................................................48 Fig. 3.2: Typical power sweep showing PAE and gain as a function of output power on several 65 nm LP standard devices. The figure also illustrates the maximum Pot, PAE and Put / W obtainable from this technology at VDs = 1.2 V and ID= 200 mA/mm - ... .................. --.... .. .. ---------... --. .............................................. 48 Fig. 3.3: Output power and PAE as a function of DC bias current illustrating the trade-off between Pout and PAE. ..49 Fig. 3.4: Measured output power, at various DC bias currents, for 65 nm LP CMOS devices of different widths.....50 Fig. 3.5: Normalized output power as a function of device width, measured at various DC bias currents on 65 nm LP 1 CM O S devices.............................................................................................................................................................5 Fig. 3.6: Measured peak PAE, on 65 nm LP CMOS devices, versus device width at different DC bias currents.......51 Fig. 3.7: Typical power sweep showing PAE and gain as a function of output power on several 0.25 gm long I/O devices in 65 nm LP CMOS technology. The maximum Pow, PAE, and Po,1, / W obtainable from this technology at VDs = 2.5 V and ID = 200 mA/mm are noted on the figure. .......................................................................... 52 Fig. 3.8: RF power measured at 8 GHz as a function of device width for I/O devices designed in 65 nm LP CMOS 53 techn o lo gy . .................................................................................................................................................................. Fig. 3.9: Peak PAE, measured at different bias currents, as a function of device width for I/O devices designed in 65 54 nm LP C M O S technology. .......................................................................................................................................... Fig. 3.10. Output characteristics (VGS VDs = 0.2 V to 1 V in steps of 0.2 V) and transfer characteristics (VDs = 1 V) on the standard 65 nm HP device (W = 96 pm).. Fig. 3.11: De-embedded fT and f., 0.05 V and = ............................................................................... 55 measured on 65 nm HP CMOS devices, are shown as a function of device w id th ............................................................................................................................................................................ 55 Fig. 3.12: Typical power sweep on several 65 nm HP devices. The maximum Pow, PAE and Pu, / W are also shown in th e fig ure .................................................................................................................................................................. 56 Fig. 3.13: Output power, measured at different frequencies, as a function of device width, for 65 nm HP CMOS devices. VDD = IV , ID= 25 mA /m m ............................................................................................................................ 57 Fig. 3.14: PAE measured at different frequencies, as a function of device width for 65 nm HP CMOS devices. VDD 1V , ID= 25 mA /m m . ................................................................................................................................................... 58 Fig. 3.15: Output power, measured at different frequencies, as a function of device width, for 65 nm HP CMOS devices. VDD = V, ID= 200 m A/m m .......................................................................................................................... 59 Fig. 3.16: PAE measured at different frequencies, as a function of device width for 65 nm HP CMOS devices. 1V , ID= 200 m A/mm . . --------------........ Fig. 3.17: Output characteristics (VGS and VDS = Fig. 3.18: - = --........................... ---. . ----....................................................................... 59 0.4 V to 1.1 V in steps of 0.1 V) and transfer characteristics (VDs = 0.05 V 1.1 V) on the standard 45 nm device (W = 90 ptm). fT VDD = .......................... ........... 60 and fma as a function of drain current density for single and double gate contact structures in 45 nm C M O S technology.......................................................................................................................................................6 1 Fig. 3.19: PAE and normalized output power as a function of measurement frequency for single gate and double gate contact structures in 45 nm CMOS technology. ............................................ 62 Fig. 3.20: De-embedded fmax as a function of unit finger width and number of fingers for all the test structures of the "old test site". WF = 1.5 pim was found to be the optimum finger width for the 45 nm CMOS technology. ......... 63 Fig. 3.21: Typical power sweep on several 45 nm CMOS devices, designed as part of the old test site. The maximum Pt, PAE and Po,, / W obtained from this test site at f = 2 GHz, VDD = 1.1 V and ID= 200 mA/mm are also shown in th e fig u re......................................................................................................................................................................6 4 Fig. 3.22: Peak PAE, measured at different frequencies, versus total device width. For the figure on the left, device width is increased by keeping NF constant at 60 and increasing WF from 0.5 jim to 5 jm. For the figure on the right, device width is increased by keeping WF constant at 1.5 jm and increasing NF from 20 to 120..........................65 Fig. 3.23: Normalized output power, measured at different frequencies, versus total device width. For the figure on the left, device width is increased by keeping NF constant at 60 and increasing WF from 0.5 gm to 5 jm. For the figure on the right, device width is increased by keeping WF constant at 1.5 jim and increasing NF from 20 to 120. 65 Fig. 3.24: Output characteristics (VGs = 0.4 V to 1.1 V in steps of 0.1 V) and transfer characteristics (VDs = 0.05 V and VDS = 1.1 V ) on 45nm devices (W = 80 ptm)........................................................................................................67 Fig. 3.25: De-embedded fT and f, as a function of device width for the new 45 nm test structures, measured at VDD = 1.1 V and ID= 200 m A /mm ...................................................................................................................................... 67 Fig. 3.26: Power sweep showing PAE and P0ut for devices with different widths. The figure also notes the maximum PAE, Po 0 t and Pout /W obtained the new set of structures in 45 nm CMOS technology...............................................69 Fig. 3.27: Normalized output power as a function of device width for the two impedance matching conditions. Optimizing for maximum power gives a higher P 111/W at a lower PAE.....................................................................70 Fig. 3.28: Peak PAE as a function of device width for the two impedance matching conditions. Optimizing for maxim um PAE gives a higher PAE at a lower Pout IW . ....................................................................................... 70 Fig. 3.29: Normalized output power as a function of measurement frequency. Pout /W is relatively independent of frequency for both impedance matching conditions ............................................................................................... 71 Fig. 3.30: Peak PAE as a function of measurement frequency. PAE decreases significantly with increasing frequency for both impedance matching conditions ............................................................................................... Fig. 3.31: PAE versus f/f. 71 for 45 nm CMOS devices of different widths and measured at various frequencies from 2 G H z to 18 G H z.........................................................................................................................................................72 Fig. 3.32: Output power versus device width for 65 nm low-power and 65 nm high-performance devices, measured at VDS = VDD, ID = 50 mA/mm, and f = 8 GHz.......................... ................ ................................. Fig. 3.33: PAE versus device width for 65 nm low-power and 65 nm high-performance devices, measured at V DD, ID = 73 VDS = 50 mA/m m , and f = 8 G H z..........................................................................................................................74 Fig. 3.34:De-embedded f. versus device width for 65 nm devices and old and new 45 nm device structures.........75 Fig. 3.35: Output power versus device width for 65 nm devices and old and new 45 nm device structures...........76 Fig. 3.36: Extracted normalized on-resistance as a function of device width for 65 nm devices and old and new 45 nm device structures....................................................................................................................................................77 Fig. 3.37: Output power versus on-resistance for 65 nm devices and 45 nm old and new test structures...............78 Fig. 3.38: Output power versus 1/ (RL+Rofl) illustrating the universal behavior of measured data for 65 nm and 45 nm (old and new ) devices..................................................................................................................................................78 Fig. 3.39: PAE versus f/fma, illustrating universal behavior of measured data for 65 nm and 45 nm (old and new) d ev ices.........................................................................................................................................................................7 9 Fig. 3.40: Normalized output power versus device width. The figure compares power data presented in this thesis to the values reported in literature. .................................................................................................................................. 80 Fig. 3.41: Output power versus operating frequency. The figure compares power data presented in this thesis to the values reported in literature.........................................................................................................................................8 Fig. 4.1: Circuit diagram of a reduced conduction angle power amplifier. .............................. 1 86 Fig. 4.2: Waveforms for input voltage and output current and voltage in the reduced conduction angle amplifier show n in Fig . 4.1.........................................................................................................................................................88 Fig. 4.3: Illustration of difference between DC and AC load line for RL = 100 Q. The DC operating point is at VDQ = 1.1 V, IDQ = 200 mA/mm. The AC operating point is at VDC = 1.1 V, IDC Fig. 4.4: Modeled locus of drain efficiency versus output power. Put and = 212 mA/mm. ............... 71D ............. 94 measurements at different RL are shown as symbols in the figure and show good agreement with the model........................................................................96 Fig. 4.5: Modeled locus of output power versus load resistance. Output power measurements at different RL are shown as symbols in the figure and show good agreement with the model. ................................................... 97 Fig. 4.6: Modeled locus of drain efficiency versus load resistance. Drain efficiency measured at different RL is ....... 98 shown as symbols in the figure and shows good agreement with the model.......................... Fig. 4.7: Drain Efficiency as a function of device width. The solid line represents 1D specified in model for predicting maximum Pu1 and optimum RL. The symbols are the measured 1D 99 at 2 GHz. ..................................... Fig. 4.8: Output power as a function of device width. The solid line represents the maximum output power predicted by the model. The symbols show the measured Pouk at 2 GHz...................................................................................100 Fig. 4.9: Average drain current as a function of device width. The solid line represents the average drain current predicted by the model. The symbols show the measured IDC at 2 GHz. .................................................................. 101 Fig. 4.10: Load resistance needed to obtain maximum Pu,. Solid line is optimum R predicted by model. Symbols are m easured data at 2 G H z.......................................................................................................................................102 Fig. 5.1: De-embedded fT and f., measured at VDs=1 V and at ID=1 0 0 mA/mm, as a function of total device width. 10 6 ................................................................................................................................................................................... Fig. 5.2: Small-signal equivalent circuit of a MOSFET including parasitic resistances and the substrate network..107 Fig. 5.3: Measured and Modeled s-parameters at VDD=l V, ID=100 mA/mm. Measured data is in symbols and the 109 m odel is the solid line. W =2x48 tm.......................................................................................................................... Fig. 5.4: Short circuit current gain and unilateral power gain as a function of frequency. W = 2x48 gtm, VDD = 110 = 100 m A/m m ............................................................................................................................................................ Fig. 5.5: Normalized intrinsic parameters (gm, gds, Cgs, Cgd) as a function of device width. VDD = 1 V, ID 1 V, m A/m m ...................................................................................................................................................................... ID = 100 111 Fig. 5.6: Normalized parasitic resistances vs. device width. Parasitic resistances are extracted from s-parameters m easured at VDD=1 V , ID= 10 0 mA/m m ..................................................................................................................... 112 Fig. 5.7: Schematic layout of a 2 cell (W = 96 im) and 8 cell (W = 384 pm) device, showing the need for additional wiring to connect the 8 cells in parallel. This leads to an increase in parasitic resistance as the device width is in creased .................................................................................................................................................................... 112 Fig. 5.8: Sensitivity of f1 and fa to the various small-signal equivalent circuit parameters. The figure plots the percentage decrease in fT and fm, for a 100% change in each of the parameters. W=96 ptm, VDD=1 V. m A mm.......................................................................................................................................................................1 ID=100 14 Fig. 5.9: Measured and calculated fT and fma as a function of device width. Measured data in solid symbols and lines and calculated data in open symbols and dashed lines. ....................................................................................... 117 List of Tables Table 1.1: Comparison of state-of-the-art 60 GHz CMOS power amplifiers ...................................................... 26 Table 2.1. Key device and process parameters for the device types studied in this thesis ........................ 32 Table 2.2: Definition of bias currents, voltages and figures of merit of an RF power amplifier. ............... 43 Table 4.1: Measured and modeled values of DC and RF power data on two different test structures designed in 45 nm CMOS technology. Details of the standard device layout can be found in Chapter 2. The spacing between gate fingers in the new test structure is twice that in the standard structure......................................................................103 20 Chapter 1 Introduction 1.1 RF and Millimeter-wave Applications Wireless and mobile communication systems have become ubiquitous in our daily life to the point where we cannot imagine life without them. Some of the more prevalent applications include cellular and cordless phones, pagers, RFID tags, Bluetooth, wireless local area networks (WLAN) and wireless personal area networks (WPAN) [1-2]. These applications operate over a wide range of frequency from 900 MHz to 5.6 GHz. The need for higher bandwidth, and thus higher speed and data rate, in wireless communications has prompted scientists and engineers to explore millimeter wave frequencies (30 GHz to 300 GHz range). Within this frequency range, the 60 GHz band has received a lot of attention in the past few years because of the availability of 7 GHz of unlicensed spectrum around that frequency. At 60 GHz, there is an unusually high attenuation of signals due to atmospheric oxygen adsorption [3-5]. This attenuation precludes long-range communications, but provides an extra spatial isolation for secure short-range point-to-point links, while reducing ~jjj~ co-channel interference and allowing frequency reuse in an indoor dense local network. Moving to higher frequencies also reduces the form factor of antennas, as antenna dimensions are inversely proportional to carrier frequency. Therefore, for a fixed area, more antennas can be used, thus increasing antenna gain, which in turn helps direct the electromagnetic energy to the intended target [6]. 40 (a) (b) (c) (d) Fig. 1. 