Nanometer CMOS RFICs for Mobile TV Applications
ANALOG CIRCUITS AND SIGNAL PROCESSING SERIES Consulting Editor: Mohammed Ismail. Ohio State University
For other titles published in this series, go to http://www.springer.com/series/7381
Ahmed A. Youssef James Haslett ●
Nanometer CMOS RFICs for Mobile TV Applications
Ahmed A. Youssef University of Calgary Department of Electrical & Computer Engineering Theoretical & Empirical Software 2500 University Drive Calgary AB T2N 1N4 Canada
[email protected]
James Haslett Department of Electrical and Computer Engineering University of Calgary Calgary Canada
[email protected]
ISBN 978-90-481-8603-7 e-ISBN 978-90-481-8604-4 DOI 10.1007/978-90-481-8604-4 Springer Dordrecht Heidelberg London New York Library of Congress Control Number: 2010922879 © Springer Science+Business Media B.V. 2010 No part of this work may be reproduced, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, microfilming, recording or otherwise, without written permission from the Publisher, with the exception of any material supplied specifically for the purpose of being entered and executed on a computer system, for exclusive use by the purchaser of the work. Printed on acid-free paper Springer is part of Springer Science+Business Media (www.springer.com)
TO OUR FAMILIES
Preface
People often ask me why I chose to publish this work. Primarily, this book was motivated by my experience in industry after having been in the academic world for many years. We often hear about the need to bridge the gap between industry and academia, but one can easily intellectually understand something without having experienced it first hand. I lived in the gap between the academic and industrial worlds for some time and saw how people on both sides can target the same application and work toward the same goal, but do so from completely different angles. Radio frequency (RF) IC designers in the industry do not have the luxury of time to dig through the extensive publications that flood their field to understand the latest developments related to their own circuits or systems. Researching the many areas related to RF circuit design such as device models, fabrication technology, signal analysis, and communication theory proves to be an insurmountable task. Therefore, their circuit analysis, and thus optimization, can be easily curtailed because of the time restraints inherent in the semiconductor industry. On the other hand, university students of all levels are often so flooded in their own studies that they lose the practical sense needed to make their research seem applicable to realworld products – the world of RFIC design becomes dry and intangible. The main purpose of this book is thus to help bridge the gap between academia and industry in the area of nanometer CMOS RFIC design for one of the latest, most dynamic wireless applications, the mobile TV. In 2005, South Korea became the first country in the world to offer mobile TV reception. Japan and several European countries soon followed, and the United States and Canada are not far behind. This book explores the design, implementation, and demonstration of highly linear, low power, RFICs that facilitate the integration of TV service into cellular phones using nanometer CMOS technologies. Emphasis is made on how to break the trade-off between power consumption and performance (linearity and noise figure) by optimizing the mobile TV front-end dynamic range in three hierarchical levels: the intrinsic MOSFET level, the circuit level, and the architectural level.
vii
viii
Preface
The objectives of this book are as follows: –– Develop a complete system analysis to extract the target specifications of the mobile TV RFIC signal path. The RF front-end architectures and circuits described in this book are based on this analysis. –– Discuss the challenges associated with designing wide dynamic range, lowpower, broadband CMOS low-noise amplifiers (LNAs). The main CMOS LNA topologies are analyzed in the context of their dynamic range. The modern noise-cancelling CMOS amplifiers are discussed and their limitations in achieving the requirements of the mobile TV receiver are described. The book also presents several new circuit design techniques proposed to improve the dynamic range of current noise-cancelling LNAs. –– Provide methods for implementing the gain variation function in CMOS amplifiers. Both the analog and digital approaches are considered. –– Present the RF passive gain control technique as a low-power solution to help mobile TV receivers achieve their stringent linearity requirements and present the linearization circuits needed for this technique. –– Expose students in academia to the practical steps required to develop successful mobile TV products: the physical realization of three 65 nm CMOS test chips is described and the associated challenges are highlighted, the bias generator circuits that allow the developed circuits to withstand process and temperature variations are discussed, and lastly, the test setup required to characterize the designed test boards in the lab is presented. Although the RFIC specifications developed in this book target the DVB-H mobile TV standard, the solutions described can be scaled to any other digital TV standard or any standard requiring a wide dynamic range receiver. Finally, I am much obliged to mention the various individuals who helped make this book possible. I would first like to thank James Haslett, my co-author, advisor and mentor for his innumerable contributions to my knowledge of this field. I will also always be indebted to Professor Hani Ragai, who introduced me to VLSI circuit design while I was an undergraduate student and inspired me to pursue my study of it. Additionally, I would like to thank my colleagues Edward Youssoufian and Janakan Siva (Newport Media Inc.), Aly Ismail (Ultrawave Labs), and Ahmet Tekin (Aydeekay Inc.). All of them have contributed to the material in this book in so many ways. Very special thanks go to Hassan Elwan for transferring some of his baseband circuit design knowledge to me. The text was reviewed and edited by Carey Williamson (University of Calgary), Igor Filanovsky (University of Alberta), and Mohammed Ismail (Ohio State University). I am indebted to all of them for their kind assistance. I must also thank the staff at Springer, particularly Cindy Zitter for their assistance. Finally, this work would not have been possible without the support of my wife, Marwa. I thank her for her encouragement and patience. November 2009
Ahmed Youssef
Contents
1 Introduction and Overview....................................................................... 1.1 Mobile TV Architectures.................................................................... 1.2 DVB-H Mobile TV System Definitions............................................. 1.3 Scope of This Book.............................................................................
1 3 9 12
2 Wideband CMOS LNA Design Techniques............................................. 2.1 Dynamic Range Limits in MOSFETs................................................. 2.1.1 The Noise Limit...................................................................... 2.1.2 The Distortion Limit............................................................... 2.1.3 Dynamic Range Trade-offs in CMOS..................................... 2.2 Traditional CMOS LNA Topologies................................................... 2.2.1 R-CS Amplifier....................................................................... 2.2.2 CG Amplifier.......................................................................... 2.2.3 SFB Amplifier......................................................................... 2.2.4 L-Degenerate Amplifier.......................................................... 2.3 Recent Trends in Wideband CMOS LNAs......................................... 2.3.1 Current Reuse Amplifiers....................................................... 2.3.2 L-Degenerate Wideband Amplifiers....................................... 2.3.3 Capacitive Cross-Coupled CG Amplifiers.............................. 2.3.4 Noise and Distortion Cancelling Amplifiers........................... 2.4 Techniques to Improve the Wideband LNA Dynamic Range............. 2.4.1 Wideband CMOS LNA State-of-the-Art................................ 2.4.2 New Low-Power Noise-Cancelling Technique....................... 2.5 Chapter Summary...............................................................................
15 16 16 19 25 26 26 28 28 31 33 34 36 37 39 43 43 45 56
3 Nanometer CMOS LNAs for Mobile TV Receivers................................ 3.1 Requirements of the LNA in Mobile TV Receivers........................... 3.1.1 DVB-H RF Front-End Specifications..................................... 3.1.2 DVB-H LNA Performance Requirements.............................. 3.2 A 65 nm CMOS Wideband LNA Prototype....................................... 3.2.1 LNA Core Circuit.................................................................... 3.2.2 DC Bias Generator Circuits.................................................... 3.2.3 Multi-Mode Test Buffer Circuits............................................
59 59 60 61 63 64 71 77 ix
x
Contents
3.3 Experimental Results.......................................................................... 3.3.1 Test Environment Descriptions............................................... 3.3.2 Measurement Results.............................................................. 3.4 Chapter Summary...............................................................................
81 81 84 91
4 RF Attenuator Linearization Circuits..................................................... 4.1 The Necessity of RF Automatic Gain Control.................................... 4.1.1 RF Gain Control in Mobile TV Receivers.............................. 4.1.2 Gain Control Circuit Techniques............................................ 4.2 RF Gain Control System Analysis...................................................... 4.2.1 Case One: DR Is Limited by the Clipping Level.................... 4.2.2 Case Two: DR Is Limited by the IIP3 Level........................... 4.3 Highly-Linear RF Front-End Architectures........................................ 4.3.1 Linear RF Architectures.......................................................... 4.3.2 Gain Step Size......................................................................... 4.4 Design of the Binary-Weighted RF Attenuator.................................. 4.4.1 Topology Evolution................................................................. 4.4.2 Binary-Weighted RF Attenuator Design................................. 4.4.3 Gain Control Logic Circuits.................................................... 4.5 Practical Considerations...................................................................... 4.5.1 RF Attenuator & LNA Integration.......................................... 4.5.2 Package Bond Wire Coupling................................................. 4.6 A 65 nm CMOS RF Passive Attenuator............................................. 4.6.1 The Fabrication....................................................................... 4.6.2 Measurement Results.............................................................. 4.6.3 Comparison with Simulations................................................. 4.7 Chapter Summary...............................................................................
95 95 95 96 98 99 100 103 103 104 106 106 108 112 117 117 120 123 123 124 127 130
5 Wide Dynamic Range Mobile TV Front-End Architecture................... 5.1 Mobile TV Front-End with Automatic Gain Control......................... 5.1.1 Self-Contained RF AGC Control............................................ 5.1.2 DVB-H RF Front-End with AGC Algorithm.......................... 5.1.3 AGC RF Level Indicator Circuit............................................. 5.2 A 65 nm CMOS RF Front-End Prototype.......................................... 5.2.1 The Fabrication....................................................................... 5.2.2 Measurement Results of the AGC Test................................... 5.3 Chapter Summary...............................................................................
131 131 131 132 133 135 135 136 138
6 Summary and Conclusions....................................................................... 6.1 Summary and Conclusions................................................................. 6.1.1 Digitally-Controlled Variable-Gain LNA............................... 6.1.2 Digitally-Programmed RF Passive Attenuator........................ 6.1.3 Wide Dynamic Range Mobile TV Front-End.........................
139 139 139 140 142
Contents
xi
6.2 Further Research Areas....................................................................... 143 6.2.1 System Studies........................................................................ 143 6.2.2 Circuit Studies......................................................................... 144 References......................................................................................................... 145 Index.................................................................................................................. 153 Author Biographies.......................................................................................... 157
List of Symbols and Abbreviations
Acronym
Definition
AC ADC AGC BiCMOS BPF BSIM CG amplifier CMMB CMOS CSP dB dBm DC DCR DMB-T DR DTV DVB-H DVB-T f fs ft FB FE FOM GHz GSM Hz IF IIP3 IP3
Alternating Current Analog-to-Digital Converter Automatic gain control Bipolar and CMOS capable process Band-Pass Filter Berkeley Short-Channel IGFET Model Common-gate amplifier China Multimedia Mobile Broadcasting Complementary Metal Oxide Semiconductor Chip scale package Logarithmic ratio of power Logarithmic ratio of power referenced to 1 mW Direct Current Direct Conversion Receiver Digital Multimedia Broadcast-Terrestrial Dynamic range Digital TV Digital Video Broadcasting-Handheld Digital Video Broadcasting-Terrestrial Frequency Sampling frequency Unity current gain frequency Feedback Front-end circuit Figure of merit Gigahertz Global System Mobile communications Hertz Intermediate frequency Input-referred IP3 Third-order intercept point xiii
xiv
IM3 IR ISDB-T low_th_n low_th_p kHz LC LG LNA LO LPF MER MIM MOSFET NF NMOS OFDM Opamp OTA PMA PMOS Q QAM QFN R-2R R-CS amplifier RF RFLI RL rms S11 S12 S21 S22 SAW SFB amplifier SMA SNR SOC up_th_n up_th_p U/D UHF
List of Symbols and Abbreviations
Third-order intermodulation distortion Image rejection Integrated Services Digital Broadcasting-Terrestrial Negative lower threshold voltage of the RFLI Positive lower threshold voltage of the RFLI kilohertz Inductor–capacitor circuit Loop gain Low-noise amplifier Local Oscillator Low-pass filter Modulation Error Rate Metal–Insulator–Metal capacitor Metal-Oxide-Semiconductor Field-Effect Transistor Noise figure N-channel MOSFET Orthogonal Frequency Division Multiplexing Operational amplifier Operational transconductance amplifier Post-mixer amplifier P-channel MOSFET Quality factor Quadrature amplitude modulation Quad Flat No-Lead Package Analog-to-digital converter topology Resistive terminated common-source amplifier Radio frequency RF level indicator Load resistor Root-mean-square Input reflection coefficient Reverse transmission coefficient Forward transmission coefficient Output reflection coefficient Surface acoustic wave Shunt feedback amplifier Sub-Miniature version A Signal-to-Noise Ratio System On Chip Negative upper threshold voltage of the RFLI Positive upper threshold voltage of the RFLI Undesired-to-desired-signal Ultra High Frequency
List of Symbols and Abbreviations
Vgs vgs VGA VHF WLAN
DC gate–source voltage Small signal AC gate-source voltage Variable Gain Amplifier Very High Frequency Wireless Local Area Network
xv
Chapter 1
Introduction and Overview
Digital TV (DTV) has currently been deployed in several countries around the world [1]. Where in use, digital TV has many advantages over traditional analog TV such as better image and sound quality, increased number of channels, and interactive multimedia services. Additionally, digital TV standards would facilitate the implementation of video reception (i.e., TV) in cellular phones and multimedia enabled handheld devices [2]. In broadcast DTV, the two most dominant standards are DVB-T (Digital Video Broadcasting-Terrestrial) in Europe and ISDB-T (Integrated Services Digital Broadcasting-Terrestrial) in Japan and Brazil. Both of them are based on OFDM (Orthogonal Frequency-Division Multiplexing) modulation. Using OFDM, the broadcast tolerates both fading and multipath [3, 4], which until now have generated unacceptable obstacles for mobile reception of conventional analog TV. Since both ISDB-T and DVB-T standards are for full screen TV, the data transmission rates are too large for mobile TV use. Each of these standards was thus modified differently to try to make it applicable for use in mobile TV applications. In the Japanese standard, a mobile TV standard called ISDB-Tss (ISDB-T single segment), which uses a lower bandwidth than that of the broadcast ISDB-T, was developed. It uses 430 KHz bandwidth in a 6 MHz channel divided into 13 segments as shown in Fig. 1.1. The standard is defined mainly for a portion of the UHF band, 440–740 MHz, to allow interoperability with the 900 MHz band for GSM (Global System Mobile). In contrast, the European standard was modified to develop a mobile version called DVB-H (DVB – Handheld), which uses time divided transmission called “time slicing”, where each burst transmission is followed by a longer off period. Time slicing operation is shown in Fig. 1.2. The first DVB-H field trials were held in Europe and used the UHF band (470–862 MHz). Since then, DVB-H has also been targeted for deployment in North America using the L-band spectrum between 1670 and 1675 MHz. There has also been discussion of reallocating European L-band frequencies for DVB-H services. Finally, several other mobile TV standards, both terrestrial and satellite, are emerging world-wide such as MediaFLO [2] and ATSC-mobile in the USA, DMB-T and satellite based CMMB in China and T-DMB [1] in Korea.
A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_1, © Springer Science + Business Media B.V. 2010
1
2
1 Introduction and Overview 1 segment = 430 KHz
6 MHz Channel
12 10
8
6
4
2
0
1
3
5
Guard band
7
9
11
Guard band
Segment 0 is dedicated for mobile use. The other 12 segments are used for high bit rate broadcast
Fig. 1.1 Transmissions of ISDB-Tss broadcast
0.25s 0.75s (OFF)
Burst Transmission
Data Rate 5 Mbps (QPSK 1/2)
Effective Data Rate = 400 kbps Fig. 1.2 Time slicing in DVB-H
A key component of the next generation of mobile TV-enabled cellular phones is the silicon receiver that integrates all the TV tuner functions into a silicon die. Low-cost, low-power and small physical size are required by cellular phones and other multimedia mobile products. These features necessitate moving into a low-cost technology and to a smaller geometry. Nanometer CMOS technology is thus essential to achieve these requirements by facilitating the integration of more high-performance RF (Radio Frequency) functions into a single piece of silicon. In addition to the power, cost, and size requirements, all sensitivity, blocking
1.1 Mobile TV Architectures
3
8 Economic recession period
6 Paper numbers at ISSCC
4
WLAN application Mobile TV application
2 0 2003
2005
2007
Year
Fig. 1.3 Comparison between the numbers of research papers at the ISSCC conference for the last five years for WLAN and mobile TV applications
and intermodulation specifications of different standards should be met without sacrificing the optimality of the solution. For the past few years, the development of mobile TV systems has been under extensive study in both industry and university research communities. The first paper addressing the challenges of implementing mobile TV reception in cell phone front-ends was published in 2004 at the world-class International Solid-State Circuit Conference (ISSCC). Since then, the number of papers on this topic has increased rapidly. Figure 1.3 shows a comparison between the number of published papers in the area of WLAN (Wireless Local Area Network) applications and the number of published papers in the area of mobile TV applications in the last 5 years at ISSCC. As shown in this chart, since 2006 the topic of mobile TV has started to attract research interests and is expected to continue doing so in the future.
1.1 Mobile TV Architectures The broadband nature of the TV band introduces many technical challenges in the design of silicon tuners needed for mobile TV. To support digital TV reception, typical broadcast TV tuners need to address issues such as coverage of several octaves, rejection of interference at integer multiples of the wanted signal, and inband and out-of-band image rejection. These issues have been extensively addressed in the technical literature [5–10]. The traditional TV-can tuner architecture uses a passive LC tracking filter and an IF SAW (surface acoustic wave) filter as shown in Fig. 1.4. The tracking filter prevents image mixing and harmonic mixing which are described in Fig. 1.5 [8]. Image mixing is the process whereby the negative frequency component folds into the baseband in the down-conversion mixing process. Harmonic mixing is the
4
1 Introduction and Overview on-chip
RF input
RF-VGA Tracking Filter
x
SAW IF output
Recently, an elaborate varactor was proposed to facilitate the tracking filter integration [21]. Also, a selectable on-chip bipolar filter was used in [14] to suppress the odd harmonics of LO.
VGA: Variable Gain Amplifier Channel Selection
Fig. 1.4 Conventional TV-can tuner architecture
LO harmonics
Harmonic Conversion
DC
LO
2LO
3LO
Desired signal Image
Blockers
Image Conversion
DC IF
LO-IF
LO
LO+IF
48-862 MHz
Fig. 1.5 Harmonic and image conversion in wideband receiver
process whereby undesired bands fold into the desired signal due to LO (Local Oscillators) harmonics. Harmonic mixing should be addressed by conventional tuners since the conventional TV broadcast uses a wide spectrum range from VHF to UHF. Therefore, the undesired TV bands may fold into the desired TV signal. Although TV-can tuner architecture meets the broadcast TV standard specifications, it needs many discrete components such as coils, varactors and SAW filters that run contrary to the mobile TV application requirements. However, it has a great advantage of reducing the linearity requirement of the RF front-end circuit (i.e., low-noise amplifiers (LNAs) and down-conversion mixers).
1.1 Mobile TV Architectures
5
RF input
On-chip ~ 1GHz
RF-VGA BPF
x
SAW
x
IF output
Recently in [10], this filter was replaced by a digital IR algorithm in the baseband and the SAW filter (1GHz) was also replaced by a digital filter
Channel Selection
Fig. 1.6 Up-down type double-conversion architecture
Furthermore, the high IF frequency of TV-can architecture requires high sampling rate at the following ADC (Analog-to-Digital Converter), which tends to increase the power consumption of the ADC. Thus, this solution is not considered power efficient. In an effort to eliminate the LC tracking filter, up-down double-conversion approaches have been proposed [9–13]. Shown in Fig. 1.6 is the up-down type double conversion architecture, where the first IF frequency is chosen to be 1 GHz or above so that the LO harmonics will be out of the TV band. Since the image is two times the IF frequency away from the desired signal, an LPF (low-pass filter) in front of the tuner is used to attenuate it. The SAW filter, at the first IF path, is used to remove the second mixer image. This architecture is more amenable to integration than the first one. However, increased complexity of circuits and the necessity of using a SAW filter are still obstacles for low-cost and low-power solution tuners. Also, removing the RF tracking filter in this architecture usually complicates the front-end circuit design (it needs to handle the whole TV spectrum as an input). The discussion above illustrates how the choice of a TV architecture along with understanding the limitations of current technologies influences the RF front-end performance requirements (see Fig. 1.7). Integrated TV tuners should meet a set of requirements quite different from those of the existing can-tuners. The task of suppressing harmonic mixing and images on-chip contributes greatly to the complexity of the mobile TV tuner architecture. A more integrated TV tuner architecture uses a selectable on-chip bandpass filter for suppression of the odd harmonics of the LO and a polyphase filter for image rejection [14]. However, this filter requires high power for active filtering, and the in-band image rejection is still limited by the gain and phase imbalance. A recent architecture, published last year, overcomes these challenges by using the polyphase mixer for harmonic rejection and a digital filter for image rejection [15]. In order to clearly and succinctly quantify the impact of the solutions described in this book on the overall performance of mobile TV systems, all recent published mobile TV receivers are simplified in two main architectures. The first one is the direct-conversion architecture (DCR) (also called zero-IF) and the second
6
1 Introduction and Overview Both covered in section 1.1
Tuner Architecture
Technology Limitation
+
covered in section 1.2
+
Mobile TV Standard
RF Front-end Specs
Fig. 1.7 RF front-end specification control factors On-chip
RF input
RF-VGA
x
LPF I
Baseband output
BPF Q
x
LPF
Channel Selection
Fig. 1.8 Direct-conversion architecture
one is the low intermediate frequency (low-IF) double conversion architecture. The direct-conversion architecture, considered one of the most efficient architectures that can be used for mobile TV application, is shown in Fig. 1.8. Since there is no image problem (the IF frequency is zero), simple on-chip RF circuits can be used. However, this architecture suffers from the well-known flicker noise and the DC offsets problems [16]. The information around the center frequency can be lost due to these problems, with the degree of loss depending on the type of modulation scheme. In other words, some modulation schemes inherently can provide a robust solution for the DC offset problem in the DCR architecture. Womac et al. [17] have shown recently that a high pass pole at 300 Hz removes only one OFDM subcarrier from the DVB-H signal, which means that the information
1.1 Mobile TV Architectures
7 on-chip
RF input
RF-VGA
x
0
o
I
+
BPF
LPF
Q
x
o
90
IF output
Image rejection mixer [30]
Channel Selection
Fig. 1.9 Low-IF architecture
can still be recovered using digital signal processing in the demodulator. Accordingly, we can say that although this architecture cannot be considered a good candidate for the traditional analog TV tuner, it may be used in some of the digital TV tuners that are based on the OFDM modulation. A low-IF receiver (shown in Fig. 1.9) with an image rejection mixer [18] is a favorable architecture when the information loss in direct conversion is not tolerable. In this case, usually the image rejection ratio is limited to less than 40 dB due to component mismatch in the image rejection mixer. One of the main advantages of this architecture is that the low IF (~4 MHz) facilitates the digital implementation of the IF channel filtering so that it can be programmed for different channel bandwidth. Additionally, the image can be suppressed in the digital domain. The mobile TV receivers reported in the literature indicate that the DCR architecture is the chosen candidate for DVB-H tuners [17, 19–22]. In the meantime, the low-IF architecture is often used for ISDB-Tss tuners as reported in [23–29]. In this book, the DCR architecture is selected to be the candidate that best facilitates implementing the TV reception function into DVB-H based cell phones. As mentioned earlier, this architecture has the advantages of reduced complexity, fewer external components, and no need for image suppression. A summary of the recent published papers on receivers (RF front-end and baseband) for analog/digital broadcast and mobile TV application is shown in Table 1.1. The main features of each receiver including the harmonics, and image rejection techniques, the dissipated power consumption, and the noise performance are briefly described. The table gives a good overview of the research efforts that have been made to produce low-cost, high-performance tuner chip sets. Knowledge of these systems helps in creating new solutions to implement the RF tuner functions. Unless the system requirements are well defined and well understood, coming up
8
1 Introduction and Overview
Table 1.1 State-of-the-art of analog/digital TV receivers
Mobile TV Standards
Research paper
Technology
Architecture
Image rejection
Harmonic rejection
Noise figure (dB)
Power (mW)
Stevenson 2007 [19]
0.35 µm BiCMOS
DCR
IR mixer a (75 dB)
Up/down
6
330
Gupta 2007 [23]
0.18 µm CMOS
Low-IF
Digitally (61 dB)
Polyphase b mixer
4
540
Fillâtre 2007 [24]
0.25 µm BiCMOS
Low-IF
IR mixer and polyphase filter c (60 dB)
LC tracking filter
5
750
0.5 µm BiCMOS
Low-IF
IR mixer and Polyphase filter
Not mentioned
2.7
87f
0.25 µm BiCMOS
DCR Zero-IF
SAW
2.6
61
0.18 µm CMOS
DCR Zero-IF
Not mentioned
3.5
259
0.35 µm BiCMOS
DCR Zero-IF
SAW
3.6
340
0.18 µm CMOS
Low-IF
IR mixer and polyphase filter (56 dB)
Not mentioned
3
100
Zero-IF
Not mentioned
4.3
185
Digitally (60 dB)
Up/down
6.7
750
Selective LNA
9
150
SAW
~11
160
d
Sakai 2007 [25] g
Peluso 2006 [20] h
Vassiliou 2006 [21] i
Womac 2006 [17] j
Kim 2006 [26] k
Kim 2006 [22] l
Heng 2005 [27] m
Rumpt 2005 [28] o
Azuma 2004 [29]
a
0.18 µm CMOS
DCR
0.25 µm CMOS
Low-IF
BiCMOS
Low-IF
BiCMOS
Low-IF
e
n
(FPHP) filter
p
IR mixer and polyphase filter (37 dB)
Achieved 75 dB IR with using 2 IF-SAWs. On-chip RF tracking filter has been used also for harmonic rejection 72 dB. An auto-calibration technique is implemented to boost the IR. d Dedicated to the ISDB-T Mobile TV standard. e Adaptive control technique has been proposed for power reduction. f IR mixer is based on Hartley architecture [30]. g Dedicated for Media-Flo Mobile TV standard (698−746 MHz). h Dedicated for DVB-T broadcast TV standard. i Dedicated for DVB-H Mobile TV standard. j Dedicated for ISDB-T/T-DMB Mobile TV standard. k Dedicated for DVB-H Mobile TV standard. l Dedicated for NTSC, PAL, and SECAM standards. m Dedicated for PAL, NTSC, SECAM, DVB-H standards. n FPHP: fully integrated programmable high precision filter with a dynamic image suppression system. o Dedicated for ISDB-T Mobile TV receiver. p IR mixer is based on Hartley architecture [30]. b c
1.2 DVB-H Mobile TV System Definitions
9
with innovative ways that enable the multimedia reception to cell phones becomes difficult and risky. The following section presents the system definitions of the DVB-H mobile TV standard and discusses the associated challenges.
1.2 DVB-H Mobile TV System Definitions Among mobile TV standards, DVB-H poses the most stringent sensitivity (the minimum signal that a receiver can detect) and blocking requirements (the robustness of a receiver to interferers). The standard specifies a set of interferer patterns to test the conformity of the tuner [31, 32]. These reception conditions have significant impacts on the tuner specifications as will be described below. The interferer patterns are classified into two categories. One specifies the selectivity test patterns to check the immunity to a single analog or digital TV interferer (shown in Table 1.2). The other category defines the linearity test patterns (shown in Table 1.3), which check the immunity to two analog or digital interferers. In those test patterns, there are some main mechanisms that can create distortion or noise components on the desired channel: 1. The intermodulation among sub-carriers on the interferer channel 2. The clipping that might happen due to a large interferer channel 3. The cross-modulation among sub-carriers on the interferer channel and on the desired channel [33] 4. The decrease in the small-signal gain (gain compression) due to the third order nonlinearity [16] 5. The mixing between interferer components and noise components of the local oscillator output, harmonic mixing [16]. The receiver dynamic range (DR) is generally defined as the ratio of the maximum input level that the receiver can tolerate (DR upper end) to the minimum input level at which the receiver provides the required sensitivity (DR lower end). All of the above described distortion mechanisms can limit the receiver dynamic range in one way or another depending on the system design parameters and its performance. Mechanisms One through Four are discussed in detail throughout this book. Most of the design recommendations that have been presented in this book to optimize the DVB-H receiver DR were based on these four mechanisms. Since mechanism Table 1.2 Selectivity test patterns of the DVB-H standard Pattern Modulation of Interferer Interferer Location U/D a (dB) S1 Analog N+1 38 N + K (K ¹ 0, 1) 40 N+1 29 S2 Digital 40 N + K (K ¹ 0, 1) a U: undesired interferer power (on each channel), D: desired channel power.
U (dBm) −35 −28 −35 −28
10
1 Introduction and Overview
Table 1.3 Linearity test patterns of the DVB-H standard Pattern Modulation of Interferer Interferer Location U/D a (dB) L1 Digital and analog N+2 45 N+4 L2 Analog N + 2, N+4 45 L3 Digital N+2, N+4 40 Desired channel modulation: 16QAM. a U: undesired interferer power (on each channel), D: desired channel power.
U (dBm) −35 −35 −35
Five is controlled by other receiver blocks (the frequency synthesizer) that are out of the scope of this book, it is highlighted very briefly. The lower end of a tuner DR is the sensitivity level, which depends on the tuner noise figure (NF) (a measure of degradation of the output (signal-to-noise) ratio caused by the tuner circuits). In other words, the sensitivity requirements always set the target noise performance of a receiver. For DVB-H standard, the sensitivity of the tuner for 16QAM modulation (Quadrature amplitude modulation) is −86.6 dBm and hence the tuner NF is required to be below 5 dB. Typically, a receiver NF is controlled by the RF front-end noise performance and specifically by the first stage of the RF front-end, which is the low noise amplifier (LNA). This can be seen using Friis formulas as in [16]. The tuner NF is given by
NFtuner = NFLNA +
NFMixer − 1 GLNA
(1.1)
where NFLNA and GLNA represent the noise figure and available power gain of the LNA respectively. NFMixer is the mixer noise figure. As will be shown, achieving a competitive noise performance (NF < 3 dB) for an LNA that consumes low power and supports the broadband nature of TV bands is one of the biggest challenges. The other limit (upper end) that sets the receiver DR is the linearity figure of merit. For the DVB-H receiver, this figure of merit is set by the linearity pattern (L3) [25] as shown in Table 1.3. A graphical illustration of this pattern is given in Fig. 1.10. Due to the third-order nonlinearity, two digital TV interferers that are 40 dB larger than the desired signal can generate a third-order intermodulation product on top of the desired signal. The corruption of the desired signal due to third order intermodulation of two nearby interferers is quantified by “third order intercept point” (IP3). This parameter is measured by a two-tone test as shown in Fig. 1.11. The required input IP3 (IIP3) according to the interferer level can be calculated by [16]
IIP3 =
1 (3Pundesired − Pdesired + S / N ) 2
(1.2)
where Pundesired is the interferer power level, Pdesired is the desired signal power, S is the signal power and N is the noise power. Using L3 pattern parameters and
1.2 DVB-H Mobile TV System Definitions
Desired Signal
11
Digital Interferer
Digital Interferer
−35 dBm
−35 dBm
N+2 (N+16MHz)
N+4 (N+32MHz)
f1
f2
−75 dBm U/D 40 dB N IM3
2f1-f2
Freq.
f
Fig. 1.10 DVB-H L3 pattern requirements
Vout (dB)
DR is limited by the distortion Low-distortion region (0.1%-0.001%)
DR is limited by the noise 3 dB/dB
IM3 1 dB/dB
DR and IM3 is maximum here
IP3 is the IM3 Intercept point
Noise
IM3=1
IP3
Vin (dB)
Fig. 1.11 Third-order intermodulation distortion
Eq. 1.2 the IIP3 of DVB-H tuners is found to be equal to −8 dBm, where the required S/N is supposed to be 13.7 dB for 16QAM modulation. One of the important characteristics of mobile TV standards is that achieving this level of linearity is not required at the receiver sensitivity level. This can be further clarified by finding the NF of the DVB-H tuner in the L3 pattern reception scenario. The NF can be given by:
NF = Psignal − 10 log(kTB) − (S / N )
(1.3)
where k, T, and B represent Boltzman’s constant, temperature, and channel bandwidth, respectively.
12
1 Introduction and Overview
Since the desired channel power of L3 pattern is −75 dBm, the required tuner NF can be estimated to be 15 dB. Therefore, achieving the required −8 dBm IIP3 can come at the expense of the tuner noise performance. As will be shown, this has a great impact on the design of the mobile TV RF front-end architectures. Additional challenges include the presence of GSM interference from a cell phone up-link in DVB-H. GSM handsets incorporating DVB-H may suffer from increased noise floor if the transmitter power amplifier noise leaks into the DVB-H signal band. Also, a +33 dBm GSM carrier may block the desired TV signal and desensitize the tuner. Given the requirements for low cost and low power the above specifications call for careful design and innovation techniques.
