CMOS POWER DEVICE MODELING AND AMPLIFIER CIRCUITS BY DORIS A. CHAN DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Electrical and Computer Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 2010 Urbana, Illinois Doctoral Committee: Professor Milton Feng, Chair Assistant Professor Yun Chiu Professor Elyse Rosenbaum Professor Jose E. Schutt-Aine ABSTRACT A power amplifier (PA) is a key part of the RF front-end in transmitters for a local broadband network. Today, commercial PAs are made of III-V HEMT and HBT technology with excellent results. An integrated system-on-chip power amplifier circuit using CMOS technology for cost-effective and spectrum-efficient high-speed wireless communication presents major challenges because power amplifiers have been the limiting components in RF CMOS transmitter integrated circuits (ICs). At high frequencies, the distributed effect and power device-scaling issues put other constraints on PA design such as the trade-off between output power (Pout) and power added efficiency (PAE). Recently, CMOS has become attractive for low-cost and high-level integration due to the advancement of NMOS performance with ft and fmax > 100 GHz and is available from commercial CMOS foundries. However, the foundry-provided BSIM-RF model is unable to accurately predict the I-V characteristics and RF behaviors (ft and fmax) of power devices with widths of several hundred microns. Therefore, an advanced large-signal model which is able to predict distributed nonlinear effects is crucial for the successful design of high-frequency PAs. The microwave lumped and distributed layout parasitic effect in the 130 nm (BSIM3v3-RF) and 90 nm (BSIM4-RF) models to accurately predict gain, output power, and harmonic distortions of power MOSFETs at millimeter wave frequencies. The proposed power device model is verified for single devices as well as for the integrated power amplifier circuits in S-band and W-band applications. For S-band WiMAX application, we have developed an accurate modeling with layout parasitic of power CMOS devices and designed lossless matching networks to achieve single-end PA performance of 31 dB gain, 21.4 dBm output power, and 14.5% PAE at the 1 dB compression point. The measured maximum output power is 25.5 dBm and the associated PAE is 32%. For W-band application, a compact two-stage CMOS power amplifier is designed ii with gain boosting at the common gate transistor, source degeneration for the cascode devices and LC short stub matching networks. The amplifier was fabricated and demonstrated with excellent RF performance of 18 dB gain, 10.8 dBm linear output power, 13.3 dBm saturated power, and 11.8% PAE at 80 GHz with a minimum chip area of 0.35 mm2 in 90 nm CMOS technology. Monolithic power-combining techniques are attractive for delivering linear power over 20 dBm at W-band range due to the size reduction of the combiner. A W-band monolithic CPW Wilkinson power combiner of two CMOS power amplifiers is implemented in 90 nm CMOS technology. The 77 to 83 GHz CMOS PA achieved the 17 dB small-signal gain, 4.5 GHz 3 dB bandwidth, 10.6 dBm linear output power, 12.3 dBm saturated power, and 3.9% PAE at 80 GHz. iii To my family iv ACKNOWLEDGMENTS First, I would like to thank my adviser, Prof. Milton Feng for his guidance throughout my graduate school career and completion of this work. I greatly benefited from a research environment that allowed me to broaden my horizons, and the opportunity to work in the High Speed Integrated Circuits (HSIC) group has been a great privilege. I would like to thank Prof. Yun Chiu, Prof. Elyse Rosenbaum, and Prof. José Schutt-Ainé for serving on my doctoral committee and for their advice on this dissertation. I am grateful to Dr. Henry Pao for providing a Yunni Pao Family Fellowship in my graduate career. I would also like to thank Prof. Nick Holonyak Jr. for giving some space to work in his lab. Thanks to United Microelectronics Corp. (UMC) for providing the 130 nm and 90 nm CMOS foundry process for chip fabrication. Special thanks to Dr. Richard Chan for technical discussions throughout my career. Thanks to the past and present circuit design group members, Dr. Giang Nguyen, Mark Stuenkel, Kurt Cimino, Eric Iverson, Sean Graham, and Huiming Xu. Also, thanks to the processing group members, Dr. William Snodgrass, Dr. Han-Wui Then, Dr. Forest Dixon, Adam James, Wayne Wu, Donald Chen, Rohan Bambery, and Fei Tan. All these people have provided invaluable assistance and friendship. I am very appreciative of Prof. Milton Feng and Prof. José Schutt-Ainé for my first industry experiences and the opportunity to work at Xindium Technologies, Inc. I would like to thank Dr. David Caruth, Dr. Shyh-Chiang Shen, Dr. Gabriel Walter, Jeff Feng, and Aunt Thu for mentoring in my early industry and graduate career. Also, thanks to Dr. Ken Bower, Dr. Shahid Yousaf, Dr. Aleksandr Kavetskiy, Galena Yakubova, and Heather Socarras for the opportunity to work at Trace Photonics, Inc. and providing friendly working experiences. I am also grateful for the help of Mr. James Hutchinson and Mr. Jerome Colburn with v editing manuscripts throughout the journal and conference proceeding publication process. I would also like to thank Prof. Yan N. Lwin of Western Illinois University for his advice and guidance during my undergraduate and graduate career. Above all, I would like to thank my family and relatives for their endless support. The love and encouragement of my parents George and Maisie, my sisters and brother Harriet, Theresa, Richard, and Angel have made all my endeavors possible. I cannot express enough gratitude to my family, and I dedicate this dissertation to them. vi TABLE OF CONTENTS Page 1. INTRODUCTION .........................................................................................................1 1.1. Wireless Communication Technologies .................................................................1 1.2. Linear CMOS Power Amplifiers ............................................................................3 1.3. Organization of this Work ......................................................................................3 2. THEORY AND SPECIFICATION OF POWER AMPLIFIERS..................................6 2.1. Specifications of the Power Amplifier ...................................................................6 2.2. Classes of Linear Power Amplifiers .......................................................................7 3. POWER DEVICE MODELING FOR SCALING POWER MOSFET.........................9 3.1. High-Frequency Power MOSFET Issues ...............................................................9 3.2. Intrinsic and Extrinsic DC/RF Characteristics of a MOSFET Device .................11 3.3. Layout of Parasitic RC Lumped Modeling...........................................................16 4. DISTRIBUTED MODELING OF LAYOUT PARASITIC EFFECTS IN POWER DEVICE.......................................................................................................................21 4.1. Distributed Power Device Model (ICF-D1) .........................................................21 4.2. High-Frequency Scalable Distributed Modeling (ICF-D2)..................................31 5. A 2.5 GHZ CMOS POWER AMPLIFIER FOR WIMAX APPLICATION .............38 5.1. Power Amplifier Circuit Topology.......................................................................38 5.2. Power Inductor Model and Measurements ..........................................................39 5.3. Load-Pull and Source-Pull Simulation with Matching Networks Design............40 5.4. Single-End Power Amplifier Measurement Results ............................................43 6. An 80 to 85 GHz CMOS POWER AMPLIFIER FOR W-BAND APPLICATION ...48 6.1. Overview of W-Band Power Amplifier Design ...................................................48 6.2. A Two-Stage Power Amplifier Circuit Design ....................................................49 6.3. Wilkinson Power Amplifier..................................................................................61 7. CONCLUSIONS ........................................................................................................70 REFERENCES ..................................................................................................................72 AUTHOR’S BIOGRAPHY ...............................................................................................75 vii 1 INTRODUCTION 1.1 Wireless Communication Technologies Broadband wireless access fourth-generation communication systems (4G) enable innovations that take advantage of much higher data rates, allowing users to seamlessly reconnect to different networks even within the same session. Using 4G will in principle allow high-quality transmission systems such as Bluetooth (IEEE 802.15.4), ZigBee (IEEE 802.15.1), UWB (IEEE 802.15.3), WLAN/Wi-Fi (IEEE 802.11), and WiMAX (IEEE 802.16). These standards cover the range of distances illustrated in Fig. 1.1. The WiMAX is used in last mile wireless broadband access as an alternative to cable and DSL. It will work with other shorter-range wireless standards, including Wi-Fi, which has taken off as an easy way to provide Internet access throughout a home or business [1]. For wireless broadband communication, the power amplifier (PA) is a key part of the RF front-end in any transmitter (TX) system (Fig. 1.2). The PA is usually the last stage of the transmitter end and boosts the signal power high enough that it can propagate the required distance over the wireless medium. The 4G required for an optimal implementation of the complete system, including different digital and analog technologies, must be combined in a system-on-chip (SoC) integration for high-performance wireless. Today, almost all power amplifiers on the market are manufactured with III-V compound semiconductors because high output power and high power efficiency are required in various applications. Currently, there is no complete system for wireless communication with on-package integration because of the RF components. The power amplifier, RF filters, digital software-defined radio, and antenna are different technologies [2]. Therefore, an integrated analog power amplifier circuit using CMOS technology for cost-effective and spectrum-efficient high-speed wireless under the WiMAX 1 1-500Mbps PAN:UWB or Bluetooth 54Mbps WLAN: 802.11/ WiFi 10m 70Mbps WWAN: 802.16/ WiMAX 5km 100m Distance Mobile Wireless (2.5 GHz) 1-2 GHz L band 2-4 GHz S band Space Radar (77 GHz) 4-8 GHz C band 8-12 GHz X band 12-40 GHz Kau band 40-60 GHz U band 50-75 GHz V band 75-110 GHz W band Fig. 1.1. Wireless standard application distances. synth BB PA synth Fig. 1.2. Block diagram of wireless TX/RX system. standard is developed in this dissertation enabling fast local wireless connection to the network. Similarly, in W-band applications like phased array radar, wideband communication systems, and automotive sensors, a power amplifier is the key component in millimeter wave integrated circuits (MMICs). Usually, power amplifier MMICs with transistors in GaAs, InP, and HEMT technologies have been published with excellent results at W-band frequencies. Due to the high losses in silicon at millimeter wavelengths, when combining several monolithic power amplifier circuits in microstrip modules, it is desirable to obtain high output power from a single chip [3]. 