1: Illustration of some of the applicationsin the millimeter-wave regime. (a) High-speed wireless local area network, (b) High data rate wireless personal area network, (c) Automatic cruise control and collision avoidance radarin automobiles, (d) Imagingfor detection of hidden weaponsfor airportsecurity applications. Some of the applications in the millimeter-wave regime are enumerated below [3-5, 7-10] and illustrated in Fig. 1.1. " The 60 GHz band allows bit rates of several Gb/s, with a range of up to 10 meters. This could be used to create wireless personal area networks (WPANS) and local area networks (WLANS), allowing wireless Gigabit Ethernet, laptop docking stations, and mobile device synchronization (cellphones, ipods, MP3 players, cameras, handhelds/PDAs, etc) (Fig 1.1 (a) and (b)) e Fig. 1.1 (c) illustrates the use of millimeter-wave frequencies for automotive radar applications. The FCC has allocated 76-77 GHz frequency band for long range (100 m) automatic cruise control and automotive collision avoidance systems and the 22 - 29 GHz range for short range automotive radar applications. * Millimeter-wave imaging in the 94-95 GHz band is being used for radar cloud detection and all-weather landing, military surveillance, navigation and precision targeting, detection of hidden weapons for security applications (Fig. 1.1 (d)), and medical imaging for tumor detection, temperature measurements, blood flow, and water/oxygen content measurements. * Another application where high data rate is a necessity is the wireless transmission of uncompressed high-definition television (HDTV) signals. Compressed HDTV signals need data rates of 50 Mbps, while uncompressed signals could need data rates as high as 1.5 Gbps. 1.2 Advantages and Challenges in using CMOS Most of the millimeter-wave applications described in the previous section are consumer applications and hence are very cost sensitive. Traditionally these systems were designed using a 23 combination of Silicon CMOS, bipolar, III-V HEMT, and HBT technologies making them prohibitively expensive. Recent advances in silicon technology, driven by high performance digital applications, have made CMOS the ideal candidate for consumer applications in the millimeter-wave regime. CMOS allows for higher levels of integration (and thus smaller system size), lower cost and lower power dissipation, all of which are important considerations in portable applications. However, designing wireless transceivers in CMOS presents quite a few challenges. The most difficult component to design in CMOS is the power amplifier. Some of the factors affecting power amplifier design in CMOS technology are listed below [2, 11-12]: (a) The low breakdown voltage of deep sub-micron technologies limits the maximum gatedrain voltage and the output power. This is because the output voltage at the transistor's drain normally reaches 2 times the supply voltage for classes B and F, and around 3 times the supply for class E operation. Thus, transistors have to operate at a lower supply voltage, delivering low power. Additionally CMOS technology has lower current drive and fmax compared to III-V devices, which means that the gain provided by a single stage is very low necessitating the need for multiple stages. (b) Silicon CMOS generally uses a highly doped conductive substrate. This results in substrate interaction in a highly integrated CMOS IC. The leakage from an integrated power amplifier might affect the stability of, for example the VCO in a transceiver chain. (c) Since the inherent output device impedance in CMOS is very low, impedance matching becomes very difficult, requiring higher impedance transformation ratios. Additionally, the output matching elements require low loss, and good thermal properties since there are usually significant RF currents flowing in these elements. If CMOS technology is used, the losses in the substrate will decrease the quality factor of the passive elements in the matching network. (d) Conventional transistor models for CMOS devices have been found to be only moderately accurate for RFICs, and need to be improved for analog operation at radio frequencies. Accurate large signal CMOS RF models and substrate modeling are critical to the successful design and operation of integrated CMOS RF power amplifiers. 1.3 Literature survey of 60 GHz power amplifiers Despite the challenges outlined in the previous section, universities and companies around the world have been working on designing millimeter-wave CMOS power amplifiers. Table 1.1 shows a comparison of 60 GHz CMOS power amplifiers published over the past two years. The most popular technology for designing 60 GHz power amplifiers is the 90 nm CMOS technology, though a few recent designs have also been demonstrated in 65 nm and 45 nm CMOS technologies. The power amplifiers typically contain 2 to 4 stages of common-source or cascode configurations. The cascode topology achieves good isolation, high gain, good stability and low power consumption, but requires a large power supply voltage. The common-source configuration maximizes the output voltage and current swings yielding high output power and good linearity. The first stage of the power amplifier is often designed as a cascode stage to improve isolation, and subsequent stages are usually common-source configurations. Transformer coupling and power combining techniques are often employed to further increase the output power and PAE. Saturated power levels as high as 19.9 dBm [13] and power added efficiency (PAE) of 25.7% [14] have been demonstrated at 60 GHz. Tech Node Gain (dB) P1dB Psat (dBm) PAE (%) VDD (dBm) 45 nm 19 - 7.9 19.4 45 nm 6 11 13.8 45 nm 16 7.6 45 nm 20 65 nm (SoI) 65 nm Stages Topology Reference 1.2 3 Common source [15] 7 1.1 2 Push-pull [16] 12 12.3 1.8 2 Cascode [17] 11.2 14.5 14.4 2 2 Differential cascode [18] 16 12.7 14.5 25.7 1.8 2 Cascode [14] 13.4 12.2 13.8 7.6 1.2 4 Common source [19] 65 nm 15.8 2.5 11.5 11 1 3 Differential CS [20] 65 nm 4.5 6 9 8.5 1.2 1 Common source [21] 90 nm 20 8.2 12 9 1.2 4 Cascode + power combining 22] 90 nm 26.1 10.5 14.5 10.2 1.8 3 Cascode + DAT power combining [23] 90 nm 15 10.2 12.2 19.3 1.2 3 Transformer coupled CS [24] 90 nm 10.1 5.1 8.5 7.7 1 2 coupled differtial [25] 90 nm 30 10.3 13.8 12.6 1.8 3 Cascode [26] 90 nm 10 8.8 12.6 6.9 1 3 Common source [27] 90 nm 20.6 18.2 19.9 14.2 1.2 2 4-way power combining 13] 90 nm 4.2 12.1 14.2 5.8 1 2 4-way current combining [28] 90 nm - 9 12.3 8.8 1 2 Differential CS, transformer coupled [29] 90 nm 9.8 7.5 11.2 5 1.2 3 Common source [30] 90 nm 16.3 10.5 11.5 8.5 1 3 Cascode + CS [31] 90 nmn 11 8.3 11.2 3.6 1.2 4 Differential CS + Power combining 32] (V) Table 1.1: Comparison of state-of-the-art60 GHz CMOS power amplifiers. I.. .... -V-............. ............................. . ....... ....... 1.4 CMOS technologies for RF power applications The previous section has shown that 60 GHz CMOS power amplifiers exhibit relatively low power levels (<15 dBm) at very low efficiencies (<20 %). The best way to improve the power amplifier performance is by improving the power performance of the transistor. However, very little work has been published on the RF power performance of CMOS devices [33-37]. The published device data is summarized in Fig. 1.2 and Fig. 1.3. 180 160 N 90 nm [37] 140 VDD=1.5V 120 100 80 * 60 406540 20 -VDD 2090 nm [33] 45 nm [36] = 1.2 V , 10 = 1V nm [34] VoQ 0' VDD = 1 V 100 1000 Device Width (gm) 10000 Fig. 1. 2: Output power, normalized to device width and supply voltage, as a function of device width. All previously published device data is shown in thisfigure. Fig. 1.2 shows all the previously published data for normalized output power (normalized to device width and supply voltage) as a function of device width for CMOS technologies beyond 90 nm. The operating frequency was different for each of these measurements and will be discussed further in Fig. 1.3. The most comprehensive study over a wide range of device widths was published in [33]. However, the bias currents in this study were quite low (25 mA/mm) and ........... ........................... .................. the source and load impedances were optimized for maximum PAE as opposed to maximum power, resulting in low overall normalized power levels. The 45 nm device shows lower power density because it was measured at a much higher frequency (35 GHz). The most common technique used when designing for high power levels is to use a device with large device width. However, Fig. 1.2 clearly shows that the normalized output power decreases with increasing device width. This limits the obtainable power level is CMOS devices. The reasons for the decrease in normalized output power with device width will be investigated in this thesis. 90 80 ,70 E 60 50 0 040 65 nm [33] L 30 0 45 nm [36] VDD = 1 V 20 VDD= 1.2 V 10 90 nm [37] VDD= 0 0 1.5 V 10 20 30 40 Frequency (GHz) Fig. 1.3: Output power as a function of measurementfrequency. All previously published device data is shown in thisfigure. Fig. 1.3 shows the published data for output power as a function of measurement frequency. The figure shows that the maximum power delivered by a given CMOS device is strongly dependent on the frequency of operation and decreases significantly with increasing frequency. The main reason for this decrease is that the parasitic resistances and coupling capacitances, 28 associated with wiring multiple fingers in wide devices, become significant at high frequencies thus limiting the output power and efficiency. At first glance, it might look like the data in Fig. 1.3 is in contradiction to the data in Table 1.1. However, when the 60 GHz power amplifier circuit data from Table 1.1 is normalized to obtain the power per device, we find that most of the designs show an output power of less than 8 mW per device. Further, the output power drops to less than 5 mW per device when the supply voltage is normalized to 1 V. This is consistent with the data in Fig. 1.3 and illustrates the main problem with using CMOS devices at high frequencies. In this thesis, we will explore various device design options to understand the main detractors to power performance at high frequencies and suggest improvements to device layout that would enable CMOS use at millimeter-wave frequencies. 1.5 Thesis Goals The main goal of this thesis is to explore the performance limits of RF power CMOS. Through a combination of detailed RF characterization, analysis and modeling, we hope to answer the following questions: (a) What is the limiting power-frequency locus for 45 nm and 65 nm CMOS generations? (b) What limits the power in the millimeter-wave regime? (c) Is there a way to break these limits on a CMOS platform? The thesis is organized as follows. Chapter 2 describes the technologies and device structures studied in this thesis and the types of experiments carried out on them. Chapter 3 presents measured RF power data as a function of device width and operating frequency for a variety of devices designed in 65 nm high-performance, 45 nm high-performance and 45 nm low-power technologies. Chapter 4 describes a first order model that predicts output power - drain efficiency 29 locus for any given bias point and device on-resistance. This model will be shown to be consistent with the measured data of chapter 3. Chapter 5 investigates the reasons for the decrease in output power with increasing device width and frequency through small-signal equivalent circuit extractions. The final chapter (chapter 6) draws key conclusions about the RF power capability of CMOS technologies. Chapter 2 Experimental This chapter describes the technologies and devices studied in this thesis, as well as the types of experiments carried out on them. I will first describe the device technologies used in this work and the various test structures designed in each of these technologies. I will then broadly describe the characterization carried out on these devices: DC, S-parameter, and RF power characterization. 2.1 Technology Description The device technologies studied in this work are foundry 65 nm high-performance (65 nm HP), 65 nm low-power (65 nm LP) and 45 nm low-power (45 nm LP) bulk CMOS technologies manufactured at IBM [38-41]. Each of these technologies offers thin oxide and 1/0 device options, as well as multiple VT options in the thin oxide devices. This thesis focuses mostly on the regular VT thin oxide device (which we will call the standard device), with some data presented on the 2.5 V 1/0 device in 65 nm LP technology. The devices wil-l be labeled according to their ITRS equivalent technology node: Lgnn = 45 nm, 65 nm, and 250 nm, with the understanding that the physical gate length is different from those values. Table 2.1 shows the key device and process parameters for the different devices studied in this thesis. Parameter ITRS Equivalent Node 65 nm 65 nm 45 nm High Performance Low Power Low Power 65 nm 0.25 [tm 65 nm 0.25 im 45 nm 50 230 60 230 40 Oxide Thickness (A) 12.5 52 19.5 53.5 18.2 Nominal Voltage (VDD) (V) 1.0 2.5 1.2 2.5 1.1 0.29 0.41 0.39 0.41 0.46 780 650 600 640 540 10 0.2 0.3 0.2 0.15 Ldesign(nm) Vtsat (VDS=VDD) (V) Ion (VDS=VGS=VDD) (gA/gm) Ioff (VDS=VDD,VGS=O) (nA/gm) Table 2.1. Key device andprocess parametersfor the device types studied in this thesis The 65 nm high-performance technology is a foundry technology developed for SRAM, logic, mixed signal and mixed voltage input/output (1/0) applications [38]. The standard device in this technology has a nominal gate oxide thickness (tox) of 12.5 50 nm and a nominal power supply voltage 1/0 device with oxide thickness of 52 (VDD) A, a minimum gate length (Ldesign) of of 1.0 V. The technology also offers a 2.5 V A and a minimum design length of 0.23 gim. The standard device delivers a drive current (In) of 780 gA/gm with an off-state leakage (Ioff) of 10 nA/gm, while the 1/0 device delivers Ion = 650 gA/gm at Ioff = 0.2 nA/gm. The 65 nm low-power technology is primarily targeted towards mobile applications and focuses on process simplicity, low cost, and reduced power consumption [39]. The back-end-ofthe-line (BEOL) metallization for this technology is identical to that of the 65 nm highperformance technology. The standard device has tox = 19.5 A, Ldesign = 60 nm and VDD = 1.2V. The thicker gate oxide and longer gate length, as compared to 65 nm HP technology, help in reducing off (0.3 nA/gm), but also reduce Ion (600 gA/tm). The 2.5 V 1/0 devices in this technology have very similar specifications to those in the high-performance technology. The next generation for the low cost, low power applications is the 45 nm low-power CMOS technology [40-41]. Only the thin oxide, regular VT device was studied in this technology. The device has tox = 18 A, Ldesign = 40 nm and VDD = 1.1 V and exhibits an extremely low Ioff (0.15 nA/gm) at Io=540 gA/gm. 2.2 Device Layout A variety of device layouts were designed to provide a broad range of total device width. The designs were fabricated at IBM Microelectronics as part of their regular foundry shuttles. The details of the device layout in each of the technologies are described below. 2.2.1 65 nm Technologies: The same set of test structures were laid out and studied in both 65 nm HP and 65 nm LP technologies. The test structures were designed using ten metallization levels. Ml, M2, M3, M4 are thin (iX) Cu levels, BI, B2, B3 are 2X Cu levels, EA, EB are 4X Cu levels and LB is the final thick Al level. All devices were designed at the minimum gate length (50 nm for the HP technology and 60 nm for the LP technology). The total device width (W) is the product of the number of parallel device cells (Nc), times the number of device fingers per cell (NF), times the unit finger width (WF). We will label devices as W = Nc x NF x WF. Device structures with total device width ranging from 48 gm (Nc =1) to 38 mm (Nc = 792) were designed by connecting multiple unit cells in parallel. A schematic of the unit cell can be seen in Fig. 2.1. Each unit cell contains 24 gate fingers of 2 pm finger width. The gate poly fingers were connected together on both sides of the device to reduce gate resistance. Substrate contacts were placed on either side of the device to minimize substrate resistance. The two gate contacts were connected together on M2 and M3 levels. The individual source fingers and drain fingers were connected together at M4 level. Thicker metallization levels were then used to route the gate, source and drain to the RF pads. Fig. 2.1: Illustration of a unit cell in 65 nm devices containing 24 fingers of 2 pm finger width. Gate is contacted on both sides of the device and is wired out at M3 level. Source and drain fingers are wired out at M4 level. Fig. 2.2 shows the layout of an 8x24x2 m device, with 8 unit cells connected in parallel. The layout uses a double-sided source access to the device on the thick Al level (LB). The gates of the individual cells were connected together at M3 level, the sources at M4 level, and the drains at B 1 level. The gate, source, and drain were then wired up to LB level and connected to the signal and ground pads. In connecting the intrinsic device to the RF pads, metal layers were strapped together, wherever possible, to reduce parasitic wiring resistance. ........................................................ A set of 1/0 devices with thicker gate oxide and gate length of 0.25 gm were also designed. The I/O devices have the same layout variations as the standard devices. A detailed list of the various layout variations designed for 65 nm can be seen in appendix A. Fig. 2.2: Schematic of a W=8x24x2 pn device showing 8 unit cells connected in parallel. The diffusion area of the unit cells is shown as the black outline. Gate, Source and Drain are wired out to the last metal level (LB) before connecting them to the RF pads. In order to facilitate accurate de-embedding of measured S-parameters, open and short deembedding structures, specific to a given device layout, were custom designed for some of the devices. e The open structures do not contain source or drain diffusions or gate polysilicon regions in the device area. The substrate contact regions and the metallization on the source fingers (which are eventually connected to the ground pad) are identical to the corresponding device layout. However, there is no metallization (Ml, M2, M3 levels) on the drain fingers. The metallization needed to connect the unit cells in parallel (levels M4 and up) was kept identical to the original device layout. .................. ....................................................................................................................... * The short structures are identical to the open structures, but contain an M4 plate that shorts out the gate, source and drain terminals. 2.2.2 45 nm Technology: -~W- Substrate Ring (M1+M2) M- M M- M--M- M- M- M Fig 2.3 Layout of a unit cell in 45 nm technology containing 20 fingers of 2 Um finger width. Gate is connected on both sides of the device to reduce gate resistance. Gate, source and drain are wired out on alternate metal levels to reduce coupling capacitancebetween terminals. Two sets of test structures were designed in the 45 nm technology. Both sets were designed using 7 layers of metallization. M1, M2, M3, M4, M5 are thin (iX) Cu levels, FA is a 6X Cu level and LB is thick Al level. In the first set of structures, the device width was increased by either (a) keeping the number of fingers (NF) constant (NF=60) and increasing the unit finger width (WF) from 0.5 lm to 5 gm, or (b) keeping the finger width constant (WF -. 5 gm) and increasing NF from 20 to 120. All devices in the set have a gate length of 40 nm and contain a substrate contact ring that surrounds the device and reduces the substrate resistance. In all other respects, the layout of each device is similar to the 65 nm unit cell layout described above. Details of the various layout variations can be found in appendix A. In the second set of devices, the device width was increased by designing an optimum unit cell and connecting multiple unit cells in parallel. The layout of the unit cell can be seen in Fig. 2.3. Each unit cell contains 20 fingers of 2 tm finger width. The choice of 2 ptm for finger width was found to be the optimum in a study of the RF power performance on the first set of 45 nm test structures. The gate was connected on both sides of the device to reduce gate resistance. When connecting the gate poly fingers to the first metal level (Ml), the overlap between poly and Ml was minimized to reduce parasitic gate capacitance. Substrate contacts were placed in a ring around the device to reduce substrate resistance. To minimize the coupling capacitance between the device terminals, the gate, source, and drain were wired out on alternate metal levels: the source was wired out at M3 level, the gate at M5 level, and the drain at FA level. To further reduce coupling capacitance, M2 level was eliminated on the lower part of the drain fingers, under the source connection. Fig. 2.4 shows the layout of a 160 .imwide device with 4 unit cells connected in parallel. The signal lines from the RF pads were designed in the thick Al level (LB). A ground plane containing a mesh of Ml and M2 levels was designed to be present under the signal lines and all around the device to reduce substrate coupling. The source and substrate terminals of the device are shorted together and connected to the ground pads through the Ml+M2 ground plane. In connecting the intrinsic device to the RF pads, metal layers were strapped together wherever possible to reduce parasitic wiring resistance. ....................................... .......... .. . ......... S Plane _Calibration Fig 2.4 Schematic of a W=4x20x2 pm device in 45 nm technology containing 4 unit cells connected in parallel.A ground plane on M1 and M2 levels is placed under the gate and drain signal lines to reduce substrate coupling. Source and substrate are shorted together and connected to the M1+M2 ground plane. In order to facilitate accurate de-embedding of measured S-parameters, on-wafer open and short structures are required. Due to space crunch on the test site, we could not design individual open and short structures for each device variation. Instead, a common set of open, short, load, and thru de-embedding structures was designed. The de-embedding structures de-embed all parasitics up to the end of the signal lines as denoted by the calibration plane shown in Fig. 2.4. The open structure simply removes the active device from between the signal lines, while the short structure removes the active device and shorts out the gate, drain, and source terminals. A complete list of the various layout variations in this set is provided in appendix A. 2.3 DC and S-Parameter Characterization DC Measurements on all the devices were performed using an Agilent 4155B parameter analyzer. All measurements were performed on-wafer using standard RF probes (ground-signal38 ...... _'N ....... ...... .... .... ............. ground). Sample output characteristics and transfer characteristics on a standard 45 nm highperformance device with W = 80 tm are shown in Fig. 2.5. 