1.3 Scope of This Book As previously mentioned, the goal of this book is the exploration, design, implementation, and demonstration of wide dynamic range, low-power, RFICs that facilitate integrating TV reception into cellular phones and other hand-held devices. Towards this goal, a mobile TV receiver that can meet the DVB-H mobile TV standard specifications is proposed as a case study to illustrate the step-by-step design progression and implementation of the RFICs described in the following chapters. As shown in Fig. 1.12, an RF front-end, emphasized in the dashed box, is composed of a variable-gain LNA, a programmable RF passive attenuator and an RF power detector. The front-end amplifies the input RF signal (RFin) over frequencies ranging from 470 to 862 MHz and sends them to a quadrature passive mixer. A detector performs automatic gain control (AGC) to meet the necessary level of carrier-to-noise ratio or carrier-to-intermodulation products (C/I) over the desired bandwidth. The baseband chain consists of a post-mixer amplifier (PMA), and a SD analog-to-digital converter (ADC). The baseband circuits’ performance [34, 35] facilitates extracting the RF front-end circuit specifications needed for DVB-H mobile TV receivers. 65 nm CMOS technology was chosen for the implementation as it allows an exploration of the impact of using nanometer CMOS processes on designing wide dynamic range RFICs. Since the book deals specifically with the circuit design techniques needed to improve receiver dynamic range, the dynamic range physical limits have to be defined. Basically, a receiver’s dynamic range can be limited by the third-order intermodulation distortion, the clipping level, or the phase noise performance. The dynamic range optimization techniques discussed in this book will focus only on the first two. Hence, the RF building blocks which might affect the receiver phase noise will be highlighted very briefly (i.e., the mixer and the voltage controlled oscillator (VCO)). This book contains six chapters: Chapter 2 presents the challenges of achieving low noise figure in wideband CMOS LNAs. It begins by reviewing the MOSFET
1.3 Scope of This Book
13
RF Detector
Quadrature Passive Mixer
Post Mixer Amplifier
PMA
Σ∆ADC
ADC I
RFin LNA RF Attenuator
1/4
RF Front-End
Q PMA
ADC
VCO
Gain
6dB
21dB
-
Noise
1.5dB
10nv/Hz 0.5
60nv/Hz 0.5
IIP3
+20dB
-
-
Fig. 1.12 Block diagram of a DVB-H mobile TV receiver
noise mechanisms at RF frequencies and the different sources of transistor distortion with the purpose of defining the MOSFET dynamic range. Then, the main CMOS LNA topologies including the resistor-terminated common-source amplifier (R-CS), shunt feedback (SFB) amplifier, common-gate (CG) amplifier, and wideband inductively degenerated (L-degenerate) amplifier are analyzed and discussed in the context of their dynamic range and their applicability for mobile TV application. The recent noise-cancelling CMOS LNAs are also presented and solutions to design an LNA that can meet the DVB-H standard requirements are given. Chapter 3 presents a silicon prototype of a noise-cancelling LNA that can meet the mobile TV standard requirements. This chapter introduces the reader to the challenges associated with the practical realization of a mobile TV LNA in 65 nm CMOS technology, including the MOSFET low output resistance, the high resistivity of the polysilicon material, the increased substrate coupling, and the use of a digital transistor layout that is not optimized for RF operation. The bias circuit generators as well as the measurement environment are also discussed in this chapter. Chapter 4 continues by describing the use of an RF automatic gain control technique to enhance receiver linearity. This chapter discusses the two possible ways needed to achieve gain control in RF front-ends; namely, by active gain control through a variable-gain LNA or by passive gain control through a programmable passive attenuator. The chapter ends by presenting measurement results of a prototype of
14
1 Introduction and Overview
an RF passive linearization circuit in 65 nm CMOS technology. Chapter 5 describes a wide dynamic range RF front-end architecture implementation with an AGC algorithm. Chapter 6 concludes by summarizing the RFICs developed in this book and outlining further areas of research required to finalize the mobile TV tuner. Finally, it should be stated that some of the discussions in these chapters rely on knowledge of areas outside the scope of this book. Readers are therefore encouraged to refer to the references recommended throughout the text for further study.
Chapter 2
Wideband CMOS LNA Design Techniques
The low-noise amplifier (LNA) is the backbone of any radio frequency (RF) communication receiver. Its specifications define the overall receiver noise performance and can have deleterious effects on the overall linearity. CMOS LNAs specifically receive intense attention because they help in achieving a one-chip solution by integrating the LNA with the receiver’s baseband digital signal processing blocks that are inherently realized in CMOS technology. The one-chip solution reduces overall package cost and form factor. Moreover, it saves the power required to drive package pins in the multi-chip solution [36, 37]. Wideband LNAs are used in the receiver processing of several signal channels such as TV cable modems, multi-band mobile terminals [38], software defined radio [39, 40], and ultra-wideband applications [41, 42]. The high sensitivity required for these applications demands a noise figure below 3 dB over a wide frequency range. In the meantime these LNAs have to provide a wideband input match to a 50 W source impedance to limit reflections on a cable or to avoid changing the characteristics of the RF filter preceding the LNA. Also, they have to provide sufficient gain and good isolation. Achieving all these requirements with moderate power consumption is a challenge especially without having the benefit of using resonance input circuits and loads as is the case with narrowband LNAs. An additional requirement that was previously ignored or considered as a minor constraint in LNA design is the nonlinearity factor. In some applications like mobile TV, large blocking interferers exist nearby the desired channel requiring a very linear LNA. An LNA with a large dynamic range (DR) is key for such applications. The dynamic range of wideband amplifiers that works well in other applications like high-speed optical transceivers may not be good enough for mobile TV systems. In fact, the LNA design is one of the main challenges. In this chapter, the foundation for the design of a wideband LNA that can meet the DVB-H mobile TV specifications has been defined. To provide some background, Section 2.1 describes the fundamental physical limits of MOSFET dynamic range in CMOS. Based on this, the four traditional topologies of CMOS LNAs are
A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_2, © Springer Science + Business Media B.V. 2010
15
16
2 Wideband CMOS LNA Design Techniques
discussed. Their limitations in achieving the required noise performance over several octaves of frequency with low-power consumption are discussed in Section 2.2. The chapter continues by presenting most of the research efforts that have been made to push the state-of-the-art of wideband CMOS LNA performance. Section 2.3 presents different LNA design techniques proposed recently in the literature to improve the noise and linearity performance and to decrease the power consumption. LNAs based on noise-cancelling approaches are selected as a candidate that can achieve mobile TV requirements. New techniques are described to overcome the limitation in existing noise-cancelling techniques and hence improve the dynamic range of the current wideband LNAs. The associated theory as well as the design methodology is given in Section 2.4.
2.1 Dynamic Range Limits in MOSFETs The LNA must be able to simultaneously handle very small signals close to the noise level and very high signal levels close to the transmitter antenna. Therefore, the noise and the distortion are both important and both define the dynamic range in any device. Its lower bound is limited by the noise and the upper bound is limited by the distortion as explained in Chapter 1. Understanding different mechanisms that can control the MOSFET dynamic range helps in optimizing the CMOS LNA’s performance. In this section, the MOSFET noise mechanisms in RF are discussed with the goal of identifying the potential noise sources and presenting ways to either minimize them or minimize their contribution. The minimum noise figure NFmin that a MOSFET can achieve is briefly discussed and is used in this book as a reference to compare between different LNA topologies. This section then reviews the different sources of nonlinear distortion in a MOSFET and presents the feedback mechanism as a common technique used to reduce the transistor distortion. Throughout the discussion, the implications of using 65 nm CMOS technology are highlighted.
2.1.1 The Noise Limit This section describes different noise sources in a MOSFET including, the thermal noise, flicker noise, and substrate noise. The classical noise optimization theory is also discussed. 2.1.1.1 Noise Mechanisms in MOSFETs There are different sources of noise in a MOSFET that can control its noise performance. Since these devices are basically voltage-controlled resistors, they exhibit a channel thermal noise known as a drain noise current.
2.1 Dynamic Range Limits in MOSFETs
17
This drain noise current can be capacitively coupled through the gate-channel capacitance, and results in what is called induced gate noise. The induced gate noise current is obviously correlated with the drain noise current. However, the induced gate noise current is only relevant at very high frequencies, beyond ft /5 [43], where ft is the frequency at which the small-signal short circuit current gain is unity. For 65 nm CMOS technology where ft equals about 160 GHz, it is not expected for the induced gate noise to have that great of an impact in the overall MOSFET noise performance in commercial applications at much lower frequencies. The noise current due to drain noise and induced gate noise can be modeled as current sources as shown in Fig. 2.1, where Cgs is the gate-source capacitance, Cgd is the gate-drain capacitance, gds is the channel conductance, and gm is the transconductance. The mean-squared noise current spectral densities for these sources are
2 ind = 4 kTggdo Df
(2.1)
2 ing = 4 kTdgg Df
(2.2)
where gdo is the drain-source conductance of the transistor when the drain-source voltage vds = 0, gg = (wCgs)2/5gdo, Df is the noise bandwidth, k is Boltzman’s constant, g and d are the coefficients of the drain noise and the induced gate noise respectively (g = 2/3 and d = 4/3 for long channel devices) and T is the absolute temperature [44, 45]. Despite the fact that g and d values in nanometer CMOS have been under extensive research for the last 20 years [46–49], their exact values have not yet accurately been defined. However, most experimental papers have largely confirmed that their values increase with smaller transistor gate lengths. Some of them indicate that this is due to the channel length modulation and carrier heating because of the high electric field associated with decreasing the channel length [50, 51]. Others proclaim that their increase is due to a diffusion current that flows in the channel on the source side with associated shot noise [52, 53]. It has been reported that they are about 50% larger for 180 nm CMOS and double for 130 nm CMOS technology [49]. In this work, by comparing simulation results with measured values, we were able
Cgd
Gate
2
ing
Drain gmvgs
Cgs
2
ind
gds
Source 2 Fig. 2.1 The small-signal AC model of a MOSFET device with its noise generators. ind repre2 sents the drain noise current and ing represents the induced gate noise current
18
2 Wideband CMOS LNA Design Techniques
to show that their values increased by about 65% in 65 nm CMOS technology. g = 1.45 and d = 3 are used in this work. Another source of thermal noise in a MOSFET is the poly-gate resistor RG. Even if the gate material is highly doped, it can make a large contribution to the overall noise performance in the MOSFET, depending on the width of the gate fingers and on how far the fingers extend beyond the active area of the channel. The question that then arises is how does one quantify the contribution of each noise source to the overall noise performance of a MOSFET? Furthermore, how does one then minimize each source contribution? Referring these noise sources to the input of a MOS device as a single noise voltage generator helps in quantifying each noise generator’s effect on the overall noise behavior. The input noise voltage is called 2 = 4 kTReff Df ). the equivalent input noise voltage ( vieq This voltage can be minimized by decreasing the MOSFET overdrive voltage (Vgs − Vth) for a fixed current, where Vgs is the DC gate-source voltage and Vth is the device threshold voltage or by increasing W/L ( W is the transistor width and L is the transistor length) and the drain current. This holds true so long as the gate resistance is small. The gate resistance can be decreased by contacting the gate on both sides and using multiple gate fingers [54, 55]. Another noise source that can eventually dominate the MOSFET noise is the substrate back-gate resistance (RB). This noise is always present even if the substrate contact is shorted to the source contacts. The RB noise can be referred to the input by multiplying its contribution by (n − 1)2 where n − 1 is the ratio between the bulkback gate transconductance (gmb) to the MOS transconductance (gm) as shown in Fig. 2.2. Using sufficient substrate contacts around the transistor gate fingers can greatly eliminate the contribution of RB noise [56]. The last contribution to the equivalent input noise voltage is given by the series resistor in the source itself. Normally, this is quite small, but depends on the effective channel length. The realization of low-noise MOSFETs requires the simultaneous minimization of the four contributions. In addition to these types of noise discussed above, MOSFETs generate a 1/f flicker noise. This is due to the surface states. The flicker noise can be modeled as an input-referred noise voltage generator [57]
vieq
Reff
Cgd
RB +
Cgs gmvgs
gmbvbs
gds
vbs −
Reff = γ /gm+RG+RB(n-1)
2
(n-1)=gmb /gm ~ 0.2-0.5
Fig. 2.2 The small-signal AC model of a MOSFET with the equivalent input noise voltage generator. The total equivalent input noise voltage includes the contribution of the substrate resis2 2 tance R , the channel noise current, and the gate noise R . At high frequency ing =(Cgsw)2 vieq B G should be added. It is correlated with the noise voltage
2.1 Dynamic Range Limits in MOSFETs 2 vieqf =
19
KF Df WLCox2 f
(2.3)
where KF is a constant dependent on the device type, W and L are the transistor dimensions, Cox is the gate capacitance per unit area, f is the frequency, and Df is the bandwidth. Although most LNA designers ignore the flicker noise because of the LNA high frequency of operation, this cannot be done in cases where the LNA’s nonlinearity is taken into account. In fact due to the LNA nonlinearity, a large interferer blocker can upconvert the low frequency noise (i.e., 1/f noise) to the RF bands. Therefore, large signal noise analysis is mandatory in some cases of LNA design. In the examples described in this book, such analysis has been conducted to ensure the robustness of achieved noise performance in all wireless environment scenarios. One can note that a MOSFET with thin oxide and a large area minimize the 1/f noise. 2.1.1.2 Minimum Noise Figure In order to identify the minimum noise figure that a MOSFET can achieve, the classical noise optimization technique is reviewed very briefly. The theory shows that there is an optimum signal-source admittance (Yopt=Gopt + jBopt) for a given MOSFET that leads to the lowest noise figure. The optimum source admittance, is given by [58]
g Yopt = m × wCgs K1 − jwCgs (1 − K 2) gdo
(2.4)
and the corresponding minimum noise figure NFmin is given by NFmin = 1 + 2g
w K1 wT
(2.5)
where K1 and K2 are technological parameters and wT = gm/Cgs. The expression in (2.4) shows that the optimum susceptance required to achieve NFmin is basically an inductor. Hence, achieving a broadband minimum noise figure is fundamentally difficult. The optimization also shows that the optimum conductance is that which balances the contribution of drain and gate noise generators. As a signal source admittance Ys = Gs + jBs deviates from Yopt, the noise figure degrades according to
NF = NFmin +
ggdo 1 (Gs − Gopt )2 + ( Bs − Bopt )2 . gm2 Gs
(2.6)
2.1.2 The Distortion Limit In this section, we describe the main sources of distortion in MOSFETs. The linear distortion that is a result of filter action (i.e, nonlinear distortion) is not considered here as it is not a concern in LNA designs. Also, since LNAs usually deal with
20
2 Wideband CMOS LNA Design Techniques
small signal levels, we limit our discussion to low levels of distortion, on the order of −60 dB down to −100 dB. The high level of distortion that is a result of clipping is not considered. We are therefore only interested in what are called soft nonlinearities, which give a small amount of distortion. The main sources that generate soft distortions in CMOS are the transconductance (gm) nonlinear behavior and the MOSFET output resistance (ro=1/gds) nonlinearity. As will be shown, the latter has a great impact in specifying the level of distortion in nanometer CMOS. The other sources such as bulk and depletion capacitance nonlinearity are also briefly mentioned. 2.1.2.1 The Distortion Mechanisms in MOSFETs As in any nonlinear system, a MOSFET transfer function can be described by a power series: io = ao+ a1vin + a2vin2 + a3vin3 +¼ . The coefficients of this power series (ai) determine the amount of distortion that a MOSFET can generate. Among these coefficients, a3 is particularly important because it controls the third-order intermodulation distortion (IM3) at low signal levels and, thus, determines the IIP3 value. In our discussion, we will intuitively specify the main factors that can control the magnitude of these coefficients and hence can control the MOSFET nonlinearity. However, more detailed analysis and derivations about these coefficients can be found in [59]. Now, let us discuss the distortion in MOSFETs that result from the gm nonlinearity. Figure 2.3 illustrates the relation between the transconductance (gm) and the overdrive voltage (Vod) for a MOSFET with a length of 65 nm and width of 100 µm. It shows that the overdrive voltage range that is available to bias a 65 nm MOSFET in strong inversion (the square-law region) is between 200 and 320 mV. This is quite small compared to the one that is available for a 0.18 µm MOSFET, which is about 250 mV, as reported in [60]. This difference is due to the velocity gm mS 0.4
Vod = 0.32V Weak Inversion
Strong Inversion
Velocity Saturation gmsat
Vod = 0.2V
square-law region becomes smaller in nanometer CMOS
0.2 Vth
0
0.2
0.32
Vgs-Vth
Vsat
Fig. 2.3 Transconductance (gm) versus overdrive voltage (Vod = Vgs − Vth) for a MOSFET with a minimum feature length of 65 nm. Sixty-five nm CMOS models are used in this simulation (the transistor width = 100 µm)
2.1 Dynamic Range Limits in MOSFETs
21
saturation voltage Vsat scaling (in Fig. 2.3) as the channel length decreases. Shrinking the strong inversion region makes the simple square-law model invalid and inaccurate to represent the MOSFET operation. Furthermore, the curvature of gm curve causes a lot of distortion. The strong inversion region can be seen as where the gm curvature is inverted. The exponential of the weak inversion region curves upwards whereas flattening in velocity saturation gives a downwards curvature. As a result of this, the second derivative of the gm would go through zero and hence makes IM3 equal to zero. As an example, we measured IIP3 for a MOSFET with L = 65 nm and W = 150 µm for different values of gate-source voltages Vgs. As can be seen in Fig. 2.4, there is a Vgs value (Vgsopt) at which the IM3 would equal zero and IIP3 is maximum. However, this peak in IIP3 is very narrow and requires an accurate biasing within ±10 mV of the optimum voltage. Manually tuned bias circuits are not capable of such precision in the presence of process and temperature variations. A research effort has been made by Aparin and Larson in [60] to improve the robustness in generating Vgsopt. However, their bias circuit that can keep track of process and temperature variation while cancelling IM3 is very sensitive to the device mismatch and requires an expensive thin-oxide technology.
IIP3 (dBm)
15
Optimum gate bias (Vgsopt) required to cancel IM3
5.0
−5.0
0.25
1
Vgs (V) Fig. 2.4 Simulation results of a MOSFET IIP3 in 65 nm CMOS technology. The gate-source voltage (Vgs) has been swept between 0.1 and 1.2 V. The MOSFET has a length of L=65 nm and a width of W=150 µm. The peak IIP3 is located at the transition from the weak inversion to the strong inversion mode of operation point
22
2 Wideband CMOS LNA Design Techniques
In addition to the gm nonlinearity, a MOSFET has many more distortion components. Voltage variations at the drain or at the bulk also generate distortion. In general, the AC small-signal drain current (ids) can be expanded into the following power series in terms of the small-signal gate-source voltage (vgs) around the bias point [61]: ids = gmvgs+ K2gmvgs2 + K3gmvgs3 + govds + K2govds2 + K3govds3+ gmbvbs + K2gmbvbs2 + K3gmbvbs3 + K2gm&gmbvgsvbs + K3,2gm&gmbvgs2vbs+ K3,gm&2gmbvgsvbs2 +.... K3gm&gmb&govgsvdsvbs
(2.7)
where Kigm are the coefficients of the transconductance nonlinearity, Kigo are the coefficients of the output resistance nonlinearity and, Kigmb are the coefficients of the bulk transconductance (gmb) nonlinearity. A natural question would be how can we control these coefficients? Kigm coefficients depend on the value of the DC overdrive voltage used to bias the device (Vgs − Vth). For low distortion, a large overdrive voltage must be used. As for Kigo coefficients, they depend on the DC drain-source voltage bias Vds. These are really
Strong inversion
IIP3 (dBm)
15
5.0
RLoad = 50 Ohms
−5.0
RLoad = 15 Ohms 0.2
Vds (V)
0.8
Fig. 2.5 Simulation results for a MOSFET IIP3 with the drain-source voltage Vds sweep. The output swing varies by changing the load choke RLoad. The MOSFET has an L = 65 nm and W = 150 µm. The Vds is biased through an RF choke with ideal power supply. The power supply varies from 0.1 to 1.1 V
2.1 Dynamic Range Limits in MOSFETs
23
important as the output resistance of a nanometer transistor has become quite small. Figure 2.5 shows the simulated IIP3 of a MOSFET with L = 65 nm and W = 150 µm with varying drain-source voltage Vds. Two curves have been generated by changing the output voltage swing applied on the device by using two different loads. As can be seen, for relatively low output voltage swing (Rload = 15 W), the IIP3 is independent of the Vds variations. This can be explained by the fact that the IIP3 is not dominated by the ro nonlinearity. However, in the case of relatively large output voltage swing (Rload = 50 W), ro nonlinearity is brought into the picture. The fact that IIP3 improves with increasing Vds values in the strong inversion region demonstrates that ro nonlinearity heavily impacts the distortion level in this case. Another implication of dominating ro nonlinearity is that it makes the scaled transistors experience a shift of the IIP3 peak to higher current densities requiring large power consumptions to achieve high linearity. This can be seen in the plot shown in Fig. 2.6. In this plot, the Vds bias is swept and IIP3 is measured for two different transistors with L1 = 65 nm and L2 = 130 nm. Both of these have the same W/L. Also, it is noted that the longer channel provides better IIP3 value as it provides better ro linearity. As a result, moving to the 65 nanometer technology makes the design trade-offs much more difficult and very tight. In summary, as in the case of LNA design, the designer should know if the design is limited by the input nonlinearity (gm) or the output nonlinearity (ro). This will facilitate the process of IIP3 optimization.
L= 130 nm L= 65 nm
IIP3 (dBm)
5.0
L=130 nm
−5.0
L=65 nm
0.2
Vds (V)
0.8
Fig. 2.6 Simulation results for a MOSFET IIP3 with the drain-source voltage Vds sweep for two different transistors. The first has an L = 65 nm and the second has an L = 130 nm. Both transistors have the same current swing and the same W/L ratio. 65 nm CMOS models are used in both cases
24
2 Wideband CMOS LNA Design Techniques
Other sources of distortions such as the bulk and depletion capacitance distortion will not be discussed in detail as they are expected to have a very limited effect on the mobile TV LNA IIP3 described in this book. The bulk nonlinearity is governed by (gmb) and source depletion capacitances nonlinearity is governed by the squareroot characteristic in this equation [61]: Cj =
Co 1−
vin j
(2.8)
where Cj is the junction capacitance, Co is Cj at zero volts, vin is the reverse bias voltage and, F is the junction built-in voltage (0.6–0.8 V). 2.1.2.2 Distortion Reduction by Feedback The most common technique to reduce the distortion is by application of feedback. As explained in the previous section, one can decrease the distortion in a MOSFET by increasing either Vgs or Vds depending on the type of the nonlinearity that dominates. Also, smaller signals cause less distortion. Thus, one can say that in general the distortion in MOSFETs improves by decreasing the relative signal swing (i.e., vgs/Vod for the input nonlinearity), and as will be discussed that is exactly what the feedback does. Feedback reduces the relative signal swing by the loop gain value. Moreover, the distortion should also be decreased by the feedback loop gain [59]. To study the distortion while taking into account the effect of the feedback on the signal swing, let us take a single MOSFET amplifier with a series resistor Rs in the source. The amplifier is shown in Fig. 2.7. It is biased so that its IIP3 is limited by the input nonlinearity. A plot that shows the effect of changing the amplifier loop gain (gmRs) on the distortion generated by this amplifier is also shown in the same figure. The DC current is kept constant and so is the relative current swing. This means that the input voltage Vin is constant until the value of the loop gain is equal to unity and then increases with the loop gain. As can be seen in this plot for values of the loop gain larger than unity, both IM2 and IM3 decrease with the same slope of −20 dB/ decade. This is expected since the distortion should be reduced by the loop gain. A very interesting conclusion can now be made. Contrary to what we have already mentioned about the distortion in MOSFETs, the overdrive voltage (Vgs − Vth) should decrease for a fixed current to reduce the distortion for a MOSFET that is in a feedback loop. This is because in this case the effect of the increase in gm, and hence in loop gain is more important than the increase of the overdrive voltage. Although this technique can be used to improve the MOSFET IIP3 value, the improvement achieved in IIP3 won’t benefit the MOSFET dynamic range. Adding Rs leads to an increase in the thermal noise of this amplifier and hence just causes a shift in the dynamic range. This technique is used in the LNA described in the next chapter after describing a way to cancel the noise added by Rs so that the amplifier can benefit from the IIP3 boosting.
2.1 Dynamic Range Limits in MOSFETs
a
25
b Iout
IM
After this point VRS > Vod / 2
% 10
Vin
VRS
IM2
1.0
+ −
IM3
0.1
Rs
Same slopes
0.01 1 0.01 0.1
10
1.0
100
10
1K 100
Rs
Loop Gain
Fig. 2.7 A common-source amplifier with a series feedback resistor Rs is shown in (a) and its distortion plot is shown in (b). In this plot the second-order intermodulation distortion IM2 and third-order intermodulation distortion IM3 are calculated at different values of Rs. The relative current swing remains constant by increasing Vin at all points where loop gain is >1. The IM2 and IM3 values decrease with the same slope of −20 dB/decade after the loop gain values become larger than unity
2.1.3 Dynamic Range Trade-offs in CMOS In this section, guidelines to optimize the MOSFET dynamic range (DR) are presented. As it has been shown earlier in this chapter, there is a trade-off between the noise performance of a MOSFET and the amount of distortion that it generates. Improving the noise figure by decreasing the overdrive voltage (Vgs − Vth) at fixed current consumption directly hurts the IIP3 value (assuming that the distortion is limited by the MOSFET input nonlinearity). Although this might not be a problem when the LNAs IIP3 is relaxed according to the targeted application, it becomes a real challenge in cases where linear LNAs are required. The only factor that can be used to increase the dynamic range and relax this trade-off is the power dissipation factor. The dynamic range of a MOSFET amplifier can be given by [62]:
DR =
IIP3 µIxZ o NF − 1
(2.9)
where I is the current consumed by the amplifier and Zo is the amplifier input impedance. As shown in (2.9), increasing the MOSFET DR requires consuming more current. This can be noted by exploring the research efforts that have been made to push the state-of-the-art of LNA performance in the last 20 years. Most of the efforts were focused on inventing new circuit design techniques to reduce the power consumption in LNAs. In other words, researchers were trying to improve LNA performance in terms of dynamic range while keeping the power consumption at a
26
2 Wideband CMOS LNA Design Techniques
modest level. Although there have been many outstanding LNAs proposed, there is still a need for novel LNA designs that can satisfy the stringent requirements of some of the new emerging applications such as mobile TV. In the past couple of years LNA research has shifted to other approaches. Instead of trying to minimize the magnitude of the noise and distortion generators, efforts have focused on trying to keep them as they are but cancel their contributions. Before going into detail about this approach, the main conventional LNA topologies are reviewed in the next section to provide the background foundation needed to discuss the recent trends in CMOS LNA design.
2.2 Traditional CMOS LNA Topologies In the design of low-noise amplifiers (LNAs) in wireless receivers, there are several common goals. These include low noise figure (NF < 3 dB), reasonable gain with sufficient linearity, a stable 50 W input impedance and low-power consumption, which is needed in portable systems. Satisfying all the design goals with the broad bandwidth required by some of the wireless applications like the mobile TV is particularly difficult compared to conventional wireless receivers. To demonstrate this, the main CMOS LNA topologies including the resistor-terminated commonsource amplifier (R-CS), shunt feedback (SFB) amplifier, common-gate (CG) amplifier and, inductively degenerated (L-degenerate) amplifier are analyzed and discussed in the context of the mobile TV application. We show that adding a 50 W input matching requirement sets the power consumption level in some of these amplifiers (i.e., CG & SFB amplifiers) and therefore limits the dynamic range that they can achieve. Moreover, trying to decouple the input matching from the power consumption as in the R-CS amplifier merely slides the dynamic range by raising the noise figure and IIP3 together. The L-degenerate topology can offer the right properties by its capabilities of achieving the NFmin of a MOSFET as prescribed by the classical optimization (Section 2.1.1.2) but unfortunately just in a narrow band around a single frequency. A complete dynamic range study for these amplifiers is presented in the following subsections. This study also shows that adding more design constraints other than the dynamic range (i.e., input matching) complicates the compromise and tightens the trade-offs even more. In the described analysis, the induced gate noise of MOSFETs (Section 2.1.1.1) is ignored because of the relatively low operating frequency of the LNA designed for mobile TV application.
2.2.1 R-CS Amplifier The first topology uses a resistive termination at the input of a common-source amplifier to provide a 50 W input matching [63, 64]. To quantify the effect of adding this resistor, consider a transistor to which a resistor aRs is added, as shown in Fig. 2.8.
2.2 Traditional CMOS LNA Topologies
27
Fig. 2.8 A resistive terminated commonsource amplifier
RFout
Rs
RFin
αRs Zin
The NF of this amplifier can be written as a function of the source resistance Rs and the transconductance gm to be: 1+ a 1+ a g + . a gm Rs a 2
NFR −CS =
(2.10)
Indeed the NF can be examined in two cases, with perfect matching (a = 1) and without (a = •). These are as follows
NFR −CS = 1 +
g ,a = ∞ gm Rs
(2.11)
NFR −CS = 2 +
4g , a = 1 . gm Rs
(2.12)
As can be seen from (2.11) and (2.12), the resistive termination degrades the achieved NF. Two effects are responsible for this degradation. First, the added resistor contributes its own thermal noise to the output, which is equal to the contribution of Rs. This results in a factor of two difference in the first terms of (2.11) and (2.12). Second, the input is attenuated leading to the factor of four difference in the second term of (2.11) and (2.12). The large noise penalty resulting from these effects pushes the R-CS amplifier noise figure further from NFmin of the transistor. Therefore, although this topology offers the possibilities of having very lowpower design (gm is decoupled from Zin) across broad bandwidth, the achieved noise figure (NFR−CS > 3dB) makes it unattractive for mobile TV application. Also, adding this resistor would not benefit the DR at all since the NF and IIP3 both shift in the same direction.
28
2 Wideband CMOS LNA Design Techniques
Fig. 2.9 A common-gate amplifier
Rs RFout
RFin Zin
2.2.2 CG Amplifier Another topology that can provide a 50 W input matching is the common-gate amplifier (shown in Fig. 2.9) [65]. The input impedance of this LNA is controlled by the transistor transconductance (Zin ~ 1/gm). Hence, gm of this amplifier has to be at least 20 mA/V. In other words, the matching determines the power consumption of this amplifier. Moreover, the matching also has an effect on the achieved NF. The noise figure of this topology at matching can be given as follows:
NFCG = 1 + g .
(2.13)
The achieved NF of this amplifier is usually larger than 3 dB for short channel MOSFET devices (NFCG > 3 dB). The only thing preferable about this amplifier is that the achieved matching has a broadband nature. However, one can note that it suffers from the low-voltage gain associated with this wideband feature. For a given capacitive load (i.e., from a mixer), the load resistance must be lowered to push out the output pole (~200 W), this leads to an LNA with a voltage gain of 12 dB. Moreover, the load resistor noise raises the NF to about 5 dB. In terms of linearity, this amplifier offers higher values of IIP3 compared to the R-CS amplifier because it is current driven. So, as a conclusion, the features that the CG amplifier can offer don’t fit with the target application figure of merits.