2 1.2 Linear CMOS Power Amplifiers Wireless technology has created market demand for highly integrated circuits, such as a transmitter, receiver, and frequency synthesizer on a single chip. Silicon CMOS technology has made such integration possible with the exception of the power amplifier (PA), which is still typically implemented in non-CMOS technologies. Ideally, silicon CMOS PAs can be developed for tight integration with other wireless building blocks. Power amplifiers involve a balancing of many different parameters, including poweradded efficiency (PAE), linearity, maximum output power, maximum stable gain, input/output matching, stability, heat dissipation, and breakdown voltage [4]. As with many RF component designs, these requirements are often in conflict with one another. For example, achieving good linearity usually comes at a cost in PAE. Linearity is typically evaluated in terms of output third-order intercept (OIP3), 1 dB compression point (P1dB). Improved linearity is usually achieved by backing off an amplifier’s output power from its saturated output level; more DC power will be consumed in order to meet a given linearity requirement. Although many such trade-offs face a PA designer, amplifier circuits have been well researched over the years with many different design approaches. There are many interesting topologies at a designer’s disposal. Various methods have been used to divide and combine RF signals, such as on-chip transformers, distributed active transformers (DATs), and the Doherty, Balanced, and Wilkinson power combiners. In order to explore the possibilities of PAs fabricated with silicon CMOS, single-end CMOS PAs will be explored, followed by power-combining Wilkinson CMOS PAs. 1.3 Organization of this Work The focus of this work is to develop system-on-chip CMOS power amplifiers to be 3 used in WiMAX (S-band) and millimeter wave (W-band) transmitter applications. Therefore, we model and build an integrated power amplifier circuit using CMOS technology for cost-effective and system-on-chip integration for high-speed wireless in the WiMAX (2.5 GHz) and millimeter wave W-band (75 to 110 GHz) ranges. The major challenges in CMOS power amplifiers are nonlinear effects, power gain under stable operating conditions with minimum amplifier stages, and lossy passive networks on the silicon substrate that cause power loss at the transmitter output; furthermore, power amplifier design involves providing accurate active device modeling where CMOS device models are currently inaccurate. In Chapter 2, we will start with a brief background on the theory and specifications of linear power amplifiers. In Chapter 3, we will discuss BSIM-RF modelling in power devices. The theory of MOSFET device extrinsic and intrinsic parasitic characteristics and its smallsignal behavior will be explained. This leads to an Illinois Chan-Feng (ICF) model with the layout parasitic of power CMOS devices for 1 to 10 GHz application. Chapter 4 presents a distributed modelling in the high-frequency application. The effects of distributed power device model optimization and extraction techniques are incorporated to predict the current gain frequency (ft), power gain frequency (fmax), and transducer power gain and output power in 130 nm CMOS. The Illinois Chan-Feng Distributed lumped (ICF-D1) and scalable (ICF-D2) models are developed using small-signal and large-signal BSIM-RF. The model is verified by comparison with measured power device results. In Chapter 5, a 2.5 GHz CMOS power amplifier design for WiMAX application is discussed. The single-end PA design includes an input driver stage, a second output cascoded stage, and matching networks. The feedback of the first stage is designed to stabilize the gain. The impedance-matching networks between stages are designed to maximize linearity and power delivered to the load. We have designed lossless matching 4 networks to achieve better performance in a single-end PA. Experimental results are compared with the other published CMOS power amplifiers. A millimeter wave CMOS power amplifier design for W-band application is discussed in Chapter 6. A novel two-stage, 80 to 85GHz CMOS power amplifier is designed with gain boosting at the common gate transistor, source degeneration for cascode device and LC short stub matching networks. A Wilkinson power divider/combiner is discussed with its amplifier circuits. The two-stage W-band and Wilkinson amplifiers’ experimental data are reported and compared with the previous published results to validate the design methodologies. 5 2 THEORY AND SPECIFICATION OF POWER AMPLIFIERS 2.1 Specifications of the Power Amplifier In WiMAX application, the power amplifier is designed for a particular frequency and all the parameters are measured at that frequency. There are two different operating frequency ranges: 3.3 to 3.8 GHz (International) and 2.3 to 3.7 GHz (North America). The WiMAX application is focused on 2.5 GHz, 3.5 GHz, and 3.8 GHz [5]. 2.1.1 Frequency of 2.5 GHz is utilized. 2.1.2 Output power is the amount of power that needs to be delivered to the load. A WiMAX power amplifier requires high output power, approximately < 24 dBm, for long-distance applications [1]. 2.1.3 Efficiency of output power is defined as a measure of how efficiently the supply power is translated to output power. η= Power delivered to load Power drawn from the source (2.1) 2.1.4 Power added efficiency is a metric used commonly to compare PAs with different input power levels. PAE = Pout − Pin PDC (2.2) 2.1.5 Power gain/Voltage gain is the ratio of output power/voltage delivered to the load to the input power/voltage available from the source. The power gain will be equal to the voltage gain of the amplifier only if the input and output impedances are the same. 2.1.6 Linearity is an important metric of any power amplifier. It is desired that the amplifier operate with high linearity, i.e., that the output power scale linearly with input power. Compression of 1 dB and two-tone third-order intercept points are typically used to measure linearity in Fig. 2.1. 6 One decibel (dB) compression is the input power at which the linear gain of the amplifier has been compressed by 1 dB. The output referred 1 dB compression point (in dBm) would then be given by the sum of the input referred 1 dB point (in dBm) and the gain of the amplifier (in dB) Pout(dBm) Pout(dBm) POIP3 -- Ideal Pout-1dB 1dB -- Fundamental power -- IM3 Product power -- Actual Pd 1 3 Pin-1dB Pin(dBm) PIM PIIP3 Pin(dBm) (b) (a) Fig. 2.1 Nonlinear characteristics measurement. (a) 1 dB compression characteristics. (b) Two-tone third-order intersection point of power amplifier circuit. The two-tone third-order intercept point is defined as the point where the intermodulation (IM3) product power is equal to the fundamental power (PIM = Pd), and it is independent of the input power levels. Assuming two interferers very close to the desired frequency, a nonlinear output from the amplifier will generate inter-modulation products. The most important of the products is the third-order product since it falls directly in the frequency band of interest. This IM3 product term increases in amplitude on the order of the cube of the fundamental amplitude and, beyond a certain input power, it can be as significant as the fundamental tone. 2.2 Classes of Linear Power Amplifiers In classifying power amplifiers, the most widely used distinction is how a device handles the trade-off between linearity and efficiency. In amplifiers, the output amplitude of the signal is a linear function of the input amplitude. In Class A, B, and AB amplifiers, the output transistor acts as a current source, and the average output impedance during the operation is relatively high. The current and voltage waveforms through and across the 7 output device are often full or partial sinusoids. Class A, B, AB, and C power amplifiers depend on bias conditions which are shown in Fig. 2.2. ID(mA) Linear IMAX Class A Cutoff Saturation Class AB Class B VGS(V) Class C Fig. 2.2. Class A, B, AB, and C amplifiers with bias points. In Class A operation, the amplifier is always (100%) conducting current, which results in a maximum efficiency (η) of 50%. However, the linearity is excellent as it preserves the input and output waveforms without any distortion. In Class B operation, the bias is arranged to shut off the output device for half of every cycle so that the current conducted is 50% of the input. Therefore, power consumption is lower than that of the Class A type, while theoretical efficiency is approximately 78%. In Class AB operation, the amplifier conducts for 50% to 100% of a cycle, depending on the bias levels chosen. Good linearity can be achieved with devices in this regime, with efficiency intermediate between the Class A and B amplifiers (50% to 78%). In Class C operation, the gate bias is arranged to cause the transistor to conduct for 0% to 50% of a cycle, which is very low power consumption, and the efficiency can reach up to 100%. However, the output power levels will be unsuitably low for most applications. 8 3 POWER DEVICE MODELING FOR SCALING POWER MOSFET 3.1 High-Frequency Power MOSFET Issues Over the years, power amplifiers (PAs) have been the limiting components in RF CMOS transmitter integrated circuits (IC) due to the low breakdown voltages and nonlinearity problems of nano-MOSFETs. At high frequencies, the distributed effect and power device-scaling issues put other constraints on PA design such as the trade-off between output power (Pout) and power added efficiency (PAE). The foundry-provided BSIM3v3-RF model is unable to accurately predict the I-V characteristics and RF behaviors (ft and fmax) of power devices with widths of several hundred microns as in Fig 3.1. Therefore, an advanced large-signal model which is able to predict distributed nonlinear effects is crucial for the successful design of high-frequency PAs. Our proposed approach will use the microwave distributed effect in the 130 nm model to accurately predict output power and harmonic distortions of power MOSFETs at high frequencies. Device size L120nm x W7.2x16x1 (115 µm) Device size L120nm x W7.2x16x1 (1843 µm) 120 100 BSIM3v3-RF Measured 80 L 120 nm x W Vgs = 0.3 - 100 ft (GHz) ft (GHz) 120 L 120 nm x W Vgs = 0.3 - 60 40 BSIM3v3-RF Measured 80 60 40 20 20 0 0 1 10 100 Id/W (μA/μm) 1 1000 10 100 Id/W (μA/μm) 1000 Fig. 3.1. Layout model comparison of device size 115 µm and 1843 µm and its RF prediction and measured results for different devices. 9 MOSFET compact models are developed based on device characteristics at DC or low frequencies and therefore lack the robust non-quasi-static descriptions needed at high frequency. Recent modeling approaches, such as the BSIM3v3-RF model [6-7] shown in Fig. 3.2, for RF CMOS applications add lumped components to the DC core model to account for high-frequency extrinsic parasitics. Such approaches show reasonable DC/RF fitting results for MOSFETs with gate widths up to 100 microns. However, for a measured power device with a large gate width of 1843 µm (capable of Pout > 80 mW), there are significant discrepancies between the measured data and BSIM3v3-RF model results for output drain current IDS (over-prediction of ~25%) as well as for the current cut-off frequency ft and Ids (mA) maximum frequency of oscillation fmax at high current density (over-prediction of ~50%). 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 nMOS L=120 nm W=1843 µm Vg=0.4-0.8V BSIM-RF Measured 0 0.2 0.4 (a) 120 60 BSIM-RF Measured 40 80 60 20 0 0 10 1.2 40 20 1 1 nMOS L=120 nm W=1843 µm Vg=0.3-0.8 V Vd=0.3-1.2 V 100 fmax (GHz) ft (GHz) 80 0.8 (b) 120 nMOS L=120 nm W=1843 µm Vg=0.3-0.8 V Vd=0.3-1.2 V 100 0.6 Vds (V) 100 Id/W (μA/μm) 1000 BSIM-RF Measured 1 10 100 Id/W (μA/μm) 1000 (d) (c) Fig. 3.2. (a) Schematic of BSIM3v3-RF compact model. (b)–(d) Its DC / RF prediction and measured results for power device with a large gate width of 1843 µm. Clearly, in the high-frequency region, the effect of parasitics as well as other highfrequency mechanisms of the power devices become significant, and need to be modeled to 10 describe their dynamic performance accurately. To deliver the required output power, device widths have to be scaled up, and eventually can be comparable to the wavelengths of the signals in the high-frequency region. Hence, a distributed large-signal model is a must for PA design. In addition to the power device scaling issue, precise knowledge of nonlinear characteristics is also very important since the harmonic distortions determine the error of predicted PAE derived using maximally-flat-waveform analysis. 3.2 Intrinsic and Extrinsic DC/RF Characteristics of a MOSFET Device In theory, the DC I-V characteristics of the MOSFET device in the active and triode regions are defined using Eqs. (3.1) and (3.2). Active region Id = uC ox ⎛ W ⎞ 2 ⎜ ⎟(V gs − Vt ) 2 ⎝L⎠ Triode region Id = uCox ⎛ W ⎞ 2 ⎜ ⎟ 2(Vgs − Vt )Vds − Vds 2 ⎝L⎠ [ (3.1) ] (3.2) From the DC bias conditions, the small-signal transconductance, intrinsic input of the resistances, and capacitances can be determined. The transconductance follows from the standard definition of a MOSFET device as gm = uCox ⎛ W ⎞ ⎜ ⎟(Vgs − Vt ) 2 ⎝L⎠ The small-signal equivalent circuit of a MOSFET is shown in Fig. 3.3. (3.3) The intrinsic elements include gm, Cgs, Cgd, Cds, Rgs, and τi, which are functionally dependent on biasing conditions. The extrinsic elements include Lg, Rg, Cgp, Ls, Rs, Rd, Cdp, and Ld, which are all independent of the biasing conditions. 