600 600 5nmCMOS 500 45nm CMOS W=2x20x2 jm 500 W=2x2x22 sm L=40 nm L 40 nm E 400 E - E 400 E E300 E 300 200 200 100 100 a 0 0 0.2 0.4 0.8 0.6 VDs (V) Fig. 2.5: VDS= 1-V 1 1.2 VDS 50 mV 0 0 0.2 0.6 0.4 0.8 1 1.2 Vs (V) Output characteristics (VGS = 0.4 V to 1.] V in steps of 0.] V) and transfer characteristics(VDs = 0.05 V and VDS = 1.1 V) on 45 nm devices (W = 80 pm). S-parameter measurements, from 0.5 GHz to 40 GHz, were performed using an Agilent 8510C network analyzer. DC bias, during S-parameter measurements, was provided through bias-tees using Agilent 4155B. S-parameters were often measured at VDS = VDD and VGS Set to ensure a constant drain current density across all devices. The measurement system was calibrated till the probe tips by measuring short, open, load and through structures (SOLT) on a standard calibration substrate. S-parameters were then measured on each device and, if available, on the corresponding open and short de-embedding structures. The two-port parameters of the intrinsic transistor were then obtained by de-embedding the on-wafer parasitics according to the industry standard open and short de-embedding technique described in [42]. Fig. 2.6 shows the S-parameters of a standard 65 nm CMOS device with W = 96 gm, measured at VDS = 1 V and ID = 200 mA/mm. The solid blue lines show the measured S- parameters before de-embedding and the dashed red lines show the S-parameters after open and short de-embedding. It is clear that the on-wafer parasitics significantly impact the measured Sparameters and hence proper de-embedding is essential for extracting the frequency metrics and small-signal parameters of the intrinsic device. +jl.0 +j2.0 +jO. 5 S21/10 S22*5 +j0.2 +j5.0 0.0 00 J-50 -jo.2 Si11 -jO. 5 -j2.0 -ji.0 Fig 2.6 Smith chartshowing the S-parameters on a 65 nm standarddevice (W=96 Um) measured from 0.5 GHz to 40 GHz at VDD=J V, ID= 2 0 0 mA/mm. The blue solid lines show the Sparametersbefore de-embedding and the red dashed lines show the de-embedded S-parameters. The measured S-parameters allow us to extract the unity current gain frequency (fT) and maximum oscillation frequency (fma) of the device. Fig. 2.7 shows the current gain (h21) and the unilateral power gain (U) as a function of frequency for a typical 65 nm device measured at VDS = 1 V and ID= 200 mA-mm. The figure shows that both h2 1 and U increase after de-embedding, thus resulting in an improved fT and fma. The extracted fT increased from 112 GHz to 189 GHz, while the extracted fma increased from 160 GHz to 216 GHz after open and short de-embedding. 40 ---- ........... -------:-.. ................. .......... 70 50 --- 60 - - Before Deembedding -After Deembedding i40 50 9 40 35 40 I 45 0 300 25 1 C .c:i 30 20 10 20 -' 65 nm CMOS W = 2x24x2 gm 15 10 %%%%%%%1 VDD =1 V ID= 5 % 200 m/ 0 0 110 100 Frequency (GHz) T f '"" 1000 Fig 2.7 Current gain and unilateral power gain as a function of frequency for a 96 pn wide device. Blue solid lines show data before de-embedding, while red dashed lines show the data after open and short de-embedding. fT andfaux improve with de-embedding. 2.4 RF Power Characterization RF power measurements were made in the 2 - 18 GHz range using a Maury Microwave loadpull system. A schematic diagram of the system can be seen in Fig. 2.8. The RF source provides a single frequency RF signal to the gate of the device. The input power is measured through a directional coupler. DC bias at the gate and drain is provided through bias tees. Mechanical tuners set the source (gate side) and load (drain side) impedances to the device. The RF input and output power as well as the bias currents and voltages are measured, and allow computation of all the relevant figures of merit. These are defined in Table 2.2. More details of measurement setup and procedure can be found in [43-44]. Fig. 2.8 Schematic diagram of Maury Microwave Load pull system at MIT. Parameter Definition VGQ DC bias at the gate of the device VDQ DC bias at the drain of the device ID DC drain current in the absence of RF input IG Average current in the gate of the device IDC Average current in the drain of the device pi RF power available at the input of the device. IDQ Expression power Pout RF output power that leaves the device PDC DC power into the device PDC = IGVGS + IDCVDS (2.1) GT flD Gr = Power Gain Drain efficiency measures the amount of Pout Pout -- - IGVG + IGVGS + IDCVDS PDc _ riD - DC power converted to RF output power PAE Power-added PAE efficiency measures (2.2) out Pin = (2.3) Pout - Pin PDC the amount to DC power converted to RF out power, taking into account the amount of RF power that enters the device PDC ( n (2.4) Pout (1 ) GT Table 2.2: Definition of bias currents, voltages andfigures of merit of an RF power amplifier. Most of the power measurements were performed with VDQ=VDD and VGQ set to ensure a constant drain current density across all devices at low input power conditions. Throughout this thesis, we will be quoting the bias condition as a combination of VDQ and IDQ (which is achieved with a gate bias VGQ). During the power measurement, the gate bias is held constant, rather than the actual IDvalue. Power measurements were optimized to provide either peak output power or peak PAE, by alternating source- and load-pull measurements at an input power level that is a few dB below the level corresponding to peak PAE. Power sweeps were made at the optimum load and source impedances, and the maximum PAE and associated output power were noted. This optimization was repeated for each bias point and frequency on each device. A typical power measurement sweep at 2 GHz is shown in Fig. 2.9. The output power and PAE initially increase with increasing input power. Once the device enters compression, the gain drops, leading to an eventual saturation of the output power and a peaking and drop in the PAE. The peak PAE point is selected as the optimum, and the optimum PAE and associated Pout are reported. ................. 80 25 70 20 60 Peak PAE E +-lc 5 50 XiS 0 015 40 j 30 S10 0 o 5 ' PAE- 45 nm CMOS 20 W = 2x20x2 gm VDD =.1 V 10 ID 0 - U 0 2 200 6 4 mA/mm 0 1 8 10 0 Output Power (dBm) Fig. 2.9 Sample RF power sweep at 2 GHz showing power gain, PAE and average drain current as a function of output power. In reduced conduction angle amplifiers, it is common to observe that as the input power increases, the measured DC drain current changes from its low-power value (the bias current). This is known as "self-biasing." The origin of self biasing is the non-linearity that enters when the input signal swings below the device's threshold voltage or above the voltage that leads to the device entering the linear regime [45]. While the average vGs remains at its bias value (VGQ), swinging the input below Vt will not result in a negative drain current, but rather in zero current. Similarly, swinging the input above VGSsat will clamp the drain current at IDsat. This leads to an asymmetry in the drain current waveform, thus changing the average value of iD from the DC value. If a device is biased at a vGS close to Vt, the average drain current will increase with increasing input power. On the other hand, if the device is biased at vGS close to VGSsat, the average drain current will decrease with increasing input power as is the case in Fig. 2.9. Self44 biasing is important because it affects the measured PAE. For instance, in Fig. 2.9, the average current decreases because of self-biasing, and this results in an increased value of peak PAE. 2.5 Summary In this chapter, we have described the layout of device test structures used to study RF power performance as well as the technologies that were used to fabricate these test structures. A detailed description of the measurement setup used to characterize the DC, S-Parameter and RF power characteristics was also provided. The next chapter will show extensive RF power measurement data on 65 nm HP, 65 nm LP, and 45 nm LP devices. 46 Chapter 3 RF Power Performance This chapter presents extensive data on the DC, S-parameter and RF power characteristics of devices in 65 nm low-power (65 nm LP), 65 nm high-performance (65 nm HP), and 45 nm lowpower (45 nm LP) technologies. This data allows us to estimate the maximum power capability of each of these technologies. Comparison among the three technologies, as well as comparison to published literature will also be presented. 3.1 65 nm Low Power Technology We studied both the standard and 1/0 devices in 65 nm low-power technology. The standard device has a gate length of 60 nm and operates at length of 0.25 gm and operates at VDD = 2.5 VDD = 1.2 V, while the 1/0 device has a gate V. 3.1.1 1.2 V Standard Device The output characteristics and transfer characteristics on a 65 nm LP standard device with W = 2x24x2 gm are shown in Fig. 3.1. The normalized drive current of the device is around 653 mA/mm at VGS=VDS=l.2 V. ...... .. .... 700 600 500 600 65 nm LP CMOS W=2x24x2 pm 500 L= 60 nm VDS=1.2 V E E 400 400 300 E 300 200 200 DS=50 100 mV 0 0 0.2 0.4 0.6 1 0.8 0 1.2 0.2 0.4 0.8 0.6 1 1.2 VGS (V) VDS (V) Fig. 3.1. Output characteristics (VGS = 0.2 V to 1.2 V in steps of 0.2 V) and transfer characteristics(VDS = 0.05 V and VDS = 1.2 V) on the standard65 nm LP device (W = 96jum). 70 40 Max. P.,/ =150.8 mW mn 35 60 50 30 4 25 0R40 20 W=481 im < 30 15 ( 20 L 10 10 5 0 0 5 10 15 Output Power (dBm) Fig. 3.2: Typical power sweep showing PAE and gain as a function of output power on several 65 nm LP standard devices. The figure also illustrates the maximum Pout, PAE and Pau, / W obtainablefrom this technology at VDS = 1.2 V and ID = 200 mA/mm. .............. 16 14 70 68 L 66 12 6 64 E ~ 10 O -I~062 8 -60 y 58 56 65 nm LP CMOS Q 4 W = 2x24x2 gm VDS=1 V 2 f=8GHz 0 0 .50 50 100 150 200 250 52 300 350 Bias Current (mA/mm) Fig. 3.3: Output power and PAE as a function of DC bias current illustrating the trade-off between Pont and PAE. Power measurements were performed at 8 GHz on devices with different widths at various bias conditions. When making each of these measurements, the source and load impedances were optimized for maximum power-added efficiency. The input power was then swept and the maximum PAE and the corresponding output power were noted. Fig. 3.2 shows such power sweeps on three different devices, with the measured PAE and power gain plotted as a function of the measured output power. At a DC bias of VDS = 1.2 V and ID = 200 mA/mm, a maximum normalized output power of 150.8 mW/mm was measured on the 48 gm wide device, a maximum PAE of 62.8 % was measured on the 96 gm wide device, and a maximum output power of 87.7 mW was measured on the device with W = 1536 gm. Fig. 3.3 examines the dependence of output power and PAE on the DC bias current. The output power keeps increasing as the bias current increases, while the PAE decreases with ............ . .... ....................... .. .... . .............. .. ..... .. . .. ...... .. ...... ........... . increasing bias current. For most measurements in this thesis, a bias current of 200 mA/mm was chosen, to ensure a reasonably high output power at a decent PAE. 25 65 nm LP CMOS VDS=1.2V 20 E f=8GHz 15 0 0 -- ID= 25 mA/mm + lD = 100 mAmm -D = 200 mAmm 5 -' = 300 mA/mm 0' 10 100 1000 10000 Device Width (gm) Fig. 3.4: Measured output power, at various DC bias currents,for 65 nm LP CMOS devices of different widths. The measured output power at various bias currents is shown as a function of device width in Fig. 3.4. The output power scales ideally with device width for small device widths, but starts to saturate in larger width devices (W > 400 jim). The non-ideal scaling of output power at large widths is more evident when the normalized output power (Po, / W) is plotted as a function of device width (Fig. 3.5). Another observation is that the non-ideality increases with increasing bias currents. Fig. 3.6 shows the measured peak PAE as a function of device width, for 65 nm LP CMOS devices, at various bias currents. The peak PAE shows a marked decrease with increasing width. Also, for any given device width, the peak PAE decreases as the bias current increases (as was . ..... .... .... . -- ---------- 180 - E 160 - + 140 ~a +D = 100 mA/mm D= 200 mA/mm S120 + = ID 25 mA/mm = 300 mA/mm S100 80 - 0 60 * 0 40 20 - - 65 nm LP CMOS VDS= 1.2 V f = 8 GHz 0 100 1000 10000 Device Width (gm) Fig. 3.5: Normalized output power as a function of device width, measured at various DC bias currents on 65 nm LP CMOS devices. 80 - 65 nm LP CMOS 70 - VDS = 1.2 V f = 8 GHz 60 - qO50 0 w 40 30 20 += 100 = 10 mA/mm mA/mm 200 mA/mm mA/mm 25 D +ID = 3 0 0 0 100 1000 10000 Device Width (gm) Fig. 3.6: Measuredpeak PAE, on 65 nm LP CMOS devices, versus device width at different DC bias currents. ..... . .......... .. ..... . also seen earlier in Fig. 3.3). The reasons for non-ideal scaling of output power and decrease in PAE with device width form the crux of this thesis and are presented in sections 3.3.2 and 3.4 and in chapters 4 and 5. 3.1.2 2.5 V 1/0 Device 60 VDD= 50 ID= 200 mA/mm f= 8 GHz Max. P/W = 267 mW/m 40 2.5 V Max. PAE =51.6 % 30 40 4 25 20 % w3 0 - Max. Pot" =9259 mW g W0=48 pm 20 15 ( 10 W =1536 p 10 5 0 0 0 5 10 15 20 25 30 Output Power (dBm) Fig. 3.7: Typical power sweep showing PAE and gain as a function of output power on several 0.25 pm long I/O devices in 65 nm LP CMOS technology. The maximum Pout, PAE, and Pout / W obtainablefrom this technology at VDS = 2.5 V and ID = 200 mA/mm are noted on the figure. RF power measurements were also made at 8 GHz on the thick oxide 1/0 devices at VDS = 2.5 V. Typical power sweeps at DC bias current of 200 mA/mm are shown for three different device widths in Fig. 3.7 and illustrate the maximum output power and PAE obtainable from this technology at the given DC bias and frequency. A maximum normalized output power of 267 mW/mm was observed on the 48 gm wide device, a maximum peak PAE of 51.6% was measured on the 192 gm wide device, and a maximum output power of 259 mW was measured 52 .. .... .... ....... ...... .. . - ...... . V-W... ... ...... .... ..... ....... . ... . .... ..................... ................ on the widest device (W=1536 gm). The maximum output power measured on the 1/0 device is approximately 3 times that measured on the standard device (Fig. 3.2) because of the higher drain voltage (2.5 V vs 1.2 V). However, the higher output power comes with a significantly lower PAE, because of the lower frequency response (fT and fmax) of the 1/0 devices compared to the standard devices. Hence, at relatively low RF frequencies (< 10 GHz), a significant improvement in RF power can be achieved by using the 2.5 V 1/0 devices instead of the standard devices. For frequencies greater than 10 GHz, the decrease in PAE might be too drastic to justify the use of 1/0 devices for increased output power. 30 65 nm LP CMOS VDS= 2.5 V 25 f=8GHz E o 20 0 15 0. 0.10 ID 50 0 mA/mm ID=100mAmm ID=200mA/mm * ID= 300 mA/mm + 5 0 10 100 1000 10000 Device Width (gm) Fig. 3.8: RF power measured at 8 GHz as a function of device width for I/0 devices designed in 65 nm LP CMOS technology. Fig. 3.8 and Fig. 3.9 show the output power and peak PAE as a function of device width for 1/0 devices in 65 nm LP CMOS technology. Measurements at bias currents of 50, 100, 200 and 300 mA/mm are included. The behavior of output power and PAE with device width is similar to that seen in the standard devices (Fig. 3.4 and Fig. 3.6). The output power scales non-ideally with device width for W > 400 gim, while the PAE shows a consistent decrease with increasing device width. 70 65nmLP CMOS VDS= 60 2 5 . V f=8 GHz 50 S40 30 -0- 20 10 ID= 50 mA/mm + tD =100 mA/mm +D = 200 mA/mm D= 1 300 mA/mm 0 1 10 100 1000 10000 Device Width (gm) Fig. 3.9: Peak PAE, measured at different bias currents, as a function of device width for I/O devices designed in 65 nm LP CMOS technology. 3.2 65 nm High Performance Technology In this section, DC, S-parameter and RF power measurements on standard devices, designed in 65 nm high-performance technology, will be presented. All devices were designed with a gate length of 50 nm and were measured at a drain voltage of 1 V, unless otherwise noted. Fig. 3.10 shows output characteristics and transfer characteristics on a standard 65 nm highperformance device with W = 96 pm. The normalized drain current of the device is around 771 mA/mm at VGs = VDS = 1 V. ............... ........... - .-....... ...... ..... .. 700 700 65 nm HP CMOS 600 600 W= 2x24x2 pm L= 50 nm __% E 500 E 500 E VDS=1 V E 400 S400 E E * 300 "300 200 200 100 100 50 mV 0 0 0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 VGS VDS(V) 0.8 1 ) Fig. 3.10. Output characteristics (VGS = 0.2 V to I V in steps of 0.2 V) and transfer characteristics(VDS = 0.05 V and VDs = 1 V) on the standard65 nm HP device (W = 96 pm). 300 N M 0 250 Is 250 N 200 X 0 x 200 150 a) E 1) 150 M 100 -D) E 100 rn .a E 50 C) 50 10 100 1000 10000 - 0 100000 Device Width (gm) Fig. 3.11: De-embedded fT and fax, measured on 65 nm HP CMOS devices, are shown as a function of device width. ........... S-parameter measurements were performed at various DC bias conditions in the 0.5 GHz to 40 GHz range. These S-parameters were then de-embedded to the intrinsic device using measurements on on-wafer de-embedding structures that were custom designed for each device. Fig. 3.11 plots the extracted de-embedded fT and fma as a function of device width for 65 nm HP CMOS devices. The figure shows that fT is fairly constant with device width at low bias current (ID = 25 mA/mm), but decreases with device width at the higher bias current (ID= 200 mAlmm). On the other hand, fmax shows a significant decrease with device width at both bias currents. The reasons for this degradation in the frequency response of wide devices will be explained in chapter 5. 