2.2.3 SFB Amplifier Another example of an amplifier that can provide broadband matching to a 50 W source resistor Rs is the resistive shunt feedback amplifier that is shown in Fig. 2.10 [66]. Since the added resistor RF is in the feedback path, its value is larger than 50 W due to the advantage of Miller effect and therefore its noise contribution is less than that of Rs. Hence, in terms of noise figure, this topology provides better NF
2.2 Traditional CMOS LNA Topologies
29
Fig. 2.10 A resistive shunt feedback amplifier RL RF
RFout
Rs Av RFin Zin
than that of the R-CS. Moreover, this amplifier doesn’t suffer from the fundamental trade-off between its NF and input matching like the CG amplifier does. To better visualize this, let us examine the amplifier input impedance ZinSFB, the gain AvSFB, and noise figure NFSFB. The input impedance can be derived to be ZinSFB =
RF + RL 1 + gm RL
(2.14)
where RF is the feedback resistor, gm is the MOSFET transconductance, and RL is the load resistor. The MOSFET channel length modulation and the parasitic capacitances (Cgs and Cgd) are ignored in this expression. The amplifier voltage gain without taking the matching into account is given by AvSFB =
1 / RF − gm 1 / RL + 1 / RF
(2.15)
and the noise figure under the constraint of input match can be derived to be NFSFB ≈ 1 +
1 (1 − gm RF )2
RF 2 1 1 1 ( gm RF − 1)2 (2.16) + + ggm + + Rs ggm + . RF RL RL RF Rs
In this derivation, in addition to neglecting the transistor channel length modulation, it has been assumed that RL ≫ RF so that the amplifier gain is controlled mainly by the feedback resistor RF. Now looking at (2.14), (2.15), and (2.16), one can note that since the amplifier is not constrained by the input matching, its contribution to the noise figure NFSFB can be arbitrarily small by increasing gm of the transistors at the expense of the power consumption. However, although the matching/noise figure trade-off is broken in this topology, they are still directly coupled as shown in (2.14) and (2.16); both
30
2 Wideband CMOS LNA Design Techniques
depend on RL and RF. Because of this coupling, it is generally difficult to achieve an arbitrarily low noise figure for an input impedance of 50 W with reasonable power consumption [58, 67]. However, we still haven’t quantified exactly what is the minimum noise figure that this amplifier can achieve. Can this amplifier achieve a NF < 3 dB required by the mobile TV receiver? To give an answer to this question some assumptions have been made to simplify the noise figure expression in (2.16). The drain noise contribution of the main transistor is neglected. To examine the validity of this simplifying assumption, let us find the noise contribution of the main transistor in the SFB amplifier. The output noise current of the transistor can be given by
2 inout =
in2 4 kTg Df ≈ (1 + gm Rs )2 gm Rs 2
(2.17)
2 where inout is the mean-squared noise current output from the transistor that flows through resistors Rs and RF, and in2 is the mean squared channel thermal noise of the transistor. At reasonably high values of gm the transistor doesn’t contribute to the total output noise. Given that value of gm, the load resistor RL noise current can be also neglected. Therefore, the amplifier noise figure is only limited by the feedback resistor RF. Now, looking at (2.16), one can assume that at reasonably high value of RF, the amount of the noise current that leaks from this resistor to the input can be neglected as well and hence the amount that leaks to the output determines the minimum noise figure that can be achieved by this topology. The minimum noise figure NFminSFB can be derived to be
NFmin SFB = 1 +
1 1 + Av
(2.18)
which can be below 2 (i.e., 3 dB), providing adequate gain Av = gmRL is available. Despite its good noise performance, this amplifier suffers from an insufficient amount of gain and thus requires multiple cascaded stages to be used within the feedback loop. This makes its operation prone to instability [68, 69]. As for nonlinearity factor, one can expect that this amplifier would have better IIP3 values than that of the R-CS and CG amplifiers due to the negative FB. However, the loop gain expression implies different results. The loop gain of the SFB amplifier under the matching constraint can be derived assuming ZL ≫ ZF to be
Av . 2 + Av
(2.19)
As can be seen in (2.19), the open loop gain at ZinSFB = 50 W condition is below 1, therefore the closed loop linearity is not much better than that of the transistor itself. In fact it would get worse at higher frequencies due to decreasing the loop gain. In conclusion, the power consumption as well as the gain associated with the SFB amplifier makes it an unlikely candidate for use in mobile TV applications.
2.2 Traditional CMOS LNA Topologies
31
2.2.4 L-Degenerate Amplifier Inductor degeneration in common-source amplifiers was introduced by Van Der Ziel and Strutt to generate the real part needed to match the input impedance [70]. By decoupling the input impedance from the noise, these topologies allow the optimization of the dynamic range with reasonable power consumption. To better visualize this, consider the circuit shown in Fig. 2.11. The degeneration inductor Ls results in an equivalent input resistance given by: Ri = w T Ls (2.20) where wT is the short circuit unity current gain frequency of the transistor. The resulting equivalent circuit of the degenerated transistor consists of Cgs, Ri, and Ls as also shown in the same figure. The input impedance is purely resistive at the resonance frequency (wo) of Ls and Cgs. In practice, an inductor Lg is added in series to align the series resonance frequency with the desired frequency of operation. The unique property of this amplifier is that it can achieve the minimum noise figure NFmin of a MOSFET proposed by the classical optimization theory (see Section 2.1.1.2). In fact, a simulated NF of about 0.5 dB has been achieved at 1.8 GHz while gmRs = 1, Cgs = 0.5 pF, (Lg + Ls =15 nH), and Lg and Ls are lossless. This is an incredible result compared with any other amplifier noise figure. Assuming ideal inductors, this amplifier produces a noise figure NFL−deg equal to 2
w NFL − deg = 1 + ggm Rs o . wT
(2.21)
It is noted that the noise figure improves quadratically with the transistor unity gain frequency (wT) for a given resonance frequency (wo). To obtain the best noise figure, the transistor should be biased at the maximum of wT. One can note also that due to the noiseless inductive degeneration Ls, (2.20) and (2.21) are decoupled.
Iout Iin
Rs
Lg
Ls
Cgs
Zeq
RFin
Ri = ω TLs Zin = Rs (at resonanace )
Zeq
Ls
Degenerated transistor equivalent circuit
Fig. 2.11 Equivalent circuit for the inductively degenerated transistor
32
2 Wideband CMOS LNA Design Techniques
Therefore, the input-matching noise-figure trade-off limitation that exists in all other amplifiers discussed so far (R-CS, CG, SFB amplifiers) is broken here. However, in some cases, the minimum degeneration inductance is limited because of the packaging considerations. For a wirebond package, the minimum degeneration inductor even in the case of multiple parallel downbonds cannot be lower than 0.5 nH because of the mutual inductance. As can be seen in (2.20), this would limit the maximum allowable wT needed to achieve a 50 W input resistance and would thus affect the noise figure. A passive LC network can be used to relax this coupling. The network shown in Fig. 2.12 can be used to transfer down a real resistance in a narrow band of frequencies. A transformation ratio (a) is used in the matching network to lower the input resistance to 50 W. This ratio is given by a=
Rs ' . Rs
(2.22)
As a result, for a given Ls degeneration, the resulting NF can be given by 2
NFL − deg_ a
w 2 w = 1 + ggm Rs o a 2 = 1 + ggm Rs ' o Ls. w w T
(2.23)
T
Although this topology can achieve superior noise performance, adding the reactive component Ls does not lead to any improvement in the amplifier dynamic range. This can be seen with the aid of the IIP3 expression
IIP3L − deg ≈
IIP3MOSFET vgs v RFin
2
2
2 w ≈ IIP3MOSFET (gm Rs ) o wT
(2.24)
where vgs is the small-signal gate-source voltage and vRFin is the RF input voltage. Equations 2.21 and 2.24 indicate that this topology suffers severely from the common trade-off between the NF and IIP3 limiting any improvement in the achieved
Rs
Rs
Rs C1 RFin
Fig. 2.12 Narrowband impedance transformation using passive elements
L1
Ls
Rs >Rs
2.3 Recent Trends in Wideband CMOS LNAs
33
dynamic range. As can be seen from (2.23), decreasing the device transconductance gm improves the NF but also degrades IIP3. Moreover, scaling the device leads to the improvement in the unity gain frequency of the transistor, which results in improvement in NF, but again degrades IIP3. This can be intuitively described by considering the passive gain of the input matching Lg, Ls, and Cgs network.
1
Passive _ GainL − deg = 2 Rs
Cgs
.
(2.25)
( Lg + Ls )
Increasing the passive gain by decreasing Cgs and therefore decreasing gm improves the NF but unfortunately degrades IIP3 because of the larger signal swing at the transistor input. The NF and IIP3 values become strongly dependent on the passive elements rather than on the transistor itself [71–75]. However, at some point the amplifier performance is going to be limited by the losses of the passive network itself, mainly the Lg losses. Detailed optimization procedures for the L-degenerate amplifier taking into account the losses of the passive network can be found in [76]. Another trade-off arises while trying to achieve the broadband matching from this topology. The NF expression in (2.21) can be rewritten as a function of the quality factor Q of the passive matching network as
NFL − deg = 1 +
g . gm RsQ 2
(2.26)
Basically, the L-degenerate amplifier provides real impedance only in a narrow bandwidth (wo/Q) around the resonance frequency (wo). To achieve a wideband impedance matching, the Q of the matching circuit should be significantly lowered. This will largely degrade the noise figure, which defeats the purpose. As a result, this type of amplifier cannot be used in the mobile TV receiver. However, several attempts have been made to extend the L-degenerate amplifier bandwidth as will be presented and discussed in the next section.
2.3 Recent Trends in Wideband CMOS LNAs This section discusses the progress that has been made to push the performance of wideband LNAs and examines the possibility of using some of these new techniques towards the design of a mobile TV LNA. As described in the introduction of Section 2.2, in order to improve an LNA’s performance, one should either try to minimize the magnitude of the MOSFET noise or distortion generators or try to cancel their contribution at the amplifier output. These two main factors govern the discussion in this section. The section begins by presenting the current reuse techniques developed to boost the device transconductance (gm) in order to minimize the MOSFET noise. Then, the section continues by demonstrating different methodologies proposed to extend
34
2 Wideband CMOS LNA Design Techniques
the bandwidth of the L-degenerate amplifier. Although these techniques help in improving the noise performance of an LNA, they do not offer a solution to improve the IIP3 as well. Research has thus been redirected to other approaches in order to improve both the noise and linearity and therefore improve the dynamic range. The concept of cancelling the noise of the matching transistor in the CG and SFB amplifiers has been reported in 2002 at ISSCC [92]. Since then, several published results showed the feasibility of designing a CG and SFB amplifier in CMOS technology with NF well below 3 dB, regardless of the input matching constraint. Some of these techniques also allow cancelling the distortion and hence improve the amplifier dynamic range.
2.3.1 Current Reuse Amplifiers The MOSFET device suffers intrinsically from the lower driving efficiency. The value of (gm/Id) depends on how the device is biased, which is usually constrained by the NF and IIP3 optimization. A circuit technique that can be adopted to improve this ratio, and therefore lower the power consumption for a fixed DR, is to reuse the current. By simply stacking two NMOS devices (or one NMOS and one PMOS device) as amplifying devices, one can facilitate the current reuse [77]. This enhances the overall transconductance (Gm) from (gm1) to (gm1 + gm2), where gm1 and gm2 are the transconductances of the first and second devices, respectively. Doing this therefore allows one to halve the current for the same input impedance and DR. This technique is implemented in [78] as three stacked NMOS devices that form three gain blocks (shown in Fig. 2.13a). The effective Gm is almost three times higher for the same current in this topology. Large on-chip decoupling capacitors provide stable intermediate nodes between the stacked gain blocks. The output nodes of each of the gain blocks are combined with capacitors C2 and C3 to form one effective signal node at RFout. Lowering the power in this amplifier is achieved at the expense of the bias circuit complications and the use of expensive on-chip capacitors. The LNA design in [79] avoids using decoupling capacitors by using different bias techniques. As shown in Fig. 2.13b, a resistor R1 is used to bias the stacked NMOS device. The author went even further by implementing an active common-mode feedback technique to permit the amplifier to operate reliably on a 1.5 V supply. Another example of implementing the current reuse in LNAs is shown in Fig. 2.14a [80]. In this amplifier the stacked device is a PMOS transistor. This eliminates on-chip coupling capacitors and greatly simplifies the biasing circuit. The interstage transformer T2 acts as a high impedance for ac signals and low impedance for DC signals, making the reuse of bias current feasible. The only disadvantage is the smaller ft of the PMOS device. The PMOS is also used as a stacked device in (Fig. 2.14b) [81] in a simple design where no transformers or capacitors are used. The PMOS device is stacked on top on the NMOS device directly at the expense of generating a stability problem at the intermediate point. An active feedback amplifier is used to set the DC output voltage. This current reuse technique has been used also in [82] for designing a broadband CG amplifier.
2.3 Recent Trends in Wideband CMOS LNAs
35
a
b Vbias
R1
R2 Current reuse
R3
+
R1
-
R4
RFout Cc3
Cc2
Vbias
R1
M5
RFout C2
M6
M2
M4
M1
M3
C3
Vbias
R1
Rs
Ls1
Cc1
RFin
Ls1 I
RFin
Rs
Fig. 2.13 Amplifiers implementing current reuse using NMOS devices: (a) is a wideband amplifier based on [78] and (b) is a narrowband amplifier based on [79]
a
b +
RFout
−
M5
M6
current reuse devices Vbias
T2
RFout M2
M4 Rx
Rs M1
Cx
M3
RFin I
Vref
T1
RFin
Rs
Fig. 2.14 Amplifiers implementing the current reuse using PMOS devices: (a) is a narrowband amplifier based on [80] and (b) is a narrowband amplifier based on [81]
36
2 Wideband CMOS LNA Design Techniques
2.3.2 L-Degenerate Wideband Amplifiers The L-degenerate amplifier is usually the optimum amplifier in terms of the achieved NF, as it can come closer to the MOSFET NFmin as was shown in Section 2.2.4. A new type of amplifier is introduced to extend the bandwidth of the conventional narrowband L-degenerate amplifier. As shown in Fig. 2.15, an LC ladder filter has been used to match the input impedance of the L-degenerate amplifier across a wide bandwidth. It can be shown that in such a circuit the transistor Cgs/MCgd determines the maximum ratio of the upper frequency (fU) to the lower frequency (fL) across which the impedance is matched, where MCgd is the Miller multiplied feedback gate-drain capacitance and Cgs is the gate-source capacitance. The fU/fL ratio is limited to about 2 in CMOS technology because of the relatively large feedback capacitance in MOSFETs [83]. This limits the achievable bandwidth. An external capacitance is added (Cext) in parallel with the gate capacitance to overcome this limitation, but this directly compromises gain and noise. Therefore, this topology might not be feasible to cover the VHF/UHF band required for mobile TV applications. The LNA designers kept trying to extend the L-degenerate amplifier because of its superior performance. The design in [84] used the gate-drain capacitance to create another feedback loop around the main transistor in addition to the source inductor Ls feedback loop. This additional feedback loop resonates at a different frequency from the series inductor Ls loop. The overall structure of this amplifier and the input equivalent circuit taking Miller effect into account are shown in Fig. 2.16. If the cascode transistor M2 is designed properly, a wideband input match with minimum noise figure can be achieved.
Ls
RL
M2
Rs
L1
CL
CL RFout
Lg M1
RFin
L2
C2
Cext
Ls
Fig. 2.15 Broadband L-degenerate amplifier with input matching using LC ladder
2.3 Recent Trends in Wideband CMOS LNAs
Ls
37
RL
M2
CL RFout
Cgd
Rs
Lg
Lg
R2
Ls
Cgs
M1
C2 RFin
Ls
L1
Ri = ω TLs C1
R1
Degenerated transistor equivalent circuit with Miller effect
Fig. 2.16 Broadband L-degenerate amplifier with input matching using Miller effect based on [84]. The equivalent circuit for the input matching is also shown
Drain-gate capacitance has also been used to provide the wideband matching in [85, 86]. However, this wideband matching technique strongly depends on process and temperature variation making it an unattractive solution. The same concept has been used to design the wideband amplifier shown in Fig. 2.17 [87]. This topology is implemented by inserting a feedback capacitor CL between the two transistors M1 and M2 of a narrowband inductively degenerated cascode LNA. The design is robust against temperature and process variations. However, it requires two extra pins to provide the DC path to the cascode transistor, which increases the packaging cost.
2.3.3 Capacitive Cross-Coupled CG Amplifiers The CG amplifier discussed in Section 2.2.2 requires a gm = 1/Rs to provide a wideband input match. Given this condition, the CG amplifier has a NF that is larger than 1 + g (>3 dB in nanometer CMOS). This trade-off is somewhat relaxed for the balanced CG amplifier, which uses capacitive input cross-coupling as shown in Fig. 2.18. In 1999, Cho proposed the cross-coupling idea as a technique that can be used for gain enhancement in CG amplifiers [88]. Subsequently, the authors in [89] reported the benefit of using this technique on the NF of the CG amplifiers for the first time. The noise figure of the capacitive cross-coupled CG amplifier NFCG−CC under the input matching constraint can be derived to be equal to
38
2 Wideband CMOS LNA Design Techniques
Fig. 2.17 Broadband L-degenerate amplifier based on [87]
Ls
RL
Cascode device M2
RFout
CL
RF choke
Rs
CL
Lg M1
RFin
Ls
Fig. 2.18 Capacitive cross-coupled common-gate amplifier
RFout M1
M2
Cross-coupling capacitors
C1
C2
RFin
NFCG −CC = 1 +
g . 2
Rs
(2.27)
The cross-coupling technique does not add much cost or complexity. It doubles the effective Gm, thus, the power consumption can be reduced. Also, in terms of IIP3, the CG amplifier utilizing a capacitive cross-coupling is expected to achieve a higher IIP3 than that of the conventional differential CG amplifier. The only disadvantage is that the capacitors limit the amplifier bandwidth. The same idea has been used in [90] to design a shunt feedback common-gate (SFBCG) amplifier. As shown in Fig. 2.19, this amplifier is no more than a GC with cross-coupled
2.3 Recent Trends in Wideband CMOS LNAs
39
current reuse
C1
C2
M2 Rf1
M4 Rf2
RFout M1
M3
C1
RFin
C2
Rs
Fig. 2.19 Broadband SFBCG amplifier based on [90]
capacitors incorporating the current reuse technique. The current reuse technique further pushes the effective Gm to 4gm where gm is the transconductance of a single transistor M1. In this case, the SFBCG amplifier produces a NFSFBCG equal to
NFSFBCG = 1 +
g . 4
(2.28)
Although these efforts have helped in pushing the noise performance of CG amplifiers with resulting dynamic range improvements in wideband LNAs, these amplifiers still suffer from the input match-noise figure trade-off. New techniques are thus still needed to break this trade-off and hence increase the degree of freedom to allow the dynamic range to be pushed even further while keeping the power consumption to a minimum. The authors in [91] used the cross-coupling idea to boost the gm, but used a transformer instead of capacitors. They reported a 2.5 dB NF amplifier. However, the impedance matching constraint remained unchanged.
2.3.4 Noise and Distortion Cancelling Amplifiers In 2002 at ISSCC, Bruccoleri reported a SFB amplifier with a 2.4 dB NF [92]. He was able to break the SFB amplifier input matching NF trade-off described in Section 2.2.3 by proposing a thermal noise-cancelling technique. A simplified schematic diagram of his amplifier is shown in Fig. 2.20. His idea was based on
40
2 Wideband CMOS LNA Design Techniques
The noise current output from M1 leading to noise voltages at X and Y that are fully correlated
"A" RF
M3 Y
Rs
RFout
M1
M2
X
RFin
noise curent
Amplifier plus adder to cancel the noise of M1
Fig. 2.20 A resistive shunt feedback amplifier using noise cancelling
exploiting the difference in sign for the noise and the signal gain at the output (see point Y in the same figure) making it possible to cancel the output noise contribution of the main transistor M1 while adding the signal. This was done by creating a new output, where the voltage at node Y is added to a scaled negative replica of the voltage at X, using a second stage amplifier “A”. It can be shown that the output noise cancellation is achieved if
gm 2 RF = 1+ gm 3 Rs
(2.29)
where gm2 and gm3 are the transconductance of M2 and M3, respectively. With this condition, the overall gain of the SFB amplifier with noise cancellation can be found to be
GainSFB _ noise _ cancelled = −2
RF . Rs
(2.30)
At this gain, the noise of M1 is totally cancelled. However, this is not the case with the feedback resistor RF. This can be seen by splitting its noise current in two correlated sources to ground, at the output node Y and the input node X. The former is cancelled but the latter is not. Therefore, the minimum noise figure of the conventional SFB amplifier expressed in (2.18) can be reformulated to include the noise cancellation as shown in (2.31)
NFmin SFB _ noise _ cancelled = 1 +
Rs . RF
(2.31)
This expression assumes that gm2 is high enough that it would not affect the total NFminSFB_noise_cancelled. Therefore, to lower the NF even further, the power consumption
2.3 Recent Trends in Wideband CMOS LNAs
41
R1
R2 RFout
noise current
CG noise
M1
CG noise
Rs M2
RFin
Fig. 2.21 A common-gate amplifier using noise cancelling
of the second stage should be high and RF should be large. In terms of the amplifier DR, the added amplifier “A” would degrade the overall linearity and therefore limit the IIP3 value. If these limitations could be avoided, a better version of a SFB amplifier can be realized. The concept of noise cancelling has been implemented in a CG amplifier as shown in Fig. 2.21 [83]. An interesting property of this circuit is that the fraction of M1 noise current that flows into R1 and Rs induces an in-phase amplified noise current in R2 by driving the gate of M2. Therefore, if the resulting noise voltages at the two output terminals are equal, then M1’s noise is common-mode and can be cancelled by differential sensing. It can be proven that the noise cancellation occurs when
gm1 R1 = gm 2 R2.
(2.32)
At this condition, assuming balanced loads, the new noise figure of the CG amplifier exploiting noise cancelling (NFCG_noise_cancelled) is derived to be
NFCG _ noise _ cancelled = 1 + g.
(2.33)
Comparing with the NF of the conventional CG amplifier expressed in (2.13), the noise-cancellation technique transferred the problem of the high noise level associated with the CG transconductance needed for matching to the commonsource amplifier that is used for noise cancelling. In other words, for this noisecancelling technique to be efficient, unbalanced loads are required to achieve NF < 3 dB with low-power consumption. In the presence of unbalanced loads the NFCG_noise_cancelled is
42
2 Wideband CMOS LNA Design Techniques
NFCG _ noise _ cancelled = 1 +
gm 2 R 2 2 4g Rs (gm1 R1 + gm 2 R 2 )2
(2.34)
where gm1 and gm2 are the transconductances of the common-gate (M1) and common-source (M2) amplifiers, respectively. As can been seen in (2.25), the load R2 can be traded with power consumption of the common-source amplifier (gm2). Another advantage of this amplifier is that the gain can be increased by up to 6 dB when the output swings are balanced. Also, the load resistor noise is less significant. The only limitation of this amplifier is that the common-source amplifier M2 used to cancel the noise of M1 limits the overall noise figure of the new CG amplifier. Not only that, but its transconductance value is somehow coupled with the input matching requirement in the case where balanced loads are present. If these limitations can be avoided, a very high dynamic range common-gate wideband LNA can be realized. Other circuits recently proposed in the literature exploiting noise-cancellation mechanisms are shown in Fig. 2.22 [93, 94]. A CG amplifier in a feedback path is used to achieve 50 W wideband input matching and partial noise cancellation at the same time. The noise current of the common-source transistor M1 produces a noise voltage at the output. A scaled version of this noise is fed back to its gate again, where it is amplified and inverted. As a result, at the output (summing node), some of this correlated noise is cancelled and the total output noise is reduced. Some limitations also exist in these amplifiers. The major noise contribution now originates from transistor M3. Because M3 is outside the feedback loop, its noise is not subjected to the cancellation mechanism. Specifically, the noise contribution of M3 is (1 + gmRo)2 times larger than that of the CG transistor M2, assuming the same W/L ratio, where Ro is the equivalent resistance seen by the source of M2. a
b Currect reuse
M oad
R oad
Rload
RFout
RFout
M2
M2 Noise cancellation by active feedback
Ro
Ro
Mcascode
R1
Rs
Rbias
Mcascode
Rs M1
RFin
M1
RFin M3
C M3 Currect reuse
Fig. 2.22 A common-gate amplifier exploiting noise cancelling by active feedback. Amplifier (a) is reported in [93] while amplifier (b) is reported in [94]
2.4 Techniques to Improve the Wideband LNA Dynamic Range
43
Another disadvantage is that these circuits degrade substantially the IIP3 value because of the large signal usually applied at the gate of the CG transistor M2. Another effect that also contributes to the IIP3 degradation is the feedback loop. The feedback can combine the second-order distortion with the fundamentals of the output signal of the circuit, producing third-order distortion at the circuit input, which worsens the IIP3 value. The same mechanism leading to cancellation of the output noise can also be exploited to cancel the distortion. However, since the IIP3 will always be limited by the other added amplifier, specific distortion cancellation circuits have been investigated. In [95], the LNA design uses a circuit that can sense the distortion as an inherently generated signal, amplified through a parallel path, and cancel it at the output. The large overhead of the auxiliary amplifier makes this technique unattractive. Other designers went even further by trying to cancel the third-order nonlinearity of a MOSFET by adding a parallel transistor that is biased in the triode region as in [96] or in the subthreshold region as in [97]. However, this concept did not catch on due to this approach’s sensitivity to process and temperature variations.
2.4 Techniques to Improve the Wideband LNA Dynamic Range In this section, the LNA topology that can best meet the stringent dynamic range (DR) requirements of the mobile TV application has been introduced. Two main factors are considered here: the minimum noise figure that a topology can achieve, and its DR.
2.4.1 Wideband CMOS LNA State-of-the-Art As has been shown earlier, the inductor-based wideband LNAs have superior noise performance as their NF can reach the NFmin of a MOSFET. Moreover, with a good design, they can achieve a very good dynamic range that is limited by the amount of current they consume. However, these topologies have received less interest with respect to their inductorless counterparts. In the modern nanometer CMOS technology, the use of area consuming on-chip inductors must be avoided as the cost per area for such processes is extremely high. Also, high quality factor (Q) inductors do not easily lend themselves to integration in a digital CMOS process, due to their need for special process enhancements such as high substrate resistivity for implementation. The CMOS process with RF enhancements usually lags one to two generations behind the digital one. Moreover, these types of amplifiers do not fit with the trend toward flexible multi-mode and multi-standard radios as they have limited bandwidths. As shown in Table 2.1, inductorless type amplifiers have been
CGf CGh SFB Noise Cancel CG
Inductorless CMOS Wideband LNAs 0.6 µm CMOS Choe 1999 [88] Zhuog 2000 [89] 0.5 µm CMOS Van Zeijli 2002 [99] 0.18 µm CMOS
Bruccolerik 2002 [92] Rogin 2003 [100] 2.4 4
0 +10
N/A +5 N/A
−9 −3
IIP3 (dBm)
13 10
15 12 N/A
15 18
Voltage gain (dB)
35 7
N/A 20 7j
7.5c 50
Power (mW)
Chehrazi 2005 [83] 0.13 µm CMOS Noise Cancel 3 +1 19 12.5 Zhan 2006 [101] 90 nm CMOS SFBl 2 −14 25 42 Bagheri 2006 [40] 90 nm CMOS Noise Cancel 3 +1 19 12.5 Ramzan 2007 [94] 0.13 µm CMOS Noise Cancel 2.7m −4n 17 25 90 nm CMOS Noise Cancel 2.5dB −15 17 10 Borremanso2007 [93] Grey: Narrow band applications. a The amplifier supports an ultra-wideband receiver (3–10 GHz). b The drain-gate capacitance helps in achieving a very low NF. c This amplifier is optimized for the power consumption. d The amplifier supports 0.7–1.4 GHz band. e The amplifier supports a 900 MHz spread-spectrum cordless phone receiver. f Capacitive cross-coupling technique has been used to boost the CG gm. g The amplifier supports a 900 MHz narrowband receiver. h Capacitive cross-coupling techniques have been used to push the NF of the CG amplifier. i The amplifier supports Bluetooth radio. j This power has been achieved for a source resistance of 150 W. k The LNA supports 2–1,600 MHz band. l A source follower feedback amplifier has been proposed to tune the input matching, this topology benefits from high ft of the 90 nm CMOS process. m Measured differentially from 1 to 7 GHz. n High IP3 value achieved due to the differential operation and the feedback technique. o The LNA covers from DC to 6 GHz.
0.25 µm CMOS 0.13 µm CMOS
2.3b 0.5
L-degenerate L-degenerate
Inductor-based CMOS Wideband LNAs 0.18 µm CMOS Leea 2006 [84] Belostotskid 2007 [98] 0.18 µm CMOS 2 3 3.5
NF (dB)
Table 2.1 Current state-of-the-art of wideband CMOS LNAs Research paper Technology Topology
44 2 Wideband CMOS LNA Design Techniques
2.4 Techniques to Improve the Wideband LNA Dynamic Range
45
implemented several times in narrowband receivers [88–100] because of the above mentioned reasons. For the mobile TV LNA, low cost, small area, inductorless topologies should only be considered. High performance wideband inductorless topologies such as noise-cancelling amplifiers [40, 63] or feedback type designs [99, 101] suffer from high power, or inadequate NF as shown in Table 2.1. As shown in Section 2.3.4, the noise-cancelling amplifiers incorporate another amplifier to cancel the noise of the main transistor of an SFB or CG amplifier. This added amplifier unfortunately dominates the noise performance and makes achieving low NF (<3 dB) at low power consumption extremely difficult. More research thus is needed in this approach to improve the noise performance of these amplifiers and improve their dynamic range.
2.4.2 New Low-Power Noise-Cancelling Technique A CG amplifier, shown in Fig. 2.23a, was chosen as the starting topology to provide the input matching. The resistor RF is added for two reasons. First, it makes the CG amplifier that has a g of 1.45 in nanometer CMOS to act as it has a g of 1. In other words, the noise performance of the CG stage is going to be dominated by the feedback resistor RF given that the loop gain gm1RF is larger than unity. At this point, the noise of the CG transistor M1 can be neglected. This technique has been known for the bipolar transistor current source since 1975. Second, this feedback resistor boosts the IIP3 value as it decreases the relative current swing applied on the gate of the CG transistor (see Section 2.1.2.2). In Fig. 2.23b, the noise-cancelling technique proposed by Chehrazi (see Section 2.3.4) is applied primarily to cancel the noise contribution of the feedback resistor. The current in the CG and commonsource (CS) branches should be scaled in a way that facilitates cancelling the RF noise at the output differentially. At this stage this amplifier suffers from the same trade-off as the case of Chehrazi’s LNA. The noise of the amplifier shown in Fig. 2.23b is limited by the CS amplifier transconductance (gm2). In Fig. 2.23c, connecting a feedback loop around the common-source amplifier M2 is proposed. The CG M1 transistor with the RF resistor has been used to eliminate the M2 noise contribution at the output (+RFout). The idea of applying a negative active feedback around a CS amplifier was previously presented before in [57]. However, cancelling the noise of the feedback elements (M1 and RF) was apparently not achieved in this amplifier. Also, as will be shown, the noise contribution of M2 at (–RFout) node is almost zero. Consequently, the power consumption needed in Chehrazi’s amplifier can be reduced by more than 50% depending on the loop gain applied on the CS amplifier. The only concern in this topology is that the CG amplifier linearity might limit the overall IIP3 value due to the large-signal swing at its gate. A cascode transistor M3 is proposed to lower the signal swing at the gate of M1 and hence improve the overall linearity of the amplifier. Generally, the cascode transistor M3 should not affect the overall NF of this amplifier since its noise contribution is reduced by the common-source amplifier (M2) intrinsic gain.
46
2 Wideband CMOS LNA Design Techniques
a
b The M1 noise is reduced by the feedback resistor RF
RF resistor noise
R1
R1
RFout
+ The RF noise appears as a common-mode at +RFout and -RFout
CG stage
M1
M1
(a) RF
Rs
inRF
RFin
Rs
RF resistor noise
R2
RFout −
M2
RF
RFin noise current
d
c R1 The M3 noise is also canceled differentially
R2 −
RFout
M2 noise
Cascode
+
R1
M3 noise
R2 −
M3
RFout +
+
M3 noise
FB M1
M1 M2
Rs RFin
M2
RF
Rs
RF
The M2 noise is reduced by the feedback (FB) inCS
RFin
Fig. 2.23 New noise-cancelled LNA realization. The DC biases are not shown
In nanometer CMOS processes, like 65 nm, the output resistance of the common-source transistor M2 can be low enough to make the noise of M3 more pronounced. However, a careful look at Fig. 2.23d shows that the noise of the cascode transistor M3 can simply be treated as an added noise source to the CG transistor M1, since it only modulates its gate voltage. As a result, the mechanism that is applied to cancel the noise of the feedback resistor RF is by default applied to M3 as well. Therefore, M3 noise is not a concern in this topology. Simulation results show that this topology can achieve a NF as low as 1.4 dB, IIP3 of −5 dBm with a power consumption of 18 mW across VHF/UHF bands. The 130 nm CMOS models are used in this simulation. The exact noise behavior of this wideband CMOS LNA topology as well as the performance trade-offs are discussed in the next section.