11 g m e jωτ Fig. 3.3. Determination of the small-signal intrinsic and extrinsic parasitic resistances, capacitances, and inductances. From the RF characterization, the input capacitance can be found from a determination of the maximum current gain frequency, f t , from S-parameter measurements. Current gain is extrapolated from |H21(ωt)| = 1, H 21 = 2 ⋅ S 21 S12 ⋅ S 21 + (1 − S11 ) ⋅ (1 + S 22 ) (3.4) The unity current gain cut-off frequency, ft, is determined from an extrapolation of H21 versus frequency. f t = freq ⋅ mag (H 21 ) (3.5) The small-signal current gain frequency relative to extrinsic and intrinsic MOSFET parameters [8-10] is as follows: ft = gm 2πC gs ⎧⎡⎛ Rd + R s ⎪ ⎨⎢⎜⎜1 + Rds ⎢ ⎝ ⎪⎩⎣ ⎞⎛ C dg + C gp + C dp ⎟⎟⎜1 + ⎜ C gs ⎠⎝ (3.6) 2 ⎞ C ⎤ ⎡ ⎛ ⎞C ⎤ ⎟ + g m (Rd + Rs ) dg ⎥ − ⎢ Rs + ⎜1 + Rs + g m Rs ⎟ dg ⎥ ⎜ R ⎟C ⎟ C R ⎢ ⎥ gs ds ds ⎝ ⎠ gs ⎥⎦ ⎠ ⎦ ⎣ 2 1/ 2 ⎫ ⎪ ⎬ ⎪⎭ For the case C gp = 0 and Cdp = 0 ft = gm 2πC gs ⎧⎡⎛ Rd + R s ⎪ ⎨⎢⎜⎜1 + Rds ⎪⎩⎣⎢⎝ ⎞⎛ C dg ⎟⎟⎜1 + ⎜ ⎠⎝ C gs (3.7) 2 ⎞ C ⎤ ⎡ ⎛ ⎞C ⎤ ⎟ + g m (Rd + Rs ) dg ⎥ − ⎢ Rs + ⎜1 + Rs + g m Rs ⎟ dg ⎥ ⎜ R ⎟C ⎟ C R ⎥ ⎣⎢ ds ⎝ gs ⎦ ds ⎠ gs ⎦⎥ ⎠ 12 2 1/ 2 ⎫ ⎪ ⎬ ⎪⎭ For the case Cgp = 0, Cdp = 0, and Rds = ∞ ft = gm 2πC gs (3.8) 1/ 2 2 2 ⎧⎡ C dg ⎤ ⎡ Rs C dg ⎤ ⎫⎪ ⎪ ( ) ( ) 1 1 g R R g R − + + + + ⎥ ⎬ ⎥ ⎢ ⎨⎢ m d s m s C gs ⎥⎦ ⎢⎣ Rds C gs ⎥⎦ ⎪ ⎪⎩⎢⎣ ⎭ For the first-order approximation with Cgp = 0, Cdp = 0, and Rds = ∞ C gs R 1 = + (Rd + Rs )C dg + s (1 + g m Rs )C dg 2πf t gm Rds (3.9) If the parasitic series resistances, inductance, and transmission lines are considered to have minimal effects at moderate frequencies (2 to 15 GHz), it follows that the high-frequency current gain is determined from a simple analysis of the remaining elements as follows: gm gm = 2π (C in ) 2π (C gs + C gd ) ft = (3.10) Similarly, the unilateral power gain can be calculated from the S-parameter measurement. The maximum frequency of oscillation of the MOSFET device is extrapolated at U(ωmax) = 1, U= 2 where k= S 21 S12 − 1 2 (3.11) 2 ⋅ k S 21 S12 − 2 ⋅ Re(S 21 S12 ) 2 1 − S11 − S 22 + S11 ⋅ S 22 − S12 ⋅ S 21 2 2 S12 ⋅ S 21 The maximum current gain frequency is determined from an extrapolation of U versus frequency. f max = freq ⋅ mag (U ) 13 (3.12) The maximum power gain is achieved when the device is complex conjugate matched at both the input and the output. For the simplified small-signal equivalent circuit, the maximum power gain is given by Gmax f max = R 1 g m2 = ⋅ 2 2 ⋅ out . 4 ω C in Rin Rout 1 Rout 1 gm ⋅ ⋅ = ⋅ ft ⋅ 2 2πC in 2 Rin Rin (3.13) The maximum intrinsic gm and ft occur just when the drain voltage is sufficient to saturate the electron velocity in Fig. 3.4, because the electric field limited velocity saturated current and also the mobility decrease at the high apply bias Vgs as shown in Fig. 3.5 (a). The gm degradation in Fig. 3.5 (b) is the effective channel mobility as a function of increasing transverse electric field across the gate oxide and the source-drain series resistances [11-12]. The ft follows the intrinsic behavior of the bias dependent and transient delays of both intrinsic and external resistances and capacitances (gm, Cgs, Cgd, and Rds) in Eq. 3.14. As gate voltage increases, the gate-to-drain capacitance, Cgd, decreases and the gate-to-source capacitance, Cgs, increases while the total capacitance (Cgs + Cgd) is dominated by Cgs. The drain source output resistance, Rds, is only a few times greater than Rd and Rs and cannot be neglected. Therefore, the parasitics effect modeling in large-scale power devices is important. ft = gm 2πC gs ⎧⎡⎛ Rd + R s ⎪ ⎨⎢⎜⎜1 + Rds ⎪⎩⎣⎢⎝ ⎞⎛ C dg + C gp + C dp ⎟⎟⎜1 + ⎜ C gs ⎠⎝ (3.14) 2 ⎞ ⎡ C ⎤ ⎛ ⎞C ⎤ ⎟ + g m (Rd + Rs ) dg ⎥ − ⎢ Rs + ⎜1 + Rs + g m Rs ⎟ dg ⎥ ⎜ ⎟C ⎟ C R R ⎥ gs ⎦ ds ⎠ gs ⎦⎥ ⎠ ⎣⎢ ds ⎝ C ⎛ L ⎞ ⎛ R + Rs ⎞ 1 ⎟⎟ + Cgd ⋅ (Rd + Rs ) + gs = ⎜⎜ g ⎟⎟ ⋅ ⎜⎜1 + d 14 4244 3 gm 2πft ⎝ ν sat ⎠ ⎝ R { τ RC 1444244ds43⎠ τ C gs τ Lg where g m = vsat (Cgs + Cgd ) Lg 14 2 1/ 2 ⎫ ⎪ ⎬ ⎪⎭ (3.15) 60 gm Measured f t ft ≈ 2πC gs Calculated ft 50 Measured C gs Measured g m 1.4 4.0 1.2 3.5 1.0 0.8 40 0.6 30 3.0 2.0 0.4 20 Vds =1.2 V Vgs =0.3-0.8 10 10 Jd (mA/mm) 1.5 0.2 0.0 100 2.5 Cgs (pF) Nonlinear transconductance and nonlinear stored charge g (mS) m ft (GHz) 70 1.0 Fig. 3.4. Nonlinear characteristics of the unity current gain (ft), the transconductance, and the gate source capacitance depend on applied gate source voltage. 700 600 g m (m S/m m ) 500 gm = ueff Ci ⎛ W ⎞ ⎜ ⎟(Vgs − Vt )ν sat 2 ⎝L⎠ Vds =1.2 ν ε c = sat μ eff Vds =0.9 Vds =0.7 V 400 300 Vds =0.5 V 200 ξ eff = 1 ⎛ 1 ⎞ ⎜ Qd + Qi ⎟ 3 ⎠ ε si ⎝ Vds =0.3 V 100 Vds =0.1 V 0 0.2 0.4 0.6 Vgs (V) 0.8 1.0 Fig. 3.5. (a) Measured hole mobility at 300 K and 77 K vs. effective normal field for several substrate doping concentrations [11]. (b) Measured characteristics of transconductance, gm, depend on applied bias voltage. 15 3.3 Layout of Parasitic RC Lumped Modeling For the power device modeling with the core of the BSIM3v3, we add an additional capacitor connected parallel with Cgs and a series gate resistor, which compensates the error between model and measured values in [13]. Also, a lumped resistance network, which includes a gate resistor, a drain-to-body series resistor and capacitor, and a source-to-body series resistor and capacitor, are used for GHz communication [14]. In our first modeling approach, we have developed the Illinois Chan-Feng (ICF) model by incorporating the external extrinsic parameters of overlap capacitances (Cgsx, Cdsx) and wire resistances (Rgx, Rdx, Rsx) into the BSIM3v3 for accurate DC I-V and RF linearity prediction of a large-width power CMOS device as shown in Fig. 3.6. L (nm) W (µm) Cgsx (pF) Cdsx (pF) Rgx (Ω) Rdx (Ω) Rsx (Ω) 340 1536 0.6 3.5 0.2 0.2 0.6 340 3840 1.5 10 0.2 0.2 0.3 Fig. 3.6. Illinois Chan-Feng power device model schematic. We employed a UMC 130 nm process and triple n-well devices with a wider gate length of 340 nm in this power amplifier design because the deep n-well contacts improve current-handling capabilities at the source and drain contacts. We chose devices with widths of 1536 µm and 3840 µm and characterized the S-parameters to model the external parameters of the ICF model. Compared to the measured data of an L 0.34 × W 1536 µm device, the BSIM3v3 model DC I-V results over-predict by ~20% at high current (250 mA), 16 as shown in Fig. 3.7 (b); however, the ICF model predicts measured data well over a wide bias range. 0.35 BSIM3 ICF Measure 0.3 Vgs = 1.6 V IDS (A) 0.25 0.2 Vgs = 1.2 V 0.15 0.1 Vgs = 0.8 V 0.05 L 340 nm x W 1536 μm 0 0 0.5 (a) 1 1.5 2 2.5 VDS (V) 3 (b) Fig. 3.7. (a) Power device L 0.34 × W 1536 µm. (b) DC I-V of the driver stage power device for ICF and BSIM3v3 models, and measurement. 30 20 50 40 15 10 30 Vds = 3.3V Vgs = 0.4-1.6 V 20 5 0 10-5 L 340 nm x W 1536 μm 10-4 10-3 10-2 10-1 fmax (GHz) ft (GHz) 25 60 ft Meas ft ICF ft BSIM3v3-RF fmax Meas fmax ICF fmax BSIM3v3-RF 10 0 100 J (mA/μm) Fig. 3.8. Frequencies ft and fmax of the driver stage power device for ICF and BSIM3v3 model simulations, and measurement. For RF comparison, ft and fmax are extracted at 5 GHz in Fig. 3.8. The BSIM3v3 model’s ft and fmax over-predict by ~20% and 50%, respectively, at the design current, while the ICF model closely agrees with the measured RF results. Similarly, the DC and RF comparisons of the larger gate width device 3840 µm are shown in Fig. 3.9 (a) and (b), respectively. 17 0.7 BSIM3 ICF Measure 0.6 L 340 nm x W 3840 μm Vgs = 1.2 V IDS (A) 0.5 0.4 Vgs = 1 V 0.3 0.2 Vgs = 0.6 V 0.1 0 0 0.5 1 1.5 VDS (V) 2 2.5 3 (a) 100 30 20 15 10 80 60 Vds = 3.3V Vgs = 0.4-1.4 V 40 20 5 0 10-3 fmax (GHz) ft (GHz) 25 ft Meas ft ICF ft BSIM3v3-RF fmax Meas fmax ICF fmax BSIM3v3-RF L 340 nm x W 3840 μm 10-2 10-1 100 0 101 J (mA/μm) (b) Fig. 3.9. (a) DC I-V of the output stage power device for ICF and BSIM3v3 models, and measurement. (b) Frequencies ft and fmax of the output stage power device for ICF and BSIM3v3 model simulations, and measurement. A device nonlinearity model can be analyzed with the Volterra series representation. The Volterra transfer functions clearly bring out the frequency-dependent nature of transistor distortion [15-17]. The third-order Volterra kernel appearing in Eq. (3.16) is shown where the intermodulation distortion of various mixing frequency κ 1 = ω1 − ω 2 , κ 2 = 2ω1 , and κ 3 = 2ω1 − ω2 , the corresponding output-intercept point is i ds = G1 (ω a ) o v s + G1 (ω a , ω b ) o v s2 + G1 (ω a , ω b , ω d ) o v s3 OIP3 (2ω1 − ω 2 ) = G1 (ω1 ) 3 ⋅ 4 G1 (ω1 , ω1 − ω 2 ) 1 / 2 (3.16) 3/ 2 18 (3.17) The second derivative of the ft versus Id curve is related to third-order distortion (HD3)—i.e., the more linear ft versus Id, the smaller the third-order distortion—the high-frequency distortion at a DC operating point (Id, Vgs) can be written as depending on the ft,: 10 −3 ∂ 2ω ⎡ ω '' ⎤ and ωT'' = 2 T HD3 = 20 log10 ⎢ g T ⎥ , where g = 2 RL ∂I DS ⎣ 6 ωT ⎦ (3.18) The calculated HD3 versus measured results are shown in Fig. 3.10 (a). The output intercept point, OIP3, can be written as a function of ft and estimated from fmax [12]: OIP3 = OIP3 ≈ -80 L 340 nm x W 1536 μm 12 ωT'' -105 -110 BSIM3v3-RF ICF Measure 8 7 6 OIP3 = 5 4 8ωT 12 ωT'' ω T' = 0 f OIP3 ≈ 2 max ∂ f max 2 ∂I DS 3 2 10 −3 ∂ 2ω and ωT'' = 2 T 2 RL ∂I DS 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 L 340 nm x W 1536 μm 9 ⎡ ω '' ⎤ HD3 = 20 log10 ⎢ g T ⎥ ⎣ 6 ωT ⎦ where g = (3.20) 10 OIP3 (dBm) -100 ωT' = 0 f max ∂ 2 f max 2 ∂I DS -90 -95 (3.19) BSIM3v3-RF ICF Measure -85 HD3 (dB) 8ωT 1 0 1 0 Id (A) 0.1 0.2 0.3 Id (A) 0.4 0.5 (a) (b) Fig. 3.10. (a) HD3 and (b) OIP3 of BSIM3v3 model and ICF model simulations, and measurement. The output intercept characteristic of an nMOS device with a gate length of 340 nm and 1536 µm width calculated using BSIM3v3 modeled device parameters is compared to that using measured device data in Fig. 3.10 (b). The complex bias dependence of OIP3 (2ω1 − ω 2 ) , including the occurrence of distinct peaks, is predicted for the magnitude and phase of device. The BSIM3v3’s scalability, linearity, and bias current estimation are 19 very inaccurate for wider-power devices, while the ICF model more closely matches measurement results. 20 4 DISTRIBUTED MODELING OF LAYOUT PARASITIC EFFECTS IN POWER DEVICE 4.1 Distributed Power Device Model (ICF-D1) A full illustration of the proposed physical layout of the Illinois Chan-Feng Distributed (ICF-D1) model, is given in Fig. 4.1. At high frequencies, both the parasitics and the distributed nature of this large-size power device layout are significant. The foundryprovided BSIM3v3-RF large-signal model fits well only for small devices with gate width approximately < 150 µm. Therefore, we have developed the ICF-D1 model by incorporating external distributed parameters into the BSIM3v3-RF model. In the ICF-D1 model, several sections of lumped components corresponding to each unit cell of a fixed number of fingers are proposed to effectively describe the distributed effects. The transistor is then separated into extrinsic and intrinsic contributions. The intrinsic elements of each unit cell depend on the bias conditions and geometry of the active area of the device, and thus they are scalable. Figures 4.1 and 4.2 show the perspective 3D and 2D layouts of a MOSFET, with circuit elements indicating the parasitic inductances and capacitances due to the input and output manifolds. These inductances and capacitances model the transmission line behavior of the manifolds as well as metallization capacitances between the three terminals of the device. The resistances of the input and output manifolds, which consist of thick metal, are negligible compared to the gate, source, and drain resistances of the MOSFET. There are additional extrinsic resistances and capacitances that vary with unit width to account for high-frequency parasitics, which are also scalable (Fig. 4.3). The outermost extrinsic inductances and capacitances, due to the input and output manifolds of unit-cell combining structures, are then added to complete the distributed model (Fig. 4.4). These structures are typically fixed, and hence are non-scalable. Devices of a 120 nm RF CMOS 21 process have been thoroughly characterized and simulated using ICF-D1 models to illustrate the model robustness. S Ls Cgsx Cdsx G M CdsixNF Cdsix Lg Rs Rg Rs Rd Cgsix Cgdix Rg L Rd W Cgsix Cgdix Cdgx Ld D Wtot=W x NF x M Fig. 4.1. 3D distributed physical layout and schematic view. Fig. 4.2. Layout of large-size power devices. Fig. 4.3. Single lump section of the distributed ICF-D1 model. Fig. 4.4. Schematic of the distributed ICF-D1 model. 22 The inductances associated with the input and output manifold transmission lines are defined as Lg, Ld, and Ls respectively. The capacitances associated with the manifolds are lumped together with the gate-source, gate-drain and drain-source metallization capacitances and are modeled as Cgsx, Cgdx, and Cdsx. The extrinsic values of Lg, Ld, Ls, Cgsx, Cgdx, and Cdsx will remain constant versus unit width for a fixed number of gate fingers. Figure 4.3 illustrates a “unit-MOSFET” cell. A single-gate unit cell consists of several parallel gate fingers. Figures 4.2 and 4.4 show parallel unit cells connected together to form a distributed power device. The device’s total width is equal to the product of the unit finger width, W, the number of gate fingers, NF, and the multiple cells, M, in Figs. 4.1 and 4.2. 