80 40 70 35 60 30 50 25 0 Uj40 20 30 15 20 10 10 5 0 0 0 5 15 10 20 Output Power (dBm) Fig. 3.12: Typical power sweep on several 65 nm HP devices. The maximum Pout, PAE and Po, / Ware also shown in the figure. Fig. 3.12 shows RF power sweeps at 2 GHz and at a DC bias of VDD = 1 V and ID = 200 mA/mm. The maximum normalized output power at this frequency and bias was measured on the 48 gm wide device as 122.2 mW/mm. The maximum PAE was 72.4% as measured on the 56 .......... ........ 192 pm wide device, and the maximum output power was 77.2 mW on the 1536 gm wide device. RF power measured in the 2 GHz to 18 GHz range, and at a DC bias of VDD = 1 V and ID = 25 mA/mm, is shown as a function of device width in Fig. 3.13. The source and load impedances were optimized for peak PAE for each of these measurements. The figure shows that the output power scales ideally with device width, and is relatively independent of frequency for W < 400 jim. However, for W > 400 jim, the output power starts to saturate, and in some cases even peaks and then starts dropping. As the frequency of operation increases, the peak of the output power happens at lower device widths. Thus, as the frequency increases, the maximum output power, at any given DC bias, decreases. This is similar to the behavior observed in [33]. 25 E -- 2 GHz -- 6 GHz -- 10 GHz 20 -+14GHz +018GHz 15 0 lob 10 65nmHPCMOS o5 , VDD =1 V ID= 25 mA/mm 0 10 100 1000 10000 100000 Device Width (gm) Fig. 3.13: Output power, measured at different frequencies, as a function of device width, for 65 nm HP CMOS devices. VDD = IV, ID = 25 mA/mm. ......... . .... .... .. ........... ............ .............. ............. .. .... ... The measured peak PAE, corresponding to the output data of Fig. 3.13, is shown as a function of device width in Fig. 3.14. PAE decreases with increasing device width and with increasing frequency. This behavior is similar to that observed in previous sections and in [33], and can be attributed to a decrease in fmax in the wider devices. Further details can be found in section 3.3.2 and in chapters 4 and 5. Higher levels of output power can be obtained by biasing the device at larger drain currents. Fig. 3.15 and Fig. 3.16 show output power and peak PAE measured at VDD = 1 V and ID 200 mA/mm. The behavior is similar to that seen at lower drain currents (Fig. 3.13 and Fig. 3.14). However, there is no discernable peak seen in the output power versus width curves because the maximum device width that could be measured was limited by the 500 mA current rating of the bias tees in the power measurement system. 80 +*2G Hz +6 GHz +*10G Hz -o-1 4 GHz -w-1 8 GHz -O70 60 5 50 4) 40 030 65 nm HP CMOS L 020 0 VDD=1V ID= -10 0 10 25 mA/mm 0 100 0 1000 10000 1 100000 Device Width (gm) Fig. 3.14: PAE measured at different frequencies, as a function of device width for 65 nm HP CMOS devices. VDD = IV, ID = 25 mA/mm. .......... 20 18 - 16 - E I 14 -- 2 GHz 12 - -+6 GHz -- 10 GHz -- 14 GHz +w18GHz 0 10 - 8 6 65 nm HP CMOS 0 VDD=l V ID= 2 00 mA/mm .1 100 10000 1000 Device Width (gm) Fig. 3.15: Output power, measured at different frequencies, as a function of device width, for 65 nm HP CMOS devices. VDD = IV, ID= 200 mA/mm. 80 -- 2 GHz -6 070 060 0 -50 GHz +10 GHz -o14 GHz -w-18 GHz - W 40 - V -30 - C 65 nm HP CMOS M20 0 VDD = 1 V 0-1 0 ID= 200 mA/mm 100 1000 10000 Device Width (gm) Fig. 3.16: PAE measured at different frequencies, as a function of device width for 65 nm HP CMOS devices. VDD = IV, ID = 200 mA/mm. ............ 45 nm Low Power Technology 3.3 This section will present DC, S-parameter, and RF power data measured on devices designed in 45 nm LP CMOS technology. There were 2 sets of designs in this technology. In the first set, the device width was increased by either increasing the number of gate fingers or the unit finger width. We will refer to the devices in this set as "old" test structures [46]. In the second set, a unit cell was created with 20 fingers of 2 jim finger width, and the device width was increased by connecting multiple unit cells in parallel. The devices in this set will be called "new" test structures. The unit cell design in the new test structures was optimized for reduced parasitic resistances and capacitances as described in Chapter 2. All devices in both the "old" and "new" test sets were designed with a gate length of 40 nm and operate at VDS 1.1 V, unless indicated otherwise. 3.3.1 Old Test Structures 500 500 450 45 nm CMOS 400 W=60 x1.5sm L=0nm 350 O 350 E 300 E 300 VDS=l11V E E 250 E 250 . 45nm CMOS W60 x1.5 pm L40 nm 450 400 , 200 200 150 -9 150 100 100 50 50 0 0 0 0.2 0.4 0.6 VDS (V) 0.8 1 1.2 Vs= 50 mV 0 0.2 0.4 0.6 0.8 1 1.2 VGS ( Fig. 3.17: Output characteristics (VGS = 0.4 V to 1.1 V in steps of 0.1 V) and transfer characteristics(VDS = 0.05 V and VDS = 1.1 V) on the standard45 nm device (W = 90 m). ............................................. .......... ::. .. ...... . .. ............... A variety of test structures were designed with varying number of gate fingers and gate finger width to explore the optimum design that provides the best frequency and power performance. Fig. 3.17 shows the output characteristics and transfer characteristics measured on a test structure with 60 fingers of 1.5 gm finger width, resulting in a total device width of 90 pm. The normalized drain current density on this device was measured to be around 466 mA/mm at VGS = VDS= 1.1 V. 300 400 45 nm CMOS 250 W = 20x1.5 gm VD= 1.1 V doublecontact 350 300 200 single contac 250 150 double contact 100 200 N " 150 100 50 50 0 1 10 100 0 1000 D(mAmm) Fig. 3.18: fT and fax as a function of drain current density for single and double gate contact structuresin 45 nm CMOS technology. It is well known that contacting the gate on both sides of the device could result in lower gate resistance and hence improved fmax. To quantify the improvement in fmax and output power, we designed test structures in which the gate was contacted on only one side of the device (single contact), or on both sides of the device (double contact). Fig. 3.18 compares the de-embedded fT and fmax of a single gate contact structure with that of a double gate contact structure as a function of bias current. A double gate contact results in a significantly lower gate resistance, .- I . - : . ZZ Z Z .. I.. II - I IIIIIII .. I ... .. II I.IIII I.III . but a slightly higher gate capacitance. This leads to a slightly lower fT for the double gate contact structure compared to single gate contact at the same drain current density. However, fma, is higher in the double gate contact structure because the reduction in gate resistance more than compensates for the increase in gate capacitance. The improvement in fmax is more pronounced at higher current densities. 160 100 140 doublecontact 90 80 g~~~ single contact 70 out atu. 120 E E double contac100 60 50 80 -- single contact PAE < C 40 60 30 W = 20x1.5 gm 20 VD=l1-1V 2 00 ID= mA/mm 45 nm CMOS 40 C0 200 10 --- 0 0 '- 5 20 - 0 - ---- 10 15 20 Frequency (GHz) Fig. 3.19: PAE and normalized output power as a function of measurementfrequency for single gate and double gate contact structures in 45 nm CMOS technology. Fig. 3.19 shows the PAE and normalized output power (Pout/W), measured at ID = 200 mA/mm and VDD =1.1 V, as a function of measurement frequency for the single and double gate contact structures. The load and source impedances were tuned for optimum PAE and the output power and PAE were measured at peak PAE. The PAE for the double gate contact structure is similar to that of the single gate contact structure. The PAE is also remarkably frequency independent. The output power is slightly higher for the double gate contact structure. All the .............. ................................................................. ......... test structures discussed in the remainder of this section were designed with double sided gate contacts. Fig. 3.20 shows the impact of number of fingers (NE) and finger width (WF) on de-embedded fmax. In general, increasing the total device width (through increase in WF or NF) results in a significant decrease in fm.. For a given NF, increasing WF initially leads to an increase in fm, but eventually fm, decreases as the gate resistance increases. The optimum unit finger width for this technology is 1.5 pim. On the other hand, for a given WF, increasing NF leads to a decrease in fma. This is because of the increased parasitic resistance resulting from the additional wiring needed to connect all the fingers in parallel. 275 275 WFot =1.5 pm 250 250 N225 225 W 0 2200 04- E 200 45nm CMOS VDD=.1 V ID= 200 mA/mm W= 30 to 90 pm 175 150 0 , 1 a 2 45 nm CMOS VDD=1,1 V ID=200 mA/mm 175 WF= 1.5 pm 150 3 4 5 6 0 WF (pm) 20 40 60 80 100 120 140 NF Fig. 3.20: De-embeddedf,n, as a function of unit finger width and number offingersfor all the test structures of the "old test site". WF = 1.5 pm was found to be the optimum finger width for the 45 nm CMOS technology. Fig. 3.21 illustrates the maximum output power and efficiency that can be obtained from the structures in the old 45 nm test site at f = 2 GHz, VDD = 1.1 V, and ID = 200 mA/mm. The source and load impedances were optimized for maximum PAE when making these measurements and all other power measurements presented in the remainder of this section. The maximum normalized output power was measured on the 40 m wide device as 140.6 mW/mm, the maximum power-added efficiency was measured on the 240 Rm wide device as 72.2% and the maximum output power, measured on the widest device (W = 640 pm) was equal to 28.8 mW. The peak PAE, measured at 2, 6, 10, 14 and 18 GHz, is shown as a function of total device width for all the test structures of the old test site in Fig. 3.22. For the figure on the left, device width is increased by keeping NF constant at 60 and increasing WF from 0.5 Rm to 5 gm. For the figure on the right, device width is increased by keeping WF constant at 1.5 pm and increasing NF from 20 to 120. The dependence of PAE on device width is similar irrespective of whether the device width is increased by increasing WF or NF. PAE increases with width for small device widths (W < 100 pLm) because of an increase in power gain as a result of lower relative 80 Max. PAE = 72.2 % VDD = 1.1 V ID= 200mAmm f = 2 GHz !6 70 W =240 gm o60 C Max. P 0 = 28.8 mW .o 50 W 40 V W =40 gm Max. Pout/W =140.6 mW/mm W =640 pm 30 2 20 0 a. 10 0 0 2 4 6 8 10 12 14 16 18 Output Power (dBm) Fig. 3.21: Typical power sweep on several 45 nm CMOS devices, designed as part of the old test site. The maximum Pout, PAE and Pou, / W obtainedfrom this test site atf = 2 GHz, VDD = 1.1 V and ID = 200 mA/mm are also shown in thefigure. 64 ... .... ....... ......... 0 w 40 30 NF= 60 10 VD= 1.1V ID= 200 mAlmm Constant WF Ju Constant NIF 20 ............ . 42GHz 70 a6GHz 6 10GHz 14 GHz 5 00 ., L0j40 18 GHz I 2GHz 6GHz 10 GHz 18GHz 14GHz WF=1.5 sm WFI VD= 1.1V 200 mA/mm 'D= NF t . 0 50 100 150 200 250 300 350 50 Total Device Width (pm) 100 150 200 Total Device Width (pm) Fig. 3.22: Peak PAE, measured at differentfrequencies, versus total device width. For the figure on the left, device width is increased by keeping NF constant at 60 and increasing WF from 0.5 pn to 5 pun. For the figure on the right, device width is increasedby keeping WF constant at 1.5 pum and increasingNFfrom 20 to 120. Constant WF Constant NF 160 140 140 10GHz ,120 E E GHz 100 100 6 GHz E80' 80 %E 14GHz s 60 60' NF= 60 0 M. AA VD= 1.1V 20 00 120 ID= 200 50 WFt mA/mm 100 150 200 Total Device Width (jim) 250 300 50 100 150 Total Device Width (pm) Fig. 3.23: Normalized output power, measured at different frequencies, versus total device width. For the figure on the left, device width is increasedby keeping NF constant at 60 and increasing WF from 0.5 pm to 5 pn. For the figure on the right, device width is increased by keeping WF constant at 1.5 pn and increasingNFfrom 20 to 120. parasitics. For W > 100 tm, PAE is constant with width at lower frequencies but tends to decrease with width at higher frequencies because of the reduction in gain with frequency. The corresponding normalized output power is shown as a function of total device width in Fig. 3.23. The output power shows no clear dependence on frequency, as expected, but shows a gentle decrease with increasing width. This decrease in output power at large widths is because of the non-ideal scaling of on-resistance in the wide devices, as described in section 3.4. 3.3.2 New Test Structures A new set of test structures were designed in 45 nm low-power CMOS technology to explore a wider range of device widths. In this set, the device width was increased by connecting multiple unit cells in parallel. The unit cell was designed with 20 fingers of 2 gm finger width (which is close to the optimum finger width found in the previous section). Further improvements to the unit cell design were made to optimize the parasitic resistances and capacitances, thus resulting in improved frequency and power performance (please see Ch. 2 for layout details). Typical output and transfer characteristics on a 2 cell device (total width of 80 mm) can be seen in Fig. 3.24. The device delivers a normalized drain current density of 513 mA/mm at VDS = 1.1 V. VGS ................................... 600 45nm CMOS 500 W= 2x20x2 pm 500 VDs=1-1 V L=40nm 400 E 400 E E E E 300 300 200 200 100 100 ~----- ------- 0 0.4 0.2 0 0.6 1 0.8 0 1.2 0 0.2 0.4 VDs (V) 0.6 VGS(VM Output characteristics (VGS = 0.4 V to 1.1 V in steps of 0.1 V) and transfer characteristics(VDs = 0.05 V and VDS = 1.1 V) on 45nm devices (W = 80 um). Fig. 3.24: 275 - 250 - 225 N 0 200x . 175 - 150 - 45 nm CMOS VDD = 1.1 V 125 100 ID= 200 mA/mm F -I-----------------u- 0 100 200 300 400 500 600 700 Device Width (pm) Fig. 3.25: De-embedded fT and f.a as a function of device width for the new 45 nm test structures,measured at VDD = 1.1 V and ID = 200 mA/mm. S-parameter measurements were made in the 0.5 GHz to 40 GHz range and were deembedded using data on a common set of on-wafer open, short, load and thru structures. The extracted de-embedded fT and fmax are shown as a function of device width in Fig. 3.25 for a DC bias of VDD = 1.1 V and ID= 200 mA/mm. fT is relatively constant with device width, but fmax decreases with increasing device width. This behavior is similar to what was presented on other technologies in the previous sections. RF power measurements were performed in the 2 GHz to 18 GHz range at VDD ID= = 1.1 V, and 200 mA/mm. Two sets of measurements were made on each device at every measurement frequency. For the first set of measurements, the source and load impedances were tuned for maximum PAE (as was the case for all the power measurements presented previously in this chapter). For the second set, the source and load impedances were tuned to obtain the maximum output power, while keeping the PAE still relatively high. In other words, it would be possible to obtain more power from these devices at the expense of PAE. Fig. 3.26 shows the PAE versus Pou, for three different 45 nm CMOS devices and illustrates the maximum Pou1 t, PAE, and Pout / W obtainable from this technology at the given frequency and DC bias. Tuning the load impedance for maximum power results in a maximum normalized output power of 210 mW/mm at a PAE of 61.6% on the narrowest device, and a maximum total power of 68.6 mW at a PAE of 48.6% on the widest device. This normalized power density is the highest reported on any single 65 nm or 45 nm CMOS device. When the load impedance is tuned for maximum PAE, a PAE in excess of 75% is obtained at 2 GHz, but Pout/W drops to 93 mW/mm. .......... ...... ....... ... ................... . ........ .... ................ ............ .......... .. ............................................................. 80 - W = 240 m VDD = 1.1 V eID = 200 mA/mm f =2GHz S70 -- Max. PAE =76 % opt. PAE A o60 ~W C 4) = 40 gm opt. Pout 50 Max. Pout/W W= = 211.5 mW/mm W 40 V) ( 30 4 a) 640 m opt. P0 ut Max. Pout =68.6 mW 20 0 o. 10 0 0 2 4 6 8 10 12 14 16 18 20 Output Power (dBm) Fig. 3.26: Power sweep showing PAE and Pourfor devices with different widths. The figure also notes the maximum PAE, Po, and Pou, 1W obtained the new set of structures in 45 nm CMOS technology. Fig. 3.27 and Fig. 3.28 show Pout/W and peak PAE measured at 6 GHz as a function of device width for the two impedance matching conditions. Both Pout/W and PAE decrease with increasing device width. The width dependence is similar across all frequencies. This dependence is similar to what was observed on other technologies in previous sections of this chapter, and will be investigated further in section 3.4 and in chapters 4 and 5. "I.-....... ... .. .... ...... . .. ...... .. .. 250 - 45 nm CMOS VDD = 1.1 V 200 ID=200mAmm f = 6 GHz E 150 Optimized for Max. Power E 100 0 Optimized for Max. PAE 50 0 0 100 200 300 400 500 600 700 Device Width (gm) Fig. 3.27: Normalized output power as a function of device width for the two impedance matching conditions. Optimizingfor maximum power gives a higher Pou,/W at a lower PAE. 80 - F 70 - Optimized for Max. PAE 60 Optimized for Max. Powe? 0 .- .50- W 40 V 30 45nmCMOS 20 0 0. 10 VDD= 1.1 V ID= 200 mAmm f=6GHz 0 ) 100 200 300 . 400 .0 500 600 700 Device Width (gm) Fig. 3.28: Peak PAE as a function of device width for the two impedance matching conditions. Optimizingfor maximum PAE gives a higherPAE at a lower PouIWW. ......... .. . .. .. .. .. ... ............ 250 Optimized for Max. Power 200 E E 150 Optimized for Max. PAE E 100 s 0 45 nm CMOS 0. 50 - VDD = 1.1 V 1D= 2 00 mA/mm W =40gm 0 0 2 4 6 12 8 Frequency (G Hz) 14 16 18 20 Fig. 3.29: Normalized output power as a function of measurementfrequency. Pou, /W is relatively independent offrequency for both impedance matching conditions. 70 0 45 nm CMOS VDD= 1.1 V lD= 200 mA/mm 65 W = 40 gm 60 p4 a) '- 0 55 V cc 50 4) 0oM 45 .3 8 10 12 Frequency (GHz) 20 Fig. 3.30: Peak PAE as a function of measurementfrequency. PAE decreases significantly with increasingfrequency for both impedance matching conditions. ............ . . . . , .............. .................. Fig. 3.29 and Fig. 3.30 illustrate the frequency dependence of the normalized output power and PAE. The normalized output power is independent of frequency, as expected. On the other hand, PAE decreases significantly with increasing frequency. The frequency dependence of Pom / W and PAE is similar across all device widths. 80 00ft%70 o60 W 50 - 609 ."Optimized for Max. PAE 50 e Wi 40 "o -30 Optimized for Max. Power - 4 20 - 45 nm CMOS o VDD OD = 1.1 V = 200 mA/mm 10 0 0 0.05 0.1 0.15 0.2 0.25 f/fmax Fig. 3.31: PAE versus f/fma for 45 nm CMOS devices of different widths and measured at variousfrequenciesfrom 2 GHz to 18 GHz. The reason for the degradation in PAE with increasing frequency can be attributed to the decrease in gain with frequency. The same argument can be used to explain the decrease in PAE as the device width increases (Fig. 3.28). As was shown in Fig. 3.25, fmax decreases with increasing device width. Hence, at the same frequency of operation, a wider device would be operating closer to its maximum frequency limit compared to a narrower device, resulting in lower gain in the wider device and thus degraded PAE. ..................................................... .. ......... ............ PAE has often be shown as a function of the ratio of the frequency of operation to the maximum oscillation frequency of the device (f/fma) [33]. As f/fm increases, gain of the device decreases and hence PAE decreases. Plotted in this way, all the data on devices with different widths, measured at various frequencies, falls on one universal curve for each of the impedance matching conditions (Fig. 3.31). 