2.4 Techniques to Improve the Wideband LNA Dynamic Range
47
2.4.2.1 Noise Analysis To quantify the noise reduction of the CS transistor M2 due to the active negative feedback, let us consider the circuit shown in Fig. 2.24. The feedback loop has been broken at point X in order to calculate the noise current that flows out of the commonsource transistor M2. The RFin source is connected to the ground for noise analysis. Since we have a differential output, the noise current that flows out of M2 into the CS branch (ino+CS) and the noise current that flows out of M2 and leaks into the CG branch (M1) (ino−CS) are both important. The block diagram shown in Fig. 2.25 is used to model the feedback system to facilitate the noise analysis. Using this model the output noise current that flows out of the CS transistor M2 and leaks into the CG branch can be expressed as1
ino −CS =
1 g m1 gm 3 1 + gm1 RT ( g Rs ) gm1 1 + m2 gm 3 1 + gm1 RT
inCS
(2.35)
where gm1, gm2, and gm3 are the transconductances of the transistors M1, M2 and M3 respectively and RT = RF + Rs. In the same manner, the output noise current that flows out of the CS transistor M2 into the CS branch can be expressed as ino + CS =
1 inCS . ( gm 2 Rs ) gm1 1+ 1 + g RT g m3
(2.36)
m1
To get some insight, (2.35) and (2.36) can be written as a function of the loop gain (LG) of the active feedback loop as follows
1
ino −CS
LG g Rs i . = m2 1 + LG nCS
1 ino + CS = i . 1 + LG nCS
(2.37)
(2.38)
The channel noise current of M2 can be modeled at low frequencies as an input-referred voltage equal to vinCS = inCS /gm2.
48
2 Wideband CMOS LNA Design Techniques
noise
R1
vo-cs
R2
ino-CS
noise
vo+cs M3
1/gm3 M1
ino+cs
vs1
RF
x M2
Rs
incs
vinCS
Fig. 2.24 The contribution of the common-source transistor (M2) noise at the output. ino+CS is the part of inCS that flows out of M2 into the CS branch while ino−CS is the part that leaks to the output node of the CG branch (vo−CS) ino+cs vinCS
+ −
+
gm2
vo+cs
R2 1/gm3
voltage at source of M3
Rs RT=Rs+RF
(gm1RT)/(1+gm1RT) voltage at source of M1
ino-cs
R1
vs1
1/RT
vo-cs
Fig. 2.25 Feedback loop block diagram used to find the output noise current that flows out the common-source transistor M2
As can be seen in (2.38), the output noise current that flows out of M2 into R2 is inversely proportional to (1 + LG). The larger the LG, the lower the noise contribution of M2 at the (vo+CS ) node as shown in Fig. 2.26. However, as can been seen in the same figure and also from (2.37), the output noise current that flows out of M2 and leaks into R1 is a weaker function of the loop gain (LG). In fact, it increases if the LG increases. Since the noise of the CS transistor M2 cannot be cancelled differentially (both are out-of-phase), ino−CS will slightly degrade the overall noise reduction of M2. However, at reasonable values of gm2, this effect can be neglected especially at lower LG values. For example, assuming a feedback loop with an LG = 1, the noise that leaks from M2 to the (vo−CS) is just
M2 noise contribution, normalized
2.4 Techniques to Improve the Wideband LNA Dynamic Range
49
ino+ CS 75%
ino−CS 25%
0
1
2
3
Loop Gain Fig. 2.26 Effect of negative feedback loop gain (LG) on the output noise contribution of the common-source transistor M2. As shown, the amount of noise current normalized to the total noise current (inCS) that flows out of M2 to the CS branch (ino+CS) decreases with increasing LG. The other curve illustrates the amount of noise current of M2 also normalized to the total noise current (inCS) that leaks to the CG branch (ino−CS). It saturates at higher LG values depending on gm2 and Rs. In this plot gm2 is assumed to be equal to 0.1 S and Rs equal to 50 W
noise
R1
vo-CG
noise
R2
ino-CG
vo+CG M3
x inCG inRF
M1
1/gm3 vinCG
ino+CG
RF M2
Rs
Fig. 2.27 The contribution of the common-gate branch noise at the output. This noise appears as common-mode at differential output. ino+CG is the part of inCG that leaks into the CS branch while ino−CG is the part that leaks into the output node of the CG branch
50
2 Wideband CMOS LNA Design Techniques ino-CG
R1
vo+CG
R2 vinCG
−
+
(gm1RT)/(1+gm1RT)
1/RT
voltage at source of M1
Rs
gm2
vo+CG
ino+CG
−
Diffrential output noise voltage = zero
voltage at gate of M2
1/gm3 RT=Rs+RF
Fig. 2.28 Feedback loop block diagram used to find the output noise current contribution of the common-gate transistor M1
10% of its total noise, which results in a total noise reduction of 40% when the output is sensed differentially. Therefore, using the active feedback loop can greatly improve the noise performance of wideband LNAs and the part of the noise that leaks to the CG branch and reaches vo−CS is negligible. This improvement in the NF can be traded for the reduction in the power consumption. The main source of noise in the CG branch comes from the feedback resistor RF as long as gm1RF > 1. Let us consider the circuit diagram shown in Fig. 2.27 in order to discuss the feedback loop elements (CG transistor M1 and the feedback resistor RF). To find the noise contribution of M1, the feedback system is modeled as shown in the block diagram of Fig. 2.28. Using this model, the output noise current that leaks from the M1 to the load resistor R2 (ino+CG) can be found as follows2
ino + CG =
(gm 2 Rs ) (1 + gm1 RT ) g g m1 1 + m2 Rs gm 3 1 + gm1 RT
inCG .
(2.39)
Similarly, the output noise current that leaks to the load resistor R1 in the common-gate branch (ino−CG) from the M1 is
ino −CG =
1 (1 + gm1 RT ) g g m1 1 + m2 Rs gm 3 1 + gm1 RT
inCG .
(2.40)
As can be seen from (2.39) and (2.40), the output noise current of the M1 branch can be totally cancelled with differential sensing (ino+CG and ino−CG are inphase) given that gm2 = 1/Rs. However, this would limit the power consumption of 2
The channel noise current of M1 can be modeled at low frequencies as an input-referred voltage vinCG = inCG/gm1.
2.4 Techniques to Improve the Wideband LNA Dynamic Range
51
M1 noise contribution, normalized
100%
60%
R2=80 Ω
R2=40 Ω R2=120 Ω 20%
0
0.02
0.06
0.1
gm2 (S) Fig. 2.29 Dependence of the common-gate noise-cancellation mechanism on gm2 and R2. The active feedback loop is broken to decouple the CS branch from the CG branch. Otherwise, the amplifier input matching cannot be maintained. The R1 is set to 250 W
the CS transistor. Using unbalanced loads can resolve this by cancelling the differential output noise voltage at the output such that (ino+CG R2 = ino−CG R1). Therefore, the noise of M1 can be completely cancelled given that R1 = gm 2 × R2 × Rs.
(2.41)
This condition is satisfied at infinite possibilities of gm2 and R2. The dependence of the cancellation mechanism on these parameters is shown in Fig. 2.29. The choice of gm2 depends on the power-noise tradeoff. The same noise-cancelling mechanism applied on M1 is also applied on the feedback resistor RF. Using Fig. 2.27, the output noise current that leaks to the load resistor R1 in the common-gate branch (io−RF) from the resistor RF is
ino − RF =
(gm1 RF ) (1 + gm1 RT ) g g m1 1 + m2 Rs gm 3 1 + gm1 RT
inRF .
(2.42)
Similarly, the output noise current that leaks from the feedback resistor RF to the load resistor R2 (ino+RF) can be found as follows
52
2 Wideband CMOS LNA Design Techniques
ino + RF =
(gm1 gm 2 RsRF ) (1 + gm1 RT )i g g m1 1 + m2 Rs gm 3 1 + gm1 RT
nRF
.
(2.43)
As can been seen from (2.42) and (2.43), the output noise current of RF can be totally cancelled with differential sensing giving the same noise-cancelling condition of the CG transistor M1 (Eq. 2.41). Given the derivations of the noise contribution of M1, M2, and the feedback resistor RF, the noise figure of the new LNA topology (NFCG_noise_cancelled_NEW) can be derived (assuming that the noise of the CG transistor M1 as well as the feedback resistor RF is cancelled) to be g 1 R2 gm 2 . = 1+ 2 (2.44) R1 ( g Rs ) Rs g m1 1 + m2 gm 3 1 + gm1 ( Rs + RF ) 2
NFCG _ noise _ cancelled _ NEW
Comparing this expression with the one reported for Chehrazi’s noise-cancelling amplifier [83] shows that the NF of the new amplifier has improved by the square of the loop gain. The simulated NF value mentioned in Section 2.4.2 is very close to the value predicted from (2.44) and is similar to values of counterpart L-degenerate amplifiers. To determine the minimum achievable NF value of this topology, remaining performance parameters for the amplifier have to be defined. These parameters, namely input matching, gain, and linearity are discussed in the following sections. 2.4.2.2 Input Matching and Gain Analysis A common-gate amplifier with a source resistance RF (shown in Fig. 2.30a) has an input resistance approximately equal to
RinCG = RF + 1 / gm1 .
(2.45)
This expression has to be modified in the presence of feedback around the CG amplifier as shown in Fig. 2.30b. In this figure, the transistor M1 acts as a commondrain amplifier for the feedback signal, making M1 look like a shunt-shunt feedback type amplifier. Because of this type of feedback, the input impedance of the CG amplifier (RinCG) is decreased by the amount of the loop gain (“A” in Fig. 2.30). The input resistance of the CG amplifier with an active feedback loop (RinCG_FB) is expressed by
RinCG _ FB =
RF + 1 / gm1 . 1 + gm 2 / gm 3
(2.46)
2.4 Techniques to Improve the Wideband LNA Dynamic Range
a
53
b
R1
ro1
R1
M1
ro1
CS amplifier
M1
RinCG
A
ro>>R1
RF
RF
RinCG_FB
A=gm2/gm3
Fig. 2.30 Input resistance calculation of the common-gate amplifier. Amplifier (a) without feedback (b) with feedback. The output intrinsic resistance ro1 of M1 is assumed to be large enough so that the effect of the load resistance R1 on the total input resistance is negligible
Two important observations can be made by comparing (2.45) and (2.46). First, adding the loop gain around the CG amplifier allows the matching to be adjusted through (gm2/gm3). This makes it possible to achieve the input match requirements at lower power consumption. Second, feedback helps in saving some voltage headroom that can then be used to increase the amplifier gain (by decreasing the value of RF required to achieve the matching). The question that remains is what is the ultimate achievable NF for this topology? As discussed in Section 2.4.2.1, the NF can be improved by increasing the feedback loop gain (LG). However, as can be seen in (2.46), this demands an increase in the resistance value of the feedback resistor RF to maintain the matching. Since the upper limit of this value is bound by the available voltage headroom, the minimum NF that can be achieved for this amplifier is now limited by the supply voltage. Therefore, decreasing the NF value of this amplifier requires an increase in the supply voltage.3 The gain in this topology also differs from that of a conventional CG amplifier. This topology includes a single-to-differential conversion which results in an increase in the gain. The gain can be increased up to 6 dB when the output swings are balanced. The differential gain of this topology under the input matching constraint can be expressed by
Differential _ Gain =
gm 2 R2 Rs + R1 . 2 Rs
(2.47)
3 In our previous discussion, the output resistance (ro1) was neglected. However, this value can be very small in nanometer CMOS and hence its effect cannot be ignored. In fact, it tightens the design trade-offs.
54
2 Wideband CMOS LNA Design Techniques
The gain can be further increased by increasing the load resistors R2 and gm2. However, this degrades the NF since it affects the cancellation mechanism and also increases the power consumption. Therefore, under the constraint of the noise cancellation, the gain can best be increased by increasing the supply voltage. 2.4.2.3 Linearity Analysis A wideband LNA requires high linearity to suppress the cross-modulation and intermodulation distortions that result from the increased co-existence of adjacent blockers or the on-chip transmitter leakage (see Chapter 1). While the noise of nanometer CMOS improves with scaling, the linearity deteriorates with supply voltage, high mobility effects, and the transistor velocity saturation effect. In general, the amplifier linearity can be controlled either by changing the characteristic of the intrinsic MOSFET device itself (see Section 2.1.2.1) or by using linearization circuit techniques (see Section 2.1.2.2). Neither of these techniques should affect the noise performance of the circuit – i.e., shifting of the dynamic range has to be avoided. Even though the nonlinearity of the MOSFET drain current can be generated either from the nonlinear transconductance or from the nonlinear drain conductance, the latter will be ignored assuming that the load resistors R1 and R2 are relatively small. Consider the circuit diagram of the described noise-cancelling LNA topology shown in Fig. 2.31. The input signal at the gate of M1 is large in amplitude relative
R1
To cancel the second order distortion of M2
R2
RFout
−
gm1(v x)
3
+
Vbias ο RBias
M3
g''m1(vx)
M4
M1 +
Vx RFin
Rs
ro1 −
RF
g'm1(v x )
2
M2
C
Fig. 2.31 Noise-cancelling CG LNA topology incorporating a second order distortion free circuit technique in the common-source stage M2. The nonlinearity transconductance of M1 is demonstrated by g¢m and g¢¢m
2.4 Techniques to Improve the Wideband LNA Dynamic Range
55
to the other transistors due to the feedback. Therefore, M1 nonlinearity plays a significant role in controlling the overall IIP3 value. As discussed in Section 2.1.2.2, feedback can be utilized to reduce the distortion. Resistor RF is used as a negative feedback for M1 in order to decrease the input signal swing and hence improve the linearity of the feedback path. Cancelling the RF noise in this topology allows for a wide dynamic range to be reached. The common-source transistor M2 also uses negative feedback for noise cancellation as well as to cancel the distortion. Therefore, M2 linearity improves as the loop gain increases. Second-order nonlinearity is usually not of primary concern in the LNA design because it results in distortion that is outside of the frequency of interest. However, due to the feedback configuration of the described amplifier, secondorder nonlinearity may contribute third-order distortion. The common-source transistor M2 generates second order distortion at the circuit output, which propagates linearly to the input through the feedback path. The second-order nonlinearity of M2 combines this second-order distortion with the fundamentals at the amplifier input, producing third-order distortion. Finally, this third-order distortion propagates linearly to the output. Concurrent cancellation of the intrinsic thirdorder distortion from the CS and CG stages then limits IIP3 by the second-order interaction between them. Therefore, in order to achieve very low third-order distortion, it becomes vitally important to linearize both second and third-order nonlinearities. Although the LNA topology serves as a single-to-differential converter, it does not suppress second-order nonlinearity, as would a fully-balanced circuit. However, this can be compensated by using a PMOS transistor (M4) as a common-source amplifier with the NMOS CS transistor as shown in Fig. 2.31 [102]. The output current from the PMOS and NMOS transistor can be expressed as
g' g '' iout = in + i p = gmn × vin + mn × v 2 in + mn × v3in 2 6 g 'mp g '' (2.48) − gmp × (−vin ) + × (−vin )2 + mn × (−vin )3 2 6 ( g 'mn − g 'mp ) 2 ( g ''mn + g ''mp ) 3 = ( gmn + gmp ) × vin + × v in + × v in 2 6
where gmn and gmp are the NMOS and PMOS transconductances respectively and g¢mn, g¢mp, g¢¢mn, and g¢¢mp are the second and third derivatives of the NMOS and PMOS transconductances. As can be seen in (2.48), the single-ended input current iout is free from second-order distortion if g¢m is well matched. As for dynamic range, the added PMOS transistor M4 doesn’t affect the noise of the amplifier since its noise can be cancelled in a similar way as the M2 noise was cancelled. Moreover, the bias current of NMOS transistor M2 is reused in PMOS transistor M4, thereby all the current reuse technique benefits can be
56
2 Wideband CMOS LNA Design Techniques
utilized (see Section 2.3.1). Another advantage of M4 is that the current reduction in the load resistor R2 facilitates using high resistance values, allowing for a smaller chip area.
2.5 Chapter Summary This chapter has examined the problem of designing wideband CMOS low-noise amplifiers (LNA) with wide dynamic range (DR) and has presented solutions to improve the state-of-the-art LNA performance. The intrinsic dynamic range (DRMOS) in CMOS is proportional to the amount of current consumed. Thus, the more current MOSFETs consume, the wider DR they can achieve. Additionally, this DR can be shifted up or down using the gate bias voltage (Vgs) for constant current consumption. Since MOSFET-based LNAs have to provide a 50 W input match to their antennas, their dynamic range would shift from DRMOS. Hence, several LNA topologies have emerged based on the technique used to provide the matching, each with its own distinct dynamic range. For fixed current, the R-CS topology shifts up the DRMOS because of the 50 W resistor added at the transistor gate for matching. The current consumption and the gate bias can still be used to control the DR of this amplifier. However, the achieved DR would be limited by the thermal noise of the matching resistor, a problem that can further be avoided by using reactive components for matching. The L-degenerate topology shifts down the DRMOS due to the LC network added at the transistor gate for matching. Adding these matching elements to the MOSFET transistor provides an additional degree of freedom in controlling the DR of these amplifiers. In addition to the current consumption and gate bias, the passive gain of the LC matching network can be used to shift the DR up and down. In fact, this DR can be shifted down until its lower limit hits the NFmin value of a MOSFET. Despite these beautiful features, the expensive on-chip coils make this topology unattractive for low-cost production. Naturally, wideband amplifiers such as CG and SFB amplifiers are good alternatives, especially in nanometer CMOS technologies. The CG topology can provide a wideband input match through the intrinsic properties of the MOSFET, the transconductance (gm). This makes the DR of these amplifiers fixed and dependent on the DC power consumption that is constrained by the input matching requirement. Moreover, this fixed DR (i.e., NF) is inadequate to achieve the sensitivity required by most modern communication systems (−100 dBm). The SFB amplifiers can break this trade-off and therefore allow flexibility in the DR optimization, but their matching mechanism prevents them from achieving the DRMOS at low-power consumption. The feedback resistor must be increased in order to decrease its noise contribution. This unfortunately requires high current consumption to preserve the matching. The amount of power needed to widen the DR makes SFB amplifiers an unattractive solution for portable receivers. In most CMOS amplifiers, the current reuse technique can be used to
2.5 Chapter Summary
57
cause a downward shift of the DR. However, even with the use of this technique, R-CS, SFB, and CG amplifiers cannot fulfill the performance requirements of modern wireless receivers. Several noise and distortion cancelling techniques have been described to overcome the shortcomings discussed above. The cross-coupling capacitive technique has been applied to the CG amplifier to partially cancel the transistor noise contribution at the output. The dynamic range is improved with respect to the conventional CG topology but the input match constraint remains, and DR is still fixed. Recently, Chehrazi proposed a technique to completely cancel the noise of the matching transistor in the CG amplifier and therefore break the matching-noise figure trade-offs. Breaking this trade-off allows the possibility of controlling the amplifier DR by varying the current consumption without any constraints. However, this flexibility comes at the expense of the power consumption of the added amplifier “A” needed to cancel the CG transistor noise. Therefore, the current noise-cancelling CG amplifier can achieve the same DRMOS but at higher power consumption. This chapter presents a noise-cancelling technique to decrease the power consumption of current noise-cancelling CG amplifiers. By using active negative feedback, the noise of the added amplifier “A” can be eliminated. This allows the CG amplifiers to achieve the DRMOS with lower power consumption than the current state-of-the-art. The described amplifier can also be seen as an SFB amplifier with noise-cancelling feedback resistor (shown in Fig. 2.32) and with a more relaxed power requirement on the matching MOSFET (Mmain) compared with the current state-of-the-art SFB amplifiers. The physical fundamental noise limit of the described solution has been examined by exploring the power, linearity, input match, and gain trade-offs. The next chapter presents the IC design of a wideband CMOS LNA using the cancelling techniques introduced in this chapter and looks at the challenges associated with the LNA implementation.
RFB eliminates the noise of A RFB noise is cancelled by the differential sensing
RFin
RFB
A
Mmain noise is eliminated by A
Rs Mmain
Fig. 2.32 The proposed noise-cancelling solution in this work can be seen as an SFB amplifier that implements noise free feedback resistor RFB
Chapter 3
Nanometer CMOS LNAs for Mobile TV Receivers
This chapter presents the design of a wideband low-noise amplifier in 65 nm digital CMOS technology that takes advantage of the noise and distortion cancelling techniques described in Chapter 2. In addition to examining the practicality of these techniques, this chapter also discusses the challenges associated with using a nanoscale technology such as the 65 nm CMOS process, including the MOSFET low output resistance, the high resistivity of the polysilicon material, the increased substrate coupling, and the use of a digital transistor layout that is not optimized for radio frequency (RF) operation. Furthermore, this chapter examines ways to achieve the nonlinearity requirement of the mobile TV application while providing RF gain control in the LNA. Lastly, this chapter looks at biasing techniques that allow the LNA to withstand process and temperature variations, and verifies the design’s performance using lab measurements from tests conducted on a prototype of the designed LNA.
3.1 Requirements of the LNA in Mobile TV Receivers This section presents the LNA specifications that allow a receiver to meet the requirements of the mobile TV standard, DVB-H. Figure 3.1 shows a DVB-H receiver that has been introduced to help extract the DVB-H LNA performance requirements. The receiver baseband circuits are based on the state-of-the-art solutions reported in [34, 35] (see Chapter 1). A post-mixer amplifier (PMA) is used to provide the gain and filtering needed to relax the dynamic range of the ADC (Analog-to-Digital converter). The gain and noise specs of these circuits are shown in Fig. 3.1. A maximum signal level of −6 dBm is required at the input and the output of the PMA circuit in order to allow the baseband chain to achieve the desired linearity of the DVB-H mobile TV standard. While the PMA circuit provides a variable gain
A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_3, © Springer Science + Business Media B.V. 2010
59
60
3 Nanometer CMOS LNAs for Mobile TV Receivers
−6 dBm max
−6 dBm max
Baseband Circuits PMA
ADC
RFin RF Front-end Circuit
Post Mixer Amplifier
PMA
Gain
0 dB-21 dB
Noise
10 nv/Hz0.5
I
Σ∆ ADC
ADC
Q
60 nv/Hz0.5
Fig. 3.1 The mobile TV DVB-H receiver
function to maintain the signal level at −6 dBm at its output, the RF front-end (FE) has to provide the same function (gain control) in order to adjust the signal level at the PMA input. The RF front-end circuit requirements (i.e., Gain, NF, IIP3) are derived taking the baseband circuit specifications into account. These front-end specifications then determine the LNA performance goals.
3.1.1 DVB-H RF Front-End Specifications Chapter 1 defined the NF and IIP3 values required for a receiver to be compatible with the DVB-H standard. A receiver NF of less than 5 dB is needed to achieve the sensitivity requirement (−86.6 dBm for 16QAM modulation). However, the lowest possible NF value is desirable (NF < 3 dB) to ensure satisfactory TV signal reception. This LNA design example targets a NF of 2 dB. The IIP3 value should be set to comply with all interferer patterns included in the standard. To simplify the analysis, only the most stringent interferer’s patterns are considered here (Table 3.1). The DVB-H front-end circuit has to provide a certain amount of gain in order to minimize the baseband noise contribution to the overall cascaded NF. Using Friis equation, 42 dB of gain is required to limit the noise contribution of the baseband circuits (Fig. 3.1) to 0.1 dB. This would then result in a cascaded NF of 2.1 dB (in case if the LNANF = 2 dB). Although a gain of 42 dB helps in achieving the sensitivity requirement, it can cause a lot of distortion if there is a large interferer
3.1 Requirements of the LNA in Mobile TV Receivers Table 3.1 Interferer signal patterns of the DVB-H standard [31] Modulation Undesired/ Pattern of interferer Interferer location desired (dB) S1 Analog 40 N + K (K ¹ 0, 1) L3 Digital N + 2b and N+4c 40 a N: desired channel (8 MHz bandwidth). b N + 2 blocker locates at 16 MHz. c N + 4 blocker locates at 32 MHz.
61
Desireda (dBm) −68 −75
Undesired (dBm) −28 −35
level at the RF input. For example, in an S1 pattern reception scenario, in order to prevent the receiver from clipping (maximum of −6 dBm at the PMA input), the gain of the front-end has to be adjusted to 22 instead of 42 dB. At this gain, the required NF for the DVB-H receiver is equal to 18 dB (carrier-to-noise ratio C/N is set to 19.2 dB for 64QAM signal). The 18 dB NF will limit the dynamic range of the DVB-H receiver since the S1: N + 2 interferer third-order distortion IM3 does not fall within the desired channel bandwidth. The IIP3 value does not affect the dynamic range in this case. On the other hand, in the case of receiving an L3 pattern, the interferers are two digital signals located at N + 2 and N + 4, i.e., 16 and 32 MHz away for the desired signal respectively. In this case, IM3 products fall within the desired channel and can cause third-order intermodulation distortion. If we set the signal level at the PMA input at −6 dBm while receiving a −35 dBm digital interferer (see L3 pattern in Table 3.1), the gain of the front-end should be adjusted to 26 dB. At this gain, an IIP3 value of −8 dBm is required (see Chapter 1). From the above discussion, it can be deduced that to achieve both L3 and S1 requirements, high linearity at high gain setting and low noise figure at lower gain setting are required. The trade-off between the linearity and noise is inevitable, which makes the design of the variable-gain DVB-H front-end circuit challenging. The target performance of the DVB-H front-end (FE) is given in Table 3.2. Additionally, it must be noted that a minimum gain of −6 dB is required so as not to saturate the baseband circuits at a maximum desired input signal level in the range of 0 dBm.
3.1.2 DVB-H LNA Performance Requirements An RF front-end circuit consisting of an RF passive attenuator, wideband LNA, and a passive quadrature mixer is introduced as a candidate for the DVB-H receiver (shown in Fig. 3.2). The high linearity of the passive mixer compared with the active mixer makes the former a better choice for the DVB-H application [40]. A quadrature version of the mixer is used to facilitate the generation of the in-phase (I) and quadrature- phase (Q) signals. The mixer’s specs are shown in Fig. 3.2. Based on these specifications, the LNA has to provide 36 dB of gain across the UHF band and an NF of 2 dB in order to achieve the sensitivity requirement of the DVB-H receiver.
62
3 Nanometer CMOS LNAs for Mobile TV Receivers Table 3.2 RF front-end target specifications for the DVB-H front-end receiver UHF: 47 0-862 MH z
Frequency range
Maximum gain
42 dB
NF @ Maximum gain
2 dB
Minimum gain
- 6 dBa
NF @ FE gain of 22 dB
18 dBb
IIP3 @ FE gain of 26 dB
>-8 dBmc
Supply voltage
2.5 V
Technology
65 nm CMOS
To achieve the sensitivity requirement
To achieve the maximum input level requirement To achieve the selectivity requirement To achieve the linearity requirement
aBaseband
circuit gain =0-21 dB. 64 QAM C R = 2/3, C/N = 19.2 dB, 8 MHz channel bandwidth (noise floor = -105 dBm). cAssuming -35 dBm digital interferers at 6 and 32 MHz. bAssuming
RF FE: RF Front-end
Quadrature Passive Mixer
RFin
I
Baseband Circuit Chain
LNA
Q Gain: 6 dB (Differential) Noise: 1.5 dB RF attenuator
IIP3: +20 dB
Fig. 3.2 The variable gain RF front-end for the DVB-H receiver
In the case of S-pattern reception, the LNA has to drop 20 dB from its maximum gain (36 dB) to avoid any kind of degradation in the overall linearity. However, system analysis show that in order to maximize the receiver dynamic range, a gain drop of only 6 dB should be provided by the LNA while the remainder of the 14 dB gain drop is to be provided by a passive RF attenuator preceding the LNA (see Fig. 3.2). A detailed discussion about the effect of using an RF attenuator to optimize receiver dynamic range is given in Chapter 4. According to this discussion, the minimum gain of the LNA is set to 30 dB. At this gain, the NF looking at the LNA input has to be 4 dB to achieve the dynamic range requirement at S-pattern interferer reception. Using the cascaded IIP3 expression:
1 IIP32 Front − end
≈
1 ( LNA _ Gain)2 + IIP 32 LNA IIP 32 Mixer
(3.1)
3.2 A 65 nm CMOS Wideband LNA Prototype
63
Table 3.3 The target specifications of the DVB-H wideband LNA Frequency range Maximum gain
UHF: 47 0-862 MH z 36 dBa
NF @ Maximum gain
2 dB
Minimum gain
30 dB
NF @ Mini mum gain
4 dBb
IIP3 @ Minimum gain
>-17 dBmc
Supply voltage
2.5 V
Technology
65 nm CMOS
To achieve the frontend NF specs To achieve the selectivity requirement and to maximize the receiver dynamic range To achieve the linearity requirement
a
The mixer has a differential gain of 6 dB. To allow a 18 dB cascaded NF at the S1 interferer pattern reception. Mixer IIP3= +20 dBm, LNA gain is 30 dB, a 10 dB gain reduction is provided by the RF attenuator. b c
the LNA IIP3 has to equal at least −17 dBm to help the DVB-H receiver to achieve the L3 linearity pattern requirement (in this case: the IIP3Front-end = −8 dBm, IIP3Mixer = +20 dBm, and the LNA_Gain = 30 dB). The summary of the DVB-H LNA requirements is given in Table 3.3. It is noted that the IIP3 value requirement of the LNA is relaxed by the amount of the passive attenuator attenuation (IIP3LNA = −7 dBm is required without using the passive attenuator to control the RF gain).
3.2 A 65 nm CMOS Wideband LNA Prototype This section describes the design of a wideband LNA that meets the specifications of the DVB-H mobile TV standard. The noise and distortion cancelling techniques described in Chapter 2 have been used in order to widen the LNA’s dynamic range. A variable-gain control mechanism is implemented in this LNA in order to improve the overall receiver linearity. The success of the fabricated LNA in achieving the target performance depends strongly on the techniques used to optimize the silicon layout. Additionally, the layout parasitics as well as the package parasitics must be taken into account before taping out the design. The practicality of the design demands reliable bias generator circuits that can help the LNA circuit to achieve the performance goals independent of the amount of power supply noise, and process and temperature variations. Furthermore, to facilitate the LNA prototype characterization in the lab, auxiliary circuits are needed to interface the LNA to the 50 W environment of the measurement equipment. Two multi-mode buffers are designed and integrated to allow successful noise figure and linearity measurements for the DVB-H LNA prototype.
64
3 Nanometer CMOS LNAs for Mobile TV Receivers
3.2.1 LNA Core Circuit 3.2.1.1 LNA Circuit Description The schematic circuit diagram of the LNA is shown in Fig. 3.3. The common-gate (CG) transistor M1 is used to provide a wideband input matching (across the UHF band) to a 50 W source impedance. The feedback resistor RF is used to eliminate M1 noise as well as to cancel part of its distortion as described in Section 2.4.2. The inductor Lbias is used to provide a DC path for the M1 device. The use of an inductor instead of the conventional bias method of using a current source helps in avoiding the added noise of the latter. The inductance value is selected so any resonance that might occur at the LNA input is out of its frequency band of operation. 2.5 V
RFin 2.5 V
Cc2
Common-gate branch
M4 20K
+ OTA −
20K
Ο
R1
Vbias_CS_P
Vbias_cas_P
Ο +
RFout
M5 CMOS3
−
2.5 V
Ο
Ο
M6
M3
Cc3
20K
M1 Common-source branch
RF
Cc1
RFin
M2
Lbias
5K
Ο
5K CMOS2 CMOS1
10µA
2.5 V
Vbias_cas_N
Vbias_CS_N
Ο Vbias_CG
Fig. 3.3 The schematic diagram of the DVB-H wideband low noise amplifier
VREF
3.2 A 65 nm CMOS Wideband LNA Prototype
65
The common-source (CS) transistors M2 and M4 generate another path for the signal that allows cancelling the CG branch noise, M1 & RF noise (Section 2.4.2.1), when the output signal is sensed differentially. The current ratio in both branches (CG and CS branches) is adjusted to facilitate achieving the condition of the noise cancellation mentioned in Chapter 2. The noise of M2 and M4 is eliminated by the negative active feedback loop provided by the CG transistor M1 and the feedback resistor RF. The bias current of M2 is reused to bias M4 to save the current consumption. In addition to the power saving, the second-order distortion that M2 generates cancels some of the second-order distortion of M4. Since M1, M2, and M4 devices have negative feedback loops around them, the LNA IIP3 is expected to be limited by the second-order interaction between the CG and CS transistors (see Section 2.4.2.3). Cascode transistors M3 and M5 offer the capability of boosting the gain in order to overcome the low intrinsic gain of a MOSFET in 65 nm CMOS technology. However, this comes at the expense of increasing the voltage headroom and generating an unstable DC point at the (−RFout) node. To mitigate this, an operational transconductance amplifier (OTA) is introduced to stabilize this node. The OTA compares the voltage of the (−RFout) node to a reference voltage VREF and then uses the output to control the bias of the CS transistor M4. This guarantees that the DC voltage at the (−RFout) node is equal to a value very close to VREF depending on the precision of the transconductance amplifier (OTA gain). Two coupling capacitors (Cc1 and Cc2) are thus required to decouple the bias of M2 from M4 bias so that the OTA does not affect the M2 bias. Another coupling capacitor Cc3 is also used to facilitate the CG transistor M1 biasing. These coupling capacitors provide flexibility in optimizing the LNA design. The noise coupled through the power supplies as well as the noise generated from the bias generator circuits used to bias the CS and CG transistors (M1, M2, and M4) can have a deleterious effect on the LNA noise figure. In fact, their noise contributions can dominate the LNA NF. Hence, all the bias voltages are filtered by using low pass filters. Although the capacitors used in these filters are based on MOS devices (MOS-capacitors) to save silicon area, the amount of filtering is still limited by the area budget constraints. CMOS1 and CMOS2 are based on N-channel devices while CMOS3 is based on a P-channel device. This kind of implementation helps in rejecting any noise that might couple to the LNA through the supplies and ground traces.