4.1.1 Extrinsic inductances and scalable extrinsic resistances extraction In order to determine the scalable extrinsic contributions of the gate-source, gatedrain, and drain-source capacitances, inductances, and resistances, it is reasonable to accurately de-embed the non-scalable extrinsic MOSFET. At higher frequencies, the transmission line extrinsic inductor can be determined by measuring the S-parameters of the MOSFET versus frequency under the bias conductions Vds = 0 V and Vgs > Vbi, where Vbi is the Shottky diode forward-biased turn-on voltage. Since the MOSFET has no small-signal gain at Vds = 0 V, this measurement is called a “cold-FET” measurement [18]. To determine the frequency dependence of the elements in T-topology, such as the circuit in Fig. 4.5, only cold z-parameters of first-order terms are considered. The inductors Lg, Ld, and Ls can be calculated from the imaginary part of the z-parameters at high frequencies using the following equations: Z11 ≅ R gx + Rsx + nkT + jω (Lg + Ls ) qI g (4.1) Z12 ≅ Z 21 ≅ Rsx + jωLs (4.2) Z 22 ≅ Rdx + Rsx + jω (Ld + Ls ) (4.3) 23 Fig. 4.5. Simplified schematics of extrinsic parasitic using T-model extraction at high frequency. The resistances Rg, Rd, and Rs can be calculated from the real part of the z-parameters. The scalable distributed resistance (Rgx, Rdx, and Rsx) Rix is based on the physical layout structure and scales as Rix = Ri ⋅ Wtot with parallel connections from the top of the thick metal layers and via holes to the bottom of the active device area. The measured resistances scale in a directly proportional manner to the inverse of total device width, as shown in Fig. 4.6. From the slopes of the lines, the gate, drain, and source resistances values are Rg = 0.335 Ω*mm, Rd = 0.308 Ω*mm, and Rs = 0.41 Ω*mm. 3.5 Rg Rd Rs Resistance (Ω) 3 2.5 2 1.5 nMOS L=130nm Vds=0V Vgs=0.8V 1 0.5 0 0 0.002 0.004 0.006 0.008 0.01 1/Wtot (µm) Fig. 4.6. Measured extrinsic resistances vs. inverse of total width for three devices. 24 4.1.2 Scalable and non-scalable extrinsic capacitances extraction To determine the extrinsic contributions of the gate-source, gate-drain, and drainsource manifold and metallization capacitances, the total capacitances measured between the terminals of the devices, Cgst, Cgdt, and Cdst, can be separated into those which are constant versus the total gate width, Wtot (Cgsx, Cgdx, and Cdsx), and those which scale proportionally to Wtot (Cgs, Cgd, and Cds ). The three total capacitances can be determined by measuring the S-parameters of the MOSFET versus frequency under the bias conditions, Vds = 0 V and Vbd < Vgs < Vp , where Vbd is the reverse-biased diode breakdown voltage and Vp is the channel pinch-off voltage. At low frequencies, the impedances of the inductances Lg, Ld, and Ls and resistances Rgx, Rdx, and Rsx will be negligible compared to the impedance of the capacitances. The simplified capacitor π topology is shown in Fig. 4.7 and total extrinsic capacitances are extracted at Vgs = –0.5 V with the following equations: Im (Y11 + Y12 ) = ω (C gsixWtot + C gsx ) = ωC gst (4.4) Im(Y12 ) = −ω (C gdixWtot + C gdx ) = −ωC gdt (4.5) Im(Y22 + Y12 ) = ω (C dsixWtot + C dsx ) = ωC dst (4.6) Fig. 4.7. Simplified schematic of extrinsic parasitic using π model extraction at low frequency. 25 Capicatance (pF) 4 3.5 Cgst (pF) Cgdt (pF) Cdst (pF) 3 2.5 nMOS L=130nm Vds=0V Vgs=-0.5V 2 1.5 1 Cdst = 0.0019Wtot + 0.062 Cgst = 0.0005Wtot + 0.0357 0.5 Cgdt = 0.0004Wtot - 0.0094 0 0 500 1000 1500 2000 Wtot (µm) Fig. 4.8. Measured extrinsic capacitance vs. total width for three devices. The three total capacitances Cgst, Cgdt, and Cdst of each FET are directly proportional to total device widths in Fig. 4.8. As shown in Eqs. (4.4)–(4.6), the three y-intercepts of large-width PA devices determine the parasitic manifold capacitances Cgsx = 36 fF, Cgdx = 9.4 fF, and Cdsx = 62 fF, which are non-scalable extrinsic capacitances. The slopes of the lines are equal to the normalized scalable extrinsic capacitances along the total width of the finger. The three scalable capacitor values are Cgsix = 0.5 pF/mm, Cgdix = 0.4 pF/mm, and Cdsix = 1.9 pF/mm. 4.1.3 Experimental validation High-frequency on-wafer SOLT calibration measurement is carried out with an E8364 network analyzer and 4142 DC supply. On-wafer SOLT calibration standards are used in the DC and RF measurements, which allows the measurement reference planes to be shifted to inside the test set, past the probe tips. The MOSFET devices are measured from 0.5 to 40 GHz with 0.1 GHz steps. The ICF-D1 model is optimized in the Agilent Design System (ADS). We chose devices with widths of 115 µm, 921 µm, and 1843 µm and characterized the S-parameters to model the external parameters of the ICF-D1 model. Detailed DC-IV characteristic results are measured for a 120 nm MOSFET (W = 1843 µm) in Fig.4.9. Compared to the BSIM3v3-RF model, the ICF-D1 model accurately predicts the DC 26 I-V curves with less than 2% error. From small-signal S-parameter measurements, ft and fmax predictions are shown in Figs. 4.10 and 4.11. The accuracy of the ICF-D1 model is within Ids (mA) 10% across the bias points [19]. 0.5 0.45 0.4 0.35 0.3 0.25 0.2 0.15 0.1 0.05 0 nMOS L=120 nm W=1843 µm Vg=0.4-0.8V BSIM-RF ICF-D1 Measured 0 0.2 0.4 0.6 Vds (V) 0.8 1 1.2 Fig. 4.9. DC power device I-V characteristics for BSIM3v3-RF and ICF-D1 models vs. measured results. 120 nMOS L=120 nm W=1843 µm Vg=0.3-0.8 V Vd=0.3-1.2 V BSIM-RF ICF-D1 Measured ft (GHz) 100 80 60 40 Vd=1.2 Vd=0.9 Vd=0.6 20 Vd=0.3 0 1 10 100 Id/W (μA/μm) 1000 Fig. 4.10. Power device ft characteristics for BSIM3v3-RF and ICF-D1 models vs. measured results. 27 120 fmax (GHz) 100 80 nMOS L=120 nm W=1843 µm Vg=0.3-0.8 V Vd=0.3-1.2 V 60 Vd=1.2 40 Vd=0.9 BSIM-RF ICF-D1 Measured 20 0 1 10 100 Id/W (μA/μm) Vd=0.6 Vd=0.3 1000 Fig. 4.11. Power device fmax characteristics for BSIM3v3-RF and ICF-D1 models vs. measured results. For large-signal model verification, an Agilent E8364 network analyzer was used to sweep the input power to the device from –25 to 0 dBm in 1 dB steps at fo = 3.5 GHz. The output powers into a 50 Ω load at fo and its harmonics were measured using an HP 8565E spectrum analyzer. The loss between the sweeper and the device input at 3.5 GHz and the loss between the device output and the spectrum analyzer at 3.5 GHz and its harmonics were determined and removed from the measurements. A one-tone harmonic balance simulation was performed at the same bias points using the BSIM3v3-RF and ICF-D1 models for a device of length 120 nm and width 1843 µm. Figure 4.12 shows the measured and simulated one-tone results at class A bias point at 3.5 GHz. The power device transducer power gain comparison is shown in Fig. 4.13. The BSIM3v3-RF model over-predicts by ~ 1.5 dB, and the ICF-D1 agrees with measurement. 28 7.5 Gain (dB) 7 nMOS L=120 nm W=1843 µm 6.5 6 5.5 5 Vg=0.6 V Vd=1.2 V freq=3.5 GHz BSIM-RF ICF-D1 Measured 4.5 4 -25 -20 -15 -10 -5 0 5 10 Pout (dBm) Fig. 4.12. BSIM3v3-RF and ICF-D1 models, and measured results, of transducer power gain vs. output power. At 0 dBm input power, the output power of the measured device shows 4.77 dBm (~3 mW), the BSIM3v3-RF model shows 6.91 dBm (~ 4.91 mW), and the ICF-D1 model shows 5.34 dBm (~ 3.41 mW) in Fig. 4.13. The BSIM3v3-RF model over predicted an ~ 2.4 dBm error in the fundamental output power. Therefore, the BSIM3v3-RF model error is ~ 63% and the ICF-D1 model error is ~13% compared to measured results. L=120 nm W=1843 µm f0 Vg=0.6 V Vd=1.2 V freq=3.5 GHz f2 -20 -80 f3 -100 -12 -16 -20 -24 -28 -120 BSIM-RF ICF-D1 Measured 0 -60 -4 -40 -8 Pout (dBm) 0 Pin (dBm) Fig. 4.13. BSIM3v3-RF and ICF-D1 models, and measured results, of output power harmonics characteristics of power devices (L = 120 nm). 29 Similarly, a device gate length 340 nm comparison of BSIM3v3-RF, ICF, and ICF-D1 models and measurement results are shown in Fig. 4.14. The BSIM3v3-RF model over predicted for ~ 5 dBm and the ICF-D1 model is close to measurement results. The ICF-D1 model predicts well for fundamental and second harmonics compared to BSIM3v3-RF. However, the ICF-D1 model did not predict well for the third-order harmonics because the model did not modify intrinsic parts of the small-signal BSIM3v3-RF model. 20 BSIM ICF ICF-D1 Measured 10 0 -10 f0 Pout (dBm) -20 -30 -40 f2 -50 -60 -70 nMOS L=340 nm W=1538 µm Vg=1 V Vd=3.3 V -80 -90 f3 -100 0 -2 -4 -6 -8 -10 -12 -14 -16 -18 -20 -22 -24 -26 -28 -110 Pin (dBm) Fig. 4.14. BSIM3v3-RF, ICF-L, and ICF-D1 models, and measured results, of output power harmonic characteristics of power devices (L = 340 nm, freq = 2.5 GHz). 30 4.2 High-Frequency Scalable Distributed Modeling (ICF-D2) MOSFET compact models are developed based on the device characteristics at DC or low frequencies. The extrinsic layout parasitic effect is very important for high frequency RF MOSFET applications. The scalable ICF-D2 distributed layout parasitic effects model is extended for W-band amplifier design. The device test structure is shown in Fig. 4.15. The ICF-D2 model is implemented with the BSIM4-RF model in Fig. 4.15 with scalable and nonscalable parts. The distributed model of multiple unit cell connection is shown in Fig. 4.16. Fig. 4.15. A 90 nm MOSFET device (W= 64 µm) and a single unit of the ICF-D2 model integrated with the BSIM4-RF model. Fig. 4.16. High-frequency distributed ICF-D2 model integrated with the BSIM4-RF model. 31 The device’s total width is equal to the product of the unit finger width, W, the number of gate fingers, NF, and the multiple cells, M. The parallel multiple unit cells, M, are connected together to form a high-frequency distributed power device. A single-gate unit cell consists of several parallel gate fingers. It is separated into the extrinsic device cell which is scalable with the device’s unit width (UW) and finger (F). The extrinsic schematics of the capacitances and resistances (Cgsix, Cgdix, Cdsix, Rg, Rd, and Rs) are scalable with the UW and F in Fig. 4.16. The extrinsic pad capacitances (Cgsx, Cgdx, and Cdsx) are scalable with the finger and the extrinsic inductances (Lg, Ld, and Ls) are nonscalable with device parameters. The model parameter equations are the following: C gsix = (C gsi ⋅ 0.05 ⋅ (UW ⋅ F )) fF (4.7) C gdix = (C gdi ⋅ 0.05 ⋅ (UW ⋅ F )) fF (4.8) C dsix = (C dsi ⋅ 0.05 ⋅ (UW ⋅ F ) + 10 ) fF (4.9) C gsx = (C gsp ⋅ (F 2 )) fF (4.10) C gdx = (C gdp ⋅ (F 2 )) fF (4.11) C dsx = (C dsp ⋅ (F 2 )) fF (4.12) R g = (⋅ R gi ⋅ UW (F )) Ω (4.13) R s = (R si (UW ⋅ F )) Ω (4.14) R d = (R di (UW ⋅ F )) Ω (4.15) 4.2.1 Experimental model verification from 1 to 50 GHz, V-band, and Wband We set up an automatic DC and RF measurement program in Agilent VEE for data collection. The power devices are measured for S-parameters from 1 to 50 GHz, V-band (50 32 GHz to 75 GHz), and W-band (75 GHz to 110 GHz) with the vector network analyzer (VNA 8510C) and the synthesized sweepers (HP 83651A and HP83621A) as in Figs. 4.17 and 4.18. HP 83621A 45 MHz to 20 GHz Synthesized sweeper HP 8510C Vector network analyzer LO RF HP 83651A 45 MHz to 50 GHz Synthesized sweeper LO RF HP 85106A Millimeter wave controller LO HP 85104A 1-50 GHz 50-75 GHz 75-110 GHz mm wave test set HP 85104A 1-50 GHz 50-75 GHz 75-110 GHz mm wave test set 1-mm cable DC HP 4142B Modular DC source RF GSG / V-band / W-band probe LO DC 1-mm cable RF Device under test GSG / V-band / W-band probe Fig. 4.17. S-parameters measurement setup (1 to 50 GHz, V-band, and W-band). 8510C Mixer Mixer Fig. 4.18. S-parameters measurement setup (1 to 50 GHz, V-band, and W-band). Calibration for VNA can be carried out with the “on-wafer” SOLT calibration standard which is on the same substrate as in Fig. 4.19. The standards have identical pad structures to the devices measured, which allows the measurement reference planes to be shifted from inside the S-parameter test set in Fig. 4.20. The coplanar waveguide (CPW) transmission line (L = 400 µm) in Fig. 4.21 is measured at three separate frequencies bands for the 33 verification of the calibration and measurement systems. The CPW insertion loss is less than –1 dB and isolation is better than –15 dB in Fig. 4.22. Short Open Load Thru Fig. 4.19. On-wafer calibration standards. Measurement Reference Planes Fig. 4.20. Measurement reference planes after on-wafer calibration. Fig. 4.21. CPW test structure (W = 10 µm, G = 5 µm, L = 400 µm). The resulting S-parameters are created for the distributed two-port models of the power devices and transmission lines. High-frequency distributed model parameters are measured and extracted in physically based intrinsic and extrinsic parameters from 1 to 110 GHz. The BSIM4-RF model, ICF-D2 model, and measurement comparison is shown in Fig. 4.23. The BSIM4-RF model over predicted the 2.5 dB gain compared to measured S-parameters. The ICF-D2 model agrees with measured S-parameters for all three band (1 to 50 GHz, V-band, and W-band) measurements. 34 0.0 -10 -15 -0.2 S12 and S21 (dB) S11 and S22 (dB) -20 -25 -30 -35 -0.4 -0.6 -0.8 -1.0 -40 -1.2 -45 freq (Hz) 0 (0.5-50 GHz) S12 and S21 (degree) -20 S11 and S22 (degree) 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 freq (Hz) 200 100 0 -100 V band (50-75 GHz) -40 -60 W band (75-110 GHz) -80 -100 -120 -140 -200 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 freq (Hz) freq (Hz) Fig. 4.22. S-parameter measurement of coplanar waveguide transmission line (400 µm) after on-wafer calibration in three separate frequency bands. 