3.4 Comparison among different technologies So far we have presented a vast amount of frequency and power data on devices designed in the 65 nm low-power, 65 nm high-performance and 45 nm low-power technologies. Now, we will compare the power and frequency response of these technologies. 120 50 mA/mm f =8 GHz ID= E 100%E E 80 65 nm LP ~(1-2)2 VDD=1.2V a)60 0 40 ,& 65 nm HP VDD= 1 V o 20 10 100 1000 10000 Device Width (gm) Fig. 3.32: Output power versus device width for 65 nm low-power and 65 nm high-performance devices, measured at VDs = VDD, ID = 50 mA/mm, andf = 8 GHz. ....... ................. We will start with a comparison between the 65 nm low-power and high-performance technologies. Fig. 3.32 and Fig. 3.33 show the output power and peak PAE as a function of device width for devices designed in both 65 nm low-power and 65 nm high-performance technologies, and measured at VDS identical in both technologies. = VDD, ID = 50 mA/mm, and f = 8 GHz.. The device layout is It is clear from the figures that, while the measured PAE is similar between the two technologies (Fig. 3.33), the measured output power is higher in the 65 nm LP technology (Fig. 3.32). The higher output power is expected because the supply voltage is 1.2X higher in the 65 nm LP technology. 70 60 65 nm HP 50 65 nm LP VDD VDD = 40 0- 40 1.2 V 30 20 ID = 50 mA/mm f = 8 GHz 10 0 10 100 1000 10000 Device Width (gm) Fig. 3.33: PAE versus device width for 65 nm low-power and 65 nm high-performance devices, measured at VDS = VDD, ID = 50 mA/mm, andf = 8 GHz. Fig. 3.34 compares the de-embedded fma versus device width for 65 nm devices and 45 nm devices (both old and new structures). fma is extracted from S-parameters measured at mA/mm and VDS = 1 V on the 65 nm devices, and VDs = ID= 200 1.1 V on the 45 nm devices. The figure .............................. ... .. shows that the 45 nm devices have a higher fmax than 65 nm devices, as expected. For the 45 nm devices in the old test site, fmax is higher than the corresponding 65 nm devices for small device widths, but drops sharply in the wider devices due to the presence of large interconnect parasitics. A significant improvement in fax can be observed in the new 45 nm test structures, where the device layout was optimized to reduce interconnect parasitic. The layout differences between the 45 nm old and new test structures were described in Chapter 2. 300 VDS = VDD N 250 N 250= 0 20 0 mA/mm E 200 6 150 .0 E -+-65nm HP -*-45 nm old structures - -45 nm new structures 100 S100 50 10 100 1000 10000 Device Width (gm) Fig. 3.34:De-embeddedf,,a versus device width for 65 nm devices and old and new 45 nm device structures. Fig. 3.35 compares the output power versus device width for the 65 nm and 45 nm (old and new) devices. The data was measured at 6 GHz with ID = 200 mA/mm and VDS nm devices and VDS = = 1 V for the 65 1.1 V for the 45 nm devices. In general, the output power scales ideally with device width for small device widths, but starts to saturate for the larger device widths. We also observe that the measured output power on the old 45 nm devices is similar to that on the 65 nm devices, despite the 1.1X increase in VDD. The new 45 nm test structures perform much 75 I ' :::- - - -Is - ................. better, delivering a higher output power than the corresponding 65 nm devices. Even higher output power can be obtained by optimizing the source and load impedances for maximum power, instead of for maximum PAE. The reasons for this behavior can be understood by examining the on-resistance of all these devices as a function of device width (Fig. 3.36). 60 VDS = VDD 200 mA/mm f=6 GHz 50 'D= E 40 o 30 CL 45 nmnew opt. Power e 45 nm new opt. PAE 20 CL 20 3 i A45 nm old opt. PAE * 65 nm opt. PAE 10 0 0 200 400 600 800 Device Width (pm) Fig. 3.35: Output power versus device width for 65 nm devices and old and new 45 nm device structures. Fig.3.36 shows that the normalized on-resistance, for all devices in both the technologies, increases with increasing device width instead of being constant with device width as expected from ideal scaling. This non-ideal scaling of the on-resistance with device width is the main contributing factor to the non-ideal scaling of output power at large device widths. Fig. 3.36 also shows that Ron for the 45 nm devices is much higher than that for the 65 nm devices across all device widths. This increase in Ro1 for 45 nm devices can be attributed to the larger interconnect resistance associated with the thinner and narrower metal lines and smaller contacts in 45 nm technology. The higher Ron of the old 45 nm devices would explain the lower output power on 76 those devices, especially on the wider devices, where Ro. plays a more dominant role. Optimizing the device layout of the new 45 nm structures helped lower their Ro. thus improving their output power to around (1.1)2 times that on the 65 nm devices. 1 0.9 VDS 0.05 V 0.8 E0.7 E 6 0.6* 0.5 +45nm 0.4 0.3 0.2 1 0 45nm new ideal scaling ""65nm o 200 old HP 400 0 600 800 Device Width (gm) Fig. 3.36: Extracted normalized on-resistance as a function of device width for 65 nm devices and old and new 45 nm device structures. A better way of comparing the output power delivered by 65 nm and 45 nm devices would be to plot the measured output power as a function of the extracted on-resistance, as shown in Fig. 3.37. When plotted in this way, all the 45 nm device structures show the (1.1)2 times improvement compared to the 65 nm devices, when compared at the same Ro,. A further improvement in output power can be obtained by optimizing the source and load impedances for maximum power instead of for maximum PAE, as seen in Fig. 3.37. . .. I, . .. .................... .... 80 VDS -= VDD 200 mA/mm f =6 GHz 70 ID= 60- E 50- 1 45 nm new opt. Power @45 nm new opt. PAE A45 nm old opt. PAE 0 4 30 * 65 nm opt. PAE O 2010 0 0 2 4 6 8 10 12 1. Ron (Ohm) Fig. 3.37. Output p ower versus on-resistancefor 65 nm devices and 45 nm old and new test structures. 80 145 nm new opt. Power 70 P45 nm new opt. PAE 60 L45 nm old opt. PAE 0 65 nm opt. PAE -- 50 E 40 4 30 Z VDS = VDD 20 10 00.00 eA , ID ,* * = 200 mA/mm f=6 GHz + 0.05 0.10 0.15 . 0.20 1/(RL+Ron) (Ohm-1 ) Fig. 3.38: Output power versus 1/ (RL+Ron) illustratingthe universal behavior of measured data for 65 nm and 45 nm (old and new) devices. ........................... ............ . ........... .... . ...................................................................................................... ................. Output power also depends on the fundamental load resistance that the transistor drives. To first order, the output power is approximately proportional to the reciprocal of (RL + Ron). In fact, when we plot the measured output power as a function of 1/(RL + Ron), we observe that all the data fall along a universal line, irrespective of the technology, layout or measurement condition (Fig. 3.38). 80 ~70 60 2 A IA 40* * * 40 5 nm*nw 300 A 20 2 45 nm new opt. PAE A 45 nm old opt. PAE ~10 -*465 nmopt. PAE VDS =VDD ID =200mrA/rmm 0 0 0.05 0.1 0.15 0.2 f/ max Fig. 3.39: PAE versus f/fmax illustrating universal behavior of measured data for 65 nm and 45 nm (old and new) devices. Another universal relationship that was mentioned in the previous section is the dependence of PAE on f/fmax. We had shown earlier that when PAE is plotted against f/fm for 45 nm devices of different widths and measured at various operating frequencies, all the data falls along a universal curve, with the PAE decreasing with increasing f/fma, because of the decrease in gain as the operating frequency approaches the maximum oscillation frequency of the device. Now we extend this comparison to include other technologies. Fig. 3.39 shows the measured PAE at ID = 200 mA/mm and VDS = VDD for the 65 nm devices and the old and new 45 nm test 79 I I,- - I- - - - ........... :: :......... structures. All these measurements were performed with the source and load impedances optimized for maximum PAE. It is clear that all the measured data fall along the same straight line, further confirming the universal nature of this relationship. 3.5 Comparison with literature In this section, we will compare the output power data presented in this thesis with the RF power data published in literature [33-37]. Fig. 3.40 compares the normalized output power as a function of device width for 45 nm, 65 nm and 90 nm devices. The output power is normalized to both device width and supply voltage. The frequency of operation was not identical for all the cases. However, we have shown earlier that output power is independent of frequency and hence this comparison is still valid. 180 160 45 nm (this work) [37] nm 1.5 90 VDD= V 140 /'VDD =1.1 V 120 %-100 65 nm (this work) 80 0 VDD = 1 V 60'O 60 40654045 n nm [36] 20 [33] +/VDD= 90 nm [34] VDD 0 10 100 = 1 V 1000 10000 Device Width (gm) Fig. 3.40: Normalized output power versus device width. The figure compares power data presented in this thesis to the values reportedin literature. It is clear from Fig. 3.40 that the normalized output power measured on the 45 nm devices in this thesis is the highest ever reported so far. This high value for output power was achieved through a combination of high DC drain current (ID = 200 mA/mm), optimized device structures that reduce interconnect resistance and thus Ron of the transistor, and by optimizing the source and load impedance for maximum power when making the RF power measurements. The output power measured on the 65 nm devices is also higher than that reported on similar devices in [33]. The main reason for this improvement is the increased DC drain current used when making the power measurements (ID = 200 mA/mm versus ID = 25 mA/mm in [33]). 90 80 70 ,6 45 nm (this work) E060 -VDD (50 0 CL. 40 15 D- 30 =11V /m[3 6 VDD = 1 V o 20 10 65 nm (this work) VDD = 1 V 45 nm [36] VDD = 1.2 V 90 nm [37] 5 VDD = 1. V 0 0 10 20 30 40 Frequency (GHz) Fig. 3.41: Output power versus operatingfrequency. The figure compares power data presented in this thesis to the values reported in literature. Fig. 3.41 shows the dependence of maximum output power on the frequency of operation. The device width and operating voltage was different for each of the cases in the figure. The 90 nm devices in [37] had a device width of 40 jim, while the 45 nm devices in [36] had a device width of 192 ptm. The 65 nm devices in [33] had different device widths at each operating frequency: W = 12.288 mm at 2 GHz, W = 6.144 mm at 4 GHz, W = 3.072 mm at 6 and 8 GHz, W = 768 jim at 14 GHz and W = 384 gm at 18 GHz. In our work, the maximum device width that could be measured was limited by the 500 mA current rating on the bias tees in the power measurement system. Hence, W = 1536 gm for the 65 nm devices and W = 640 gm for the 45 nm devices from this work. The most extensive data among those published before is on the 65 nm devices in [33], and showed that the output power drops below 10 mW around 20 GHz. Later work on 45 nm devices showed that it is possible to obtain 10 mW of power even at frequencies as high as 35 GHz [36]. In this thesis, we presented data on 65 nm and 45 nm devices, measured at VDS = VDD and ID= 200 mA/mm, as a function of frequency. Fig. 3.41 shows that the measured output power from our work is lower than that presented in [33] at frequencies below 8 GHz because of the much higher device widths in [33]. However, at higher frequencies, our data is much higher than previous reports. The main reason for improvement in 65 nm data is measurement at higher current levels and that for improvement in 45 nm data is due to higher current levels as well as optimization of device layout. 3.6 Summary This chapter has presented extensive RF power data as a function of device width, frequency of operation, and DC bias conditions, for devices designed in 65 nm low-power, 65 nm highperformance, and 45 nm low-power technologies. In general, the normalized output power is independent of measurement frequency, but decreases with increasing device width. The main reason for the non-ideal scaling of output power in wide devices is the increase in on-resistance in wide devices because of their higher interconnect resistance. On the other hand, the PAE 82 decreases with increasing frequency, as well as with increasing device width. The reason for the decrease in PAE in wide devices can be attributed to the degradation in fma with increasing device width. Comparisons of frequency and power performance among the three technologies, as well as comparison of measured data from this thesis to published literature have been presented. Normalized output power of 210 mW/mm, peak PAE in excess of 75% and total output power as high as 68.6 mW were measured on the 45 nm devices at a DC bias of VDS = 1.1 V and ID= 200 mA/mm. The output power values are the highest reported for any 45 nm or 65 nm CMOS technologies. 84 Chapter 4 Model for RF Power In the previous chapter we showed that output power decreases with increasing device width and postulated that the main reason for this decrease was because the on-resistance (R,,) of the devices does not scale ideally with device width. In this chapter, we will attempt to prove this theory through a model that predicts the maximum output power delivered by a device for any given DC operating point and the Ron of the device. Accurate estimation of the maximum output power for a device at any given bias point is an onerous task. Power amplifier designers often rely on RF power simulations using compact device models. These simulations are very time consuming and the need for accurate compact models means the power estimations are only accurate when designing in mature device technologies. Moreover, the compact models for CMOS devices are usually tailored for digital or low power RF applications and hence do not do a good job in modeling the device for large signal swings beyond the nominal voltage of the technology. Also, power simulations at the peak PAE point often lead to convergence issues and hence do not predict the peak power very well. In view of all the challenges described above, it would be ideal to have a simple analytical model that can accurately predict the maximum output power and the trade-off between output power and PAE in a device. The analytical models presented in textbooks [45, 47-52] usually 85 ignore Ron and also assume zero VDSsat. However, to obtain high power, the device is usually biased at a high drain current, which means the output power is strongly affected by R,. In this chapter, we present an analytical model that correctly accounts for the on-resistance of the device, thus allowing us to make accurate predictions of the output power and PAE. We will also compare the power and efficiency predictions from this model to the measured loadpull data from the previous chapter to validate the accuracy of the model. 4.1 Model Description VDQ RF Choke |DC High Q "tank" (tuned@fo) -,+ VDS + VgS VG VG S (N - RLV +i '2v ~ i37 etcZLz RL+JO ZL=O 1K to 2fo, 3fo, .... etc Fig. 4.1: Circuitdiagram of a reduced conduction angle power amplifier. The model described in this section is based on the overdriven class AB amplifier discussed in [45]. Consider the power amplifier circuit shown in Fig. 4.1. The transistor is biased at the DC operating point, (VGQ, VDQ). The current through the drain in the absence of RF input signal is IDQ. The DC bias to the drain is fed through an RF choke, which is assumed to have a very high reactance at RF frequencies. The DC component of the gate-to-source voltage (VGQ) is higher than the transistor threshold voltage (Vt). The AC component of the gate-to-source voltage (vgs) is assumed to be a cosine wave with peak amplitude Vgs. The drain current (iD) would then be a clipped cosine wave containing a DC component which flows through the RF choke (IDC), a fundamental component (-ii) and harmonics (-i2 , -i3, -i4 , etc). The output (drain) of the transistor is AC coupled to an RF load through the capacitor Cc. Cc is assumed to be large, such that the voltage across the capacitor is constant and is equal to the average drain voltage, VDC. The RF load consists of a fundamental load resistor RL, and a shunt connected parallel LC "tank" circuit with its resonant frequency at the fundamental frequency of operation (fe). All harmonics in the load current are assumed to be shorted by the tank circuit and hence generate no voltage. Hence, the current flowing through RL is just the first harmonic, ii, and the voltage across RL, is thus assumed to be sinusoidal, with a magnitude set by the load resistor. Fig. 4.2 shows the waveforms for key voltages and currents in the power amplifier when operated in the maximum power condition for a certain bias point and load resistance. The gateto-source voltage waveform can be expressed as VGS = (4.1) VGQ ± gs CoS &)t When vGS < Vt, the device enters the cut-off regime and the drain current is zero. When vGS > VG~sat, the device enters the linear regime and the drain current is clipped at Iknee. For all other values of vGS, the drain current is assumed to be cosinusoidal. The shape of the drain voltage waveform is 180 degrees out of phase with the drain current waveform. When iD = lknee, pinned at the minimum value Vmin, and when iD = 0, the drain voltage is clipped at Vmax. vDS is Vssat - VGQ 0 tnee D 0 s/2 a/2 T 2Tr 3 T 2w 3w Wt Vmax YDS 0 P/2 u/2 -- Wt Fig. 4.2: Waveforms for input voltage and output current and voltage in the reduced conduction angle amplifier shown in Fig. 4.1. a denotes the conduction angle and represents the proportion of the RF cycle for which VGS > Vt and iD> 0. Since the waveform is symmetric about ot = 0, the current cut-off points are at ot = --a/2. x = 2xE defines class A operation (the device is always ON), a = a defines class B operation (the device is cut-off half of the time), a < a < 2a defines class AB operation, and 0 < a < a for class C operation. 0 denotes the proportion of the RF cycle for which VGS > VGSsat and ID = lknee. From the shape of the drain current and voltage waveforms, it is clear that the average value of iD (denoted by IDc) and the average value of vDS (denoted by VDC) depend on (x and P, along 88 with the minimum and maximum values of the waveforms. In general, IDC t IDQ and VDC # VDQ, as illustrated in Fig. 4.2. Assuming that the drain current follows a pure cosine between G/2 and P/2, iD can be expressed as: Iknee iD (Ot) IDQ + Id # 2 (4.2) COS Ot a -2 < |wt| 7 0 The values of LI and I"e can be obtained by recognizing that the waveform needs to be continuous at a/2 and P/2 and by applying the appropriate boundary conditions. iD (kt iD = fl) (Wt (4-3) IDQ + Id COS fl nee= (4.4) = 0 = IDQ + Id CoS a/2 = -) Solving equations (4.3) and (4.4) results in the following expression for the drain current: Iknee 2 COSWt - iD (WO =DQ a COS~ /2 0 where Iknee - IDQ (1 2 # l -:5 2 lWt|I a 2 (4.5) 7 cos(-/2) cos(a /2)) (4.6) The drain current can be written in terms of its Fourier series components as: iD OJ = IDC - ('1 + i2 + i3 +....) (4.7) = IDC - Yl COS t + I2 cos 2t + I3 cos 3t + ---) The DC component of the drain current, IDC, is all the current that flows through the RF choke and can be computed as a 1 fDC D d(wt) Imaxd(wt)+2f IDQ( -2 __ p_ 2 2 os a .(sin a a - Cos fl)+(a-fl) IDQ[/C co 2 n = os Cos sm cs2 2 cos) - sin . 