3.2.1.2 LNA Design Trade-Offs Most of the amplifier trade-offs have been presented in detail in Chapter 2. Because 65 nm CMOS models were still unavailable at the time when the theory discussed in Chapter 2 was developed, simulation results used to support the theory were based on 0.13 µm CMOS models. Additional design considerations emerged with the use of this newer nanometer CMOS technology (i.e., 65 nm) and it is important that these considerations be highlighted here.
66
3 Nanometer CMOS LNAs for Mobile TV Receivers
Simulation results show that the IIP3 value of the designed LNA in 65 nm CMOS is dominated by the MOSFET output resistance nonlinearity (see Section 2.1.2.1). Although the achieved IIP3 of −13 dBm using 65 nm CMOS models can still meet DVB-H receiver specifications, this value differs from the one mentioned in Chapter 2 (IIP3 = −5 dBm using 0.13 µm CMOS models). Improving the IIP3 value requires either an increase in the supply voltage or a decrease in the output signal swing on the CS transistors M2 & M4. Increasing the supply voltage comes at the expense of increasing the power consumption while decreasing the output signal swing degrades the LNA noise figure. As mentioned in Chapter 2 a higher loop gain in the feedback loop around M2 (proportional to the output signal swing of M2) results in a lower noise figure of the amplifier. Hence, decreasing the loop gain lowers the output signal swing and therefore improves the linearity and degrades the noise figure. The question that remains is what is the effect of the cascode transistor M3 on the CS transistor’s (M2) nonlinearity? It can be stated that lowering the output signal swing by adding the cascode transistor M3 does not improve the M2 nonlinearity. Adding the cascode increases the magnitude of the nonlinear coefficients due to the decrease of the DC drain-source voltage Vds, which mitigates the benefit of decreasing the output signal swing. As a result of this discussion, since IIP3 is limited by the output nonlinearity of the MOSFET, the M1 and M4 CS transistors are biased at the moderate inversion region in order to improve the current efficiency by maximizing the (gm/I) ratio. This allows lowering the power consumption for the same dynamic range for the LNA. The rest of the trade-offs (matching, gain, NF, power) are the same as described in Chapter 2. The loop gain is selected to be 1.6 by setting the ratio of (gm2/gm3) to be equal to 2. This loop gain value is found to be the best compromise between the input matching, noise figure, and the power consumption taking the 2.5 V supply voltage into account. Using this loop gain value, the noise of M1 and RF is completely eliminated and the noise of M2 and M4 is reduced by 50% relative to their noise without applying the feedback. The amount of the current noise reduction achieved in this design is shown in Fig. 3.4.
3.2.1.3 LNA Variable-Gain Control Function As discussed in Section 3.1, the LNA should be able to control the gain of the RF front-end of the DVB-H receiver. Implementing the gain control function in an LNA has been introduced in [22, 29]. In [22], the gain control was provided in the LNA by varying the gate–source bias voltage (Vgs) of the transistors, while the gain control in [29] was provided by varying the load current. A detailed discussion of these gain control techniques is given in the next chapter, where the dynamic range optimization for the DVB-H receiver is presented. In this LNA example, the first approach has been considered to provide the gain control function in the DVB-H LNA design example. However, instead of varying the gate–source voltage, the drain–source bias voltage is changed, so as to
3.2 A 65 nm CMOS Wideband LNA Prototype
a
100% reduction in RF&M1 noise R1
Ro
0.6in
Noise current
67
b
50% reduction in M1 noise R1
Ro
0.5in
RFout
RFout
3in M1
M1
in RF
M3
0.06in
RF
M3
M2 RFin
M2 RFin
in
Noise current
Fig. 3.4 The amount of noise cancellation achieved in the DVB-H LNA: (a) illustrates the amount of the noise current that leaks from the CG branch to the outputs while (b) shows the amount of noise current that leaks from M2 transistor to the outputs. Ro is the equivalent resistance at the drain of the cascode transistor M3
take advantage of the variations that exist in the nonlinear output resistance for a MOSFET in 65 nm CMOS technology when the drain–source voltage (Vds) is varied. As Vds increases and the pinch-off moves toward the source, the rate at which the depletion region around the source becomes wider decreases, resulting in a higher incremental output resistance. The output resistance for a MOSFET (roMOS) can be approximated by [106]:
roMOS =
2L 1 1 − ( DL / L ) I D
qN B (Vds − Vds ,sat ) 2e si
(3.2)
where Vds,sat is the drain-source voltage at the onset of pinch-off, L is the transistor channel length, ID is the drain current, esi is the dielectric constant of the silicon, q is the electron charge, NB is the doping concentration of the substrate, and DL is the variation in the channel length for a given increment in Vds. As shown in Fig. 3.5, since the OTA fixes the DC bias at the (−RFout) node to a value equal to VREF, the bias of this node can be varied by varying the VREF value. Changing the bias at the (−RFout) node consequently changes the total output resistance at this node (RoutCS-branch), and thus changes the LNA gain. This can be seen by writing the LNA gain expression (only the common-source branch is considered here)1: 1 In this derivation, the effect of ro2 and ro4 on the effective transconductance of this branch is ignored.
68
3 Nanometer CMOS LNAs for Mobile TV Receivers 2.5 V
RFin 2.5 V M4 +
OTA
Ο
Ο
−
Vbias_CS_P
Vbias_cas_P M5
RoutCS-branch
10µA
− RFout 2.5 V
Common-gate branch Vbias_cas_N
Ο
RFin
M3
M2
M1
Ο VREF
20K
Common-source branch
RF
Fig. 3.5 Simplified schematic diagram for the DVB-H LNA indicating the gain control mechanism
Gain _ LNACS
branch
≈ ( gm 2 + gm 4 ) RoutCS
branch
(3.3)
where, RoutCS
branch
= (1 + ( gm 3 + gmb 3 )ro2 )ro3 + ro2 / / (1 + ( gm 5 + gmb 5 )ro4 )ro5 + ro4 (3.4)
where gm2, gm3, gm4, and gm5 are the transconductances of M2, M3, M4, and M5, respectively, gmb3, gmb5 are the back gate transconductances of M3 and M5, respectively, and ro2, ro3, ro4, and ro5 are the output resistances of M2, M3, M4, and M5 respectively. As can be seen from (3.4), if the designed LNA gain is limited by ro2 and ro3, as is the case in this example, and the bias voltage at the (−RFout) node decreases, the first part of expression (3.4) decreases while the second part increases. This results in a decrease in the total output resistance (RoutCS-branch) at the output node and thus the LNA gain decreases. One can note that the gm3, gm5 values also change with Vds (gm ∝ (1 + lVds ) ) where l is the transistor channel length modulation coefficient.
3.2 A 65 nm CMOS Wideband LNA Prototype
69
The circuit needed to generate the variable bias VREF is given in Section 3.2.2.3, where the design of the OTA bias circuit is presented. VREF variations are limited by the bias condition required to keep M2, M3, M4, and M5 transistors in the saturation region. In this example, the VREF is varied just twice to provide two gain steps for the DVB-H LNA, each one equal to a 3 dB drop from the maximum gain. A decrease in the LNA gain results in a change in performance relative to that achieved at the maximum gain mode of operation. The LNA NF degrades while decreasing the gain since changing the CS branch gain violates the noise-cancelling condition (see Section 2.4.2.1) required to optimize the LNA NF. However, since the decrease in the LNA gain is a result of a large input signal at the LNA input, the noise performance required in this case is somehow relaxed (see Table 3.3). The achieved linearity at the LNA minimum gain settings also differs from that achieved at maximum gain mode. Since the nonlinearity in the designed DVB-H LNA is limited by the MOSFET output resistance nonlinearity (i.e., ro2), decreasing the bias voltage at the (−RFout) node (to decrease the LNA gain) degrades the IIP3 value. This can be explained by noting that decreasing Vds voltage for a MOSFET results in an increase in the magnitude of the nonlinear coefficients of its output resistance rMOS (see Section 2.1.2.1), which results in a degradation of the IIP3 value. The achieved IIP3 value at minimum gain mode is important for the DVB-H receiver to meet the linearity requirements of the L3 interferer pattern. Another factor to consider is the impact of changing the LNA gain on the input matching requirement. Since the transconductance of transistor M3 changes while changing the LNA gain, the loop gain around the matching transistor M1 also changes. As discussed in Section 2.4.2.2, changing this loop gain value results in changing the LNA input impedance. Therefore, if the loop gain around M1 is optimized to set the LNA input impedance to a 50 W value at maximum gain mode of operation, the input matching would degrade at the LNA minimum gain settings. To resolve this issue, the feedback resistor RF can vary to compensate for the input matching deviation caused by changing the loop gain. However, in this design example, the degradation in the input matching at the LNA minimum gain settings was found to be acceptable and hence the DVB-H LNA can still be matched to a 50 W at the minimum gain mode. 3.2.1.4 LNA Physical Implementation In order to achieve the expected LNA performance, especially the low noise figure, very careful attention must be given to the LNA layout. The MOSFET transistor provided through the digital design kit is not optimized for RF operation. First, the transistor layout has to be designed in a way that makes the effect of the MOSFET gate resistance on the LNA noise figure negligible. The transistor is broken into multiple fingers in order to overcome the high resistivity of the polysilicon material in the 65 nm CMOS process (~14 W/square) [103]. For example, a minimum feature length MOSFET in 65 nm CMOS with a 15 µm width results in 1 kW gate resistance. The same transistor results in 2.7 W gate resistance when it is broken
70
3 Nanometer CMOS LNAs for Mobile TV Receivers
into 20 fingers. This resistance can be decreased even further by contacting the gate at both ends (~0.6 W in this case). The only disadvantage of breaking the transistor into multiple fingers is the increase in the substrate resistance and routing parasitics. However, this can be avoided by subdividing the transistor into multiple unit cells [104]. Any ground current that might flow through the substrate is now soaked up by the low resistance path through the substrate contacts. Additionally, using the multiple unit cells configuration for a MOSFET allows the DC current to split such that the maximum current density for each of the metals used in the transistor layout does not violate the maximum amount allowed by the design rules of the 65 nm CMOS technology (i.e., 1.5 mA/µm for metal 1). Another issue that has to be considered with regards to the transistor RF optimization is the substrate coupling. Since the ultimate goal of the designed LNA is to be integrated with the rest of the DVB-H system, avoiding the substrate coupling is critical as it can increase the noise floor of the LNA circuit, and thus, damage the overall receiver sensitivity. To shield the NMOS transistors from any noise that might couple through the substrate, a deep N-well is used to improve the isolation. Doing so also allows connecting the source of the transistor to the substrate, thus one can avoid changing the device characteristics by the bulk-source voltage variations. An example of a digital transistor layout that is optimized for the RF operation is shown in Fig. 3.6. The coupling capacitors (Cc1,Cc2, and Cc3) are implemented using a number of shunt MIM-capacitors in the N-well in order to improve the LNA shielding (see Fig. 3.3). The N-well of Cc1 and Cc3 is connected to the ground bond while the N-well of Cc2 is connected to the power supply bond. The silicon area of these capacitors is reduced by taking the transistor sizes of M1, M2, and M4 into account. Other types of capacitors that are available in the 65 nm CMOS process are the MOS-capacitors. These capacitors are used in this design example to implement the filtering capacitors CMOS1, CMOS2, and CMOS3 because of their high capacitance density (10 fF/mm2 in the 65 nm CMOS process) [105]. Thus, a huge saving in the silicon area can be achieved by using MOS-capacitors instead of MIM-capacitors (2 fF/mm2 in the 65 nm CMOS process). Note that MOS-capacitors have not been used to implement the coupling capacitors because of the distortion they can generate in the signal path. A detailed discussion regarding the implementation
Substrate contacts One transistor unit Gate fingers
Gate is contacted at both ends
Fig. 3.6 Optimizing the layout of the transistor M4 for RF operation
3.2 A 65 nm CMOS Wideband LNA Prototype
Metal 6
71
Metal 7
M4 transistor
Gate is contacted at both ends
Fig. 3.7 Die photograph of the PMOS transistors M4 used in LNA core circuit. Transistor M4 has a length of 60 nm and a width of 136 µm. Each transistor has been broken into multiple unit cells. Each unit cell has 17 fingers. The polysilicon gate is connected at both ends to further decrease the transistor’s gate resistance. The ultra-thick metal 7 as well as the thick metal 6 are also shown
of MOS-capacitors in the 65 nm CMOS process is given in Chapter 4. The resistors used to provide filtering to the supply noise have been implemented by using P+ poly resistors without silicide material since they have a very high sheet resistance (~700 W/square). A final issue to consider in the layout of the LNA is the routing parasitics. The ultra-thick metal (metal 7, thickness = 3.4 nm) is used as much as the process design rules allow in routing the signal path traces from the LNA input pad to the LNA output pad. In fact, sometimes, a stack of metals (metal 1 through 7) is used to further decrease the parasitic resistance. During the LNA circuit simulations, the extracted parasitics have been taken into consideration. Moreover, the pad capacitance as well as package bond wire parasitics have also been taken into account. The die photograph of the RF transistors (M4) is shown in Fig. 3.7, while the fabricated die photograph of the LNA core circuit is shown in Fig. 3.8.
3.2.2 DC Bias Generator Circuits The design of the bias circuits is critical for ensuring optimal LNA performance. The bias circuits should be designed in a way that makes the LNA performance independent of the power supply voltage variations. Any coupled noise from the supply can easily increase the noise floor of the LNA and thus, degrade its noise performance. Additionally, any noise generated from the bias circuits themselves should be filtered so as not to contribute to the LNA NF.
72
3 Nanometer CMOS LNAs for Mobile TV Receivers 230 µm
180 µm
Cc1
37µm
OTA
M4
30µm
Cc2
Cc3
Fig. 3.8 Die photograph of the LNA core circuit
Another important property of these bias circuits is that they should provide a stable current independent of temperature and process variations. The MOSFET threshold voltage as well as the mobility exhibit temperature dependence [106]. The simple resistive biasing circuit shown in Fig. 3.9, for example, generates unstable drain current since the gate voltage is dependent on the supply, process, and temperature. Therefore, it can negatively affect the NF performance of the designed LNA. Current mirrors are used in this work to mirror a precisely-defined current source that is already available through a bandgap reference operating at a 1.2 V supply [106]. The bandgap reference generates bias that is independent of the supply voltage and robust against temperature and process variations. The bias scheme used for the mobile TV LNA is illustrated in Fig. 3.10. Three different bias generator circuits have been designed to bias the LNA core. Bias circuit (1) takes a 10 µA bandgap current and provides the bias for the cascode transistors M3, M5 and M6. Since transistors M3 and M6 share the same bias, only two different voltages have been generated (Vbias_cas_N, Vbias_cas_P). Bias circuit (2) also takes a 10 µA bandgap current and provides the bias for the CS transistor M4 (Vbias_CS_P). An OTA has been used to stabilize the LNA common-mode voltage. Finally, bias circuit (3) takes a 100 µA bandgap current and provides the bias for the transistors M1 and M2 (Vbias_CG, Vbias_CS_N). To resolve any chance of transistor breakdown in these bias circuits, the transistors are chosen to be 2.5 V devices that are available in the 65 nm CMOS process instead of the 1.2 V devices used in the LNA core. The following sections describe each of these bias generator circuits in detail.
3.2 A 65 nm CMOS Wideband LNA Prototype
73
Fig. 3.9 Simple resistive bias generator circuit noise
Iout R1 M1 noise
R2
1.2V Bandgap Reference Current
10 µA
Vbias_cas_N Vbias_cas_P
100 µA
Cascode Bias Generator Circuit (M3, M5, M6)
2.5V
CS & CG Bias Generator Circuit (M1, M2)
(1)
Vbias_CS_N Vbias_CG
(3) Common-Mode Stabilization Circuit (OTA-M4) (2)
Vbias_CS_P
Fig. 3.10 Biasing schemes used to generate the required bias for the LNA core
3.2.2.1 Cascode Bias Generator Circuit The circuit used to generate the bias for the cascode transistors in the LNA core is shown in Fig. 3.11. The current mirror employed by MBias_cas1 and MBias_cas2 maps the 10 µA of the bandgap reference current to 20 µA. This 20 µA current is then used to provide the bias voltages, through a variable resistor, to the cascode transistors. A variable resistor is selected to provide flexibility while conducting the lab measurements. In this case, the generated voltage is independent of the power supply noise as well as the process and temperature variations. The current mirrors employ large channel length (L = 2 µm) in order to improve the accuracy of the current mapping
74
3 Nanometer CMOS LNAs for Mobile TV Receivers 2.5 V
CBias_cas3
Bandgap Reference 1.2 V MB as_cas3
MBias_cas4
MBias_cas5
10 µA 4K
Vbias_cas_N
R MBias_cas1
MBias_cas2
4K
R
41.25K
41.25K CBias_cas1
CBias_cas4
20 µA
Vbias_cas_P
20 µA
CBias_cas2
20 µA
Fig. 3.11 Bias generator circuit of cascode transistors M3, M5, and M6
by increasing the transistor output resistance. Additionally, the same length for all of the transistors is used to minimize the errors due to side-diffusion of the source and drain area. MOS-capacitors CBias_cas1 and CBias_cas2 are used to guarantee that the cascode gates are AC grounded, hence improving their characteristics as current buffers. The variable resistor is shown in Fig. 3.12. It is digitally controlled through four switches SW1, SW2, SW3 and SW4. When the switches are closed, the output voltage is ~825 mV, while when they are open, the output voltage is ~1.8 V. The switches can turn ON and OFF in multiple combinations to generate binaryweighted voltages with a minimum increment of 60 mV. The type of the resistor units used to implement the variable resistor have been selected to match the resistor used in the bandgap circuit in order to provide voltage that is independent of the temperature variation [106]. 3.2.2.2 CS and CG Bias Generator Circuit The circuit used to generate the bias for the CS transistor M2 and CG transistor transistor M1 is shown in Fig. 3.13. The current mirror employed by MBias_CS&CG1 and MBias_CS&CG2 maps the 100 µA of the bandgap reference current to 900 µA. The CS transistor is biased through the current mirror transistor MBias_CS&CG5, while the CG transistor is biased through the voltage generated by a 2 kW resistor. These biases are independent of the supply, process, and temperature variation. However, they can be affected by the mismatch between the mirror transistors especially with
3.2 A 65 nm CMOS Wideband LNA Prototype
75
Fig. 3.12 The binary-weighted digitally-controlled variable resistor used in the cascode bias generator circuit
2.5 V
20 µA
4K
I<3:0>
SW1
SW2
SW3
SW4
Output
25K
12.5K
6.25K
3.125K
41.25K
using the 65 nm CMOS process. In general, the transistor matching is more of a concern in nanometer CMOS processes than when using other, less advanced technology [107]. In order to improve the matching in this circuit, wide transistors have been used for the mirror transistors. Additionally, dummy devices have been added to the transistor layout to mitigate any mismatch that may occur from the well proximity effect [108]. 3.2.2.3 Common-Mode Stabilization Circuit The operational transconductance amplifier (OTA) shown in Fig. 3.14 has been used to bias the CS transistor M4 of the LNA core. A feedback loop is employed around the OTA to facilitate generating a stable DC value equal to VREF at the common-mode node of the LNA CS branch. The OTA consists of a differential pair (MOTA5 & MOTA6) loaded with an active current mirror (MOTA7 & MOTA8). The tail current is provided by two current mirrors (MOTA1 & MOTA2) and (MOTA3 & MOTA4)
76
3 Nanometer CMOS LNAs for Mobile TV Receivers 2.5 V Bandgap Reference 1.2 V
MBias_CS&CG3
MBias_CS&CG4
100 µA
MBias_CS&CG6 60K 2K
Vbias_CG
60K
MBias_CS&CG1
MBias_CS&CG5 MBias_CS&CG2
900 µA
Vbias_CS
900 µA
Fig. 3.13 Bias generator circuit of the CS (M2) and CG (M1) transistors
that are biased by a 10 µA bandgap reference current. A variable resistor R is employed to provide variable values of VREF in order to facilitate the gain control function in the LNA. The same variable resistor described in Section 3.2.2.1 is reused here. The open loop gain of the OTA is chosen according to the precision required of VREF (i.e., 40 dB of gain gives 1% error in VREF). A 30 dB of gain is chosen in this design example. The sizes of the transistors have been selected to make them operate in the saturation region at different values of VREF. One of the most important issues in this bias circuit is the OTA stability. As shown in Fig. 3.14, the LNA-CS branch acts as a second stage class AB amplifier (M4 & M5) connected at the output of the OTA circuit. M4 and M5 transistors create a high impedance point at the (−RFout) node in the feedback loop, which can cause the negative feedback to be converted to positive feedback at certain frequencies, and thus create a stability problem. Moreover, the added gain of the M4 and M5 transistors increases the chance of oscillation. Therefore, a DC stability analysis has been conducted on this feedback loop. This analysis took into consideration all the non-dominant poles originating from the low-pass filters used for noise filtering along with the coupling capacitors. Additionally, the output load capacitance CL of the buffer needed to interface the LNA to the measurement equipment was also taken into account. Simulation results showing the loop gain as well as the phase margin of the feedback loop are shown in Fig. 3.15. A loop gain of 49 dB and 80° phase margin has been achieved to provide the stability of the feedback loop. The accuracy in achieving a precise value of VREF depends strongly on the physical implementation of the OTA circuit. The physical implementation of the OTA circuit
3.2 A 65 nm CMOS Wideband LNA Prototype
77
OTA Circuit 2.5 V
2.5 V
20 µA 1.2 V
MOTA3 Bandgap Reference
MOTA4
VREF
VREF >2.5 - VGS4 - VthOTA6
R
LNA-CS branch
10µA
Inp
Inn MOTA5
MOTA6 Vbias_CS_P
MOTA1
For MOTA6 to be in saturation:
MOTA2
MOTA7
2.5V
M4
MOTA8
M5
Ο - RFout
20K Feedback loop
M3
CL
Fig. 3.14 Operational transconductance amplifier used to bias the CS transistor M4. The OTA circuit with the LNA-CS branch circuit acts as a two stage opamp
is shown in Fig. 3.16. As noted before, dummy elements have been used to improve the matching between the mirror transistors. Moreover, each transistor has been broken into multiple unit cells as recommended by the foundry to avoid the oxide stress effect that might occur with using the 65 nm CMOS process [108]. One of the most important issues in the OTA physical implementation is the matching between the differential pair transistors. The interdigitating layout style has been used to implement the differential pair transistors (MOTA5 & MOTA6) in order to make them insensitive to the substrate doping variations, and thus further improve the matching.
3.2.3 Multi-Mode Test Buffer Circuits This section presents the interface circuits needed in order to characterize the mobile TV LNA 65 nm test chip. Two buffers and a transformer is used to allow the RF differential signal at the LNA outputs to drive the input port of a 50 W spectrum analyzer (see Fig. 3.17). The buffers should have a very high input
78
a
3 Nanometer CMOS LNAs for Mobile TV Receivers
60
Loop Gain (dB)
40 20 0 −20 −40 −60 100
1K
10K
100K
1M
10M
100M
Frequency (Hz)
Phase Margin (Degree)
b
200
100
0
−100
−200 100
1K
10K
100K
1M
10M
100M
Frequency (Hz) Fig. 3.15 Stability analysis of the DC stabilization loop used to bias the mobile TV LNA: (a) is the loop gain and (b) shows the phase margin results
impedance and a 100 W output impedance. The balun then transforms the output differential signal from the output of the two buffers to a single output signal. The turns’ ratio is selected to be two for this transformer to allow a 50 W impedance to be achieved at its output.
3.2 A 65 nm CMOS Wideband LNA Prototype
79 Tail current transistor is broken into multiple unit cells
D
D In(+) In(-)
Interdigitating differential pair
D
D
D Out
D
MOTA7 MOTA2
D
MOTA8
D
MOTA1
D
D: Dummy
Fig. 3.16 The OTA physical implementation
The interface circuits
Cc_buff Buffer
RFout Spectrum Analyzer
RFin
LNA
200Ω Ω
Cc_buff Buffer
50Ω
Cable
Fig. 3.17 Interface circuits needed to characterize the LNA 65 nm CMOS test chip
Most important among the figures of merit for the designed LNA are the NF at its maximum gain (36 dB) and the IIP3 at its minimum gain (30 dB) (see Section 3.1.2). The buffers have to be designed in a way that facilitates a high accuracy for these measurements. In other words, the performance of the buffers should not interfere with the LNA performance during the LNA testing. Accordingly, the buffer performance requirements are discussed first in the following section and then the buffer circuit design realization follows.
80
3 Nanometer CMOS LNAs for Mobile TV Receivers
3.2.3.1 LNA Buffer Performance Requirements As previously mentioned, the buffers should facilitate accurate performance measurements of the LNA test chip while maintaining their 100 W output resistance requirement. The LNA buffers in general should provide some gain to prevent the noise floor of the spectrum analyzer from interfering with the noise floor of the LNA. In this work, a 36 dB of LNA gain results in −137 dBm/Hz noise floor level at the LNA output (to allow an accurate measurement for NF = 1 dB). The noise floor of the spectrum analyzer used for testing (Agilent E4408B) is −168 dBm/Hz, making the LNA noise floor at the input of the spectrum analyzer 31 dB larger than the spectrum analyzer noise floor itself. Accordingly, the buffers’s gain factor in our cases is relaxed. Another factor to consider is the noise figure of these buffers. Using Friis equation (see expression 1.1), one can calculate the required NF of the buffers so that their noise would not interfere with the measured NF of the LNA. The situation is helped by the 36 dB of gain of the LNA that makes the noise requirements on these buffers very relaxed (6 dB NF – relative to the 50 W source resistance). However, this huge gain makes obtaining accurate IIP3 measurements very challenging. A −13 dBm IIP3 is expected to be measured at the LNA input at its minimum gain (30 dB). Therefore, in order for the buffers’ nonlinearity not to contribute to the measured LNA linearity, the buffers IIP3 has to be >+21 dBm, which is again very challenging. 3.2.3.2 LNA Buffer Circuit Realization The circuit schematic diagram for the LNA buffer is shown in Fig. 3.18. The buffer topology is based on the shunt-series amplifier topology. The reason for selecting this topology over more conventional buffers (i.e., source follower) is the stringent linearity requirements. The buffer shown in Fig. 3.18 consists of common-source transistors M1 and M2 and is self biased through RFBbias. A coupling capacitor (Cc_buff) is used in order to facilitate the simple self biasing technique and to help prevent the buffer from clipping at the output and damaging the LNA IIP3 measurements. The value of this capacitor is chosen so that it would not affect the LNA bandwidth. R1 and R2 act as degeneration resistors for the common-source transistors in order to improve the buffer nonlinearity. The degeneration resistor can be varied by switching ON and OFF the transistor switches M3 and M4 through a single bit called “Test mode control bit”. When M3 and M4 are ON, the buffer operates in what is called high-gain mode, while when they are OFF the buffer operates in lowgain mode. In high-gain mode the buffer consumes 13 mA and is used for LNA NF measurements, while in low-gain mode it consumes 3 mA and is used to measure the LNA IIP3. In high-gain mode, the buffer has a negligible effect on the cascaded NF measured by the spectrum analyzer (0.04 dB), therefore guaranteeing accuracy in the LNA NF measurement. In low-gain mode, increasing the value of the degeneration resistor helps the buffer to achieve an IIP3 of +22 dBm. In order to preserve
3.3 Experimental Results
81 2.5 V
R2
M4
R4 M2
Cc_buff RFin
RFout
RFBbias
M1 R3
R1
M3 Test mode control bit
Fig. 3.18 LNA-buffer schematic circuit diagram
the output matching requirement while changing the buffer test mode (i.e., buffer’s gain), the feedback resistor RFBbias must vary. A digitally-controlled resistor has been implemented for that purpose. A detailed design analysis of a shunt-series amplifier can be found in reference [58].
3.3 Experimental Results 3.3.1 Test Environment Descriptions The complete schematic diagram of the LNA test chip is shown in Fig. 3.19. The LNA core, the bias generator circuits as well as the two output buffers have been implemented in one die using 65 nm CMOS technology. The fabricated die is shown in Fig. 3.20. The die occupies (472 × 345 µm) of silicon. The test chip die is wirebonded onto a circuit board as shown in Fig. 3.21. An I2C chip (inter-integrated circuit), was used to program the CMOS test chip (i.e., LNA gain control, buffer test mode). I2C is a multi-master serial computer that can allow a microcontroller to control a test chip with just two general purpose
100 µA
60K
2K
Vbias_cas_N
MB as_cas4
MB as_CS&CG4
2.5 V
41K
R
4K
MB as_CS&CG5
MB as_cas4
MB as_CS&CG2
MB as_CS&CG3
10 µA
MB as_cas3
Vbias_CG
RFin
Vbias_cas_P
Vbias_CS
60K
MB as_CS&CG6
41K
R
4K
MB as_cas5
Lbias
RF
ΟΟ
R1
M1
M6
+
5K
5K
Cc1
Cc3
Ο
-
Ο
Fig. 3.19 The complete schematic diagram of the 65 nm CMOS LNA test chip
MB as_CS&CG1
1.2 V
MB as_cas1
1.2 V
2.5 V
Ο
20K
M2
RFin
Buffer
Buffer
Off-chip
Cc_buff
Cc_buff
20K
CMOS3
20K
2.5 V
Vbias_CS_P
M3
M5
M4
Cc2 1.2 V
10 µA
RFout
MOTA2
MOTA3
Inp
MOTA7
MOTA5
MOTA4
MOTA8
MOTA6
R
Inn
VREF
20 µA
82 3 Nanometer CMOS LNAs for Mobile TV Receivers
3.3 Experimental Results
83
345µm
Pads
RFout
LNA core
Output buffers Bias generator circuits
VDD
RFout +
472µm
-Gnd2
Gnd1
RFin
Fig. 3.20 The 65 nm DVB-H LNA test chip die photo
SMA of the RFin LNA die I2C ports
Transformer
LNA Board
1.2V
SMA of the RFout 2.5V
Gnd
Fig. 3.21 LNA die bonded to a circuit board for testing
I/O (input/output) pins and software. The Aardvark host adapter was used to interface a Windows OS via USB (Universal Serial Bus) to a downstream embedded system environment and transfer serial messages using the I2C protocol. The 1.2 V supply is used to bias the I2C chip as well as the bandgap reference generator circuit, while the 2.5 V supply is used to bias the LNA test chip (LNA core, bias
84
3 Nanometer CMOS LNAs for Mobile TV Receivers
Windows OS
Aardvark Adapter
Power Supplies
Circuit Board
Spectrum Analyzer
Function Generators
Network Analyzer
Fig. 3.22 Test setup in the laboratory
circuits, and buffers). A balun transformer, B4F type, provided by TOKO Inc. [109], was selected to drive the input port of the measurement equipment since it can achieve a very low insertion loss across the UHF band. The test setup used to characterize the LNA circuit board is shown in Fig. 3.22.