0.5 -10 Vd = 1.2V Vg = 1 V 0.0 -0.5 BSIM4-RF model ICF-D2 model Measured -1.5 -2.0 -2.5 -25 -30 -35 -3.0 -40 -3.5 -45 -4.0 -50 S22 (dB) 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 Measured -5.5 Vd = 1.2V Vg = 1 V 0 4.0E10 -5.0 4 -2 Vd = 1.2V BSIM4-RF model Vg = 1 V ICF-D2 model -4.5 6 2 3.0E10 8 2.0E10 10 1.0E10 12 freq (Hz) -4.0 BSIM4-RF model ICF-D2 model Measured BSIM4-RF model ICF-D2 model Measured Vd = 1.2V Vg = 1 V 0.0 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 freq (Hz) 14 S21 (dB) -20 S12 (dB) S11 (dB) -1.0 -15 -6.0 -6.5 -7.0 -7.5 -8.0 -4 -8.5 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 1.1E11 1.0E11 9.0E10 8.0E10 7.0E10 6.0E10 5.0E10 4.0E10 3.0E10 2.0E10 1.0E10 0.0 freq (Hz) freq (Hz) Fig. 4.23. Comparison of BSIM4-RF model, ICF-D2 model, and measured S-parameters of device size (L 90 nm x W 2 x NF 32 x M1) 35 The measured DC-IV curve comparison plot is shown in Fig. 4.24 and measured results agree with both BSIM4-RF and ICF-D2 models. For high-frequency amplifier operation, the RF characteristic of the device models is significant. The measured maximum ft = 90 GHz, and maximum fmax = 120 GHz, as shown in Figs. 4.25 and 4.26. The BSIM4-RF model prediction is ft ~ 130 GHz, which is >30% of the measured ft = 90 GHz. The BSIM4RF model predicts fmax = 220 GHz, which is 90% off the measured fmax = 120 GHz. On the other hand, the ICF-D2 model predicts well on both ft and fmax, within 10% across the biased current range, as shown in Fig. 4.25 and 4.26 respectively. 25 nMOS L=90 nm W=64 µm Ids (mA) 20 Vg=0.4-1 V BSIM-RF ICF-D2 Measured 15 10 5 0 0 0.2 0.4 0.6 Vds (V) 0.8 1 1.2 Fig. 4.24. DC-IV comparison of the BSIM4-RF model, ICF-D2 model, and measured device characteristics. 140 120 ft (GHz) 100 80 nMOS L=90 nm W=64 µm Vg=0-1 V Vd=0.3-1.2 V Vd 60 40 BSIM-RF ICF-D2 Measured 20 0 0.001 0.01 0.1 Id/W (mA/μm) 1 Fig. 4.25. BSIM4-RF and ICF-D2 models vs. measured results of the ft characteristic. 36 240 fmax (GHz) 200 160 120 nMOS L=90 nm W=64 µm Vg=0-1 V Vd=0.3-1.2 V Vd 80 BSIM-RF ICF-D2 Measured 40 0 0.001 0.01 0.1 1 Id/W (mA/μm) Fig. 4.26. BSIM4-RF and ICF-D2 models vs. measured results of the fmax characteristic. The measured 90 nm devices are summarized in Table 4.1. The measured threshold voltage is 0.38 V, subthreshold slope (S) is 108 mV/dec, drain induced barrier lowering (DIBL) is 133 mV/V, and Ion/Ioff is ~ 4 x 103. Those DC characteristics are approximately the same for the different gate width devices. However, for the RF characteristics, ft increases with device unit gate width and fmax decreases with total device gate width 256 µm. For MMIC power amplifier design, we can choose different size devices depending on the measured RF characteristics of the current gain, ft, and power gain, fmax, for the driver gain stage and the output power stage. TABLE 4.1 Summary of 90 nm device mesured DC/RF characteristics Device Vth (V) S (mV/dec) DIBL (mV/V) Ion/Ioff ft (GHz) fmax (GHz) L90W2x32x1 (64 µm) 0.38 108.2 133 4407 95 120 L90W4x32x1(128 µm) 0.38 107.4 133 4652 120 120 L90W8x32x1(256 µm) 0.38 110.4 133 4299 145 90 L90W2x32x2(128 µm) 0.38 110.9 133 3862 100 140 37 5 A 2.5 GHZ CMOS POWER AMPLIFIER FOR WIMAX APPLICATION 5.1 Power Amplifier Circuit Topology Class AB single-end power amplifier driver and output stages are designed for a stable input with high output power. In this frequency range, a CMOS common-source driver stage is unstable, and an RC feedback network, shown schematically in Fig. 5.1, is added to get stable inputs. The RC feedback network is optimized to provide unconditionally stable device operation at the design frequency. We use ADS simulation to find the resistance and capacitance of the feedback network that are maximized for higher gain and stable operation. The driver-stage power device uses a W = 1536 µm nMOS, and its drain current is approximately 95 mA. Γs1 Γin1 Γs 2 Γin 2 Γout1 Γl1 Γout 2Γl 2 Fig. 5.1. Schematic of driver and output cascoded stages. The output stage uses the cascode device topology to increase the gain, power, and breakdown voltage of the output power device. Since the cascoded amplifier stage is stable, an RC feedback network is not required. The W = 3840 µm nMOS is used for delivering the required output power, and its bias current is 200 mA. The driver and output stages consume 0.98 W of power from a 3.3 V DC supply. 38 5.2 Power Inductor Model and Measurements Custom power inductors are designed for the drain voltage supply, inter-stage matching, and output stage matching networks. The power inductors are key components of high-performance PAs. High Q values are optimized at design frequency, and low DC resistance is required in order to minimize supply drop. This can be achieved by increasing the metal width and using parallel multi-layer metals to increase thickness. An ADS momentum simulation is used to design the maximum width of the top three metal layers at 20 μm for the passive device model. Inductance and quality factors are calculated from Sparameter measurements using the following equations: Z in = 50 L= Q= (1 + Γin ) (1 − Γin ) Im(Z in ) ω Im(Z in ) Re(Z in ) (5.1) (5.2) (5.3) The thick metal inductor model is shown in blue, and the measurement results are shown red in Fig. 5.2. The inductor’s model and measured calculation values at 2.5 GHz are shown in Table 5.1. The inductance percentage error is 0.13% to 7.87% and the quality factor percentage error is 6.5% to 20%. The measured quality factor data are higher than the momentum model simulation at design frequency. TABLE 5.1 Summary of inductor model and measurement at 2.5 GHz Lmodel (nH) Lmeas(nH) Qmodel Qmeas 0.95 1.02 9.1 12.99 1.56 1.66 10.32 15.5 1.84 1.83 11.02 13.92 2.36 2.28 10.77 11.14 39 3.4 12 3.2 10 3 2.8 6 2.6 4 2.4 Q 8 Q mod Q meas Lmod Lmeas 2 (a) L (nH) 14 2.2 0 2 1 2 3 4 5 6 freq (GHz) (b) Fig. 5.2. (a) Inductor 2.2 nH. (b) Measured and model comparison of 2.25 nH inductor’s quality factor and inductance. 5.3 Load-Pull and Source-Pull Simulation with Matching Networks Design The purpose of using matching in the amplifier design is to obtain the highest gain and return loss at the design frequency. If there is mismatch between the amplifier source and input, some of the power will be reflected back to the source; therefore, the input signal power will not be amplified. Also, mismatch between the load and output of a power amplifier causes gain reduction. Therefore, a simultaneous complex conjugate match is obtained on both sides of the transistor to get maximum gain. The optimum load for a power amplifier is different from the maximum conjugate gain load for a small-signal amplifier. Real loads are rarely purely resistive, and the effect of variable load impedance on power delivered will vary. Load impedance at the fundamental frequency can be swept with its amplitude and phase. The output cascoded-stage transistors’ output reflection coefficients are characterized through load-pull analysis to select impedance for the maximum linear power delivered to the load at 2.5 GHz. Using 15 dBm input power with the different load impedances, the contour plots for maximum delivered power of 26.7 to 24.6 dBm are shown as thin dark blue lines in Fig. 5.3 (a). From load-pull optimization, the complex conjugate of the output cascoded-stage impedance (Zout2 = 9.19 – j12.92) is 40 matched to the source (ZO = 50 Ω) in Fig. 5.3 (b) using a ZY Smith chart. One possible solution for the matching network is shown in Fig. 5.3 (b) with series C and shunt L networks. The motion is in a counterclockwise direction along a constant-resistance circle zC = j0.41, and a constant-conductance circle gives the shunt inductor admittance (yL = –j2.12). From the normalized impedance and admittance values at 2.5 GHz, the capacitance (3.1 pF) and the inductance (1.51 nH) are plotted with thick dark gray and thick bright pink curves, respectively, in Fig. 5.3 (a). Similarly to interstage output matching of the PA, the output of the driver stage is optimized with load-pull simulation to match a 50 Ω load in Fig. 5.3 (b). Output MN2 * Γout 2 Pdel = 26.7 - 24.6 dBm @ Pin = 15 dBm * = 9.19-j1.24 Z out 2 Zo=50 Γout 2 1.51 nH 3.1 pF - Pdel contour - Series C - Shunt L (a) Output MN1 Pdel = 16.7 - 14.6 dBm @ Pin = 1 dBm Γ Γout 2 * out 2 * = 19.03-j5.06 Z out 2 Zo=50 2.5 nH 2.47 pF - Pdel contour - Series C - Shunt L (b) Fig. 5.3. (a) Load-pull and LC matching network using a ZY Smith chart for the output cascoded stage. (b). Load-pull and LC matching network using a ZY Smith chart for the driver stage. 41 The driver stage’s input reflection coefficients are characterized through source-pull analysis to select an impedance to give the maximum linear power delivered to the source at 2.5 GHz. From source-pull analysis with 1 dBm input power, contour plots for maximum delivered power of 15.7 to 13.7 dBm are shown as thin dark blue lines in Fig. 5.4 (a). The driver stage input gets maximum power at the input impedance 12.4 – j12.92. Using the complex conjugate of the normalized impedance of the driver stage, the LC matching network is designed as in the output cascoded stage. The output matching network’s series capacitance is 7.32 pF and its shunt inductance is 1.81 nH as shown in Fig. 5.4 (a). Input MN 1 Pdel = 15.7 – 13.7 dBm @ Pin = 1 dBm Γin1 Z in* 1 Zo=50 Γin* 1 7.35 pF = 12.4-j12.92 - Pdel contour - Series C - Shunt L 1.81 nH (a) Input MN 2 Pdel = 26.7 – 14.7 dBm @ Pin = 15dBm Γin1 Z in* 1 Zo=50 Γin* 1 15.9 pF = 3.99-j9.56 - Pdel contour - Series C - Shunt L 0.94 nH (b) Fig. 5.4. (a) Source-pull and LC matching network using ZY Smith chart for driver stage. (b) Source-pull and LC matching network using ZY Smith chart for input cascoded stage. 42 Similarly for interstage input matching of the PA, the input of the cascoded stage is also optimized with source-pull simulation, and LC matching networks [20] between them are designed for maximum power as shown in Fig. 5.4 (b). 5.4 Single-End Power Amplifier Measurement Results The single-end PA was implemented with 340 nm gate length in the UMC 130 nm RF CMOS process. The single-end power amplifier and fabricated PA chip occupies an area of 2.6 mm × 1.2 mm, including RF and DC pads, as shown by the superimposed schematic diagram in Fig. 5.5. At the drain gate voltage of 3.3 V, the top cascode nMOS gate is biased at 3.3 V, and the second gate is biased at 1 V; both gates draw drain currents of 295 mA. Fig. 5.5. A single-end Class AB (first feedback driver and second cascode output) power amplifier schematic and chip photo. High-frequency on-wafer SOLT calibration and S-parameter measurement are carried out with an Agilent 8364 network analyzer and an Agilent 8565 power spectrum analyzer as illustrated in Fig. 5.6 (a). Collected data is de-embedded to remove all losses of the cable and RF pads. The power amplifier is designed for a 2.5 GHz WiMAX application, and the frequency is shifted to 2.3 GHz because of the 10% process variation of the custom inductor 43 model from simulation. The measured and simulated results of S-parameters are plotted in Figs. 5.7 and 5.8. Measured and simulated input and output matching networks are better than –10 dB. The device achieves –3 dB bandwidth of 400 MHz, and the small-signal gain maximum is 32 dB. The isolation is better than –40 dB. PNA HP 8364 10Mz to 50 GHz DC HP 4142B Modular DC source DC RF RF GSG probe Device under test GSG probe Fig. 5.6. High-frequency S-parameter measurement setup. 0 -5 -10 -10 -20 -15 S11meas -20 S12 meas -30 S11sim -40 S12 sim -25 -50 5 4.5 4 3.5 3 2.5 -80 2 -40 1.5 -70 1 -35 0.5 -60 0 -30 S12 (dB) S11 (dB) 0 Freq (GHz) Fig. 5.7. Power amplifier S11 and S12 parameter models and measurements. 44 25 S21meas 30 S21sim 20 S22 sim S22 meas 15 10 5 0 -5 -10 S22 (dB) S21 (dB) 40 -15 -20 5 4.5 4 3.5 3 2.5 2 1.5 1 0.5 -25 0 -30 Freq (GHz) Fig. 5.8. Power amplifier S22 and S21 parameter models and measurements. At design frequency, 2.5 GHz, on-wafer SOLT and one-port power calibration measurements are carried out with an Agilent 8364 network analyzer and an Agilent 8565 power spectrum analyzer as illustrated in Fig. 5.9. A single-tone power measurement plot is shown in Fig. 5.10. The solid lines are simulations with the thick metal inductor model, and the symbol shapes correspond to measurements. The measured output power at 1 dB compression is >21.4 dB at –8 dBm input power, and it shows very good agreement with the simulation. The large-signal gain is 31 dB, and PAE is 14.5% at 1 dB compression. PNA HP 8364 10Mz to 50 GHz DC HP 4142B Modular DC source DC RF GSG probe Spectrum Analyzer 8565E RF GSG probe Device under test Fig. 5.9. Single-tone power calibration and measurement setup. 