2s The fundamental component of the drain current is also the current that flows through the load resistor, because we assumed that the DC component flows through the RF choke and that the second and higher order harmonics are shorted out at the fundamental frequency of operation. The magnitude of the fundamental component of the drain current is given by I1 =f _Ir iD (Wt)cos wt d(wt) - #/2 =- a/2 Imax cos wt d(wt) + 2 2 IDQ 0 1- #/2 = a s (os a +Csa /si. a si. + co sm -s Substituting a = 271 and cos td(ot) cos9) Cos 2 7 cos7- a ) 2) (a 4 P = 0 in equations p) - (sin a - sin) 4 (4.8) and (4.9) gives IDC IDQ and 11 IDQ, which are the correct values for class A operation. Also, when fl = 2w - a, the waveform is symmetric about IDQ and hence IDC = IDQ. For any other values of a or P, the average drain current that flows through the RF choke (IDC), is not equal to the drain current at low power condition (IDQ) because the drain current waveform is no longer symmetric about IDQ- For calculating output power delivered to the load, we also need expressions for the drain-tosource voltage, VDS. Our model describes the maximum power situation for any given value of a( and P. Hence, we assume a maximum voltage swing for vDS, that is symmetric about VDQ. In other words, Vmax - Vmin = 2(VDQ (4.10) - Vmin) This assumption is valid for Class A, Class B and Class AB amplifiers and is consistent with the assumptions in most power amplifier textbooks [45, 48, 50]. Thus Vmax can be written as (4.11) Vmax = 2VDQ - Vmin where Vmin can be expressed in terms of Iknee and the on-resistance of the device (Ron) as: (4.12) Vmin = Iknee Ron = IDQRon 1 - cos(fl/2) cos(a/2)) In general, the drain voltage waveform can be expressed as: V< VDS(Jt) = Iwt| 0 Vmin cs <- 2 a (max -<IJWtI: 2 a 2 (4.13) 7 The value of Vo and Vd can be determined by applying the following boundary conditions: vDS VDs (aft = Vmin Vo +VdCOS -Vmax= Vo + V CoS Solving for Vo and Vdgives the following expression for vDS: (4.15) Vmin COS VDS - 0 < |at Vmin a 2 / Vmaxcos + 0 (ymax Vmin)cosWt /a a cos -cscsos CO2 -CS2 CO (4.16) 2-2 COS 2 a- Iw !IT Vmax The magnitude of DC component of the drain voltage can be obtained by Fourier analysis of equation (4.16) and is given by: IT f VDS(wt)d(Wt) VDC a 22 f 0 E - 2 2 _ 1 - COS __Ii + )min) - d(w) + CO ft COS -COS + X [m (Vmax - Vmin) (sin C - f OS ) d(wt) Vmax d(wt) (4.17) a (- Vmi - + Vmax (7 (COs COS a COS- amin Vmind(t) + +(Vmin cos 2 + - Vmax cos) 2(cos - (a- p) cos -sin2) -COS The fundamental component of the drain voltage is the same as the voltage across the load resistor, because we assumed that the average value of VDS drops across the capacitor Cc, and that the second and higher order harmonics are shorted out at the fundamental frequency of operation. The magnitude of the fundamental component of the drain voltage is given by: 7r ffvDS(t)cos wt Vi = d(wt) - T 2 =- 2Vmin cos wt d(wt) + cos wt d(wt) 0 + (Vmax - Cos Vmin) cos t d(wt) (4.18) + Vmax cos wt d(wt) - cds 2 a a sin-+ Vminsin 2l- VmaxVmx2 (Vmax - Vmin)(a 4 (cos - (Vmin Cos 7 Vmax COS 2 a (COS 2 -COS sin Z - sin) /) 4 (cos - COS - COS The output power and efficiency can be directly computed using the expressions for the DC and fundamental components of voltage and current. The DC power is given by PDC = VDQ ' IC where IDC (4.19) is given by equation (4.8). The RF output power is given by I1 out -- - V-1 (4.20) where 11 and V1 are shown in equations (4.9) and (4.18) respectively. Finally, the drain efficiency can be computed as 7D (%) = Pot * 100 (4.21) Pnc The fundamental load resistance value that is associated with a particular choice of alpha and beta can also be readily computed as ........ ................ 2 VD -IDRon 1 Cos E (4.22) cos() RL = Vmax - Vmin 'knee cos(#/2)) 600 Ron 45 nm CMOS W=40gm VGS=1.1V 500 imax Ron = 9.59 Q VGS = 1.0 V 400 E E VGS = 0.9 V 300 E (VDQI DQ) 200 (VDC, 'DC) VGS =0.7 V 4O%% 100 % % VGS = 0.5 V --- -- 0 0 ~I. -min 0.5 1 -1/R 1.5 L 7/% %% 'I 2 max VDs (V) Fig. 4.3: Illustration of difference between DC and AC load line for RL = 100 2. The DC operatingpoint is at VDQ = 1.1 V, IDQ = 200 mA/mm. The AC operatingpoint is at VDC = 1.1 V, IDC = 212 mA/mm. An examination of the drain current and drain voltage waveforms in Fig. 4.2 and the expressions for IDC and VDC in equations (4.8) and (4.17) reveals that under large-signal AC conditions, the average value of voltage and current through the drain of the device shifts from the DC bias point under low input power conditions. This effect is called "self biasing" and arises due to the asymmetry of the voltage and current waveforms. Hence, for any given RL, the 94 AC load line could be different from the DC load line. This difference is illustrated in Fig. 4.3, drawn on the output characteristics of a typical NMOS studied in this thesis for the case when VDC > VDQ and IDC > IDQ. It should be noted that it is perfectly possible for IDC to be lower than and similarly for VDC to be lower than VDQ. IDQ 4.2 Comparison of Model with Measurements The model equations presented above can be used to generate the drain efficiency - output power locus for any given device. The inputs to the model are the DC operating point (VDQ, IDQ) and Ron of the device. Ron was calculated from measured DC data as VDS/ID with VDS = 50 mV and the device biased in the linear regime. For any given value of ax and P, describing a random maximum power condition, the value of output power, drain efficiency, average current and load resistance can be computed using the expressions in section 4.1. This exercise was repeated for multiple pairs of alpha and beta to construct the drain efficiency - output power locus. Fig. 4.4 shows the modeled 11D - Pout locus for a wide range of ai and 0, along with measured data on a 45 nm device with W = 40 Jtm. To obtain the experimental power data, an initial power sweep was performed, with the source and load conjugate matched, to determine the input power level at which peak PAE is obtained. Next, the input power was fixed at this level and the load impedance was varied across the Smith chart while keeping the source conjugate matched. From this load pull data, the load impedance that delivers the maximum output power was determined. A final power sweep was then performed to note the peak PAE and corresponding Pt. This procedure was repeated for multiple values of RL. The measured power data shows reasonable agreement with the modeled locus as seen in Fig. 4.4. The shape of the 11D - Pout locus makes intuitive sense. For any given X,as P is increased, more of the current waveform gets clipped at I"e, and more of the voltage waveform gets 95 clipped at Vmin. This leads to a decrease in the peak of the current waveform and a corresponding increase in peak of the voltage waveform, which in turn leads to a decrease in IDC and 11 and a corresponding increase in VDC and V 1. Since Pont = 2 , as P increases, P0 ut initially increases because of the increase in V1, but then starts decreasing as the decrease in I1 starts to dominate. On the other hand, since IDc decreases with increasing beta, the drain efficiency keeps increasing. As a decreases towards class B operation (a = a), the peak of the current waveform increases, thus leading to higher power. 8070a= 240 60 ~ a= 220 a= 190 =20 a= 260 a= 290 C 50QV a=320 0 a=360 4030- 45 nm CMOS W = 40 m -1V VDQ 20. IDQ = 200 mAlmm Ron = 9.59 Q 2 3 4 5 6 7 9 8 10 11 12 Output Power (mW) Fig. 4.4: Modeled locus of drain efficiency versus output power. Pow, and riD measurements at different RL are shown as symbols in the figure and show good agreement with the model. Fig. 4.5 and Fig. 4.6 show the Po 0 t - RL locus and the lD - RL locus respectively, along with the corresponding measured data. A reasonable agreement between the measured data and the model is seen. For any given x, as P is increased, more of the current waveform gets clipped at ................................................. .... ... .... ....... ................ ..... .. ... Iinee, and more of the voltage waveform gets clipped at Vmin. This leads to a decrease in the peak of the current waveform (Iee) and a corresponding increase in peak of the voltage waveform (Vmax). Since RL , RL increases for increasing values of = Vmax-Vmin 'knee 0. As a decreases, Iknee increases leading to a decrease in RL. The evolution of Pout and ID with a and P was already described in the previous paragraph. Hence, in general, Pout decreases with increasing RL, while ID increases as RL is increased. 12a= 190 11 - 45 nm CMOS W = 40 pm 10- VDQ= 1.1 V E 0 (D =0 a= 210 0 7CL 6- a= 220 a= 240 a=260 5L 0 = 200 mA/mmn DQ D Ron = 9.59 Q a = 00 a= 00 9- a=290 4 a=32 a =360 32 0 I 50 I I I 100 150 200 250 300 R (Ohm) Fig. 4.5: Modeled locus of output power versus load resistance. Output power measurements at different RL are shown as symbols in the figure and show good agreement with the model. .................... ....................... p =nC a=190 a = 200 70- a= 210 a = 220 60- a =240 a 260 5S50- x =290 40- 2 30- OW a =320 a =360 45 nm CMOS = 40 gm VDQ = 1.1 V = 200 mA/mm 20 Ro, = 9.59 Q 0 50 100 150 200 250 300 R (Ohm) Fig. 4.6: Modeled locus of drain efficiency versus load resistance. Drain efficiency measured at different RL is shown as symbols in the figure and shows good agreement with the model. It should be noted that the model does not impose any constraints on the maximum current drive of the device. In a real device, the drain current cannot exceed a certain saturated value. This explains why the measured output power is not as high as the maximum Pout predicted by the model. Fig. 4.4 suggests that for a given drain efficiency, the model can predict the maximum output power the device can deliver, or vice versa. We will use this fact to compute Pout as a function of device width and compare the modeled values with the measured data. Fig. 4.7 shows the measured TID (under both optimized power and optimized PAE conditions) as a function of width for 45 nm CMOS devices. The measured data is shown as symbols. The solid lines show the ... ............ . .... .......................... values of lD that were used as input to the model, to predict the maximum output power at that efficiency. The appropriate Ron and DC operating points were used in the model for each device. The model assumes that only the first harmonic of the drain current flows through the load resistance and that all the other harmonics are shorted out through perfect impedance matching at the fundamental frequency. However, in practice the impedance match need not be perfect and could result in non-zero values of second and higher order harmonics of the current. This would result in an increased value of the output power and drain efficiency. This explains why the maximum modeled drain efficiency was less than the measured values for some of the devices. In such cases, the maximum modeled efficiency closest to the measured value was selected. 90 80 Optimized for PAE model input 70 _0 60 Optimized for power C M 50 " 40 C S30 20 - 45nm CMOS VDD= 1-V ID 10 200 mAmm f=2GHz 0 0 100 200 300 400 500 600 700 Device Width (pm) Fig. 4.7: Drain Efficiency as a function of device width. The solid line represents 77D specified in model for predicting maximum Pout and optimum RL. The symbols are the measured 77D at 2 GHz. ... ... .. ............. ..... ..... For any given drain efficiency, the maximum possible power predicted by the model, for any combination of alpha and beta values, is recorded. The corresponding values of alpha and beta are also noted and are used to calculate the average drain current and the load resistance. Fig. 4.8 shows excellent agreement between the modeled values for maximum output power and the measured data, across all device widths. Both the modeled and the measured data show non-ideal scaling with device width, especially at large widths. Since the normalized DC operating point is the same for all device widths, the only parameter that can explain the nonideal scaling of the output power is the on-resistance. Hence, we can conclude that the main reason for the decrease in normalized output power with device width is because the onresistance of the devices does not scale ideally with device width. 80 VDD= 1-V ID200 70 mA/mm f=2 GHz 60 E model output,, Optimized for power 50 O 40 CL = 30 - Optimized for PAE 0. 02010 0 I 0 100 200 400 300 500 600 700 Device Width (gm) Fig. 4.8: Output power as a function of device width. The solid line represents the maximum 0 , at 2 GHz. output power predicted by the model. The symbols show the measured Po 100 ............ ... . .... ......... .... ................. .................. ............... .... The measured Pout on the 640 gm device is significantly lower than the modeled value (Fig. 4.8). This could be because of self-heating effects, which become dominant in the wide devices. The self heating explanation is further borne out in Fig. 4.9, where the modeled and measured values of the average drain current (IDC) are shown as a function of device width. In general, the modeled IDC closely matches the measured values, thus validating the model. The measured IDC on the 640 gm device is much lower than the modeled value, indicating the presence of self heating in the device. This could be verified by measuring Ron at the expected temperature and using that in the model to check if it predicts the measured value. 160 VDD= 1.1 V 140 ID= 200mAmm f=2GHz 120 100 E - Optimized for power model output 80 60 40 Optimized for PAE 20 0 100 200 300 400 500 600 700 Device Width (gm) Fig. 4.9: Average drain current as a function of device width. The solid line represents the average drain currentpredictedby the model. The symbols show the measuredIDC at 2 GHz. Fig. 4.10 shows the load resistance used in the power measurements (symbols), as well as the value of RL computed from the model (solid lines). The modeled values show reasonable agreement with the measurements. This means designers can use this model to not only predict 101 the maximum output power of a device for any given drain efficiency, but also to accurately compute the load resistance needed to achieve the required power. Output matching networks can then be designed to present this resistance to the load. 180 45 nm CMOS 160 - VDD=l1l V E41 E 140- 'D= 200 mA/mm f=2 GHz . 0 120 C 100 0 model output 60 0 40 - Optimized for PAE 20 0 I 0 100 200 300 400 500 I I 600 700 Device Width (pm) Fig. 4.10: Load resistanceneeded to obtain maximum Po,. Solid line is optimum RL predicted by model. Symbols are measured data at 2 GHz. 4.3 Discussion The model presented in this chapter shows that the maximum output power is strongly affected by the Ron of the device. Hence, the power capability of a device can be improved by designing it to have a lower Ro,. One way to decrease source and drain resistance and thus Rn is by placing additional contacts in the source and drain regions of the device. We designed a test structure where the distance between the gate fingers was increased to twice that in a standard device. This allowed us to place twice as many contacts in the source and drain regions. A 102 comparison of the Ron and the power parameters of the two structures is shown in the table below: Parameter Standard test structure New test structure (W = 2x20x2 gm) (W = 2x20x2 gm) 2X spacing between gate fingers Measured Modeled Measured Modeled VDQ (V) 1.1 1.1 1.1 1.1 IDQ (mA/mm) 200 200 200 200 Ron (Q) 5.3 5.3 4.7 4.7 Pout (mW) 15.3 15.3 17.9 17.4 67.6 65 66.5 65 IDC (mA) 20.6 21.4 24.5 24.4 RL (A) 38.6 39.6 30.3 33.9 1iD (%) Table 4.1: Measured and modeled values of DC and RF power data on two different test structures designed in 45 nm CMOS technology. Details of the standard device layout can be found in Chapter2. The spacing between gate fingers in the new test structure is twice that in the standardstructure. The table shows that increasing the number of source and drain contacts decreases the onresistance of the device. This in turn leads to a considerable increase in output power at the same drain efficiency. The table also shows that the modeled Pont, T'D, IDC, and RL are quite close to the measured values. Thus, the model does an excellent job at predicting the power performance of devices with different layouts. In a CMOS device, Ron can be approximated as the sum of the source resistance (Rs) and drain resistance (RD), since the channel resistance is typically much smaller than RD or Rs. Hence, if we can estimate RD and Rs for any device layout, through parasitic extractions from layout, we can predict the maximum Pout and 11D for that device. Thus, circuit designers can 103 estimate the power capability of their designs and optimize the design to achieve the necessary power or efficiency targets. 4.4 Summary A simple analytical model for the output power and drain efficiency of a CMOS device has been presented. The model uses the DC operating point and on-resistance of the device as inputs and gives out the maximum output power and the load resistance required to achieve that power for a given drain efficiency. We have shown that the modeled values of Pout, 11D, and RL show excellent agreement with the measured data on 45 nm CMOS devices, across a wide range of device widths and device layouts. Circuit designers can use this model to estimate the power and efficiency of their device, and optimize their device layout to meet performance targets. 104 Chapter 5 Small-signal Circuit Extractions The previous chapters have shown that the primary reason for the non-ideal scaling of output 0 ) of the device with power in wide devices is the increase in normalized on-resistance (Ro increasing device width. It was also shown that PAE decreases with increasing device width, mainly because the maximum oscillation frequency (fmax) of the device decreases with increasing device width. In this chapter, we will explore the reasons for the increase in Ro 0 and decrease in fmax in wide devices. This will be accomplished through small-signal equivalent circuit extractions from measured S-parameters. We will also discuss the sensitivity of fT and fmax to the different small-signal parameters and present analytical expressions for calculating fT and fma from these parameters. 5.1 Small-signal Circuit Extraction Methodology Accurate extraction of intrinsic device parameters requires appropriate de-embedding structures. We used 65 nm devices in this chapter because of the availability of custom designed open and short de-embedding structures for every device in this technology. The 45 nm devices use a common set of de-embedding structures, and hence would require a much more complicated equivalent circuit and extraction methodology to arrive at the intrinsic parameters of the device. 105 ......... S-parameter measurements, from 0.5 GHz to 40 GHz, were performed on 65 nm devices of different device widths, and the corresponding open and short de-embedding structures, at various DC bias conditions. S-parameters were then de-embedded using the standard open and short de-embedding technique, described in [42]. 200 200 max 180 150 160 wN N N -- 100 Z: 0 &I fT"a . 140 '4:WE V 120 D_= = 1V 50 100 mAlmm 100 10 Fig. 5.1: De-embeddedfT andfm, total device width. 