3.3.2 Measurement Results Several measurements were conducted to verify the LNA’s performance in terms of the DVB-H mobile TV standard requirements (Table 3.3). The HP 8753D network analyzer was used for S-parameter measurements. These measurements verify the LNA input matching (S11), the output matching (S22) and the reverse isolation (S12). To ensure the accuracy of these measurements the board traces must be taken into account while calibrating the network analyzer. A special board that does not contain the test chip has been built especially for that purpose. This board with the HP 8753D calibration kit was used to calibrate the network analyzer, thus, guaranteeing that the S11, S22, and S12 measurements accurately represent the LNA’s performance. An S11 < −18 dB was measured at the input of the LNA across the UHF band as shown in Fig. 3.23. In this measurement, the OTA bias was adjusted to make the LNA operate at the maximum gain mode. The same test was then repeated but by
3.3 Experimental Results
85
40 S11 (dB) 20
0
–20
–40
150MHz
900MHz
Fig. 3.23 Measured S11 of the LNA at maximum gain mode of operation 40 S11 (dB) 20
0
–20
–40
150MHz
900MHz
Fig. 3.24 Measured S11 of the LNA at the first gain step
varying the OTA bias to change the LNA gain. Values of S11 < −12 dB and S11 < −8 dB were measured for the first and second gain steps of the LNA, respectively across the UHF band. These measurements are shown in Figs. 3.24 and 3.25. As expected, there is degradation in the input matching of the LNA while varying the OTA
86
3 Nanometer CMOS LNAs for Mobile TV Receivers 40 S11 (dB) 20
0
–20
–40
150MHz
900MHz
Fig. 3.25 Measured S11 of the LNA at the second gain step
bias due to changing the loop gain around the CG transistor (see Section 3.2.1.3). However, S11 values still satisfy the input impedance required to terminate the transmission line that will deliver the DVB-H signal from the antenna to the LNA. One important property that should be achieved in the DVB-H LNA is the high reverse isolation (see Chapter 4). A value of S12 < −40 dB was measured across the UHF band as shown in Fig. 3.26. The very low measured S12 values show the effectiveness of using the cascode transistors to improve the isolation of the output and input ports. Having a separate downbond for the CS transistor M2 also helped in achieving that value (see Fig. 3.20). To verify the buffer output matching, S22 < −11 dB was measured across the UHF band as shown in Fig. 3.27. The Agilent E4408B spectrum analyzer was used for the gain and noise figure measurements. The loss of the SMA connectors and the coax cable ranging from 0.7 to 1.2 dB was removed from the measurement as these connectors and cables will not be present when the LNA is integrated with the mobile TV antenna. The LNA gain is found by measuring the LNA output spectrum shown in Fig. 3.28. In this figure, the buffer gain as well as the balun transformer insertion loss were de-embedded. The measured maximum LNA gain is 36 dB while two ~3 dB gain steps were measured when varying the OTA bias. Degradation in LNA gain was noted at high frequencies of operation, especially at its maximum gain mode. This is attributed to the underestimation of the parasitic capacitance at the LNA output. Such degradation will not exist when the LNA is integrated with the RF passive mixer as it has very low input parasitic capacitances. As shown in Fig. 3.28, the observed degradation is less at the LNA minimum gain mode due to a minimum value of the LNA output resistance with respect to the output resistance value at maximum gain mode.
3.3 Experimental Results
87
40 S12 (dB) 20
0
–20
–40
150MHz
900MHz
Fig. 3.26 Measured S12 of the LNA at maximum gain mode of operation 40 S22 (dB) 20
0
–20
–40
150MHz
900MHz
Fig. 3.27 Measured S22 of the LNA test chip
Using the same test setup, the noise floor at the output of the LNA test chip was measured with the noise figure mode available in the Agilent E4408B spectrum analyzer. The buffer was set to operate in high-gain test mode. The measured results of the NF are shown in Fig. 3.29. An average of 1.65 dB NF at LNA maximum gain mode was measured across the UHF band, while an average of 2.4 and 3.7 dB NF were measured at the first and second gain steps, respectively. The NF degradation
88
3 Nanometer CMOS LNAs for Mobile TV Receivers −39.00 Output Signal (dBm)
36 dB of gain
−42.00 Second gain step
3 dB gain step
−45.00
First gain step Max. gain
−48.00 −51.00
250
550
850
Frequency (MHz)
Fig. 3.28 The output LNA spectrum when the input signal is equal to −78 dBm. The frequency is varied from 250 to 850 MHz. Three different gain modes have been considered, the maximum gain, the first gain step, and the second gain step. The LNA gain was varied by changing the DC bias Vref of the OTA bias circuit
6.00
4.00 NF (dB)
NF at max. gain NF at 1st gain step NF at 2nd gain step
2.00
0.00
300
575
850
Frequency (MHz)
Fig. 3.29 Measured LNA NF at maximum gain as well as at the first and second gain steps. The frequency is varied from 300 to 850 MHz
at the first and second gain steps is not surprising since the LNA noise performance is optimized only at maximum gain mode of operation. Therefore, decreasing the LNA gain will automatically deviate the LNA operation from the noise-cancelling condition (see Chapter 2, Eq. 2.31). However, the LNA NF at minimum gain still meets the requirement of the DVB-H standard (see Table 3.3). To verify the linearity performance, a two-tone IIP3 measurement was performed on the LNA and the results are shown in Fig. 3.30. The buffer was set to the
3.3 Experimental Results
89
Output Voltage (dBm)
−19.3dBm
IM3 =44.8 dB
IM3
Center 515.00 MHz
IM3
Span 50 MHz
Fig. 3.30 Linearity measurement of the LNA at the minimum gain setting
Table 3.4 DVB-H LNA measurements summary Technology 65 nm CMOS Power supply 2.5 V Frequency range UHF: 470 –862 MHz Current consumption < 8 mA Maximum gain mode Gain NF S11 Intermediate gain mode Gain NF S11 Minimum gain mode Gain NF S11 IIP3
36 dB 1.65 dB < −18 dB 33 dB 2.4 dB < −12 dB 30 dB 3.7 dB < −8 dB −14 dBm
low-gain test mode to facilitate the LNA IIP3 measurement. The two tones were applied with equal power levels at 510 and 520 MHz. The measurement indicates an IIP3 of −14 dBm. A summary of the LNA’s measurements is shown in Table 3.4. Reliability tests were also conducted on the LNA test chip. The LNA test board was examined while the temperature (T) varied from −40°C to +100°C. A temperature humidity cycling chamber from Tenney (shown in Fig. 3.31) was used to verify the robustness test in a laboratory environment. The LNA gain was measured at three different temperature values (T = 27°C, T = −40°C, and T = +100°C) and
90
3 Nanometer CMOS LNAs for Mobile TV Receivers
Fig. 3.31 LNA test board exposed to temperature variations ranging from −40˚C to +100˚C using Tenney cycling chamber
the results are shown in Fig. 3.32. As can be seen, the LNA gain shows very small variations (<2 dB) around the nominal value (the gain at T = 27°C) when temperature changes from −40°C to +100°C. The same test was repeated for the LNA NF measurement, with results shown in Fig. 3.33. As can be seen, the LNA offers a very stable NF with temperature variations. A degradation of less than 0.5 dB in the NF value was measured at T = +100°C. In conclusion, these measurements show the effectiveness of the bias generator circuits used in this work.
3.4 Chapter Summary
91
LNA Gain (dB)
41.00
37.00 T= −40 degree C T=27 degree C
33.00
T=100 degree C
29.00
25.00
300
575
850
Frequency (MHz)
Fig. 3.32 Measured LNA gain at T = 27˚C, T = −40˚C, and T = +100˚C
3.00
2.00 NF (dB)
T= 100 degree C T= 27 degree C T=−40 degree C
1.00
0.00
300
575
850
Frequency (MHz)
Fig. 3.33 Measured LNA noise figure (NF) at T = 27˚C, T = −40˚C, and T= +100˚C
3.4 Chapter Summary This chapter presents a design example of a wideband 65 nm CMOS low noise amplifier (LNA) utilizing the noise and distortion cancelling techniques descibed in Chapter 2 and targeting the mobile TV application discussed in Chapter 1. To demonstrate how the DVB-H system requirements define the parameters of the LNA circuit and how the performance of the LNA impacts that of the overall DVB-H
92
3 Nanometer CMOS LNAs for Mobile TV Receivers
system, a receiver that can meet the DVB-H standard specifications was introduced. Blocks other than the LNA in the DVB-H receiver were chosen from state-of-the-art circuits recently published in the literature. A system analysis was conducted to extract the performance requirements of the DVB-H LNA. The LNA’s dynamic range was optimized using some of the techniques discussed in Section 2.1 in order to boost the MOSFET dynamic range. Additionally, noise and distortion cancelling circuit techniques were utilized to further improve the LNA’s dynamic range. Moreover, a new RF linearization circuit, discussed in the next chapter, will help the LNA achieve the stringent linearity requirement of the DVB-H system. Therefore, we can say that Chapters 1–3 offer a complete methodology that optimizes the DVB-H front-end dynamic range in three hierarchical levels: the intrinsic MOSFET level, the circuit level, and the architectural level. System analysis shows the necessity of having a gain control mechanism in the LNA to help the receiver meet the DVB-H dynamic range requirements. The gain control function was implemented in the mobile TV LNA design example by changing the LNA transistor bias via an operational transconductance amplifier (OTA) in a feedback loop. The stability of the feedback system has been examined and discussed. The practical realization of the designed LNA, namely, the layout used to map the LNA circuit design to the silicon and the bias generator circuits used to bias the LNA received intense attention in this chapter. The challenges associated with using the 65 nm CMOS process were explored. Special layout techniques have been implemented to overcome the well proximity effect and the oxide field stress phenomenon that might change the characteristic of the designed LNA after implementing the design into the silicon. Moreover, the layout area was optimized by taking advantage of the high capacitance density associated with using MOS-based capacitors as well as using a P+ poly resistor type without silicide material that has a very high sheet resistance. The importance of optimizing the digital MOSFET transistor layout for RF operation was also explored. Another practical issue this chapter discussed is the performance of bias generator circuits needed to bias the LNA circuit. The target LNA performance can be damaged by the noise that can couple to the LNA through the power supply or by the noise that can couple through the substrate. These issues were resolved on both the circuit design level as well as the layout level. On the circuit level, the bias generator noise was dealt with by providing filtering at the output of these circuits and by using a bandgap reference voltage generator that is independent of the supply variation. On the layout level, most of the LNA components were shielded. A deep N-well was used to shield the LNA transistors. The coupling capacitors were implemented in a self-contained N-well and biased in a way so that noise can either soak into the supply lines or into the ground lines, thus leaving the transistor characteristics unchanged. In addition to the advantage of using the bandgap bias reference circuit to reject the supply noise, this circuit also makes the fabricated LNA chip robust against any temperature variations. This chapter also examines the test setup required to characterize the LNA test board. Multi-mode buffers were designed and fabricated using 65 nm CMOS
3.4 Chapter Summary
93
technology in order to interface the designed LNA test chip to the 50 W environment of the measurement equipment. A multi-master serial computer bus based on I2C protocol with Aardvark host adapter was used to program the LNA gain as well as the buffers test mode. The LNA test board was examined in the context of the mobile TV application requirements. The results show that the designed LNA can meet the specification requirements of the DVB-H standard even at environmental temperature variations of −40°C to + 100°C.
Chapter 4
RF Attenuator Linearization Circuits
This chapter discusses the challenges associated with designing low-power receivers with large dynamic range suitable for use in mobile TV applications. It also proposes techniques to achieve a highly-linear front-end circuit with low noise figure at sensitivity. Such a front-end would be suitable for cellular phone environments that require small, low-cost, low-power components. A silicon prototype of the described solution supported with measurement results is given to demonstrate the concept.
4.1 The Necessity of RF Automatic Gain Control In all RF receivers, the design strategy of the RF front-end before the down-conversion mixer differs completely from that of the rear-end baseband. The RF signal before the down-conversion mixer is not filtered and affected by the interferer blockers while the baseband signal is filtered and amplified to a 0 dBm level using highlylinear analog RC anti-aliasing filters and opamp-based feedback amplifiers. Therefore, it is important that the linearity of the RF front-end meets the standard linearity specifications such as blocker performance and IIP3. Achieving a lowcost, highly-linear, low-noise RF front-end is very difficult in a power-conscious environment such as cellular phones.
4.1.1 RF Gain Control in Mobile TV Receivers In general, the more linear a receiver is, the less filtering it requires for the blockers; subsequently less required filtering leads to less required area and less power consumption. In the DVB-H mobile TV receiver presented in Chapter 3, the signal at the baseband input (output of the RF front-end) should not exceed −6 dBm maximum to prevent the receiver from clipping. Signal clipping generates distortion and therefore destroys the receiver output signal-to-noise (SNR) ratio. Given that the signal at the A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_4, © Springer Science + Business Media B.V. 2010
95
96
4 RF Attenuator Linearization Circuits
30-36 dB Gain
6 dB Gain
RFin
Base Band & ADC
LNA
could be 0 dBm
~ −6 dBm max
Avoiding clipping requires 40-50 dB gain range in RF
Fig. 4.1 The necessity of RF automatic gain control function
RF input could be as high as 0 dBm and that the gain ahead of the baseband chain is about 42 dB, the RF front-end circuit requires a 40–50 dB gain range to avoid losing information at large RF input signal levels. This concept is illustrated in Fig. 4.1. The gain range allows the receiver to achieve the required linearity (by decreasing the gain) and still preserve its sensitivity (at maximum gain). Another factor to consider is that the dynamic range (DR) should be kept constant even when the system’s linearity is improved by changing the front-end gain. A decrease in dynamic range might cause severe degradation in the output SNR of a receiver in cases where there is a very small desired signal with a large adjacent blocker.
4.1.2 Gain Control Circuit Techniques One of the solutions that has been used in the literature to enhance a receiver’s linearity is increasing the power consumption. The more current a receiver consumes, the more linear it will be [62]. Since increasing power consumption runs contrary to the wireless environment requirements of long lasting battery life, developing techniques that break the trade-off between power consumption and performance (linearity and noise) of silicon receivers is key for mobile TV applications. One low-power solution is to control the RF front-end gain to improve the receiver linearity. Decreasing the RF gain at large input signal levels helps the receiver pass larger signals without any degradation in the output SNR. The gain control technique was proposed before [110-113] to control the gain in the baseband chain of receivers to decrease the required dynamic range of the ADC (analog-to-digital converter). However, implementing the same circuits to control the RF front-end gain causes trouble in the output SNR because these circuits have a limited bandwidth. Another possible solution is to use the LNA to continuously control the RF gain. Continuous gain control for an RF front-end comes at the expense of inferior linearity
4.1 The Necessity of RF Automatic Gain Control
97
at high signal levels [114]. The LNA gain control technique was reported in [115] and [29] as shown in Fig. 4.2a and b respectively. Circuit (a) consists of three stages of step attenuators and four cascode common-emitter amplifiers with a combined output. The base current of the four cascode transistors is controlled by a bias generator circuit. Consequently, the gain of the amplifier changes continuously with the change in analog control voltage. The larger the current intensity diverted from the load, the more attenuated the LNA gain.
a Control devices
RL output
Input
b Main signal path
RL
Steering pair's unit
vout
N SW
N/4
N/2 SW
SW
SW
SW
Mmain
Fig. 4.2 Schematic diagram of LNAs with a gain control based on [29, 115]
SW
98
4 RF Attenuator Linearization Circuits
However, the linearity (IIP3) degrades considerably as the gain reduces, due to current-density variation in the trasconductance gm [115]. Therefore, this technique is not suitable for gain tuning under stringent linearity requirements. Another technique is the current steering technique that has been applied to provide multiple gain steps for the LNA as shown in circuit (b). After voltage-tocurrent conversion, the current generated from the main transistor Mmain is split into two paths, one being the main signal path and the other one being a dumping path. In a high-gain mode, all currents flow through the main signal path. The cascode transistors are digitally controlled by gate voltages. Constant linearity can be obtained by using constant current density in the transistors. Although this technique doesn’t suffer from problems with linearity, its complexity is still an issue that remains unsolved. Another possible solution recently published in [22] is the adaptive bias current control technique. The tuner consists of RF and baseband sections, in addition to a new block to monitor MER (Modulation Error Rate) and dynamically adjusts the bias current of RFIC sub-circuits. In the absence of strong blockers, the linearity requirement for RF circuits relaxes and thus the bias currents can be reduced. The most challenging problem with adaptive bias control is to find the optimum bias current in real time without disturbing the reception. Also, for optimum performance, this algorithm must be adjusted based on modulation scheme, code rate, and channel conditions (i.e., Doppler), which results in a significant complication. In conclusion, RF circuits used for improving a receiver’s dynamic range with low noise figure at sensitivity (lowest detectable signal level) and consuming low power are essential to have successful mobile TV products. In this chapter, an RF passive attenuator linearization circuit is presented to address this issue and to provide a robust solution that facilitates developing TV-enabled cellular phones.
4.2 RF Gain Control System Analysis There are several ways to achieve gain control in RF front-ends. The LNA itself can be used to vary the RF gain as has been discussed in the previous section. Another possible solution is to use an attenuator with variable attenuation values to vary the RF gain. This section will evaluate both solutions while taking system trade-offs into consideration. Figure 4.3a shows a situation in which the LNA is used to control receiver gain, while Fig. 4.3b shows gain controlled by an RF attenuator circuit. Basically, both techniques are capable of preventing the receiver from clipping at large input signal levels and losing the required SNR. However, to make a fair comparison between both solutions the linearity as well as the noise should be taken into consideration. Therefore, the comparison should be based on the achieved dynamic range.
4.2 RF Gain Control System Analysis
99
Fig. 4.3 Receiver gain control through (a) a variablegain LNA or through (b) RF attenuator circuit
a RFin LNA
b
RF Attenuator
RFin
LNA
40
0 Power, dBm
−20 −40
Receiver DR
Clipping level
−10
0
10
20
30
40
LNA gain control gives better sensitivity at the higher gain settings Sensitivity
Passive gain control
−80
−120
Same dynamic range in both cases
z
LNA gain control
Gain, dB Fig. 4.4 Simulated receiver dynamic range (DR) in two different scenarios, the LNA gain control and RF attenuator gain control. It has been assumed that the receiver dynamic range was limited by the signal clipping (noise). In this simulation, the output of the RF front-end was regulated at −6 dBm, the baseband input-referred noise was assumed to be 10 nv/root Hz
4.2.1 Case One: DR Is Limited by the Clipping Level From the DR point of view, having the LNA control the RF gain results in a better DR value especially for the first few gain steps (the higher gain settings). This is illustrated using simulation results that are shown in Fig. 4.4. The graph shows the simulated receiver DR in two different scenarios: LNA gain control and passive gain control. As shown in this plot, both techniques achieve the same DR starting from a certain gain setting value (“z” in Fig. 4.4). Theoretical analysis also supports these facts. The NF of a receiver with LNA controlling the gain can be given by [58]
100
4 RF Attenuator Linearization Circuits
NFtot = NFRF +
NFbb - 1 2 ARF
(4.1)
where NFtot is the total noise figure of a receiver (NF is in ratio not in dB), NFRF is the noise figure of a front-end receiver, NFbb is the baseband chain noise figure, and ARF is the front-end voltage gain before the baseband circuits. The NF of a receiver with an attenuator controlling the RF gain can be given by
NFtot = ( NFRF ´ L ) +
NFbb - 1 2 L-1 ´ ARF
(4.2)
where L is the power loss of the attenuator. Equation 4.1 indicates that in the case of the LNA gain control method, the LNA helps to keep the receiver noise roughly constant for the first few gain steps. However, this is not the case in the passive gain control method. The noise is actually worse at these gain setting values if the RF gain was controlled by an attenuator. As indicated in (4.2), NF gets 1 dB worse for every dB decrease in the receiver gain (caused by L), assuming that the receiver noise is dominated by the baseband circuits. However, this holds true only until a certain achieved gain level (“z” in Fig. 4.4). At this point, the LNA control method becomes exactly equivalent to the passive control method in terms of the achieved DR. This is because at the gain level “z”, the receiver noise is no longer limited by the LNA noise but is limited by the baseband circuit noise instead. In conclusion, the LNA gain control method is the preferred solution for enhancing the receiver linearity because it provides larger dynamic range relative to the passive gain control method at higher gain settings. This remain true so long as the receiver DR is limited by the signal clipping (noise) at these gain setting values as will be explained in the next section.
4.2.2 Case Two: DR Is Limited by the IIP3 Level Practically, receiver dynamic range can also be limited by third-order nonlinearity, the “third-order intercept point,” IIP3. Figure 4.5 shows the typical effect of changing gain settings on the total IIP3. The expression for receiver IIP3 can be given by
A2 LNA 1 1 = + 2 IIP3tot IIP32LNA IIP32RF
(4.3)
where IIP3tot is the total cascaded IIP3 of the receiver, and IIP3LNA and IIP3RF represent the IP3 points of the LNA and the rest of the RF front-end circuit, respectively. Expression (4.3) tells that the IIP3 of the second block in the receiver chain is
4.2 RF Gain Control System Analysis
101
Fig. 4.5 Simulation result of the theoretical effect of changing the gain on the receiver nonlinearity IIP3
effectively scaled down by the total gain preceding it. Therefore, for every 1 dB attenuation, the receiver IIP3 improves by 1 dB. However, these results do not take the LNA nonlinearity into account. There will be a point where the receiver IIP3 becomes limited by the first stage nonlinearity (i.e., LNA). In other words, the receiver linearity might be limited by the LNA nonlinearity in the LNA gain control method or the attenuator nonlinearity in the passive gain control method. The simulation results in Fig. 4.6 show the impact of passive gain control on the receiver DR if it is limited by the third-order nonlinearity. The IIP3 and the clipping level go to a dB for a dB change in the attenuator gain until a specific gain (“x” in Fig. 4.6) where the IIP3 is limited by the attenuator IIP3 itself and consequently the receiver DR is limited by its IIP3. As shown in Fig. 4.6, the receiver DR starts to decrease at this gain setting. Considering the third-order nonlinearity, the same analysis is conducted for a receiver for the case where an LNA controls the front-end gain. Figure 4.7 shows the impact of LNA gain control on the receiver DR. As in the passive gain control method, the DR starts to be limited by the LNA IIP3 at certain gain setting (“y” in Fig. 4.7). Since the LNA is generally less linear than the attenuator, the DR starts to decrease much earlier when the gain is controlled by an LNA. In conclusion, if receiver dynamic range is limited by third-order nonlinearity, then the passive control method is the preferred solution for enhancing receiver linearity because it provides a larger dynamic range than the active gain control method, especially at lower gain settings. In this chapter, the mobile TV RF front-end incorporates both techniques. At the higher gain setting values, the LNA controls the receiver linearity and at lower gain settings the attenuator controls the receiver linearity. Doing so allows the receiver
102
4 RF Attenuator Linearization Circuits 60
Clipping level only Limited by the resistor IIP3
30 Clipping and IIP3 level
−20 −30
DR-10dB
−10
0
10
20
30
Fixed dynamic range
−60
40
DR
Power, dBm
0
−90
IIP3 Sensitivity
−120
x
Gain, dB Fig. 4.6 Simulation results show the impact of the passive gain control technique on the receiver dynamic range. In this plot it has been assumed that the receiver dynamic range is limiting by the receiver IIP3 value at gain setting x. The baseband input-referred noise was assumed to be 10 nv/root Hz in this plot 60 Theoretical IIP3
30 LNA IIP3 could limit the receiver IIP3
Power, dBm
(b)
0 −20
−10
0
10
20
30
40
y
−30 Noise roughly constant for 1st few gain steps
−60
−90
Dynamic range decrease nearly 25dB
IIP3 Sensitivity
−120
Gain, dB Fig. 4.7 Simulation results show the impact of the LNA gain control method on the receiver dynamic range. The arrow (b) indicates the increase in the receiver dynamic range when the passive gain control technique replaces the LNA one at the lower gain settings
4.3 Highly-Linear RF Front-End Architectures
103
to achieve the required linearity with the best dynamic range along with excellent noise performance at sensitivity. One of the most important design parameters is the gain value at which the RF front-end should switch from the LNA gain control domain to the attenuator domain. This design parameter actually depends on the standard blocking profile (i.e., S-pattern in DVB-H standard), the achieved LNA IIP3, the baseband chain noise, and the required SNR at the receiver output.
4.3 Highly-Linear RF Front-End Architectures There are several ways to implement the solution that has been described and discussed in the previous section. Figure 4.8 shows different architectures that can be used to improve the receiver DR with low noise figure at sensitivity. They rely on a variable-gain LNA to control the RF gain at higher gain settings and a programmable passive attenuator to control the RF gain at lower gain settings. Switching between two RF paths improves the receiver linearity and dynamic range at high signal strength.
4.3.1 Linear RF Architectures Figure 4.8a proposes having two different paths for the RF input signal. Path 1 consists of a variable-gain LNA and path 2 consists of an attenuator. There are several things that should be highlighted here; first, this architecture needs to have two different sets of mixers. One mixer follows the variable-gain LNA and the other follows the attenuator. These two mixers help in improving the isolation between the input and output of the attenuator. Secondly, switching abruptly between path 1 and path 2 might result in severe degradation in the receiver DR as illustrated in Fig. 4.9. Switching the signal from the active path to the passive path may allow the baseband noise to degrade the receiver DR and therefore the output SNR. One way to get around this problem is to increase the gain steps of the variable-gain LNA to a level that makes switching to the passive mode safe. Doing so, however, increases the LNA complexity. The more gain steps needed by the LNA, the more complex the design becomes. Additionally, from the system perspective, this solution might be undesirable because of the LNA third-order nonlinearity limitation. Although this architecture is not the most attractive solution for improving receiver linearity, it has an advantage of saving the power consumption of the LNA when switching to the attenuator path. In some situations, the standard allows some degradation in the output SNR to save the receiver power consumption (i.e., ISDB-T Japanese mobile TV standard). Architecture (b) in Fig. 4.8b has been proposed to overcome some of the shortcomings of architecture (a). It also has two different paths: path 1 has a variable-gain
104
a
4 RF Attenuator Linearization Circuits
b
Enable path 2
Enable path 1
R SW
SW Path 2 Path 1
RFin
Path 1 RFin
LNA
c
Enable path 1
SW
R
Path 2 LNA
Enable path 1
SW
Path 1 Path 3 RFin
Path 2 R LNA
Fig. 4.8 Highly-linear RF front-end architectures
LNA alone, and path 2 has the attenuator in series with a variable-gain LNA. This configuration alleviates the problem of the abrupt change that might occur in a receiver DR when switching between both paths. The receiver DR is fixed while changing the gain of the front-end whether using the variable-gain LNA or the attenuator block. As previously mentioned, it is desirable to have an LNA controlling the RF gain until the baseband noise dominates the receiver NF. At this point, the RF front-end gain can be controlled by the passive attenuator circuit to maximize dynamic range. Another possible architecture that basically combines architectures (a) and (b) is shown in Fig. 4.8c. It provides three different paths for the RF input signal. One path, named path 3, has the attenuator only to save the LNA power consumption. Another path, named path 2, has the attenuator in series with the LNA, which provides superior receiver linearity and DR performance. The first path (path 1) has the variable-gain LNA alone to achieve the low noise figure requirement at sensitivity.
4.3.2 Gain Step Size The most important question that still needs to be answered is how big should the gain steps be? Or how many gain steps are needed for the full gain range? It is tempting to try to simplify the gain control by having one or at most two large gain steps. The problem is that the output SNR drops as much as the gain if the gain step
4.3 Highly-Linear RF Front-End Architectures
105
40 Receiver noise is dominated by the baseband noise for the Ratten mode
Clipping level
Power, dBm
0 −20 −40
−10
0
10
20
30
40
Receiver noise is dominated by the baseband noise for both Decrease in the dynamic range
Ratten gain control
−80
q
Sensitivity level LNA gain control
−120
β
Switching
Gain, dB
Fig. 4.9 Impact of switching to the attenuator gain control mode on the receiver dynamic range while the LNA is OFF. Note that the distance q depends on both the attenuator noise and the baseband noise. In this plot, the baseband input-referred noise is assumed to be 10 nv/root Hz. At certain gain setting (b) both techniques achieve a fixed dynamic range
LNA Output Before Gain Step
Clipping Level
LNA Output After Gain Step
gain step Threshold SNRout
Signal is lost
Noise Floor
Desired signal
Blocker
Desired signal
Blocker
Fig. 4.10 Effect of a large gain step on the quality of the receiving desired signal. Large gain signal may cause the signal to be lost
is caused by a large blocker as shown in Fig. 4.10, and the resulting signal may consequently be lost. This can be especially damaging to the DVB-H S (selectivity) and L (linearity) test patterns. Thus, a large step size is undesirable. However, too small a step size is difficult to implement and may cause gain changes to occur very often. Therefore, a 3 or 6 dB gain step has been selected as a design target for the attenuator, as is the case in the designed DVB-H LNA described in Chapter 3. The RF attenuator circuit attenuation is digitally programmed to provide the 48 dB attenuation range in 3 to 6 dB steps.
106
4 RF Attenuator Linearization Circuits
4.4 Design of the Binary-Weighted RF Attenuator The success of the discussed solution depends mostly on the RF passive attenuator implementation. In this section, a detailed explanation of the attenuator design is given. The logic circuits that are needed to program the RF gain are developed. Some practical considerations including the attenuator adaptation to be suitable for integration with the variable-gain LNA, and resolving the design challenges associated with RF attenuator die packaging, are discussed. Experimental measurements for a prototype in 65 nm CMOS are given at the end. The ultimate goal of this section is to facilitate development of the highly-linear front-end receiver that can meet the mobile TV standard specifications.
4.4.1 Topology Evolution Conventionally, attenuator design is based on an R-2R ladder architecture that is shown in Fig. 4.11 [116]. This configuration typically requires switches both in series as well as in shunt with the signal path. The matching requirement at the input of the attenuator (to the 50 W antenna) often constrains the value of these resistors to be quite small. The repercussions are large transistor switches, which tend to dramatically limit the bandwidth of the circuit. Accordingly, there remains a need for a novel way to design an optimum attenuator block that meets the bandwidth of the UHF band. The RF attenuator should provide from 40 to 50 dB gain range (attenuations) in steps of 3 to 6 dB. Also, it should have a 50 W input impedance to allow maximum power to transfer from the antenna. Additionally, the 50 W input matching is also essential in the case where a SAW filter precedes the attenuator to notch the GSM cell phone signal. The simplest way to design an attenuator is to use a resistive divider. Since this attenuator has only two resistors, one of them is variable to provide variable attenuation values and the other is fixed as shown in Fig. 4.12. The attenuation is given by:
Attenuation =
R2 . R1 + R2
(4.4)
In Fig. 4.12, divider (a) has several advantages over that of (b). To provide the input matching using the divider (b), a 50 W shunt resistor should be added to provide the matching at all gain settings. This will result in an increase in resistance values that are required for R1 and R2 to maintain the matching (~1 KW). These resistors might severely limit the attenuator bandwidth and might need power hungry buffers to extend the attenuator bandwidth. Also, the large resistance values make the attenuator circuit very noisy, which might limit the receiver DR. As discussed in Section 4.2, the attenuator DR at certain gain settings is limited by its third-order
4.4 Design of the Binary-Weighted RF Attenuator
107
sw
sw
sw
sw
R
R
R
ZR
Zin = 50 Ω
ZR
sw
R
ZR
sw
ZR
sw
sw
makes R ~ 50 Ω
Fig. 4.11 The schematic diagram of conventional R-2R attenuator
a
b
R1
R2
R1
R2
Fig. 4.12 Two options for the passive attenuator realization
nonlinearity. Comparing the input IIP3 for both dividers, the simulations show that divider (b) has worse IIP3 than that of divider (a). Hence, divider (a) allows the receiver to achieve larger dynamic range. Taking all these concerns into consideration, the RF attenuator designed in this chapter is based on the configuration of divider (a) shown in Fig. 4.12.