45 40 Gain 35 30 25 30 Pout 25 20 20 15 15 10 10 PAE Symbol – measure Line – simulation 5 PAE (%) Gain (dB) and Pout (dBm) 35 5 0 0 -20 -18 -16 -14 -12 -10 -8 -6 -4 -2 0 Pin (dBm) Fig. 5.10. Single-end power amplifier, single-tone measurement vs. simulation. The measured performance of the Class AB PA is summarized in Table 5.2 with other published results. The two-stage, on-chip, fully integrated power amplifier demonstrates higher power gain (31 dB) and maximum output power (25.5 dBm) compared to that of [21]. Compared to off-chip integrated PAs, the fabricated power amplifier has smaller form factor, higher gain with better linearity, and excellent linear output power. TABLE 5.2 Performance summary and comparison with previously published work CMOS process Class/stages Gain (dB) P1dB (dBm) Pmax (dBm) PAE1dB (%) PAEmax (%) Freq (GHz) Vdd (V) Integration [21] 250 nm AB/2 6 21 24.1 26 40 2.45 2 Off-chip [22] 180 nm AB&B/2 38 20.5 23.5 17 45 2.4 2.4 Off-chip [23] 180 nm AB/1 19 20.2 23 30.2 35 2.4 3.3 On-chip This work 130 nm AB/2 31 21.4 25.5 14.5 32 2.5 3.3 On-chip The parasitic (RC) incorporation of large-width nMOS is demonstrated and excellent agreement is achieved with measurement results. The design and fabrication of a single-end 46 PA for 2.5 GHz WiMAX applications based on load- and source-pull analysis to match load and source impedances simultaneously are demonstrated with a low-loss on-chip LC matching network. Using Wilkinson power-combining techniques in Fig. 5.11 (a), the singleend PA can be used in WiMAX applications, and the projected PA can achieve output power 20 18 16 14 12 10 8 6 4 2 0 + 1 P 2 P1− = S11 P1+ = 0 2 Zo P2− = S 21 P1+ = λ/4 Zo Gain (dB) and Pout (dBm) 30 2Z o 2Z o 2 P3− = S 31 P1+ = Zo P1+ 2 25 20 15 10 5 0 -20 -15 -10 -5 0 PAE (%) >24 dBm at 1 dB gain compression point, as shown in Fig. 5.11 (b). 5 Pin (dBm) (a) (b) Fig. 5.11. (a) Wilkinson power amplifier schematics. (b) Simulation results of Wilkinson powercombining technique to use with single-end power amplifiers. 47 6 An 80 to 85 GHZ CMOS POWER AMPLIFIER FOR W-BAND APPLICATION 6.1 Overview of W-Band Power Amplifier Design The power amplifier is the key component in monolithic millimeter wave integrated circuits (MMWICs) applications such as phased array radar, wideband communication systems and automotive sensors. Today, MMW amplifiers are made of III-V HEMT and HBT technology with excellent results at millimeter wave frequencies. Recently, CMOS has become attractive for low-cost and high-level integration due to advancement of nMOS performance with ft and fmax > 100 GHz and is available from commercial CMOS foundries. However, the W-band CMOS amplifier design remains as a major challenge due to lack of accurate device modeling, low breakdown voltage, and high losses in silicon substrate at millimeter wave frequency. Recently, 90 nm CMOS amplifiers have been demonstrated with good gain (> 10 dB) but with low linear output power (<10 dBm) and low power added efficiency (PAE <10%) at W-band frequency [24-28]. For millimeter wave application, as above two-stage and Wilkinson amplifiers are discussed with the experimental results. A compact two-stage CMOS power amplifier is designed with gain boosting at the common gate transistor, source degeneration for the cascode device, and LC short stub matching networks. Since the similar circuit topology is used in this circuit, only one stage is discussed more in detail. CMOS transceivers are typically implemented as differential circuits to reduce susceptibility to commonmode noise. A generic single-end two-stage PA is integrated using on-chip power-combining techniques. This design approach also improves the power gain at low input power level compared to the single-transistor stage, and it still needs to improve the power divider/combiner. A Wilkinson PA is essentially two single-end PAs in parallel, and 48 an additional gain stage is implemented to account for the power divider and combiner loss. The design has been employed with a cascode topology. The cascode approach reduces the gate-drain capacitance (Cgd) Miller multiplication effects and provides added isolation between the input and output ports, improving amplifier stability without using feedback. 6.2 A Two-Stage Power Amplifier Circuit Design In high-frequency applications, a cascode connection consists of a common-source stage driving a common gate stage. The cascode derives its advantage at high frequencies from the fact that the load for common source transistors is the low input impedance of the common gate. The common gate transistor operates as a current buffer and does not contain a feedback capacitance from drain to source to cause the Miller effect. Its useful characteristic is a small amount of reverse transmission and good isolation that is required in high-frequency amplifier circuits. Another characteristic of a cascode cell is its high output resistance, which reduces the conversion power loss to the output load resistor. Power gain of a common source circuit is approximately calculated from the small signal equivalent circuit model in the Eq. (6.1) and Fig. 6.1 (a). Similarly, the small signal equivalent circuit of a cascode cell model is shown in the Fig. 6.1 (b). The main advantage of the cascode device is to increase ideally double the output impedance of a single common source device. The output voltage of a cascode device is equal to the output voltage of each active device. Therefore, the cascode device has higher gain [29] than a common source transistor and is more attractive for power matching. For optimal power operation, additional series capacitance at the gate of transistor T2 is required to avoid early power saturation [30]. Using the small signal equivalent model of the optimized cascode device Eq. (6.2), the output power of the cascode device is twice that of the common source device in Eq. (6.3). 49 Cgs Vgs Vds gmVgs Rds Cds (a) D Vgs2 Cgs g mVgs2 Vds2 Rds Cds Rds Cds G Vgs1 Cgs Vds1 gmVgs1 S (b) Fig. 6.1. (a) Schematic of the theoretical common source device and its small-signal model. (b) Schematic of the theoretical cascode device and its small-signal model. Pout cs = Pout cas = ⎛ 1 ⎞ V ⎞⎤ 1 ⎡ ⎛ Re ⎢Vds ⎜⎜ g m ⋅ V gs + ds ⎟⎟ ⎥ , where Z = ⎜⎜ + jω (C ds )⎟⎟ 2 ⎣⎢ ⎝ Z ⎠ ⎦⎥ ⎝ Rds ⎠ 1 ⎡ Re ⎢ Vds1 + Vds 2 2 ⎣⎢ ( )⎛⎜⎜ g ⎝ m ) ( ) ⎞⎟⎤ ⎛ 1 ⎞ V + Vds 2 1 ⋅ ⋅ V gs1 + V gs 2 + ds1 2 2⋅Z ( ⎟⎥ ⎠ ⎦⎥ (6.1) (6.2) + jω (C ds )⎟⎟ ,where Vds1 = Vds 2 = V dsopt , V gs1 = V gs 2 = V gsopt , and Z = ⎜⎜ ⎠ ⎝ Rds ⎛ Vdsopt ⎞ ⎤ 1 ⎡ ⎟⎥ Re ⎢ 2 ⋅ Vdsopt ⎜ g m ⋅ V gsopt + ⎜ ⎟⎥ 2 ⎢⎣ Z ⎝ ⎠⎦ = 2 ⋅ Pout cs Pout cas = 50 (6.3) A Class A two-stage CMOS amplifier is designed for a stable input with high gain and compact area. The cascode device topology is chosen to increase the high gain and breakdown voltage of the amplifier device. In the W-band frequency range, a MOSFET cascode stage is unstable. A source degeneration CPW line is added to stabilize the input as shown in the schematic of Fig. 6.2. The source degeneration CPW lines are designed between the common source transistor and the common gate transistor to achieve the maximum stable gain. The CMOS process has inherent limitations to realize millimeter wave circuits. Extrinsic and intrinsic parasitic elements of the CMOS process also cause the gain to become even lower. Furthermore, gain is decreased by the lossy silicon substrate. A CPW inductance length, Lg, is designed at the common gate transistor for gain boosting [31] and [32]. From small-signal analysis, a conventional common gate amplifier of cascode gain is shown in Eq. (6.4) and a common gate amplifier with positive feedback network cascode gain is shown in Eq. (6.5) [30]. Vout ⎛ 1 ⎞ ⎟(R L // Rds ) = ⎜⎜ g m + Vs Rds ⎟⎠ ⎝ (6.4) Vout ⎛⎜ gm 1 ⎞⎟ (R // Rds ) = + 2 ⎜1− ω L C ⎟ L Vs R ds g gs ⎝ ⎠ (6.5) The small-signal gain is boosted to be higher than one of a conventional common gate amplifier when the operating frequency of the common gate amplifier is lower than 1 L g C gs . Figure 6.3 shows the simulated gain curves with a different length Lg. We compare and choose Lg of 100 µm for higher gain at 77 GHz. 51 Fig. 6.2. Cascode device with source degeneration and gain boosting schematic. 3 2 Gain (dB) 1 0 -1 Vg=1V Vd=3V 100 µm 40 µm 0 µm -2 -3 -4 100 98 96 94 92 90 88 86 84 82 80 78 76 74 freq, GHz Fig. 6.3. Cascode device with gain boosting characteristic curve using different gate line lengths (Lg = 0 to 100 µm). For power matching, load-pull simulations of different size devices are simulated at the design frequency. The load-pull simulation of the cascode CMOS device (W = 64 µm) is shown in Fig. 6.4. The constant output power contours (blue) from 7.4 to 9.41 dBm and the constant PAE contours (red) from 1.9% to 14% are plotted with different output impedances. The output power of 9.41 dBm with PAE of 14% of the designed amplifier can be achieved with the output impedance of 12.27 + j29.03. 52 Pdel=9.41 dBm PAE = 14% Pdel=7.4 dBm Vds=2.4 V Vgs=1 V Freq=77 GHz Z=12.27+j29.03 PAE=14.01% Pin=0 dBm Pout=9.41 dBm PAE = 1.9% Fig. 6.4. Load-pull simulation of cascode device: L 90nm x W 64 µm. The input and output impedance-matching networks are designed to maximize the power gain delivered to the load with the coplanar waveguides (CPW) in Fig. 6.5. Short stub LC impedance matching networks [33-34] are designed at the input and the output of the cascode cells, and the output short stub impedance matching network simulation result is shown for each matching step in Fig. 6.6. Fig. 6.5. LC short stub matching network. 53 LC short stub MN Zm Zload LC main line MN Short TL line Z*out Output Matching Network Fig. 6.6. LC short stub output impedance matching network simulation. A single-stage schematic diagram is shown in Fig. 6.7, and it is designed and simulated in ADS with the momentum CPW model. The simulation result of the output voltage waveform is shown in Fig. 6.8 (a) for the first stage and (b) for the second stage. Different input and output nodes are defined in the schematic. The amplifier is simulated with input power –25 dBm to 5 dBm and bias at Vd = 2.4V and Vg = 1V. The two-stage Wband amplifier uses an nMOS device width of 64 µm, and its drain current is approximately 25 mA for each stage. Vout2 RFin RFout Vinc Vout1 Voutc Vin Fig. 6.7. Schematic diagram of a one-stage amplifier. 54 3.0 First stage ts(X3.Voutc), V ts(X3.Vout2), V ts(X3.Vout1), V ts(X3.Vin), V ts(X3.Vinc), V 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 0 2 4 6 8 10 12 14 16 18 20 22 24 26 time, psec 3.5 Second stage 3.0 ts(X5.Voutc), V ts(X5.Vout2), V ts(X5.Vout1), V ts(X5.Vin), V ts(X5.Vinc), V 2.5 2.0 1.5 1.0 0.5 0.0 -0.5 0 2 4 6 8 10 12 14 16 18 20 22 24 26 time, psec Fig. 6.8. Simulation of (a) first stage amplifier and (b) second stage amplifier output waveforms at each node. 6.2.1 Experimental results The amplifier is implemented with the UMC 90 nm (gate length) RF CMOS 9-metallayers process. The chip occupies an area of L 0.5 mm × W 0.7 mm, including RF and DC pads, as shown by the superimposed schematic diagram in Fig. 6.9. The DC bias supplies are connected through matching networks. High-frequency on-wafer SOLT calibration measurements are carried out with a vector network analyzer 8510C, and measurement setup is shown in Fig. 6.10. The drain voltage and top cascode nMOS gate are biased at 2.4 and 3 V, and the second gate is biased at 1 V. The measured total drain currents are 51 and 61 mA for 2.4 and 3 V, respectively. 55 Fig. 6.9. Two-stage W-band amplifier schematics and fabricated chip. Fig. 6.10. W-band measurement setup with waveguide GSG probe and fabricated chip. The power amplifier is designed for a W-band application, and the maximum gain occurs at 80 GHz because of BSIM4 model variation and the custom CPW model from the momentum simulation. The ICFD2 model and measured results of the S-parameter are plotted in Fig. 6.11. The measured input and output matching networks are better than –10 dB. The amplifier achieved –3 dB bandwidth of 1.5 GHz, and the small-signal gain maximum is 18 dB. The reverse isolation is better than –26 dB [35]. 56 S parameters (dB) 20 16 12 8 4 0 -4 -8 -12 -16 -20 -24 -28 -32 -36 S21 S11 Vg= 1V Vd= Measured 2.4V Vd= Measured 3V Vd= ICF-D2 model 2.4V S22 75 77 79 81 83 85 87 89 91 Freq (GHz) Fig. 6.11. S-parameters measurement results of a two-stage W-band amplifier. For different process corners of slow (ss), fast (ff), and typical (tt), MOSFET devices are measured for RF characteristics with maximum ft ~ 90 GHz, 100 GHz, and 110 GHz respectively, while fmax ~ 125 GHz is approximately the same. As shown in Fig. 6.12, the amplifier gains are > 16 dB. The slow-to-fast process corners alter the impedance matching characteristics, shifting the peak gain frequency and increasing the bandwidth. The measured maximum gain varies with applied gate bias voltages from 0.8 to 1.2 V and fixed drain bias voltages of 2.4 V (red) and 3 V (blue) for three different amplifiers in Fig. 6.13. The smallsignal gain increases from 14.5 to 17.5 dB, and from 16 to 18.8 dB for drain source voltages 2.4 and 3 V, respectively, in Fig. 6.13. 