100 1000 Device Width (srm) - 0 10000 measured at VDS=l V and at ID=100mA/mm, as a function of Fig. 5.1 shows the de-embedded fr and fmax as a function of total device width at VDS and ID=100 mA/mm. We find that fT = 1V decreases from 142 GHz to 110 GHz and fmax decreases from 190 GHz to 90 GHz as the device width is increased from 96 gm to 1536 in. To understand the reasons for the degradation of fT and fmax in wide devices, we extracted the small-signal equivalent circuit for these devices from the measured S-parameters. Several different equivalent circuits and the corresponding high frequency parameter extraction methodologies have been published in the literature [53-61]. We found that the equivalent circuit topology shown in Fig. 5.2 works best for our devices. The circuit includes parasitic resistances in the gate (RG), source (Rs) and drain (RD), transconductance (gm), output resistance (ro), 106 ......... . I'll ..... .... ..... .. ..... .... ....... ............. intrinsic gate-source (Cgs) and gate-drain (Cgd) capacitances, substrate resistance (Rsx), and parasitic capacitances from the body to gate (Cgb), source (Csb) and drain (Cdb). C9d Rg Rd G oQ^^^^ cgs gV : Cdb Cgb-Fr B S 0O/ Fig. 5.2: Small-signal equivalent circuit of a MOSFET including parasitic resistances and the substrate network. The following methodology was followed to extract the small-signal parameters [62-64]: 1. The parasitic resistances were first extracted from the S-parameters measured at VGs= VDS = 0 V using the following expressions [53]: RG = real(Z1 1 - Z1 2 ) (5.1) RD = real(Z2 2 (5.2) - Z12) Rs = real(Z1 2 ) 107 (5.3) 2. Next, S-parameters measured at VDS = 1 V, ID= 100 mA/mm, were converted to Z- parameters. The parasitic resistances extracted in step 1 were subtracted from these Zparameters to obtain the intrinsic Z-parameters. 3. The intrinsic Z-parameters were then converted to Y-parameters. The rest of the equivalent circuit parameters were then extracted from the intrinsic Y-parameters using the expressions given below [53, 57]: gm = real(Y2 1 ) = Rs, = c 1 real(Y22 ) (5.4) (5.5) real(Y22 + Y1 2 ) -)2(5.6) (imag(Y22 + Y1 2 ))2 imag(Y11 + Y1 2 ) C9a = Cd = Csb - imag(Y1 2 ) imag(Y22 + Y12 ) (57) (5.8) (5.9) (0) 4. Finally, the value of Cgb was determined by fitting the equivalent circuit model to the measured S-parameter data in Agilent ADS. We found that the extracted value of Cgb was usually very small (several fF). However, inclusion of Cgb in the equivalent circuit was crucial for accurate modeling of the unilateral gain (U) versus frequency. The equivalent circuit parameters of a 96 tm wide device (2 unit cells of 48 gm width), extracted using the methodology described above, from S-parameter measurements at VDS = 1V, and ID = 100 mA/mm are listed below: 108 . .. ............. .............. .. ........ . ..... ................ . RG=2.5 Q, Rs=2 Q, RD=3.25 Q, gm=0.107 S, ro=83 Q, Rsx=75 Q, Cgs= 5 6 if, Cgd=3 4 iF, Cdb=Csb=1 20 fF, Cg= 3 8 . f. An equivalent circuit, with parameter values listed above, was input into Agilent ADS, and Sparameters were simulated in the 0.5 GHz to 40 GHz range. Fig. 5.3 compares the simulated Sparameters with measured S-parameters (VDs = 1 V, ID= 100 mA/mm) on the 96 pm wide device. It can be seen that the model shows excellent agreement with the measured data over the entire frequency range. Measred Fig. 5.3: Measured and Modeled s-parametersat VDD=] V, ID=100 mA/mm. Measureddata is in symbols and the model is the solid line. W=2x48 pm. 109 ....... ... .. .. .............. ........... Fig. 5.4 plots the modeled and measured short-circuit current gain (h21) and the unilateral power gain (U) for the same device. It is clear that the model does an admirable job at predicting not only the fT and fmax of the device, but also h2 1 and U over the entire frequency range. It should be mentioned that the figure plots the absolute value of the unilateral gain. U is actually negative at low frequencies (< 10 GHz) because of the positive feedback path between gate and drain formed by Cgb, Cdb and Rsx Our model predicts even this negative U accurately. its 1E18 IE11 Frequency (Hz) Fig. 5.4: Short circuit current gain and unilateralpower gain as a function of frequency. W = 2x48 pm, VDD = 1 V, ID 100 mA/MM- 110 .............. ':: ........ ... .. ............ 5.2 ............. ...... ...... Width Dependence of Equivalent Circuit Parameters Small-signal equivalent circuit parameters were extracted for devices with total width ranging from 96 jim to 1536 jim. The extracted gm, gds, Cgs and Cgd, normalized to device width, are shown as a function of device width in Fig. 5.5. All these extracted intrinsic parameters are relatively constant with device width. Fig. 5.6 shows the normalized parasitic resistances across device width. Rs is constant across width, but RG and RD increase with increasing width. Normalized RG increases by 120% and normalized RD increases by 180% as the device width is increased from 96 gm to 1536 gm. 1.2 1 9m 0.8 E E E0.8 0.6 E 'U 0.6 0.4 E0.4 CL .U 0.2 0 0 -- I 0 10 100 1000 10000 Device Width (gm) Fig. 5.5: Normalized intrinsicparameters (g,, 1 V, ID = 100 mA/mm. gds, 111 Cgs, Cgd) as a function of device width. VDD = .. ... ........ ..... ......... ....... ............. 800 700 1V ID=lO0mA/mm VDD = - E 600 Q 500 RD 400 RG -- '300 a 200 100 0 10 1000 100 Device Width (sm) 10000 Fig. 5.6: Normalized parasitic resistances vs. device width. Parasiticresistances are extracted from s-parametersmeasured at VDD=J V, ID=100 mA/mm. 2 cells 8 cells Fig. 5.7: Schematic layout of a 2 cell (W = 96 pm) and 8 cell (W = 384 pm) device, showing the need for additionalwiring to connect the 8 cells in parallel. This leads to an increasein parasitic resistanceas the device width is increased. 112 The reason for the increase in normalized RG and RD with device width can be understood by examining the layout of these devices. The device width in these transistors is increased by wiring multiple unit cells in parallel. Fig. 5.7 shows the layout of a 2 cell device (W = 96 gim) and an 8 cell device (W = 384 m). The additional wiring required to connect the extra unit cells in the 8 cell device leads to an increase in the parasitic resistances in this device. Also, the drain current needs to travel further to get to the unit cells on the left of the layout, and thus increases the drain resistance even further. The ideal scaling of the source resistance can be explained by the de-embedding structure design for these devices. When designing the open and short deembedding structures, the source metallization was kept identical to that in the corresponding device, thus allowing de-embedding of the source up to the intrinsic diffusion layer. Hence, the normalized source resistance is constant across device width. On the other hand, the gate and drain can only be de-embedded till the metal 3 level, and hence show an increase in normalized resistance with increasing device width. We are now in a position to explain the reason for the non-ideal scaling of output power in wide devices. We had previously shown that the reason for the power degradation is an increase in normalized Ro, in the wide devices. For very short CMOS devices, R," ~ RD + Rs. Hence, the main reason for the degradation in output power can be attributed to the increase in normalized RD in the wide devices. Thus, a parasitic-aware approach is required when laying out wide devices to avoid an increase in parasitic resistance and a corresponding decrease in the power performance of the device. 5.3 fT, fmax Sensitivity to Equivalent Circuit Parameters To examine the effect of the various small-signal parameters on fT and fmax, a sensitivity analysis was carried out in ADS. The parameters were changed one at a time by 100% in a 113 .......................... ....... ....................................... ..... .... : ....... direction that would result in degradation in fT and fmax. Thus, the parasitic resistances and all the capacitances were increased by 100%, while the transconductance and output resistance were decreased by 100%. -50 * f-r m -40 --,0 E -30 C -20 C -10 O 1 -j- 0 - 10 -RG x2 20 JI RD x2 Rs gm ro x2 2 2 C., x2 Cgd x2 II Scale ideally Rsx Csb x2 I|I| x2 Cdb Cgb x2 x2 Minor Impact Fig. 5.8: Sensitivity offTand fma to the various small-signal equivalent circuit parameters. The figure plots the percentage decrease infT and fa,,xfor a 100% change in each of the parameters. W=96 pm, VDD=l V, ID=100 mA/mm. The results are shown in Fig. 5.8. The intrinsic parameters, gm, Cgs, Cgd, and to some extent ro, show the most impact on both fT and fmax. However, Fig. 5.5 showed that these parameters scale ideally with device width and hence could not be responsible for the degradation in f with device width. The substrate parameters, Rx, Cdb, Csb, and Cgb show very minor impact (less than 5%) on either fT or fmax. A 100% increase in Rs degrades fT by around 7% and improves fmax by 8%. But again Rs scales ideally with device width (Fig. 5.6) and thus cannot explain the degradation in fma. This leaves the parasitic resistances RG and RD. fT shows no dependence on RG and a small dependence on RD (13% decrease in fT for a 100% increase in RD). However, fmax shows a much stronger dependence on both RG and RD (24% degrade in fmax with 100% increase 114 in either RG or RD). It was shown earlier in Fig. 5.6 that the normalized RG and RD increase with device width. Thus, the improper scaling of RG and RD is responsible for the degradation of fmax in wide devices. To further validate our argument, we carried out the following exercise in Agilent ADS. The equivalent circuit with small-signal parameter values equal to those extracted from the measured S-parameters on the 96 gm wide device was simulated in ADS. The simulated fT and fm were found to be very close to the measured values (142 GHz and 180 GHz versus 142 GHz and 190 GHz respectively), lending credence to the model. Now the values of RG and RD were increased by 120% and 180% respectively, keeping all other parameters constant. This increase in RG and RD correspond to the increase seen in the measured values for normalized RG and RD when the device width was increased from 96 gm to 1536 pim (Fig. 5.6). The equivalent circuit with the modified RG and RD is then simulated, resulting in respectively. These simulated values for fT fT and fmax of 112 GHz and 95 GHz and fmax are very close to the measured values of 110 GHz and 90 GHz on the 1536 jim wide device. This means the decrease in the measured fT and fmax, when the device width is increased from 96 jim to 1536 jim can be attributed entirely to the increase in normalized RG and RD. Hence it can be concluded that the main reason for the degradation in fT in the wide devices is an increase in parasitic RD and the reason for degradation in fmax is the increase in parasitic RG and RD. The degradation in fmax leads to a corresponding decrease in the output power and PAE in the wide devices. 5.4 Analytical Expressions for fT and fmax A complete y-parameter analysis of the small-signal equivalent circuit shown in Fig. 5.2 was carried out in an effort to obtain simple analytical expressions for fT and fma that correctly account for the width scaling of the relevant elements. The substrate parameters, Rsx, Cdb, 115 Csb, and Cgb were not considered in this analysis because they have negligible impact on fT or fmax (Fig. 5.8). In the absence of these elements, and ignoring o and higher order terms, the y-parameters can be approximated by: Y1 ' jo>Cgs(1 + gds(RD + Rs)) +jO>Cgd(l + (9M + gds)(RD + RS)) (5.10) D -jwOCgs(gdsRs) UW&Cgd(l + (9in + gds)RS) Ygn -j)Cgs(gdsRs) -jW)Cgd(l D 9ds + jo>Cgs(gds(RG + RS))+ (5.11) + (gm + gds)RS) (5.12) j>Cgd(1 + (9m + gds)(RG + Rs)) (5.13) D Where: D = 1+ YmRs + gds(RD + Rs + RGRs + RGRD + RDRS) +joCgs(RG + Rs + gds(RGRS + RGRD + RDRS) +jWCgd(RG (5.14) + RD + gds(RGRS + RGRD + RDRS) The short-circuit current gain, h 21 and the unilateral gain, U, can be expressed in terms of yparameters as: h2 = (5.15) 21 Y12 2 Y12 -Y211 4[real(Y 1 )real(Y22 ) - real(Y12 )real (Y2 1 )] Approximate expressions for fT [65] and fmax can then be derived as: 116 (5.16) ............... JT 2 C 1 + R + Rs) + Cyd 1 + (RD + RS) ( m ................................... .. ))] (5.17) 9mV1 + 9mRs + gds(RD + RS) CgS(RG + Rs)gds(1 + 9mRs) max 47 +Cg2d(( RG D) (gm + ds) + (g +CgsCgd(RG(grm + 2 +gds) gds) + gmgds(SRGRS + 2 (2 (5.18) RGRS + RGRD + 2RsRD)) + 2RSRD) + g2RG Rs) 3 RGRD The above expressions allow technologists and circuit designers to easily determine the frequency metrics for a given device design. The traditional derivations for U and fm, [66] only consider the effect of RG. We have also included the effect of Rs and RD to improve the accuracy of the calculated fT and fmax. 200 200 180 calc fmax'A meas fmax 0 1-140 4eN - calc fT 150 10 o- meas fT 50 120 VDD= 1 V ID 100 mAlmm 100 10 . . 100 1000 0 10000 Device Width (pm) Fig. 5.9: Measured and calculatedfT and f,,, as a function of device width. Measured data in solid symbols and lines and calculateddata in open symbols and dashed lines. 117 Fig. 5.9 shows the measured data for fr and fmax along with the values calculated using the above expressions. The calculated values show excellent agreement with the measured data over the entire range of device widths studied in this work. 5.5 Summary In this chapter, we investigate the reasons for the decrease in fmax and increase in normalized on-resistance, with increasing device width. Small-signal equivalent circuit parameter extractions across device width show that the main reason for fT and fmax degradation is the increase in parasitic gate and drain resistances with width because of the presence of nonscalable parasitics in wide devices. The increase in parasitic drain resistance is also responsible for the non-ideal scaling of on-resistance in wide devices. We have shown in previous chapters that the non-ideal scaling of Ron leads to a decrease in output power with increasing device width, and that the degradation in fmax with device width is the main contributor for the decrease in PAE in the wide devices. Thus, the key to enabling CMOS for millimeter-wave applications is a parasitic-aware approach when designing wide devices. The chapter also presents simple analytical expressions for fT and fmax that allow circuit designers and technologists to easily determine the frequency metrics of any device designed in a given technology. 118 Chapter 6 Conclusions and Suggestions for Further Work We have studied the frequency and power performance limits of RF CMOS devices in the 65 nm and 45 nm technologies. Most of the RF power measurements were performed with the source and load impedances optimized for maximum power-added efficiency at a DC bias of VDS = VDD and ID = 200 mA/mm. A maximum output power density of 122 mW/mm, a peak PAE of 72% and a maximum total output power of 77.2 mW was measured on the nominal 65 nm high-performance devices (VDD = 1 V) at 2 GHz. Higher output power levels at similar PAE values were obtained on the 65 nm low-power devices, because of the higher supply voltage (VDD = 1.2 V). A record output power density of 210 mW/mm, PAE in excess of 75%, and maximum total power of 68.6 mW was measured on the 45 nm devices at VDS = 1.1 V, ID = 200 mA/mm, and f = 2 GHz. This record power density was made possible by optimizing source and load impedances for maximum output power and through careful device layout for minimized parasitic resistances and capacitances. If supply voltage is not a constraint, even higher power levels can be obtained by using the 1/0 devices. The 2.5 V I/O devices in the 65 nm low-power 119 technology demonstrated a maximum output power density of 267 mW/mm and total power of 259 mW at ID = 200 mA/mm and f = 8 GHz. We have demonstrated that gate width scaling is an effective means to increasing total output power in a device. However, output power does not scale ideally with increasing device width. Instead, the output power density (Pout / W) decreases as the device width is increased. The reason for this decrease in power density can be attributed to the non-ideal scaling of onresistance in wide devices. Increasing device width also results in a decrease in PAE because of fmax degradation in wide devices. A decrease in fmax means that for a given operating frequency, the device is operating closer to its maximum frequency limit and hence exhibits lower power gain, which in turn leads to a decrease in the PAE. Power measurements across a broad frequency range (2 GHz to 18 GHz) indicate that output power is independent of frequency, but PAE decreases as the frequency increases. A simple analytical model was developed for estimating the output power and efficiency of a CMOS power amplifier. The model uses the DC operating point and on-resistance of the device as inputs and gives out the maximum output power and the load resistance required to achieve that power for a given drain efficiency. We have shown that the modeled values of Pout, TD, and RL show excellent agreement with the measured data on 45 nm CMOS devices, across a wide range of device widths and device layouts. Power amplifier designers often rely on RF power simulations using compact device models to predict the output power and PAE of their circuits. The compact device models are usually very complex, containing more than 100 parameters to accurately model the device characteristics. Hence the power simulations are a drain on both designer's time and computational resources. Our simple analytical model would allow designers to quickly estimate 120 the maximum power and efficiency of their devices and circuits. They could even optimize their device layouts to meet or improve on the power performance targets. The measured power data and the analytical model suggest that the maximum output power delivered by a device at any given bias point is only limited by the on-resistance of the device. The device capacitances play a role in degrading the efficiency of the device. However, as long as extra input power can be provided to charge the capacitors, one can always obtain the same maximum output power. Hence, the key to obtaining high power levels in the millimeter-wave regime is to design devices with minimized parasitic resistances. Small-signal equivalent circuit extractions were performed on 65 nm devices to understand the reason for degradation in on-resistance and fmx with increasing device width. The results indicate that the non-ideal scaling of parasitic RD is responsible for the increase in normalized Ro, in wide devices, while the non-ideal scaling of RG and RD contribute to the degradation in fmax in the wide devices. The non-ideal scaling of RG and RD is due to the parasitics associated with the additional wiring needed to connect multiple unit cells to obtain the required width. Hence, the key to obtaining high power and efficiency levels from CMOS devices at high frequencies is a parasitic aware approach when designing large devices. Several device layouts were explored in an effort to reduce parasitic resistances and capacitances. We found that designing an optimum unit cell and connecting multiple unit cells in parallel was the most effective approach to design wide devices. When designing the unit cell, it is essential to determine the optimum finger width for that technology. We found the optimum finger width to be 1.5 jim for 45 nm technology. One approach to decreasing the parasitic source and drain resistances is to increase the distance between gate fingers, thus allowing placement of additional contacts in the source and drain regions. We demonstrated that increasing the gate 121 finger pitch to two times the minimum pitch specified in the design manual for 45 nm devices, resulted in a decrease in Ron from 5.3 Q to 4.7 Q, and a corresponding improvement in output power from 15.3 mW to 18 mW. We also reduced parasitic capacitances in 45 nm devices by minimizing the overlap between gate poly silicon and metal levels and by wiring the gate, source and drain regions using alternate metal levels. These layout modifications allowed us to demonstrate significant improvement in fmax, Pout and PAE of 45 nm devices. Future research can explore many of the aspects discussed in this thesis. The maximum operating frequency in this thesis was limited by our load-pull measurement setup to 18 GHz. Given the increased interest in using CMOS devices for millimeter-wave power applications, it would be worthwhile to perform an extensive characterization of these devices in the 60 GHz to 80 GHz range. Also, all the devices explored in this thesis employ a poly silicon gate and thus exhibit significant gate resistance. It would be interesting to study metal gate devices to quantify the improvement in gate resistance and corresponding improvement in fmax, Pout and PAE in these devices. Several improvements could also be made to the analytical model for output power presented in this thesis. At present, the model allows us to estimate the maximum output power and drain efficiency for a device. Incorporating power gain and frequency into the model would allow us to determine the PAE and also the evolution of efficiency with increasing frequency. Another area to explore would be to extract parasitic RD and Rs from device layout and use those values in the analytical model to predict maximum Pout and flD for that device. That would enable circuit designers to estimate the power capability of their designs and optimize the design to achieve the necessary power or efficiency targets. 