108
4 RF Attenuator Linearization Circuits
4.4.2 Binary-Weighted RF Attenuator Design A binary-weighted algorithm is used to achieve the required gain programmability of the attenuator with 6 dB step (the divide by 2 is 6 dB). The values selected for R1 and R2 values (shown in Fig. 4.12a) depend on the attenuator specifications. In some cases, the attenuator is followed by a variable-gain LNA, which means it should provide a 50 W output matching. This enables a 50 W input termination to the LNA in all of its gain settings as shown in Fig. 4.13. Changing the termination impedance of an LNA at any time may result in degradation in receiver NF [106]. Therefore, to have a fixed DR when switching from the LNA domain to the passive domain, the attenuator output impedance should remain constant at all of its gain settings and equal the LNA source impedance. The value of resistor R2 is chosen to provide the required output matching. The resistance values of R1 and R2 for the first four gain settings are shown in Fig. 4.14. The output resistance is calculated to verify the output matching requirement. As shown in Fig. 4.14a, R1 is equal to 25 W and R2 is equal to 50 W. Taking the input impedance of the LNA into consideration, this configuration gives 6 dB attenuation. It is considered the highest gain setting of the attenuator circuit. In the second setting mode, the R1 value is equal to 50 W, which gives about 10 dB attenuation. The output is also matched to 50 W. All subsequent gain steps have been verified in the same way as shown in Fig. 4.14c and d. The complete design of the binary-weighted RF attenuator network is shown in Fig. 4.15. The resistances and the corresponding attenuation values are given in Table 4.1. There are eight control bits (vcont7–vcont0) to program the attenuator for different gain settings. These control bits represent a thermometer code [117]. The highest gain setting (6 dB) is when all control bits are set HIGH and the lowest (48 dB) is when they are set LOW. There is an enable bit to activate the passive path when it is necessary to improve the receiver linearity. All bits control the gates of NMOS switch transistors. Switch sizes (shown in Fig. 4.15) were selected to minimize the off-state capacitance while still providing a sufficiently small resistance in
R1 LNA Z2
Rs R2 RFin
Fig. 4.13 Required output impedance of the RF passive attenuator
4.4 Design of the Binary-Weighted RF Attenuator
a
binary-weighted
R1 50 Ω LNA
50 Ω
RFin
50 Ω
c
d
41 Ω 200 Ω R1
R1 R2
50 Ω LNA
50
S22=−14 dB
37 Ω
100 Ω
R2
RFin
S22=−12 dB
50 Ω
33 Ω
50 Ω
R1 R2
b
30 Ω
25 Ω
50 Ω
109
50 Ω
50 Ω LNA
50 Ω
R2
50 Ω
50 Ω LNA
RFin
RFin
S22=−17 dB
S22=−21 dB
Fig. 4.14 Design parameters of the passive attenuator and the equivalent output impedance for different attenuation modes. As shown (a) represents 6 dB attenuation mode, (b) represents 10 dB attenuation mode, (c) represents 14 dB attenuation mode, and (d) represents 19 dB attenuation mode
the on-state compared to the resistance being switched. Although this attenuator achieves the required gain range and also achieves the output matching requirement, it still needs to provide the matching to the antenna or to the coax cable that is used for signal transmission. For example, Fig. 4.16a indicates that the a is not well matched at the input for the 14 dB attenuation setting (S11 = −7 dB). A shunt 100 W resistor was added as shown in Fig. 4.16b to adjust the input matching at this mode. Hence, a programmable matching network (shown in Fig. 4.17) was added
110
4 RF Attenuator Linearization Circuits vcont7
R 128x
vcont6
2R 64x
All switches are ON (25 Ω )
vcont5
4R 32x
vcont4
8R 16x
All switches are ON except vcont7 (50 Ω )
vcont3
16R 8x
vcont2
32R 4x
vcont1
64R 2x
vcont0
128R 1x
enable
RFin
RFout
128R 1x
R=50 Ω
R
Z2 50 Ω
128x
enable
LNA Zin
Fig. 4.15 Initial design of the binary-weighted RF passive attenuator network (without an input matching)
at the input of the attenuator to provide the required input matching at its different gain settings (shown in Fig. 4.18). This network can be programmed by two bits called match0 and match1. Adding the matching network modifies attenuation values of the attenuator block. Table 4.2 shows these modified values from Table 4.1.
Table 4.1 The different gain settings of the binary weighted passive attenuator Gain settings Enable vcont7 vcont6 vcont5 vcont4 vcont3 9 0 x x x x x 8 1 1 1 1 1 1 7 1 0 1 1 1 1 6 1 0 0 1 1 1 5 1 0 0 0 1 1 4 1 0 0 0 0 1 3 1 0 0 0 0 0 2 1 0 0 0 0 0 1 1 0 0 0 0 0 0 1 0 0 0 0 0 vcont2 x 1 1 1 1 1 1 0 0 0
vcont1 x 1 1 1 1 1 1 1 0 0
vcont0 x 1 1 1 1 1 1 1 1 0
R1 (W) x 25 50 100 200 400 800 1625 3200 6400
Open −6 −10 −14 −19 −25 −30 −36 −42 −48
Gain (dB)
4.4 Design of the Binary-Weighted RF Attenuator 111
112
4 RF Attenuator Linearization Circuits Still matched at the output
a
50 Ω
b
37 Ω
100 Ω
Z1
Z2 50 Ω
50 Ω
RFin
50 Ω
35 Ω
100 Ω
Z1
Z2 100 Ω
50 Ω
50 Ω
RFin
S11= -25.6 dB
S11= -7.3 dB
Fig. 4.16 Realization of the 50 W input impedance matching (Z1) for the 14 dB attenuation mode. The circuit shown in (a) is without achieving the input matching requirements. Circuit (b) illustrates the 50 W input matching realizations. The output impedance Z2 is also shown in this figure for each case vcont7
R
128x
~ ~
~ ~
vcont0
128R 1x
enable RFout
128R
Rs
1x
R
RFin Zin
Matching Network
50 Ω 128x
enable
LNA
Fig. 4.17 RF attenuator with the input matching network
4.4.3 Gain Control Logic Circuits The front-end receiver gain varies according to the input signal strength. Control logic circuits are needed to facilitate switching between the LNA domain and the attenuator
4.4 Design of the Binary-Weighted RF Attenuator
113
Ratten Matching Network
RFin
Zin
100 Ω
match1
128x
100 Ω
128x
match0
Fig. 4.18 The schematic diagram of the programmable matching network
domain. Additionally, these circuits would be able to step through the gain range of the attenuator and the DVB-H LNA described in Chapter 3. As illustrated in Table 4.2, the control bits are based on a thermometer algorithm. Although this code gives an adequate number of gain steps, the number of these steps can be further increased by implementing the code in Table 4.2 in a nonlinear fashion [118]. The nonlinear thermometer coding (shown in Table 4.3) provides more gain steps than would a linear one. Having more gain steps improves the selectivity (S-pattern) performance of the DVB-H receiver. The look-up table that is based on the nonlinear thermometer coding is shown in Table 4.3. This table gives the different values of the fourteen gain steps and the corresponding programmed values of R1. The Boolean functions that represent the control bits (vcontx) to the attenuator are derived from this table to be as follows.
enable = vcont 8 = A4 + A3 A2 A1A0
vcont 7 = A4 + A3 A2 A1
vcont 6 = A4 + A3 A2
vcont 5 = A4 + A3( A2 + A1)( A3 A2 A1A0
vcont 4 = A4 + A3( A3 A2A1A0
vcont 3 = A4 + ( A3 + A2 A1)( A3 A2 A1A0
vcont 2 = A4 + ( A3 + A2)( A3A2 A1A0
vcont1 = A4 + ( A3 + A2 + A1)( A3A2 A1A0
vcont 0 = A4 + ( A3 + A2 + A1 + A0)( A3 A2A1A0
Table 4.2 The gain and S11 for the RF attenuator different gain settings including the input matching circuit Gain settings Enable vcont7 vcont6 vcont5 vcont4 vcont3 vcont2 vcont1 vcont0 9 0 x x x x x x x x 8 1 1 1 1 1 1 1 1 1 7 1 0 1 1 1 1 1 1 1 6 1 0 0 1 1 1 1 1 1 5 1 0 0 0 1 1 1 1 1 4 1 0 0 0 0 1 1 1 1 3 1 0 0 0 0 0 1 1 1 2 1 0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0 0 1 0 1 0 0 0 0 0 0 0 0 R1 (W) x 25 50 100 200 400 800 1625 3200 6400
S11 (dB) OPEN −30 −10 −16 −12 −13 −15 −17 −21 −23
Gain (dB) OPEN −6 −8 −13.5 −18 −25 −30 −36 −42 −48
114 4 RF Attenuator Linearization Circuits
1 1 0 1 1 0 0 0 0 0 0 0 0 0 0
1 1 1 1 0 1 1 0 0 0 0 0 0 0 0
1 1 1 1 1 1 0 1 1 0 0 0 0 0 0
1 1 1 1 1 1 1 1 0 1 1 0 0 0 0
1 1 1 1 1 1 1 1 1 1 0 1 1 0 0
1 1 1 1 1 1 1 1 1 1 1 1 0 1 0
25 50 66.6 100 133.3 200 266.6f 400 533.3 800 1066.6 1600 2133.3 3200 6400
−6 −9.5 −11.2 −13.9 −16 −19 −21.3 −24.6 −26.9 −30.3 −32.8 −36.2 −38.7 −42.2 −48.1
−3.5 −2 −2.5 −2. −3. −2.2 −3. −2. −3.3 −2.4 −3.4 −2.4 −3.4 −5.9
1 1 1 0 0 0 0 0 0 0 0 0 0 0 0
Gain-step (dB)
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0
14 13 12 11 10 9 8 7 6 5 4 3 2 1 0
0 1 1 1 1 1 1 1 1 1 1 1 1 1 1
Table 4.3 The proposed programming bits of the RF attenuator block that is based on a nonlinear thermometer decoding algorithm A < 4:0> code Enable vcont7 vcont6 vcont5 vcont4 vcont3 vcont2 vcont1 vcont0 Gain (dB) R1 (W)
4.4 Design of the Binary-Weighted RF Attenuator 115
116 A<3> A<2> A<1> A<0>
4 RF Attenuator Linearization Circuits
vcont<8>
A<3> A<2> A<1> A<0>
vcont<2>
A<4> A<3> A<2> A<1>
A<3> A<2> vcont<7>
A<3> A<2> A<1> A<0>
A<4>
A<3> A<2> A<4>
A<3> A<2> A<1> A<0> A<2> A<1> A<3> A<4>
A<3> A<2> A<1> A<0>
A<4>
vcont<6>
vcont<1>
A<3> A<2> A<1> A<4>
vcont<5> A<3> A<2> A<1> A<0>
vcont<0>
A<3> A<2> A<1> A<0> A<4> vcont<4>
A<3> A<4> A<2> A<1> vcont<3> A<3> A<3> A<2> A<1> A<0> A<4>
Fig. 4.19 The logic gate diagram of the look-up table decoder used to control the passive attenuator circuit
The logic gate diagram that represents the nonlinear thermometer decoder implementation is shown in Fig. 4.19. The look-up decoder controls the attenuator gain settings according to the input signal level strength (see Fig. 4.20). To complete the attenuator design, control logic circuits are designed to program the matching network. This logic further processes the vcont bits of the attenuator gain control settings to provide the correct matching for each value of the gain setting. An Enable bit is also included to facilitate the switching between the RF attenuator domain and the LNA domain.
4.5 Practical Considerations
117 A<4:0>
Look-up Table Decoder
vcont<8>
vcont<7:0>
7
0
Matching Network RFin
Zin
RFout
Passive Attenuator
100 Ω
100 Ω
enable
50 Ω match1
128x
128x
match0
LNA Zin
Control Logic
A<4:0> Fig. 4.20 Programming the attenuator gain modes
4.5 Practical Considerations Integrating the designed RF attenuator linearization circuit [119] with the mobile TV LNA requires some design modifications to avoid any undesirable interactions between them. Also, the performance of the RF attenuator circuit can be dramatically affected by the RF package environment. The goal of this section is to address these practical issues.
4.5.1 RF Attenuator & LNA Integration Connecting the attenuator to the variable-gain DVB-H LNA requires some adaptation for the gain control look-up table decoder and for the attenuator design as well. The look-up table should be upgraded to control the gain range of the LNA.
118
4 RF Attenuator Linearization Circuits A<4:0> vcont<10>
Look-up Table Decoder
vcont<9> vcont<8>
RFout LNA
vcont<7:0>
RFin
gnd3
2R
2R
R
enable
50 Ω match1
M1
M2
MOS cap
match0
M3
enable
LNA Zin
MOS cap gnd2
gnd1
Control Logic
A<4:0>
Fig. 4.21 The RF attenuator connected to the mobile TV LNA
As shown in Fig. 4.21, two control bits (vcont10 and vcont9) are added to support the LNA gain variation. Control bit vcont8 is used to switch between the RF attenuator and LNA domains. Another potential problem is the undesirable interaction that may occur when connecting the attenuator to the LNA. This connection might create different DC paths for the LNA through MOS switches M1, M2, and M3 of the RF attenuator. Disturbing the DC bias of the LNA’s devices results in severe degradation of its performance, especially since these DC paths vary according to the attenuator programming bits. Therefore, AC coupling capacitors are proposed to be added to the attenuator to avoid any disturbances in the operating points of LNA devices. In the meantime, these caps also help to isolate the LNA domain from the noise of the attenuator domain. They prevent the noise of the 100 W resistors of the matching network from leaking to the LNA path at high frequencies when the RF front-end switches to the LNA domain. As described in Chapter 3, AC coupling capacitors can be implemented either using a MIM (Metal Insulator Metal) capacitor or a MOS (Metal Oxide Semiconductor) capacitor.
4.5 Practical Considerations
119
Using MOS-cap saves a significant amount of receiver die area due to its large capacitance density when compared with that of the MIM-cap. The MOS-cap channel length has been optimized to trade-off between the capacitance density and the quality factor (Q). The initial MOS-cap unit layout used in this design is shown in Fig. 4.22 and the corresponding C-V curve is shown in Fig. 4.23. It is based on an NMOS transistor with channel length of L = 0.25 µm, and width W = 20 µm with 45 fingers. The C-V curve shows that the capacitance achieved at 1.2 V (which is around the common-mode voltage of the LNA) is about 3 pF.
Fig. 4.22 Layout of the NMOS-cap, the length L = 0.25 µm, the width W = 20 µm, and 45 fingers 21 µm
21 µm
3.50 3.00
Capacitance. pF
2.50 2.00 1.50 1.00 0.50 0.00 0.0
0.2
0.4
0.6
0.8
1.0
1.2
MOS Gate voltage, V
Fig. 4.23 Simulation results for capacitance versus gate voltage of the NMOS-cap. The NMOScap has a length L = 0.25 µm, a width W = 20 µm, and 45 fingers
120
4 RF Attenuator Linearization Circuits
a
b 3 pF caps N-well L=1 µm L=0.25 µm
L=0.25 µm L=1 µm
Fig. 4.24 Layouts of the NMOS cap (3 pF) with a length L = 0.25 µm and a length L = 1 µm. (a) Based on a regular NMOS, shows 30% savings in area while (b) based on NMOS in N-well as substrate, shows a 50% area savings. The layout is based on the 65 nm technology
Figure 4.24 shows two techniques to improve the capacitance density of a MOS-cap. Simulations (using real models provided by the foundry) show that increasing the MOS channel length will result in improving the capacitance density because it decreases the substrate contacts overhead. However, increasing the transistor length will degrade the cap quality factor. Therefore, in this design the channel length has been increased to a point where the cap quality factor does not degrade attenuator performance in the UHF band. Figure 4.24a illustrates a 3 pF MOS cap with two different implementations (channel lengths). Increasing the channel length from 0.25 to 1 µm leads to about 30% area saving. Another option that has been investigated is implementing the MOS transistor in an N-well instead of the P-substrate. The use of the N-well as a substrate for an NMOS transistor results in an increase in the capacitance density, especially at lower gate voltages. Figure 4.24b shows a 50% area saving relative to the initial cap design (L = 0.25 µm and using P-substrate). Simulation results illustrating the variation of the MOS-cap capacitance versus the gate voltage for four different MOS-caps are as shown in Fig. 4.25. Curves (a), (b), and (c) represent the capacitance of MOS-caps that occupy the same die area. Curve (c) represents a MOS cap with L = 0.25 µm, curve (a) represents a MOS cap with L = 1 µm and curve (b) represents a MOS cap with L = 1 µm that uses an N-well as substrate. For the same occupied area, the N-well based MOS cap achieves almost double the capacitance value. Saving such area will have a direct impact on the receiver cost as well as its size.
4.5.2 Package Bond Wire Coupling The performance of a radio frequency integrated circuit can be dramatically affected by the package environment. The lead frame parasitic can cause the package to
4.5 Practical Considerations
121
5.00
Capacitance. pF
4.00
3.00
2.00 (a) L=1u NMOS, Area=A1
1.00
(b) L=1u N-Well NMOS, Area=A1 (c) L=0.25u NMOS, Area=A1 (d) L= 1u N-Well NMOS, Area=0.5A1
0.00
0.0
0.2
0.4
0.6
0.8
1.0
1.2
MOS Gate Voltage, V
Fig. 4.25 Simulation results of the capacitance versus the gate voltage for different MOS-caps. Curves (a), (b), and (c) represent MOS caps that occupy the same die area but (c) is with L = 0.25 µm, (a) is with L = 1 µm and (b) is with L = 1 µm and using an N-well as substrate. Curve (d) represent MOS cap with L = 1 µm and using an N-well as substrate. It occupies half the die area of that in (a), (b), and (c)
seriously degrade RF integrated circuit performance [120]. The ground bond wires have been taken into consideration during the attenuator simulation as shown in Fig. 4.26. The ground bond wire is modeled as 500 pH inductance as suggested by the data sheets of the QFN (quad-flat) package [58]. Simulation results shown in Fig. 4.27 illustrate a severe degradation in the attenuator performance in its low-gain settings (higher attenuations) after modeling the ground downbonds. It has been noted that this degradation gets worse at higher frequencies. The degradation happens because the input RF signal couples to the output and prevents the attenuator from attenuating the input signal. The coupling basically happens through two different paths (shown in Fig. 4.26). Path 1 represents the coupling through the matching network downbond ground and path 2 represent the coupling through the transistor gate-source parasitic capacitor of the passive attenuator switch of R1. To solve this problem, a separate downbond has been proposed to each ground gnd1, gnd2, and gnd3 to improve the isolation, as shown in Fig. 4.28. Simulation results have demonstrated that this solution totally resolves the coupling problem that might limit the achieved RF attenuator gain range.
122
4 RF Attenuator Linearization Circuits
R1
RFin
RFout
Path 2
100 Ω
50 Ω
Path 1
100 Ω
coupling through the bondwire
Fig. 4.26 Electrical coupling through the ground bond wires of the RF attenuator
0
code 0 code 1 code 2 code 3 code 4 code 5 code 6 code 7 code 8 code 9 code 10 code 11 code 12 code 13 code 14
−5 −10 Attenuation, dB
−15 −20 −25 −30 −35
Bondwire coupling limits the passive gain range
−40 −45 400
480
560 640 Frequency, MHz
720
800
Fig. 4.27 Results of the gain of the attenuator for different gain settings codes
4.6 A 65 nm CMOS RF Passive Attenuator
123 R
RFin
2R
2R
RFout
gnd3
Separate grounds to improve the isolation
gnd1
R
gnd2
Fig. 4.28 Separation of the attenuator grounds to improve the isolation between the RFin and RFout
4.6 A 65 nm CMOS RF Passive Attenuator The RF passive attenuator linearization circuit has been implemented in silicon to verify its functionality as well as its performance in the UHF band. This is very important to facilitate the complete RF front-end architecture realization. All the specifications including gain steps, 50 W input matching (S11), and 50 W output matching (S22) are measured. The noise performance and linearity specs are also evaluated for different gain setting values.
4.6.1 The Fabrication The described RF attenuator block has been fabricated in 65 nm CMOS technology. Two different unit cells have been proposed to implement the variable resistor of the attenuator (shown in Fig. 4.29). The resistor width has been selected to optimize the layout area and to minimize the routing parasitics between the resistors. Unit (a) has been used to implement the 6.5, 3.2, 1.6 K and 800 W resistors while unit (b) has been used to implement the 400, 200, 100, 50 W resistors. The MOS switches have been designed taking their “on-resistance” and their parasitic capacitance into consideration. The N-well based MOS cap unit described in Section 4.5.1 has been used to implement the matching network AC coupling caps and the attenuator AC coupling cap as well. The cap values determine the
124 Fig. 4.29 The attenuator unit cells that are used for implementation
4 RF Attenuator Linearization Circuits
a RFin
RFout 6.5K
0.6µm / 0.06µm
b
RFin
RFout 400
9.6µm / 0.06µm
Fig. 4.30 Die photograph of the RF attenuator linearization circuit
attenuator bandwidth. To support the VHF band 70 and 30 pF capacitances have been chosen for the attenuator and the matching network caps respectively. The attenuator die photo is shown in Fig. 4.30. The fabricated attenuator consumes 0.05 mm2 of silicon.
4.6.2 Measurement Results The fabricated RF attenuator has been characterized in the lab. The HP 8753D network analyzer has been used for S-parameter measurements. The output matching of the attenuator to a 50 W resistance has been tested by measuring S22. Figure 4.31 shows the measured S22 of the attenuator for different gain code settings across the UHF band. S22 < −13 dB for all gain settings across the UHF band. The same test has been repeated to evaluate the attenuator input matching to a 50 W source resistance. The measured S11 is shown in Fig. 4.32. S11 < −12 dB for all gain settings across the UHF band. A noise analysis has been conducted using the noise mode of the Agilent E4408B spectrum analyzer. Figure 4.33 shows the NF measurements for the attenuator for all gain settings across the UHF band. The loss of the SMA connectors and the coax cable ranging from 0.7 to 1.2 dB was removed from the measurement as these connectors and cables won’t be present when the attenuator is integrated with the mobile TV antenna. The measured NF is slightly better (1.5–2 dB) than that of
4.6 A 65 nm CMOS RF Passive Attenuator
125
0
code0 code 1 code 2 code 3 code 4 code 5 code 6 code 7 code 8 code 9 code 10 code 11 code 12 code 13 code 14.
−5
S22, dB
−10 −15 −20 −25 −30 −35 400.0
480.0
560.0
640.0
720.0
800.0
Frequency, MHz
Fig. 4.31 Measured output matching S22 of the RF attenuator for different gain code settings. The frequency has been swept from 400 to 800 MHz 0
code 0 code 1 code 2 code 3 code 4.0 code 5 code code 7 code 8 code 9 code 10 code 11 code 12 code 13 code 14
−5
S11, dB
−10 −15 −20 −25 −30 400.00
480.00
560.00
640.00
720.00
800.00
Frequency, MHz
Fig. 4.32 Measured input matching S11 of the RF attenuator for different gain code settings. The frequency was swept from 400 to 800 MHz
the simulated one. The analysis that explains this discrepancy is given in the following section. One of the most critical measurements is the passive attenuator third-order nonlinearity. As explained earlier the receiver DR might be limited by the attenuator’s IIP3 at the lower gain settings. A two-tone test has been conducted by applying two tones that were spaced by 4 MHz. Figure 4.34 illustrates a two tone test for one of the gain settings.
126
4 RF Attenuator Linearization Circuits 50
40
NF, dB
Simulation 30
20
10
0
0
10
5
15
Gain Codes
Fig. 4.33 Measured NF of attenuator for different gain code settings Power (dBm) 0.00
−20.00
−40.00
Output IM3 @ -66 dBm −60.00
−80.00
−100.00 5.00E+07
7.51E+09
Freq. (Hz)
Fig. 4.34 Linearity measurement of the passive network. Measurements were performed at 750 MHz with 50 W equipment. The two fundamental frequencies were spaced by 4 MHz. This measurement corresponds to code = 14 (6 dB attenuation mode)
The measured IM3 (third-order distortion) is found to be −66 dBm. As shown in Fig. 4.35, the input-referred IIP3 is calculated to be +25 dBm. The same test has been conducted for all attenuator gain settings. The IIP3 degraded by 1–2 dB at lower gain settings as shown in Fig. 4.36. The measurement results are summarized in Table 4.4.
4.6 A 65 nm CMOS RF Passive Attenuator
127
Tone Levels [input referred], dBm
20
0
−20
IP3 ~ + 25.5dBm
−40
−60
−80 −10
0
10 Input Level, dBm
20
30
Fig. 4.35 Linearity measurement of the passive network for code = 14 26 25.5
IP3, dBm
25 24.5 24 23.5 23 22.5
1
3
5
7
9
11
13
15
Code
Fig. 4.36 Linearity measurement of the passive network for different gain settings
4.6.3 Comparison with Simulations All the measurements agree with simulations except for noise measurements. There is a discrepancy of about 2 dB between the simulation and the measured NF result. This can be explained by the presence of a noisy 25 W resistor that has been assumed
128
4 RF Attenuator Linearization Circuits
Table 4.4 Measurement results summary of the RF passive linearization circuit
Technology Die area Power supply Noise figure (NF) max min fc_low Gain steps S11 for all gain settings S22 for all gain settings In-band (IIP3) max min
65 nm CMOS 0.05 mm2 1.2 V 48 dB 5.8 dB 100 MHz ~3–6 dB <−12 dBm <−13 dBm +25.3 dBm +23 dBm
at the attenuator output during the simulation. Figure 4.37 shows two different representations of a resistive divider circuit from the noise perspective. Circuit (a) represents a regular attenuator (used in the simulation) while (b) represents an attenuator where half of its output resistor load is noiseless (R2 is noisy while RL is noiseless). From the gain point of view, both of them are equivalent. However, each circuit has different noise performance. This can be illustrated theoretically using the noise figure equation. The noise figure NF can be calculated by [10]: NF =
vn2,out A
2
´
1 4 kTRsDf
(4.5)
where vn2,out represents the total noise at the output, A is the voltage gain, and Rs is the source impedance. k is Boltzmann’s constant and T is the absolute temperature in Kelvins. To find the NF of circuit (a), the output noise voltage and the voltage gain have been derived. Figure 4.37a implies that:
vn2,out ( a ) = 4 kT
( Rs + R1) R2 Df . Rs + R1 + R2
(4.6)
Using (4.5) and (4.6), the NF is found to be:
R1ö æ Rs R1 ö æ NF = ç 1 + ÷ ç 1 + + ÷ . è Rs ø è R2 R2 ø
(4.7)
To compare between both configurations, the NF of circuit (b) is derived. The output noise for the circuit shown in Fig. 4.37b is found to be:
vn2,out ( b ) = 4 kT
( Rs + R1) R2 ( RL )2 ´ Df . 2 Rs + R1 + R2 é ( Rs + R1) R2 ù êë RL + Rs + R1 + R2 úû
(4.8)
4.6 A 65 nm CMOS RF Passive Attenuator
a
129
b
R1
R1
Vout
Rs RFin
Vout
Rs
R2 25Ω
RFin
R2
RL
50Ω
50Ω Noiseless
Fig. 4.37 Noise analysis of the attenuator in case of (a) ideal attenuator (b) attenuator using a noise-cancelled resistor RL (input impedance of the LNA)
Using (4.5) and (4.8), the NF is derived to be:
2 ( RL )2 é RL + R1 + Rs ù ( Rs + R1) R2 Rs ú . ´ ´ê 2 NF = 2 RL Rs + R1 + R2 é ú ( Rs + R1) R2 ù ê ë û 2 êë RL + Rs + R1 + R2 úû
(4.9)
The divider shown in Fig. 4.37b is equivalent to the attenuator followed by the LNA. The LNA 50 W input impedance acts as a noiseless load (RL) for the passive attenuator. Assuming the 6 dB attenuation mode of operation, the variable resistance of R1 should equal 25 W. Comparing the NF for both configurations (a) and (b) using (4.7) and (4.9) for this gain setting results in a 7.7 dB NF for (a) and 5.8 dB NF for (b). This would explain the 2 dB deviation between the measured NF and the simulated one. Intuitively, this could also be explained by looking to the main noise contributor in the attenuator. Since the main noise contributor is the load resistor R2, the noise performance of circuit (a), which has an R2 equal to 25 W, should be worse than that of (b), which has an R2 equal to 50 W. This analogy can be seen from the rms noise current generated by the resistor R2. It is given by:
in2 =
4 KT Df . R2
(4.10)
In conclusion, the noise-cancelled (noise-free) impedance of the LNA helps to improve the noise performance of the passive attenuator. However, the superior performance of this impedance does not necessarily lead to an improvement in the receiver dynamic range. It is expected that this difference is going to be small since the receiver noise is limited by the baseband noise, especially at the lower gain settings.
130
4 RF Attenuator Linearization Circuits
4.7 Chapter Summary The RF gain control technique has been described in the literature as one of the low-power solutions that can help mobile TV receivers achieve their stringent linearity requirements. Decreasing the RF gain at large input signal levels helps the receiver pass larger signals without any degradation in the output SNR (signalto-noise ratio). Although there are many mechanisms to vary the RF gain in receivers, the efficacy of any given mechanism depends on the amount of the dynamic range that can be achieved while decreasing the RF gain. This chapter introduces the necessity of having a passive gain control in the front-end circuit to meet the linearity requirement of the mobile TV standard DVB-H. It presents an RF attenuator linearization circuit used to vary the RF gain of mobile TV receivers while maximizing their dynamic range. The design and implementation of the proposed attenuator is given in detail. A 5 bit linear thermometer decoder is used to program the gain of the attenuator. The decoder sets the gain value according to the signal level received at the attenuator input. Also, an on-chip programmable matching network is used to provide a stable 50 W input resistance. The highly-linear RF architecture discussed in this chapter enables a low-power, highly-linear, receiver realization with low noise figure at sensitivity level.
Chapter 5
Wide Dynamic Range Mobile TV Front-End Architecture
This chapter presents a design example of a 65 nm CMOS chip that integrates the wide dynamic range DVB-H front-end architecture described in the previous chapter with an RF power level indicator circuit. The front-end consists of the wideband noise-cancelling LNA presented in Chapter 3 and the RF passive attenuator presented in Chapter 4. A power level indicator circuit will provide an automatic gain control function to the mobile TV front-end architecture. The chapter concludes by presenting some experimental results for this test chip.
5.1 Mobile TV Front-End with Automatic Gain Control The primary purpose of this chip is to obtain real-time measurements by verifying the performance of the described DVB-H RF front-end architecture when large interferer signals are present at its input. An RF level indicator circuit IP is used to facilitate the automatic gain control (AGC) function. The section begins by giving a very brief description of the requirements for the mobile TV RF AGC as well as for the RF level indicator circuit (RFLI). Then, the test chip architecture and implementation are described.
5.1.1 Self-Contained RF AGC Control Conventionally, both the receiver RF and IF AGC functions are controlled by the demodulator, which evaluates the desired signal only and receives no information about the interferer signals [19]. This can be suitable for some applications such as cable modems or cable TV applications, where the shape of the rest of the spectrum (i.e., channel interferers) can be quantified using the measured power of the desired
A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_5, © Springer Science + Business Media B.V. 2010
131
132
5 Wide Dynamic Range Mobile TV Front-End Architecture
signal. However, for mobile TV applications, there is a large frequency range of signals present at the input of the front-end circuit with the possibility of large undesired-to-desired-signal (U/D) power ratios. Therefore, the RF AGC of the mobile TV receiver must be able to evaluate both the desired signal and the interferer’s signals as well. In this chapter, the self-contained RF AGC approach has been selected to control the RF gain of the DVB-H front-end architecture. The goal of this RF AGC is to optimize the signal power levels at the LNA output and thus provide the best possible RF performance in the presence of larger interferer signals (see Chapter 3). Although this approach seems complicated, as additional circuits (i.e., level indicator circuit) are needed to facilitate the AGC function, the other approach (RF gain controlled by a demodulator) also shares the same complications. The AGC function that is provided by a demodulator also requires additional circuits to accommodate different DVB-H modulation schemes with different required output SNR values. In fact, the 20 dB variations in the SNR due to the modulation scheme in use make the demodulator approach much more complicated than the RF AGC approach.
5.1.2 DVB-H RF Front-End with AGC Algorithm The DVB-H RF front-end architecture with AGC function is shown in Fig. 5.1. Instead of using I2C protocol to manually program the RF gain of the front-end (as was the case in programming the circuits presented in Chapters 3 and 4), an RF level indicator (RFLI) circuit is used to automatically program the RF gain according to the power levels of the signals present at the receiver input. Specifically, it measures the signal voltage at the output of the LNA and compares it to a certain upper threshold voltage. If the measured voltage is greater than the upper threshold, the RFLI circuit generates a certain code needed to instruct the thermometer decoder (look-up decoder) to decrease the gain of the front-end circuit (both the active gain and the passive gain). On the other hand, if the measured voltage at the LNA output is less than the lower threshold voltage, the RFLI circuit instructs the thermometer decoder to increase the RF gain. While the upper threshold voltage helps in avoiding the distortion as a result of large interferer signal levels present at the receiver input, the lower threshold guarantees that the minimum required SNR is still achieved even if the input signal levels decrease. Therefore, the best possible NF given the signal power levels at the receiver input can be achieved. When the desired power levels are reached (any value between the two threshold voltages) the RF AGC shuts itself down. As shown in Fig. 5.1, five 2:1 multiplexers are used to provide flexibility in debugging the test chip. They allow the I2C to manually program the RF gain through the AGC bypass mode. The detailed design of the thermometer decoder and the control logics were given in Chapter 4. The RFLI code, the decoder output, and the corresponding RF gain values are given in Table 5.1.