57 Gain (dB) 20 16 12 8 4 0 -4 -8 -12 -16 -20 -24 -28 -32 -36 Vg=1V Vd=2.4V tt ff ss 75 77 79 81 83 85 87 89 91 Freq (GHz) Fig. 6.12. Small-signal gain comparison of a two-stage W-band amplifier with three different process corner devices: typical (tt), fast (ff), and slow (ss). 19 A1 A2 A3 Gain (dB) 18 17 Freq=80GHz Vg=1V Vd=3V Vd=2.4V 16 15 14 0.8 0.9 1 1.1 1.2 Vg (V) Fig. 6.13. Maximum gain measurement with different applied voltages. Comparison of three different amplifier circuits with same process corner (tt). A single-tone large-signal power measurement is carried out with an 8510C source. Output power is measured with an external harmonic mixer, W-band attenuator, and 8565E power spectrum analyzer. High-frequency on-wafer SOLT and one-port power calibration 58 are performed in Fig. 6.14. A single-tone power measurement of the two-stage CMOS amplifier plot is shown in Fig. 6.15. The measured amplifier at 80 GHz has demonstrated gain, Ga > 18 dB, output power at 1 dB compression, P1dB =10.8 dBm, saturated power, Psat = 13.3 dBm, and PAE = 11.8% with amplifier biased at Vd = 3 V and Vg =1 V. The gain peaking is due to the common gate inductance gain boosting as shown in Fig. 6.15. Spectrum Analyzer 8510C Harmonics Mixer Att Amp 0 dB 10dB Fig. 6.14. A W-band amplifier single tone power measurement at 80 GHz. CPW CPW 16 14 20 12 15 10 Vg=1V Vd=3V freq=80GHz 10 8 6 5 4 Pout Gain PAE 0 PAE (%) Gain (dB) & Pout (dBm) 25 2 -5 0 -22 -20 -18 -16 -14 -12 -10 -8 -6 -4 -2 Pin (dBm) 0 Fig. 6.15. A W-band amplifier single-tone power measurement at 80 GHz. The measured performance of our W-band amplifier compared well with other stateof-the-art published results, as summarized in Table 6.1. The compact two-stage on-chip fully 59 integrated W-band amplifier demonstrates high power gain (> 18 dB) with smaller chip area. This W-band power amplifier also achieves the highest linear and saturated power with the highest PAE at 80 GHz. TABLE 6.1 Performance summary and comparison of W-Band CMOS PA Tech. CMOS [24] 90 nm [25] 90 nm [26] 90 nm [27] 90 nm [28] 90 nm This 90 nm work Stages freq Vdd Id Gain BW P1dB Psat Pdc (GHz (V) (mA) (dB) (GHz) (dBm) (dBm) (mW) ) 4-stage common 77 1.2 118.5 8.5 ~5 4.7 6.3 142 source 5-stage common 80 1 176 12.2 ~25 7.5 10.3 176 source 3-stage cascode 94 2.4 75 15 20 7 10 180 Balanced PA 3 123 17 17 8 12 370 3-stage common 77 0.7 76 17.4 ~ 3 0.9 5.8 53 source 1 92 20.6 ~ 2 5 9.4 92 3-stage cascode 68-80 2.4 17.6 12 > 7 > 10.8 Broadside Coupler 68-83 3 18 15 > 8.3 > 11.8 2-stage cascode 80 2.4 51 17 1.5 8.65 12.3 122 3 60 18 1.5 10.8 13.3 180 PAE Size (%) (mm2) - 0.98 4.5 0.56 5 0.4 4.8 7.1 0.53 9.6 - 0.656 13.8 0.35 11.8 We reported a compact area for the cascode of the CMOS power amplifier, using inductance gain boosting, and short stub matching to match load-pull and source-pull impedances simultaneously: the amplifier circuit is demonstrated with a low-loss LC short stub on-chip CPW matching network. The measured compact two-stage power amplifiers demonstrated SOA results of minimum area of 0.35 mm2, Ga = 18.8 dB, and Psat > 13 dBm with PAE > 11.8%. Also, we showed that the variation in different process corner devices altered the maximum gain, bandwidth, and matching frequency. 60 6.3 Wilkinson Power Amplifier The W-band (75 to 110 GHz) power amplifier (PA) is a challenge in CMOS technology because of the low breakdown voltage and high losses in silicon at millimeter wavelengths. In 2008, a cascode-balanced CMOS PA demonstrated high performance at 94 GHz with 17 dB gain, P1dB = 8 dBm, and Psat = 12 dBm with PAE = 4.8% [26]. However, most of the W-band PA output power level is limited to P1dB < 14 dBm [31]. Hence, the monolithic power-combining techniques of PAs are attractive for delivering P1dB > 20 dBm due to the size reduction of the combiner operated at the W-band. In 2009, a balanced and wideband matching approach to a cascode CMOS PA covering 68 to 83 GHz reported outstanding power performance at Psat = 12 dBm and P1dB = 9 dBm at 80 GHz [28]. The The Wilkinson power combiner is another attractive technique; it was used in a V-band power amplifier at 60 GHz to achieve outstanding P1dB = 18.2 dBm and Psat = 20 dBm [36]. A Wband monolithic coplanar waveguide (CPW) Wilkinson power combiner with two CMOS power amplifiers in 90 nm CMOS technology is discussed in the following section. The 77 to 83 GHz CMOS PA achieved 17 dB small-signal gain, BW3dB of 4.5 GHz (from 78 to 82.5 GHz), 10.6 dBm linear output power, 12.3 dBm saturated power, and 3.9% PAE at 80 GHz. 6.3.1 Passive device CPW models Passive CPWs of different sizes are designed to reduce parasitic capacitance and resistance using ADS momentum simulation. The CPW line can be characterized by characteristic impedance Z0 and phase velocity vp. Z0 = L 1 = C v po ε eff C , where v p 0 = 61 1 μ 0ε o (6.6) From momentum S-parameter simulation, the effective dielectric constants εeff and Z0 are calculated and the results are ε eff = C c = 3.77 and Length = C0 4λ ε eff (6.7) The CPW transmission lines are designed for a 50 Ω load with 10 µm width, 5 µm gap, and 3 µm thickness in the metal 9 layer of the UMC 90 nm RF CMOS 9-metal-layer process. The measured loss in passive CPW lines is less than –0.5 dB for a 150 µm line length in Fig. 6.16 CPW L = 150 µm S-parameter (dB) (b) and less than –1 dB for a 400 µm line length. 0 -0.5 -1 -1.5 -2 -2.5 -3 -3.5 -4 -4.5 -5 -10 -12 S12 -14 S21 -16 -18 -20 -22 -24 -26 S11 S22 -28 -30 75 80 85 90 95 100 105 110 freq (GHz) (a) (b) Fig. 6.16. (a) CPW test structure. (b) Measured CPW transmission line 150 µm measurement. A Wilkinson power combiner and divider can be used for a power-combining device. It was designed with a 70.71 Ω quarter-wavelength CPW transmission line to minimize loss and a 100 Ω resistor to match the 50 Ω CPW line. The 70.71 Ω CPW was designed with the ADS momentum simulation with the signal line width of 14 µm, and the gap between signal and ground plane at 14 µm. The power divider [33-34] of RF input port 1 and output ports 2 and 3 can be measured from the S-parameters from Eqs. (6.8) to (7.0). P1− = S 11 P1+ = 0 62 (6.8) 2 P2− = S 21 P1+ = 2 P3− = S 31 P1+ = P1+ 2 (6.9) P1+ 2 (6.10) The test structure of port 1 to port 2 with port 3 is terminated with 50 Ω, as shown in Fig. 6.17 (a). The measured Wilkinson power combiner loss S21 is approximately –3.25 dB (1/2 of –6.5 dB for measured S21) in Fig. 6.17 (b). The input and output of the combiner are matched to 50 Ω with S22 = –10 dB in Fig. 6.17 (b). The total power loss in the combiner and divider is 6.5 dB from the measured results. Future improvement of the loss toward –3 dB may be realized. S12 & S22 (dB) λ/4 2Z o Zo 2Z o Zo 0 -2 -4 -6 -8 -10 -12 -14 -16 -18 -20 Zo S12 measured S22 measured S12 model S22 model 75 80 (a) 85 90 Freq (GHz) 95 100 (b) Fig. 6.17. (a) Wilkinson power combiner/divider test structure. (b) S-parameters momentum model simulation and measured results comparison. 6.3.2 Circuit design and experimental results In the W-band power amplifier circuit design, the cascode connection consists of a common-source stage driving a common-gate stage. The power gain of the cascode circuit is approximately twice that of the common-source amplifier. Cascode circuits have a small reverse transmission and good isolation, which are desired properties in high-frequency 63 amplifier circuits. Another characteristic of the cascode cell is its high output resistance, which can reduce the conversion power loss to the output matching network. The Wilkinson power amplifier in Fig. 6.18 is designed with the driver cascode gain stage, and then two cascade power stages are designed with the Wilkinson power-combining technique to increase linear power [7]. P1+ 2 P1− = S11P1+ =0 2 P2− = S21 P1+ = Zo λ /4 2Z o Zo 2 P3− = S31 P1+ = 2Zo P1+ 2 Zo Fig. 6.18. Schematics of Wilkinson power combiner with two W-band, two-stage CMOS power amplifiers. A Class A power amplifier design is chosen for stable input with high gain and a gain-boosting technique with short transmission line at a common-gate transistor. Also, a source degeneration CPW line is designed between the common-source and common-gate transistor stages in Fig. 6.18. The optimum device width of 2 µm with 32 fingers is chosen in the 90 nm CMOS device. The short stub LC matching networks are designed at the input and output of the cascode device. The DC bias supplies are connected through the impedancematching networks. The cascoded and cascaded amplifiers use a 64 µm width nMOS, and the drain current is approximately 25 mA for each stage. The Wilkinson amplifier has been implemented/fabricated with 90 nm gate lengths in the UMC 90 nm RF CMOS 9-metal-layer process. The chip occupies an area of 1.3 mm × 64 0.95 mm, including RF and DC pads, as shown by the superimposed block diagram in Fig. 6.19. High-frequency on-wafer SOLT calibration measurements were carried out with a vector network analyzer 8510C. P1+ 2 − P1 = S11 P1+ = 0 Zo 2 P2− = S 21 P1+ = λ/4 Zo 2Z o 2Z o 2 P3− = S 31 P1+ = Zo P1+ 2 Fig. 6.19. Fabricated W-band CPW Wilkinson power amplifier and its associated block diagram. The drain voltage and top cascode NMOS gates are biased at 2.4 and 3 V, and the second gate is biased at 1 V. The measured total drain currents are 130 and 150 mA for 2.4 and 3 V, respectively. The power amplifier is designed for a W-band application, and the maximum gain occurs at 80 GHz. The measured results of the S-parameters are plotted in Fig. 6.20. The measured input and output matching networks are better than –10 dB. The power amplifier has measured bandwidth (at –3dB) of 4.5 GHz (78 to 82.5 GHz), and the maximum small-signal gain is 17.8 dB. The measured gain varies with applied gate bias voltages from 0.8 to 1.2 V and drain bias voltage 2.4 V (red) and 3 V (blue) for three 65 different amplifiers with the same typical (tt) device in Fig. 6.21. The small-signal gain S parameters (dB) increases from 13.8 dB to maximum 17.8 dB at 80 GHz [37]. 20 16 12 8 4 0 -4 -8 -12 -16 -20 -24 -28 -32 S21 S22 Vg= 1V Vd= Measured 2.4V Vd= Measured 3V Vd= ICF-D2 model 2.4V S11 75 77 79 81 83 85 87 89 91 Freq (GHz) Fig. 6.20. S-parameters measurement and ICF-D2 model of Wilkinson power amplifier comparison. 18 A1 A2 A3 Gain (dB) 17 16 15 Freq=80 GHz 14 Vd=3V Vd=2.4V 13 0.8 0.9 1 1.1 1.2 Vg (V) Fig. 6.21. A Wilkinson PA maximum gain measurement with different applied gate and drain voltages for three different PA circuits. RF characteristics are measured for different process corners—slow (ss), fast (ff), and typical (tt)—MOSFET devices as in Fig. 6.22. Maximum ft values are 90 GHz, 100 GHz, 66 and 110 GHz for slow, typical, and fast devices, and fmax = 120 GHz. The Wilkinson amplifier maximum gain is 18 dB, as shown in Fig. 6.22. The slow-to-fast process corners alter the matching network characteristics, altering the peak gain frequency as well as increasing the bandwidth. For the slow device PA (green), the center frequency is around 80 GHz and bandwidth increases from 78 to 81 GHz. For the typical device PA (red), the center frequency is around 80.5 GHz with a bandwidth increase from 78 to 82.5 GHz. For the fast device PA (blue), the center frequency is around 82 GHz and bandwidth increases from 79 to Gain (dB) 85 GHz. 20 16 12 8 4 0 -4 -8 -12 -16 -20 -24 -28 -32 Vg=1V Vd=2.4V Measured tt Measured ff Measured ss ICF-D2 model 75 77 79 81 83 85 87 89 91 Freq (GHz) Fig. 6.22. S-parameter measurement results of the slow (ss), typical (tt), and fast (ff) Wilkinson power amplifiers. A single-tone power measurement was set up with an 8510C vector network analyzer, a harmonic mixer, a W-band wave guide attenuator, and an 8565 power spectrum analyzer. One-port power calibration was performed and removed all losses. A single-tone power measurement with a 3 V voltage supply is shown in Fig. 6.23. The measured output power 67 gain is approximately 20 dB, saturated power is 12.3 dBm, and PAE is approximately 3.9% for 3 V supplies. The measured performance of the CPW Wilkinson PA is summarized in Table 6.2 with other published results. Using Wilkinson power-combining techniques, the PA has higher linear output power (9.6 dBm and 10.6 dBm) and compared well with published W-band amplifiers. Further improvement in CPW combiner loss, the output stage layout, and the matching network can improve output power toward the desirable level of 20 dBm. 5 20 15 4 Vg=1V Vd=3V freq=80GHz 3 10 2 5 Pout Gain PAE 0 PAE (%) Gain (dB) & Pout (dBm) 25 1 -5 0 -22 -20 -18 -16 -14 -12 -10 -8 -6 Pin (dBm) -4 -2 0 Fig. 6.23. A Wilkinson amplifier single-tone power measurement at 80 GHz. 68 TABLE 6.2 Performance summary and comparison of CMOS power amplifier Tech. CMOS [24] 90 nm [26] 90 nm [27] 90 nm [31] 90 nm This 90 nm work Stages freq Vdd (GHz (V) ) 3-stage cascode 94 2.4 Balanced PA 3 3-stage cascode 68-80 2.4 Broadside Coupler 68-83 3 3-stage cascode 55-71 1.8 4 DAT combiner 3 4 way Wilkinson 60 1.2 PA 3-stage cascode 77-83 2.4 Wilkinson PA 3 Id Gain BW P1dB Psat Pdc PAE Size (mA) (dB) (GHz) (dBm) (dBm) (mW) (%) (mm2) 75 15 123 17 76 17.6 84 18 159 26.1 ~ 172 26 ~ 568 20 130 150 16 17 20 17 12 15 16 ~5 7 10 180 5 0.4 8 12 370 4.8 > 7 > 10.8 ~184 6.5 0.656 > 8.3 > 11.8 ~252 6 10.5 14.5 286 10.2 0.64 14.5 18 ~ 517 12.2 18.2 19.9 ~ 682 14.2 1.75 4.5 4.5 9.6 10.6 11.8 12.3 312 450 4.8 3.9 1.23 In this work, a parasitic device modeling ICF-D2 was developed to accurately predict RF behavior at a high-frequency amplifier design. A Wilkinson power-combining amplifier used a cascode device topology in a 90 nm nMOS and was designed with gain boosting, short stub matching to the power match load, and source impedances simultaneously using on-chip CPW. Also, variation in different process corner devices altered the maximum gain, bandwidth, and matching frequency. For the MMW power amplifier application, the CPW Wilkinson power divider and combiner techniques to increase the power gain to 20 dB, the linear output power to 10.6 dBm, and saturated power to 12.3 dBm were demonstrated in CMOS technology. 