122 Compact models for RF CMOS devices typically model the device behavior up to the nominal operating voltage. However, for power amplifiers, the maximum voltage of operation could be over twice the nominal voltage. Hence, it is worthwhile to study the validity of the compact models in predicting the output power and PAE at millimeter wave frequencies. This could be accomplished by simulating a single stage amplifier, that uses the compact model for the device, at millimeter-wave frequencies, and comparing the simulation results to measured load-pull data. 123 124 Appendix A Device Library A.1. 65 nm Test Structures This section gives a listing of device structures designed for fabrication in IBM's 65 nm technologies. The same device structures were fabricated in both 65 nm low-power and 65 nm high-performance technologies. The only difference was the absence of open and short deembedding structures in the 65 nm low-power test site. Each device has a unique identifier (id) to help locate it on the wafer. A map of device locations on wafer is shown at the end of the section. A number of test structures were designed with increasing device width for the standard and 1/0 devices. The device width was increased by connecting multiple unit cells in parallel. The details of the unit cell design were discussed in Chapter 2. The cell orientation column in the tables below describes the orientation of the cells in each structure (The unit cells are arranged in an X x Y array and wired together). Test structures with varying device lengths were also designed for the standard devices to study the impact of gate length on frequency and power performance. The 65 nm technology has three different threshold voltage options for the thin oxide devices. The impact of the VT implant was also studied using appropriate test structures. 125 Open and short de-embedding structures were custom designed for some of the standard and 1/0 devices. The design of these de-embedding structures was discussed in chapter 2. A list of the de-embedding structures can be found in the map at the end of this section. Device Width Scaling (Standard Device): Identifier Gate Length (nm) Unit finger width (jim) # fingers per cell # cells Cell Orientation (X x Y) Total Width (gm) N 001 50 50 50 50 50 50 50 50 50 50 50 50 50 50 2 2 2 2 2 2 2 2 2 2 2 2 2 2 24 24 24 24 24 24 24 24 24 24 24 24 24 24 1 2 4 8 16 32 64 128 216 336 440 528 640 728 1xi 2x1 2x2 2x4 2x8 4x8 4x16 4x32 6x36 8x42 10x44 12x44 16x40 18x44 48 96 192 384 768 1536 3072 6144 10368 16128 21120 25344 30720 38016 # fingers # cells Cell Orientation Total (X x Y) Width ( m) N002 N 004 N008 N016 N032 N064 *N128 N216 N336 N440 N 528 N640 N792 Device Width Scaling (1/0 Device): Identifier Gate Length (nm) DG 001 DG 002 DG 004 DG 008 DG 016 DG 032 DG 064 DG 128 DG 216 DG 336 DG 528 DG 792 230 230 230 230 230 230 230 230 230 230 230 230 Unit finger width ( m) per cell 2 2 2 2 2 2 2 2 2 2 2 2 24 24 24 24 24 24 24 24 24 24 24 24 1 2 4 8 16 32 64 128 216 336 528 728 1xI 2x1 2x2 2x4 2x8 4x8 4x16 4x32 6x36 8x42 12x44 18x44 48 96 192 384 768 1536 3072 6144 10368 16128 25344 38016 Unit finger width (gm) 2 # fingers per cell 24 # cells Cell Orientation (X x Y) 4x8 Total Width (gm) Device Length Scaling: Identifier LG 040 Gate Length (nm) 40 126 32 1536 45 65 80 130 250 330 00 LG 045 LG 065 LG 080 LG 130 LG 250 LG 330 LG 5500 24 24 24 24 24 24 24 2 2 2 2 2 2 2 4x8 4x8 4x8 4x8 4x8 4x8 4x8 32 32 32 32 32 32 32 1536 1536 1536 1536 1536 1536 1536 VT Variations: SLVT SNFET SHVT Gate Unit finger # fingers # Cell Orientation Lnmh width (gm) per cell cells 50 50 50 2 2 2 24 24 24 32 32 32 (X x Y) Total i VT Implant 4x8 4x8 4x8 1536 1536 1536 Low VT Regular VT High VT Placement of Various Test Structures on Wafer: A.2. 45 nm "Old" Test Structures This section gives a listing of device structures designed for fabrication in IBM's 45 nm low- power technology. These test structures were part of the "Old" test site and were used to investigate the impact of gate length, number of fingers, unit finger width, substrate ring placement, and number of gate contacts on the frequency and power performance of 45 nm devices. All device structures were designed containing a single cell and the gate width of the 127 devices was increased by increasing either the number of fingers or the unit finger width. Open and Short de-embedding structures were also custom designed for each of the test structures to allow proper de-embedding of S-parameter measurements to the intrinsic device. Further details on the layout of these structures can be found in chapter 2. Identifier Gate Length (pm) Unit finger width (gim) # fingers per cell STD 101 STD 102 STD 103 STD 104 STD 105 STD 106 STD 107 STD 108 STD 109 STD 110 STD 111 STD 112 STD 201 STD 202 STD 203 STD 204 STD 205 STD 206 STD 207 STD 208 STD 209 STD 210 STD 211 STD 212 STD 301 STD 302 STD 303 STD 304 STD 305 STD 306 1 0.5 0.2 0.1 0.06 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.26 0.2 0.16 0.1 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.04 0.5 0.2 0.1 0.04 0.04 0.04 1 1.5 1.5 1.5 1.5 1,5 1.5 3 1.5 3 3 9 1.5 1.5 1.5 1.5 0.5 1.5 3 3 3 5 5 9 3 3 3 1.5 1.5 3 20 20 20 20 20 20 20 20 60 10 10 10 20 20 20 20 60 40 6 6 40 6 60 10 60 60 60 80 120 60 128 Total Width (m) 20 30 30 30 30 30 30 60 90 30 30 90 30 30 30 30 30 60 18 18 120 30 300 90 180 180 180 120 180 180 Gate Contact (2 or 1 sides) 2 2 2 2 2 2 1 2 2 2 1 2 2 1 2 1 2 2 2 1 2 2 2 1 2 2 2 2 2 2 Substrate ring to device distance (gim) 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 0.53 ............... :......... STD STD STD STD STD STD 307 308 309 310 311 312 0.04 0.04 0.04 0.04 0.04 0.04 3 0.42 1 1 1.5 1.5 60 60 60 120 60 60 180 25.2 60 120 90 90 1 2 2 2 2 2 0.53 0.53 0.53 0.53 0.575 0.625 Placement of Various Test Structures on Wafer: A.3. 45 nm "New" Test Structures This section gives a listing of device structures designed as part of the "New" test site for fabrication in IBM's 45 nm low-power technology. In this set, the device width was increased by connecting multiple unit cells in parallel. The unit cell was designed with 20 fingers of 2 gm finger width (which is close to the optimum finger width found from the study on the "old" test structures). Further improvements to the unit cell design were made to optimize the parasitic resistances and capacitances, to help improve the frequency and power performance. Due to space limitation on the test vehicle, separate de-embedding structures could not be designed for 129 each of the device structures. Instead, a common set of de-embedding structures were designed to allow proper de-embedding of S-parameter measurements to the intrinsic device. Further details on the layout of these structures can be found in chapter 2. 130 Bibliography [1] H. S. Bennett, et al., "Device and technology evolution for Si-based RF integrated circuits," Electron Devices, IEEE Transactions on, vol. 52, pp. 1235-1258, 2005. [2] M. I. Mona M. Hella, RF CMOS Power Amplifiers - Theory, Design and Implementation, 2002. [3] M. Cooke, "CMOS reaches for 60 GHZ+ applications," Semiconductor Today, vol. 3, pp. 76-80, 2008. [4] D. R. Vizard, "Millimeter-wave Applications: From satellite communications to security systems," Microwave Journal,vol. 49, pp. 22-27, 2006. [5] J. Laskar, et al., "The Next Wireless Wave is a Millimeter Wave," Microwave Journal, vol. 50, pp. 22-39, 2007. [6] C. H. Doan, et al., "Design considerations for 60 GHz CMOS radios," Communications Magazine, IEEE, vol. 42, pp. 132-140, 2004. [7] A. Cathelin, et al., "Design for millimeter-wave applications in silicon technologies," in Solid State Circuits Conference, 2007. ESSCIRC 2007. 33rd European, 2007, pp. 464-471. [8] H. H. Ali M. Niknejad, Ed., mm-wave Silicon Technology - 60 GHz and Beyond. 2008, p.App. Pages. [9] A. Cathelin, et al., "Deep-submicron digital CMOS potentialities for millimeter-wave applications," in Radio FrequencyIntegrated Circuits Symposium, 2008. RFIC 2008. IEEE, 2008, pp. 53-56. [10] B. G. Duixian Liu, Ullrich Pfeiffer, Janusz Grzyb, Ed., Advanced Millimeter-wave Technologies Antennas, Packagingand Circuits.2009, p.App. Pages. [11] A. Niknejad, "Siliconization of 60 GHz," Microwave Magazine, IEEE, vol. 11, pp. 78-85, 2010. [12] B. Razavi, "Design of Millimeter-Wave CMOS Radios: A Tutorial," Circuits and Systems I: Regular Papers,IEEE Transactions on, vol. 56, pp. 4-16, 2009. 131 [13] C. Y. Law and A. V. Pham, "A high-gain 60GHz power amplifier with 20dBm output power in 90nm CMOS," in Solid-State Circuits Conference Digest of Technical Papers (ISSCC), 2010 IEEE International,2010, pp. 426-427. [14] A. Siligaris, et al., "A 60 GHz Power Amplifier With 14.5 dBm Saturation Power and 25% Peak PAE in CMOS 65 nm SOI," Solid-State Circuits,IEEEJournal of, vol. 45, pp. 1286-1294, 2010. [15] E. Cohen, et al., "60GHz 45nm PA for linear OFDM signal with predistortion correction achieving 6.1% PAE and &#x2212;28dB EVM," in Radio Frequency Integrated Circuits Symposium, 2009. RFIC 2009. IEEE,2009, pp. 35-38. [16] K. Raczkowski, et al., "50-to-67GHz ESD-protected power amplifiers in digital 45nm LP CMOS," in Solid-State Circuits Conference - Digest of Technical Papers,2009. ISSCC 2009. IEEE International, 2009, pp. 382-383,383a. [17] T. Kjellberg, et al., "A Compact Cascode Power Amplifier in 45-nm CMOS for 60-GHz Wireless Systems," in Compound Semiconductor Integrated Circuit Symposium, 2009. CISC 2009. Annual IEEE, 2009, pp. 1-4. [18] M. Abbasi, et al., "A broadband differential cascode power amplifier in 45 nm CMOS for highspeed 60 GHz system-on-chip," in Radio Frequency Integrated Circuits Symposium (RFIC), 2010 IEEE, 2010, pp. 533-536. [19] T. Quemerais, et al., "A CMOS class-A 65nm power amplifier for 60 GHz applications," in Silicon Monolithic IntegratedCircuits in RF Systems (SiRF), 2010 Topical Meeting on, 2010, pp. 120-123. [20] W. L. Chan, et al., "A 60GHz-band 1V 11.5dBm power amplifier with 11% PAE in 65nm CMOS," in Solid-State Circuits Conference - Digest of Technical Papers,2009. ISSCC 2009. IEEE International, 2009, pp. 380-381,381a. [21] A. Valdes-Garcia, et al., "60 GHz transmitter circuits in 65nm CMOS," in Radio Frequency IntegratedCircuits Symposium, 2008. RFIC 2008. IEEE, 2008, pp. 641-644. [22] D. Dawn, et al., "60GHz CMOS power amplifier with 20-dB-gain and 12dBm Psat," in Microwave Symposium Digest,2009. MIT '09. IEEE MTT-S International,2009, pp. 537-540. 132 [23] J. Yung-Nien, et al., "Design and Analysis of a 55&#x2013;71-GHz Compact and Broadband Distributed Active Transformer Power Amplifier in 90-nm CMOS Process," Microwave Theory and Techniques, IEEE Transactions on, vol. 57, pp. 1637-1646, 2009. [24] T. LaRocca, et al., "60 GHz CMOS Amplifiers Using Transformer-Coupling and Artificial Dielectric Differential Transmission Lines for Compact Design," Solid-State Circuits, IEEE Journal of vol. 44, pp. 1425-1435, 2009. [25] J. Brinkhoff, et al., "A 60 GHz transformer-based variable-gain power amplifier in 90nm CMOS," in Radio-Frequency Integration Technology, 2009. RFIT 2009. IEEE InternationalSymposium on, 2009, pp. 60-63. [26] K. Jing-Lin, et al., "A 50 to 70 GHz Power Amplifier Using 90 nm CMOS Technology," Microwave and Wireless Components Letters, IEEE, vol. 19, pp. 45-47, 2009. [27] N. Kurita and H. Kondoh, "60GHz and 80GHz wide band power amplifier MMICs in 90nm CMOS technology," in Radio Frequency Integrated Circuits Symposium, 2009. RFIC 2009. IEEE, 2009, pp. 3942. [28] M. Bohsali and A. M. Niknejad, "Current combining 60GHz CMOS power amplifiers," in Radio FrequencyIntegrated CircuitsSymposium, 2009. RFIC 2009. IEEE, 2009, pp. 31-34. [29] D. Chowdhury, et al., "A 60GHz 1V + 12.3dBm Transformer-Coupled Wideband PA in 90nm CMOS," in Solid-State Circuits Conference, 2008. ISSCC 2008. Digest of Technical Papers. IEEE International,2008, pp. 560-635. [30] L. Jau-Jr, et al., "Wideband PA and LAN for 60-GHz Radio in 90-nm LP CMOS Technology," in Compound SemiconductorIntegratedCircuits Symposium, 2008. CSIC '08. IEEE, 2008, pp. 1-4. [31] K. Maruhashi, et al., "60-GHz-Band CMOS MMIC Technology for High-Speed Wireless Personal Area Networks," in Compound Semiconductor Integrated Circuits Symposium, 2008. CSIC '08. IEEE, 2008, pp. 1-4. [32] Y. Yoshihara, et al., "A 60-GHz CMOS power amplifier with Marchand balun-based parallel power combiner," in Solid-State Circuits Conference, 2008. A-SSCC '08. IEEE Asian, 2008, pp. 121-124. 133 [33] J. Scholvin, et al., "Fundamental Power and Frequency Limits of Deeply-Scaled CMOS for RF Power Applications," in Electron Devices Meeting, 2006. IEDM '06. International,2006, pp. 1-4. [34] J. Scholvin, et al., "Performance and limitations of 65 nm CMOS for integrated RF power applications," in Electron Devices Meeting, 2005. IEDM Technical Digest. IEEE International,2005, pp. 369-372. [35] J. Scholvin, et al., "RF power potential of 90 nm CMOS: device options, performance, and reliability," in Electron Devices Meeting, 2004. IEDM Technical Digest. IEEE International, 2004, pp. 455-458. [36] M. Ferndahl, et al., "Deep Submicron CMOS for Millimeter Wave Power Applications," Microwave and Wireless Components Letters, IEEE, vol. 18, pp. 329-331, 2008. [37] M. Ferndahl, et al., "90-nm CMOS for microwave power applications," Microwave and Wireless Components Letters, IEEE, vol. 13, pp. 523.-525, 2003. [38] Z. Luo, et al., "High performance and low power transistors integrated in 65nm bulk CMOS technology," in Electron Devices Meeting, 2004. IEDM Technical Digest. IEEE International,2004, pp. 661-664. [39] A. Steegen, et al., "65nm cmos technology for low power applications," in Electron Devices Meeting, 2005. IEDM Technical Digest. IEEE International,2005, pp. 64-67. [40] J. Yuan, et al., "A 45nm Low Cost Low Power Platform by Using Integrated Dual-Stress-Liner Technology," in VLSI Technology, 2006. Digest of Technical Papers.2006 Symposium on, 2006, pp. 100101. [41] J. Wang, et al., "RF Modeling of 45nm Low-Power CMOS Technology," presented at the Workshop on Compact Modeling, 2009. [42] M. C. A. M. Koolen, et al., "An improved de-embedding technique for on-wafer high-frequency characterization," in Bipolar Circuits and Technology Meeting, 1991., Proceedingsof the 1991, 1991, pp. 188-191. [43] J. Scholvin, Deeply Scaled CMOSfor RF PowerApplications, 2006. 134 [44] "Maury Microwave Corporation Automated Tuner System: Operating Manual," ed. [45] S. C. Cripps, RF Power Amplifiers for Wireless Communications, 2006. [46] U. Gogineni, et al., "RF power potential of 45 nm CMOS technology," in Silicon Monolithic IntegratedCircuits in RF Systems (SiRF), 2010 Topical Meeting on, 2010, pp. 204-207. [47] B. J. Baliga, Siliicon RF Power MOSFETs, 2005. [48] M. K. Kazimierczuk, "RF Power Amplifiers," 2008. [49] D. M. Pozar, Microwave Engineering,1998. [50] J. L. B. Walker, Ed., High-Power GaAs FETAmplifiers. 1993, p.App. Pages. [51] M. Albulet, "RF Power Amplifiers," 2001. [52] A. Grebennikov, RF and Microwave Power Amplifier Design, 2005. [53] D. Lovelace, et al., "Extracting small-signal model parameters of silicon MOSFET transistors," in Microwave Symposium Digest, 1994., IEEE MTT-S International,1994, pp. 865-868 vol.2. [54] G. Dambrine, et al., "A new method for determining the FET small-signal equivalent circuit," Microwave Theory and Techniques, IEEE Transactionson, vol. 36, pp. 1151-1159, 1988. [55] L. Seonghearn, "Direct extraction of small-signal model parameters for nanoscale MOSFETs," in Future of Electron Devices, 2004. InternationalMeeting for, 2004, pp. 111-112. [56] L. Seonghearn, et al., "A small-signal RF model and its parameter extraction for substrate effects in RF MOSFETs," Electron Devices, IEEE Transactionson, vol. 48, pp. 1374-1379, 2001. [57] C. Yuhua, et al., "MOSFET modeling for RF IC design," Electron Devices, IEEE Transactionson, vol. 52, pp. 1286-1303, 2005. [58] Y. S. Chi, et al., "An Analytical Parameter Extraction of the Small-Signal Model for RF MOSFETs," in ElectronDevices and Solid-State Circuits,2005 IEEE Conference on, 2005, pp. 555-558. [59] K. Ickjin, et al., "A simple and analytical parameter-extraction method of a microwave MOSFET," Microwave Theory and Techniques, IEEE Transactionson, vol. 50, pp. 1503-1509, 2002. 135 [60] R. Torres-Torres and R. Murphy-Arteaga, "Straightforward determination of small-signal model parameters for bulk RF-MOSFETs," in Devices, Circuits and Systems, 2004. Proceedings of the Fifth IEEE InternationalCaracasConference on, 2004, pp. 14-18. [61] M. T. Yang, et al., "Broadband small-signal model and parameter extraction for deep sub-micron MOSFETs valid up to 110 GHz," in Radio Frequency Integrated Circuits (RFIC) Symposium, 2003 IEEE, 2003, pp. 369-372. [62] U. Gogineni, et al., "Effect of Substrate Contact Shape and Placement on RF Characteristics of 45 nm Low Power CMOS Devices," Solid-State Circuits,IEEE Journal of, vol. 45, pp. 998-1006, 2010. [63] U. Gogineni, et al., "Effect of substrate contact shape and placement on RF characteristics of 45 nm low power CMOS devices," in Radio FrequencyIntegrated Circuits Symposium, 2009. RFIC 2009. IEEE, 2009, pp. 163-166. [64] U. Gogineni, et al., "Modeling Frequency Response of 65 nm CMOS RF Power Devices," presented at the TECHCON, Austin, TX, 2009. [65] P. J. Tasker and B. Hughes, "Importance of source and drain resistance to the maximum fT of millimeter-wave MODFETs," Electron Device Letters, IEEE, vol. 10, pp. 291-293, 1989. [66] B. Razavi, et al., "Impact of distributed gate resistance on the performance of MOS devices," Circuits and Systems I: Fundamental Theory and Applications, IEEE Transactions on, vol. 41, pp. 750754, 1994. 136