69 7 CONCLUSIONS The impact of external parasitic distributed effects on the performance of a CMOS power transistor has been discussed with the extrinsic scalable and non-scalable parasitic extraction methods. It has been shown that including distributed effects is important for predicting the correct ft, fmax, transducer power gain, and output power. The proposed distributed ICF-D1 model is validated by comparing the DC, S-parameters, and one-tone measurements of the scaled CMOS power devices. The scalable distributed ICF-D2 modeling method is extended to the power CMOS devices for millimeter wave application. The power devices are measured for S-parameters from 1 to 50 GHz, V-band (50 to 75 GHz), and W-band (75 to 110 GHz) with the vector network analyzer. The resulting S-parameters are created for distributed two-port models of the power devices and transmission lines. Distributed model parameters are measured and extracted in physical layout-based intrinsic and extrinsic parameters from 1 to 110 GHz. For the model verification, we compared the lumped and distributed models with the experimental results of MMIC power amplifiers in S-band and W-band applications. The lumped parasitic (RC) incorporation of the large width nMOS is demonstrated and excellent agreement is achieved with the measured results. The design and measured results of a single-end PA for 2.5 GHz WiMAX applications based on load- and source-pull analysis to match load and source impedances simultaneously are demonstrated with a low-loss on-chip LC matching network. We proposed a Wilkinson power-combining technique; the singleend PA can be used in power amplifier applications, and the projected PA can achieve output power >24 dBm at 1 dB gain compression point. A parasitic device modelling, ICF-D2, is developed to accurately predicte RF behavior at millimeter wave amplifier designs. We reported a compact area of a cascode cell 70 with the two-stage cascode CMOS power amplifier, using the inductance gain boosting, and short stub matching to match load- and source-pull impedances simultaneously. The amplifier circuit is demonstrated with a low-loss LC short stub on-chip CPW matching network. The measured compact two-stage power amplifiers were demonstrated excellent results with a minimum area of 0.35 mm2. For the MMW power amplifier application, the Wilkinson MMW amplifier used a cascode device topology in a 90 nm nMOS, and was designed with two (two-stage) amplifiers and a driver amplifier with CPW Wilkinson power divider/combiner techniques. The Wilkinson linear power gain is increased compared to the two-stage power amplifier. Due to higher loss from silicon and inter-stage and output stage saturation, Wilkinson saturated output power is lower than the two-stage amplifier. Also, variation in different process corner devices altered the maximum gain, bandwidth, and matching frequency in both W-band power amplifiers. Our preliminary work shows that CMOS power amplifier technologies have great potential for system-on-chip integration with low cost and minimum area. Further development of low-loss power-combining techniques can improve the output power of > 100 mW, gain > 20 dB, wide bandwidth and PAE > 10% Wilkinson MMWICs. Further research in MMW amplifiers can be explored in 65 and 45 nm technology nodes. 71 REFERENCES [1] B. Balvinder, R. Eline, and L. Franca-Neto, “RF system and circuit challenges for WiMAX,” Intel Technology Journal, vol. 8, no. 3, pp. 189-198, 2004. [2] L. M. Franca-Neto, R. Eline, and B. Balvinder, “Fully integrated CMOS radios from RF to millimeter wave frequencies,” Intel Technology Journal, vol. 8, no. 3, pp. 241258, 2004. [3] T. H. Lee, The Design of CMOS Radio-Frequency Integrated Circuits. New York, NY: Cambridge, 1998. [4] A. Grebennikov, RF and Microwave Power Amplifier Design, 2nd ed. New York, NY: McGraw-Hill, 2005. [5] WiMAX Forum (2006, Aug.). Mobile WiMAX—Part I: A technical overview and performance evaluation. [Online] Available: http://www.wintegra.com/assets/application_notes%20and%20whitepapers/Mobile_ WiMAX-_Part_1.pdf [6] S. Emami, C. H. Doan, A. M. Niknejad, and R. W. Brodersen, “Large-signal millimeter-wave CMOS modeling with BSIM3,” in RFIC Symposium, 2004, pp. 163166. [7] D. Chan and M. Feng, “2.5 GHz CMOS power amplifier integrated with low loss matching network for WiMAX applications,” in Asia Pacific Microwave Conference, 2009, pp. 1108-1111. [8] M. Feng, “Active microwave circuit design,” class notes for ECE 447, Department of Electrical and Computer Engineering, University of Illinois at Urbana-Champaign, 2001. [9] G. Massobrio and P. Antognetti, Semiconductor Device Modeling with Spice, 2nd ed. New York, NY: McGraw-Hill, 1993. [10] V. Dimitrov, J. B. Heng, K. Timp, O. Dimauro, R. Chan, J. Feng, W. Hafez, T. Sorsch, W. Mansfield, J. Miner, A. Kornblit, F. Klemens, J. Bower, R. Cirelli, E. Ferry, A. Taylor, M. Feng, and G. Timp, “Small-signal performance and modeling of sub-50nm nMOSFETs with fT above 460-GHz,” Solid-State Electronics, vol. 52, pp. 899-908, 2008. [11] S. Takagi, M. Iwase, and A. Toriumi, “On universality of inversion-layer mobility in n- and p-channel MOSFETs,” in IEEE Electron Devices Meeting (IEDM) Technical Digest, 1988, pp. 398-401. [12] Y. Taur and T. H. Ning, Fundamentals of Modern VLSI Devices, 2nd ed. New York, NY: Cambridge, 2009. 72 [13] Y. Eo and K. Lee, “A fully integrated 24 dBm CMOS power amplifier for 802.11a WLAN applications,” IEEE Microwave and Wireless Components Letters, vol. 14, no. 11, pp. 504-506, November 2004. [14] J. Ou, X. Jin, I. Ma, C. Hu, and P. R. Gray, “CMOS RF modeling for GHz communication IC’s,” in Symposium on VLSI Technology Digest of Technical Papers, 1998, pp. 94-95. [15] S. A. Maas, B. L. Nelson, and D. L. Tait, “Intermodulation in heterojunction bipolar transistors,” IEEE Transactions on Microwave Theory and Techniques, vol. 40, no. 3, pp. 442-448, 1992. [16] M. Vaidyanathan, L. E. Larson, P. S. Guderm, and P. M. Asbeck, “A theory of highfrequency distortion in bipolar transistors,” IEEE Transactions on Microwave Theory and Techniques, vol. 51, pp. 448-461, 2003. [17] S. Narayanan and H. C. Poon, “An analysis of distortion in bipolar transistors using integral charge control model and volterra series,” IEEE Transactions on Circuit Theory, vol. 20, pp. 341-351, 1973. [18] D. Caruth, “Large-signal modeling of MESFET,” M.S. thesis, University of Illinois at Urbana-Champaign, Urbana, IL, 1999. [19] D. Chan and M. Feng, “Distributed modeling of layout parasitics effects in CMOS power devices,” in European Microwave Conference, September 2010, to be published. [20] S. J. Franke, “Radio communication circuits,” class notes for ECE 353, Department of Electrical and Computer Engineering, University of Illinois at Urbana-Champaign, 2000. [21] C. F. Jou, K. Cheng, C. Lin, and J. Chen, “Dual band CMOS power amplifier for 5 GHz WLAN,” in European Solid State Circuits Conference, 2003, pp. 1227-1230. [22] T. Sowlati and D. M. W. Leenaerts, “A 2.4 GHz 0.18 µm CMOS self-biased cascode power amplifier,” IEEE Journal of Solid-State Circuits, vol. 38, pp. 1318-1324, August 2003. [23] H. Oh, C. Kim, H. Yu, and C. Kim, “A fully-integrated +23 dBm CMOS triple cascode linear power amplifier with inner-parallel power control scheme,” in Radio Frequency Integrated Circuits (RFIC) Symposium, 2006, pp. 165-168. [24] T. Suzuki, Y. Kawano, M. Sato, T. Hirose, and K. Joshin, “60 and 77 GHz power amplifiers in standard 90 nm CMOS,” in ISSCC Tech. Digest, February 2008, pp. 562-563. [25] N. Kurita and H. Kondoh, “60 and 80 GHz wide band power amplifier MMICs in 90 nm CMOS technology,” in IEEE Radio Frequency Integrated Circuits (RFIC) Symposium, 2009, pp. 39-42. 73 [26] Y. S. Jiang, J. H. Tsai, and H. Wang, “A W-band medium power amplifier in 90 nm CMOS,” IEEE Microwave Wireless Component Letter, vol. 18, no. 12, pp. 562-563, December 2008. [27] Y. Hamada, M. Tanomura, M. Ito, and K. Maruhashi, “A high gain 77 GHz power amplifier operating at 0.7V based on 90 nm CMOS technology,” in IEEE Radio Frequency Integrated Circuits (RFIC) Symposium, 2009, pp. 39-42. [28] J. Lee, C.-C. Chen, J.-H. Tsai, K.-Y. Lin, and H. Wang, “A 68-83 GHz power amplifier in 90 nm CMOS,” in Microwave Symposium Digest MTT-S International, 2009, pp. 437-440. [29] P. R. Gray, P. J. Hurst, S. H. Lewis, and R. G. Meyer, Analysis and Design of Analog Integrated Circuits, 4th ed. New York, NY: John Wiley & Sons, 2001. [30] A. Martin, T. Reveyrand, M. Campovecchio, R. Aubry, S. Piotrowicz, D. Floriot, and R. Quere, “Design method of balanced AlGaN/GaN HEMT cascode cells for wideband distributed power amplifiers,” European Microwave Association, vol. 4, pp. 261-267, December 2008. [31] Y.-N. Jen, J.-H. Tsai, T.-W. Huang, and H. Wang, “Design and analysis of a 55-71 GHz compact and broadband distributed active transformer power amplifier in 90-nm CMOS process,” IEEE Transactions on Microwave Theory and Techniques, vol. 57, no. 7, pp. 1637-1646, July 2009. [32] C. Lee, W. Choi, J. Kim, and Y. Kwon, “ A 77 GHz 3-stage low noise amplifier with cascode structure utilizing positive feedback network using 0.13 µm CMOS process,” Journal of Semiconductor Technology and Science, vol. 8, no. 4, pp. 289-294, December 2008. [33] G. Gonzalez, Microwave Transistor Amplifier Analysis and Design. Upper Saddle River, NJ: Prentice Hall, 1997. [34] D. M. Pozar, Microwave Engineering, 2nd ed. New York, NY: John Wiley & Sons, 1998. [35] D. Chan and M. Feng, “A compact W-Band CMOS power amplifier with gain boosting and short stub matching for high power and high efficiency operation,” Microwave Wireless and Component Letters, submitted for publication. [36] C. Y. Law and A. V. Pham, “A high-gain 60GHz power amplifier with 20dBm output power in 90nm CMOS,” in IEEE Solid-State Circuits Conference Digest of Technical Papers (ISSCC), 2010, pp. 426-427. [37] D. Chan and M. Feng, “W-band monolithic CPW Wilkinson CMOS power amplifier,” in IEEE Topical Conference on Power Amplifiers for Wireless and Radio Applications, January 2011, submitted for publication. 74 AUTHOR’S BIOGRAPHY Doris A. Chan was born in Mawlamyine, Myanmar (formally known as Burma). She started her undergraduate studies at Western Illinois University in 1997 and transferred to the University of Illinois at Urbana-Champaign in 1999. She received the B.S. and M.S. degrees in electrical engineering from the University of Illinois at Urbana-Champaign in 2001 and 2003, respectively. Her master’s thesis focused on the design of a differential voltagecontrolled oscillator for 2.4 GHz and 5.2 GHz WLAN applications. She was with Xindium Technologies, Inc. from 2002 to 2004, where she joined in the test development of InP (SHBTs) PIN diodes and transimpedance amplifiers (TIAs), and double heterojunction bipolar transistor (DHBTs) power amplifier devices. She also fabricated InP DHBTs devices for power amplifier applications. From 2004 to 2008, she was with Trace Photonics, Inc., where she was involved in the development of Radio Isotope Micro Source (RIMS) power conversion, and designed solar cell controlled circuits for military GSM transmitters. Her Ph.D. dissertation developed a characterization and modeling of RFCMOS power devices, and designed MMIC power amplifiers for WiMAX and W-band applications. Her research work on parasitics distributed power device modeling with a MMIC power amplifier achieved state-of-the-art results of output saturated power at 13.3 dBm with PAE 11.8% with a minimum area of 0.35 mm2. She received her Ph.D. in electrical engineering in 2010 in the area of radio frequency integrated circuit design. She is a recipient of a DOE Energy Research Undergraduate fellowship and a Yunni Pao’s fellowship. She is the author or coauthor of seven journal and conference proceeding publications. 75