5.1 Mobile TV Front-End with Automatic Gain Control
133
MUX
A<4:0>
RF Level Indicator
B<4:0> AGC bypass Chapter Four vcont<10>
Look-up Table Decoder
Control Logic
vcont<9> vcont<8>
vcont<7:0>
RFin
RFout LNA 2R
2R
R Chapter Three
match1
M2
M1
match0 M3
gnd1
enable
gnd2
Matching Network
RF Passive Attenuator enable
Fig. 5.1 The architecture of the DVB-H front-end with AGC control function
5.1.3 AGC RF Level Indicator Circuit The simplified block diagram of the RFLI circuit is shown in Fig. 5.2. It includes a sub-sampling circuit, an analog-to-digital converter (ADC), and a baseband digital detector circuit. The LNA output signal is sampled using a simple MOS switch and a capacitor to relax the bandwidth requirements of the ADC and the baseband detector. The power of the RF signal is now concentrated in a narrow bandwidth depending on the sampling frequency fs. The ADC then generates a digital representation of the RF signal that can be measured using a simple baseband digital detector. The baseband detector circuit includes four digital comparators and three binary counters. The four comparators compare the digitized version of the RF signal to the positive and negative upper threshold voltages (up_th_p and up_th_n) as well as to the positive and negative lower threshold voltages (low_th_p and low_th_n). Counters (1) and (2) determine the number of samples that exceed any of the threshold voltage values and then toggle when the number of counts reaches a predetermined value. This prevents the RFLI from toggling at each variation in the RF signal levels. Counter (3) either adds one or subtracts one from the RFLI code depending on whether the RF signal level is lower or higher than the threshold voltages. The changes in the RFLI code happen very slowly – controlled by CK_SLOW – to
Passive gain steps
Max. gain LNA gain steps
Gain settings
17 16 15 14 13 12 11 10 9 8 7 6 5 4 3 2 1 0
RFLI code A <4:0>
1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0
Thermometer decoder output vcont vcont vcont vcont 10 9 8 7
Table 5.1 DVB-H RF front-end gain programming codes
1 1 1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0
vcont 6 1 1 1 1 1 0 1 1 0 0 0 0 0 0 0 0 0 0
vcont 5 1 1 1 1 1 1 1 0 1 1 0 0 0 0 0 0 0 0
vcont 4 1 1 1 1 1 1 1 1 1 0 1 1 0 0 0 0 0 0
vcont 3 1 1 1 1 1 1 1 1 1 1 1 0 1 1 0 0 0 0
vcont 2 1 1 1 1 1 1 1 1 1 1 1 1 1 0 1 1 0 0
vcont 1
1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 0 1 0
vcont 0
36 33 30 24 21 18 15 12 9 6 3 0 −3 −6 −9 −12 −15 −18
Gain
134 5 Wide Dynamic Range Mobile TV Front-End Architecture
5.2 A 65 nm CMOS RF Front-End Prototype LNA output
135
Sub-sampling circuit
Baseband Detector Circuit up_th_p
(1)
ADC
CNTR
fs
(3)
O
I
c
CK
up_th_n
CNTR
fs
SUB
low_th_p
ADD
(2)
O
out
CK
CNTR O
I
low_th_n
fs
CK
CK_SLOW
Fig. 5.2 Simplified block diagram of the RF level indicator circuit
make the variation in the RF gain occur within 100 ms period of time. The output of the baseband detector then goes to the thermometer decoder to control the front-end gain. Since the positive and negative threshold voltages are programmable, the user can choose these values so as to optimize the receiver dynamic range. In this design example, the upper threshold voltage was set to −12 dBm (peak), while the lower threshold voltage was set to −17 dBm1 (peak). The window between the upper and lower thresholds was chosen to be larger than the RF gain steps in this work (~3 dB) to prevent the RFLI from toggling continuously between the two threshold voltages.
5.2 A 65 nm CMOS RF Front-End Prototype 5.2.1 The Fabrication The front-end architecture with the RFLI circuit was implemented in 65 nm CMOS technology. The exact mobile TV LNA layout described in Chapter 3 as well as the exact RF passive attenuator layout described in Chapter 4 were reused in this test chip. The digital automation software available in Cadence was used to automatically place and route the layout of the RFLI digital standard cells. One of the important issues that was considered while doing the floor-plan and layout of this test chip was the effect of the switching noise of the RFLI circuit on the RF performance. Although the LNA components are shielded (see Chapter 3), extra precautions were still needed to avoid any interactions that might occur 1 The window between the upper threshold and the lower threshold should be larger than the gain step.
136
5 Wide Dynamic Range Mobile TV Front-End Architecture
Output buffers RF attenuator DVB-H LNA
VDD
RFout +
RFout − Gnd2
Gnd1
RFin
RFLI
Fig. 5.3 Die photograph of the 65 nm CMOS DVB-H front-end chip. The die occupies 0.21 mm2 of silicon
between the RFLI switching clock and the RF signal path. Such interaction can easily increase the noise floor of the front-end circuit (i.e., the LNA) and damage the achieved noise figure. The physical spacing between the LNA circuit and the RFLI circuit was kept large enough to improve the isolation. Moreover, the RFLI clock lines were shielded to prevent any electrical coupling to the sensitive paths of the RF signal. Also, the switching noise coupling that might occur between the RFLI power supply and the LNA input through the mutual inductance between the package pins was avoided by carefully isolating these pins. The die photograph of the mobile TV DVB-H front-end architecture with an AGC function is shown in Fig. 5.3. The die consumes 0.21 mm2 of silicon.
5.2.2 Measurement Results of the AGC Test The main objective of the AGC test (closing the loop between the RF output and the RF input) is to quantify the robustness of the DVB-H RF front-end architecture to large interferer signals. The Agilent E4408B spectrum analyzer was used first to measure the RF gain at different gain codes. The AGC loop was opened by enabling the AGC bypass mode, and the thermometer decoder gain codes were manually inserted through the I2C. The purpose of this experiment was to verify the values of the active and passive gain ranges of the front-end circuit before conducting the AGC test and closing the AGC loop. The measured RF gain is shown in Fig. 5.4. As shown in this figure, the front-end circuit can provide a 54 dB of gain range. 6 dB was due to the LNA gain steps and the rest was due to the RF passive attenuator gain control. The maximum gain of the front-end circuit was measured to be ~36 dB. The gain measurements agree with the expected values given previously in Table 3.4.
5.2 A 65 nm CMOS RF Front-End Prototype
137
40
LNA gain variations 6 dB
RF Gain (dB)
max. gain ~ 36 dB 20
~ 54 dB RF gain range
48 dB
0
Passive gain variations
−20
0
2
4
6
8
10
12
14
16
18
RFLI Gain code Fig. 5.4 Measured RF gain ranges of the DVB-H front-end test chip. The AGC loop was opened and the thermometer decoder gain codes were inserted manually by the I2C
The AGC test was conducted by closing the AGC loop and letting the RFLI circuit regulate the RF gain according to the signal level preset at the input of the front-end circuit. In this case, the AGC bypass mode was disabled. The RFLI circuit was programmed to regulate the output signal level. The output signal power was measured, while the input signal to the front-end circuit varied from −102 to −20 dBm. The AGC test measurement results are shown in Fig. 5.5. As shown in this figure, the measured output power increases as the input power increases with a slope indicating 36 dB of RF gain. The output voltage kept increasing until the input voltage reached a certain value (−48dBm). Beyond this value, the RFLI circuit engages to regulate the output signal of the front-end circuit, and thus protects the RF performance during the presence of larger interferer signals. The 3 dB gain steps of the active gain (LNA) as well as of the passive gain can be observed as soon as the AGC loop turns on and the RFLI circuit controls the RF gain. The same test was repeated but with ±10% variation in the supply voltage. This is important since any variations in the LNA common-mode voltage with respect to the RFLI common-mode voltage will result in changing the regulating threshold voltage window and thus disturbing the AGC operation. In some cases, this variation can hurt the protection process of the AGC loop and damage the output SNR of a receiver. As shown in Fig. 5.5, the AGC loop runs as expected even with the ±10% variations in the supply voltage. This is due to the biasing schemes used in this work that make the LNA circuit robust against the supply voltage variations. The DVB-H front-end receiver can achieve an NF as low as 1.6 dB (the same measured NF of the LNA presented in Chapter 3) and an IIP3 as high as +24 dBm
138
5 Wide Dynamic Range Mobile TV Front-End Architecture 0
Output Level (dBm)
SNR level = constant
−20 Low volt Mid volt High volt
−40 The RFLI starts to change the RF gain and regulates the LNA output voltage
−60
−80 −110 −100
−90
−80
−70
−60
−50
−40
−30
−20
Input Level (dBm)
Fig. 5.5 Measured results of the AGC test. The output power was measured while varying the RF input power. The test was conducted with the LNA nominal supply voltage of 2.5 V (Mid voltage) and with ±10% variation from the nominal value (Low volt for the −10% variations and High volt for the +10% variations) Table 5.2 Measurement results summary of the proposed DVB-H front-end circuit
Frequency range Channel bandwidth Maximum gain Active gain range Passive gain range Minimum noise figure Maximum IIP3 DC power consumption Supply voltage Die size Technology
UHF: 470–862 MHz 8 MHz 36 dB 6 dB 48 dB 1.6 dB +24 dBm ~19 mW 2.5 V 0.21 mm2 65 nm CMOS
(the same measured IIP3 of the RF attenuator presented in Chapter 4), both across the UHF band. The performance summary of the DVB-H front-end architecture is shown in Table 5.2.
5.3 Chapter Summary This Chapter presents a front-end architecture that integrates nanometer CMOS RFICs presented in Chapters 3 and 4 with an RF level indicator circuit to facilitate the AGC control function. The performance of the 65 nm front-end prototype was observed in the lab when large interferer signals were present at its input.
Chapter 6
Summary and Conclusions
This chapter highlights the main characteristics of the nanometer CMOS RFICs described in this book. After discussing the impact of the developed wide dynamic range techniques on the overall DVB-H mobile TV tuner products, the chapter concludes by presenting some future work needed to move the work conducted in this book forward.
6.1 Summary and Conclusions 6.1.1 Digitally-Controlled Variable-Gain LNA As presented in Chapter 3, an LNA that can meet the DVB-H standard specifications has been designed and its performance has been verified experimentally in the lab by a 65 nm CMOS prototype of this design. The LNA measured performance summary was presented previously in Table 3.4. The LNA performance not only confirms that nanometer CMOS technology is capable of meeting the mobile TV standard requirements but also shows that the achieved noise figure is low enough to replace the LNA designed using the exotic GaAs and InP semiconductor technologies. The average measured 1.6 dB NF represents the lowest measured noise figures reported to date in the literature for wideband CMOS LNAs used for wireless applications covering the UHF band (see Table 2.1). The achieved NF allows the external LNA implemented by exotic technologies [7, 22] used in the DVB-H tuner available today in the market to be avoided. Increasing the integration level also helps in decreasing the cost, area, and the power consumption of current products. Moreover, avoiding using the area-consuming on-chip inductors in the described LNA will motivate designers to use nano-scale CMOS technologies (i.e., 65 nm) as the cost per area of these technologies is extremely high and thus facilitates the feasibility of the SOC (System On Chip) realization.
A.A. Youssef and J. Haslett, Nanometer CMOS RFICs for Mobile TV Applications, Analog Circuits and Signal Processing, DOI 10.1007/978-90-481-8604-4_6, © Springer Science + Business Media B.V. 2010
139
140
6 Summary and Conclusions
The designed LNA topology serves as a very useful single-ended-to-differential converter that takes the single input from the antenna and drives the differential input of the next mixer stage. This helps avoid the external broadband balun used in [24] and its associated loss, which adds directly to the noise figure of the system. Another feature of the described LNA is that it provides the whole RF gain required to eliminate the noise of the baseband circuit’s chain. Hence, a passive type mixer can be used to replace the active mixer recently used in [22, 23]. This will further reduce the power consumption of the DVB-H receiver and improve the receiver’s noise performance. Moreover, the power-consuming buffer usually the needed before the active mixer can be avoided and therefore the system NF and linearity trade-offs can be relaxed. The designed LNA also helps widen the DVB-H receiver’s dynamic range by providing three different gain modes (two gain steps) that can be controlled using two bits. These bits can be programmed either through the DVB-H demodulator or through a separate RF AGC. In addition to the compatibility of the designed LNA with the DVB-H standard requirements, its performance at maximum gain is very competitive as a standalone wideband LNA circuit. Table 6.1 shows a comparison between the LNA performances and other wideband LNA designs that were recently reported in the literature. As discussed in Chapter 2, achieving low-cost, wideband input matching along with low noise figure and low power consumption with high gain for an LNA is extremely challenging. The commonly used FOM (Figure Of Merit) in the literature considers all these trade-offs [93]. As shown in Table 6.1, the LNA discussed in Chapter 3 has the highest FOM among the other wideband CMOS LNA designs and reaches the same noise performance of other LNAs designed using more expensive technologies.
6.1.2 Digitally-Programmed RF Passive Attenuator As discussed in Chapter 1, a mobile TV front-end receiver must be able to cope with adverse reception conditions such as strong adjacent channel interferences with the possibility of a very large undesired-to-desired-signal (U/D) ratio and increasing noise levels. Therefore, the front-end circuit must be highly linear and must also achieve a very low noise figure. This usually results in an increase in the power consumption of the front-end circuit (recently power consumption as low as 100 mW was reported [25]). An external SAW filter was used in [7, 23] to attenuate the interferers and hence reduce the power consumption of the mobile TV receiver. However, in addition to the high cost associated with using SAW filters, their losses add to the total noise figure of the receiver. The design in [100] used two LNAs, one before the SAW filter and the other after the SAW filter to improve the receiver noise performance.
65 nm CMOS
0.2–1.9 (simulated)
b
æ Gain [linear ]´ BW [GHz ] ö FOM = 20 log10 ç ÷ [93]. è Power [mW ]´ (NF [Linear ]- 1)ø
Blaakmeer [121] 65 nm CMOS 0.3–3.5 Borremans [93] 90 nm CMOS 0−6 Zhan [101] 90 nm CMOS 0.5–6 Chehrazi [83] 0.13 µm CMOS 0.4–6.5 Ramzan [94] 0.13 µm CMOS 1–7 Wang [90] 0.13 µm CMOS 0.1–0.93 Lee [84] 0.18 µm CMOS 3–5 Koutani[110] 0.5 µm BiCMOS 0.1–0.85 a A minimum of 3 dB bandwidth and S11 < −10 dB.
Youssef [122] 3.5 2.5 2 3 2.7 4 2.3 1.6
1.4 0 N/A −14 0 (simulated) −4.1 −10.2 −9 −2
−11 15 17 25 19 17 13 15 25
36
21 25 42 12.5 12 0.72 7.5 15.4
18.7
Table 6.1 Performance of the mobile TV LNA and comparison of the achieved performance with other published designs Research paper Technology Frequencya (GHz) NFmin(dB) IIP3 (dBm) Gain (dB) Power (mW)
−3 6.8 12 12.8 15.3 17.8 6.6 7.3
23.5
FOMb (dB)
6.1 Summary and Conclusions 141
142
6 Summary and Conclusions
Unfortunately, this approach increases the power consumption budget of the system. Even attempts to replace the SAW filter by an integrated solution in [21] were associated with using 15 external components mounted on a PC board. An LNA with variable-gain (~50 dB gain range) can help in achieving the stringent linearity requirement of the DVB-H system with low-power consumption [13, 24, 29]. However, this solution is very complex and can easily limit the dynamic range of the mobile TV receiver (see Section 4.2). As presented in Chapter 4, an RF passive attenuator block was used before the LNA to change the RF gain and thus enhance the DVB-H front-end’s linearity and dynamic range. Combining passive step attenuators with active gain steps helps in maximizing the receiver dynamic range (see Section 4.3). As was shown in Chapter 3, the RF attenuator also helps in relaxing the linearity requirement of the DVB-H LNA, and thus it can be traded for lower power consumption. The RF attenuator circuit was designed and fabricated using 65 nm CMOS technology. The attenuation variations were programmed using a 5 bit nonlinear thermometer decoder designed and integrated in the same test chip. A programmable matching network was also implemented to provide a stable 50 W input resistance to the DVB-H antenna at all attenuation settings. The RF attenuator prototype was characterized in the lab and the measurement results were shown previously in Table 4.4. The performance of the designed block meets the requirement needed (see Section 4.4) to maximize the dynamic range of the DVB-H front-end receiver.1
6.1.3 Wide Dynamic Range Mobile TV Front-End A front-end architecture that integrates the circuits discussed above with an RF level indicator circuit to facilitate the automatic gain control (AGC) function was implemented in 65 nm CMOS technology. The performance of the described DVB-H front-end circuit when large interferer signals are present at its input was quantified in the lab. The measurement results show the effectiveness of incorporating the intrinsic device properties, noise and distortion cancellation circuit techniques, and passive gain control mechanisms in optimizing the dynamic range of the DVB-H receiver. Although the techniques described in this book target the DVB-H mobile TV standard, they can be scaled to any other wireless communication standard that needs a wide dynamic range front-end architecture.
In this discussion, other factors that can limit the dynamic range of the DVB-H receiver such as the local oscillator (LO) phase noise were ignored since they are out of the scope of this book. The rejection required for the harmonics of the LO was also ignored. SAW filters may still be needed in these cases.
1
6.2 Further Research Areas
143
6.2 Further Research Areas There are a number of possible avenues for continuing the work initiated in this book. These can be divided into two main areas: system studies and circuit studies.
6.2.1 System Studies The work in this book can be taken one step further by physically integrating the passive quadrature mixer as well as the local oscillator (LO) chain circuits with the wide dynamic range front-end circuit proposed in Chapter 5. This integration would help to verify the whole RF front-end receiver performance in the context of the DVB-H standard requirements given in Table 3.2. Additionally, the mixer and LO chain integration will introduce some other research areas for further study. One such research area deals with the suitability of using the direct-conversion (DCR) architecture for mobile TV applications. The assumption made by Womac [17] that a high-pass pole can be used to filter out the DC offset generated from using the DCR architecture without damaging the DVB-H OFDM information must be examined to determine the feasibility of this architecture in achieving the power requirement of the DVB-H system. The mixer and LO chain integration will also facilitate the study of the “onfrequency” phase noise that couples from the LO path to the RF input and its effect on the overall noise performance of the DVB-H receiver. This noise can easily limit the DVB-H front-end receiver dynamic range, and thus limit the benefits of implementing the techniques described in this book to optimize the DVB-H receiver dynamic range. As was stated in Chapter 1, this book did not discuss the distortion originating from the mixing that occurs between interferer components and noise components of the LO (i.e., harmonic mixing problem), as the absence of the mixer integration prevents the need for such examination. A study is therefore needed to evaluate the effect of the harmonic mixing problem on the DVB-H receiver dynamic range. The complex I/Q three-phase mixer proposed in [20] might be required to limit the effect of the harmonic mixing on the overall DVB-H receiver dynamic range. Finally, the derived specifications of the DVB-H front-end receiver were based on the state-of-the-art baseband circuits that were available when this work was started [34, 35]. An on-going evaluation of the new baseband circuit performances should be carried out in order to further relax the noise-linearity trade-offs of the DVB-H front-end circuit, and thus improve the achieved dynamic range and further decrease the overall power consumption of the DVB-H system.
144
6 Summary and Conclusions
6.2.2 Circuit Studies Even though the designed mobile TV LNA [123] can achieve the DVB-H system specification requirements given in Table 3.3, research is still needed to further improve the dynamic range. LNA designers are encouraged to come up with new mechanisms to minimize the MOSFET output resistance nonlinearity that is associated with the move toward nano-scale technologies. All the research efforts to improve the LNA IIP3 values so far have been focused on minimizing the MOSFET transconductance nonlinearity [75, 96, 97]. Such studies will have a great impact on the power budget of the next generation wireless products. As for the LNA presented in this book, the IIP3 can be improved by replacing the active load used for current reuse in the common-source branch by a polysilicon resistor. However, removing the PMOS stage will limit the achievable gain of this topology to 24 dB instead of 36 dB achieved in this work. This gain reduction will result in an increase in the baseband noise contribution to the cascaded NF of the DVB-H receiver to 1.3 dB instead of 0.1 dB. However, replacing the active load by a polysilicon resistor could be a very promising idea if the baseband circuit’s noise performance improved in the future. A new mechanism of gain control would then have to be investigated in this case. As was demonstrated in Chapter 4, the RF passive attenuator gain ranges were limited by the coupling between the package downbonds. Therefore, a study is needed to improve the isolation between the QFN downbonds which would lead to an increase in the attenuator circuit gain ranges. It is anticipated that a high performance package such as a chip scale package (CSP) would be a good candidate for attenuator packaging. A CSP enables good RF ground as well as minimum parasitics when it is connected to the circuit board.
References
1. G. Faria, J. Henriksson, E. Stare, P. Talmola, DVB-H: digital broadcast services to handheld devices. IEEE J. Solid State Circuits 94(1), 198–209 (January 2006) 2. M. Chari, F. Ling, A. Mantravadi, R. Krishnamoorthi, R. Vijayan, G. Walker, R. Chandhok, FLO physical layer: an overview. IEEE Trans. Broadcast. 53(1, part 2), 145–160 (March 2007) 3. V. Nee, R. Prasad, OFDM for Wireless Multimedia Communications (Artech House, Boston, MA, 1998) 4. X.D. Yang, Y.H. Song, T.J. Owens, J. Cosmas, T. Itagaki, Performance analysis of the OFDM scheme in DVB-T, in IEEE Circuits and Systems Symposium on Emerging Technologies: Frontiers of Mobile and Wireless Communication, Shanghai, China, June 2004, pp. 489–492 5. R. Makowitz, A. Turner, J. Gledhill, M. Mayr, A single-chip DVB-T receiver. IEEE Trans. Consum. Electron. 44(3), 990–993 (August 1998) 6. T. Kim, B. Kim, Y. Cho, B. Kim, K. Lee, A 13 dB IIP3 improved low-power CMOS RF programmable gain amplifier using differential circuit transconductance linearization for various terrestrial mobile D-TV applications, in IEEE Symposium on VLSI Circuits, Kyoto, Japan, June 2005, pp. 344–347 7. P. Antoine, P. Bauser, H. Beaulaton, M. Buchholz, D. Carey, T. Cassagnes, T.K. Chan, S. Colomines, F. Hurley, D.T. Jobling, N. Kearney, A.C. Murphy, J. Rock, D. Salle, T. CaoThong, A direct conversion receiver for DVB-H. IEEE J. Solid State Circuits 40(12), 2536–2546 (December 2005) 8. I. Vassiliou, K. Vavelidis, N. Haralabidis, A. Kyranas, Y. Kokolakis, S. Bouras, G. Kamoulakos, C. Kapnistis, S. Kavadias, N. Kanakaris, A 65 nm CMOS multistandard, multiband TV tuner for mobile and multimedia applications. IEEE J. Solid State Circuits 43(7), 1522–1533 (July 2008) 9. M. Dawkins, A.P. Burdett, N. Cowley, A single-chip tuner for DVB-T. IEEE J. Solid State Circuits 38(8), 1307–1317 (August 2003) 10. C.H. Heng, M. Gupta, S.H. Lee, D. Kang, B.S. Song, A CMOS TV tuner/demodulator IC with digital image rejection. IEEE J. Solid State Circuits 40(12), 2525–2535 (December 2005) 11. L. Connell, N. Hollenbeck, M. Bushman, D. McCarthy, S. Bergstedt, R. Cieslak, J. Caldwell, A CMOS broadband tuner IC, in IEEE International Solid-State Circuits Conference, San Francisco, February 2002, pp. 400–401 12. R. Montemayor, A 410-mW 1.22-GHz downconverter in a dual-conversion tuner IC for open cable applications. IEEE J. Solid State Circuits 39(4), 714–718 (April 2004) 13. G. Retz, P. Burton A CMOS up-conversion receiver front-end for cable and terrestrial DTV application, in IEEE International Solid-State Circuits Conference, San Francisco, February 2003, pp. 442–443 14. J. Sinderen, F. Seneschal, E. Stikvoort, F. Mounaim, M. Notten, H. Brekelmans, O. Crand, F. Singh, M. Bernard, A. Tombeur, A 48–860 MHz digital cable tuner IC with integrated RF and IF selectivity, in IEEE International Solid-State Circuits Conference, San Francisco, February 2003, pp. 444–445
145
146
References
15. S. Lerstaveesin, M. Gupta, D. Kang, B.S. Song, A 48–860 MHz CMOS low-IF direct-conversion DTV tuner. IEEE J. Solid State Circuits 43(9), 2013–2024 (September 2008) 16. B. Razavi, RF Microelectronics (Prentice-Hall, Upper Saddle River, NJ, 1998) 17. M. Womac, A. Deiss, T. Davis, R. Spencer, B. Abesingha, P. Hisayasu, Dual-band singleended input direct-conversion DVB-H receiver, in IEEE International of Solid-State Circuits Conference, San Francisco, February 2006, pp. 2514–2523 18. Y.H. Ka, C.J. Yu, A design of image reject mixer for DTV tuner, in IEEE Asia-Pacific Microwave Conference, Suzhou, China, December 2005, pp. 4–7 19. J.M. Stevenson, P. Hisayasu, A. Deiss, B. Abesingha, K. Beumer, J. Esquivel, A multi-standard analog and digital TV tuner for cable and terrestrial applications, in IEEE International SolidState Circuits Conference, San Francisco, February 2007, pp. 210–211 20. V. Peluso, Y. Xu, P. Gazzerro, Y. Tang, L. Liu, Z. Li, W. Xiong, C. Persico, A dual-channel direct-conversion CMOS receiver for mobile multimedia broadcasting, in IEEE International Solid-State Circuits Conference, San Francisco, February 2006, pp. 2524–2533 21. I. Vassiliou, K. Vavelidis, S. Bouras, S. Kavadias, Y. Kokolakis, G. Kamoulakos, A. Kyranas, C. Kapnistis, N. Haralabidis, A 0.18 mm CMOS dual-band direct-conversion DVB-H receiver, in IEEE International Solid-State Circuits Conference, San Francisco, February 2006, pp. 2494–2503 22. Y.J. Kim, J.W. Kim, V.N. Parkhomenko, D. Baek, J.H. Lee, E.Y. Sung, I. Nam, A multi-band multi-mode CMOS direct-conversion DVB-H tuner, in IEEE International Solid-State Circuits Conference, San Francisco, February 2006, pp. 2504–2513 23. M. Gupta, S. Lerstaveesin, D. Kang, B.S. Song, A 48-to-860 MHz CMOS direct-conversion TV tuner, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 2013–2024 24. V. Fillâtre, J.R. Tourret, S. Amiot, M. Bernard, M. Bouhamame, C. Caron, O. Crand, A. Daubenfeld, G. Denise, T. Kervaon, M. Kristen, L. Coco, F. Mercier, J.M. Paris, S. Prouet, V. Rambeau, S. Robert, F. Seneschal, J. van Sinderen, O. Susplugas, A SiP tuner with integrated LC tracking filter for both cable and terrestrial TV reception, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 2809–2821 25. T. Shinyaitos, O. AlsushiSakail, M. Okazaki, M. Nagsumim, A. Saitow, K. Kioi, M. Koutani, A digital TV receiver RF and BB chipset with adaptive bias-current control for mobile applications, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 212–213 26. B. Kim, T. Kim, Y. Cho, M.S. Jeong, S. Kim, H. Yoo, S.M. Moon, T.J. Lee, A 100 mW dualband CMOS mobile-TV tuner IC for T-DMB/DAB and ISDB-T, in IEEE International SolidState Circuits Conference, San Francisco, February 2006, pp. 2534–2543 27. C.H. Heng, M. Gupta, S.H. Lee, D. Kang, B.S. Song, A CMOS TV tuner/demodulator IC with digital image rejection, in IEEE International Solid-State Circuits Conference, San Francisco, February 2005, pp. 2525–2535 28. H. van Rumpt, D. Kasperkovitz, J. van der Tang, B. Nauta, UMTV: a single chip TV receiver for PDAs, PCs, and cell phone, in IEEE International Solid-State Circuits Conference, San Francisco, February 2005, pp. 428–608 29. S. Azuma, H. Kawamura, S. Kawama, S. Toyoyama, T. Hasegawa, K. Kagoshima, M. Koutani, H. Kijima, K. Sakuno, K. Iizuka, A digital terrestrial television (ISDB-T) tuner for mobile applications, in IEEE International Solid-State Circuits Conference, San Francisco, February 2004, pp. 278–528 30. M. Haruoka, Y. Utsurogi, T. Matsuoka, K. Taniguchi, A dual-band image-reject mixer for GPS with 64 dB image rejection, in IEEE Topical Conference on Wireless Communication Technology, Honolulu, Hawaii, October 2003, pp. 168–169 31. Mobile and Portable DVB-T/H Radio Access – Part 1: Interface Specification, Int. Standard IEC 62002-1, IEC, Oct. 2005 32. Mobile and Portable DVB-T/H Radio Access – Part 2: Interface Conformance Testing, Int. Standard IEC 62002-2, IEC, Oct. 2005
References
147
33. V. Aparin, L.E. Larson, Analysis and reduction of cross-modulation distortion in CDMA receivers. IEEE Trans. Microw. Theory Tech. 51(5), 1591–1602 (May 2003) 34. A. Tekin, H. Elwan, K. Pedrotti, A universal low-noise analog receiver baseband with noiseshaping blocker filtering. IEEE J. Solid State Circuits, in press 35. A. Tekin, A Low Analog Radio Baseband with Novel Noise Shaping Circuit Techniques. Ph.D. Dissertation, Electrical Engineering Department, University of California, Santa Cruz, 2008 36. M. Zargari, S. Jen, B. Kaczynski, M. Lee, M. Mack, S. Mehta, S. Mendis, K. Onodera, H. Samavati, W. Si, K. Singh A. Tabatabaei, M Terrovitis, D. Weber, D. Su, B. Wooley, A singlechip dual-band, tri-mode CMOS transceiver for IEEE 802.11a/b/g WLAN, in IEEE International Solid-State Circuits Conference, San Francisco, February 2004, pp. 96–97 37. M. Zargari, D.K. Su, C.P. Yue, S. Rabii, D. Weber, B.J. Kaczynski, S.S. Mehta, K. Singh, S. Mendis, B.A. Wooley, A 5-GHz CMOS transceiver for IEEE 802.11a wireless LAN, in IEEE International Solid-State Circuits Conference, San Francisco, February 2002, pp. 126–127 38. S. Andersson, O. Drugge, C. Svensson, Wideband LNA for a multistandard wireless receiver in 0.18 mm CMOS, in IEEE European Solid-State Circuits Conference, Estoril, Potugal, September 2003, pp. 655–658 39. J. Craninckx, M. Liu, D. Hauspie, V. Giannini, T. Kim, J. Lee, M. Libois, D. Debaillie, C. Soens, M. lngels, A. Baschirotto, J. Van Driessche, L. Van der Perre, P. Vanbekbergen, A fully reconfigurable software-defined radio transceiver in 0.l3 mm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 346–347 40. R. Bagheri, A. Mirzaei, S. Chehrazi, M.E. Heidari, M. Lee, M. Mikhemar, W. Tang, A. Abidi, An 800-MHz 6-GHz software-defined wireless receiver in 90-nm CMOS. IEEE J. Solid State Circuits 41(12), 2860–2876 (December 2006) 41. A. Youssef, J. Haslett, S. Magierowski, Design issues for sensor network RF receivers, in IEEE Canadian Conference on Electrical and Computer Engineering, Vancouver, BC, April 2007, pp. 1535–1538 42. A.Youssef, J. Haslett, J. Nielsen, 4G wireless systems: multi-media over wireless, in Wireless Technologies and Circuits Workshop, European Microwave Week, Manchester, UK, September 2006, pp. 8–12 43. J.S. Goo, H.T. Ahn, D.J. Ladwig, Z. Yu, T.H. Lee, R.W. Dutton, A noise optimization technique for integrated low-noise amplifiers. IEEE J. Solid State Circuits 37(8), 994–1002 (August 2002) 44. A. van der Ziel, Noise in Solid State Devices and Circuits (Wiley, New York, NY, 1986) 45. J. Haslett, F. Trofimenkoff, Thermal noise in field-effect devices. Proc. Inst. Electr. Eng. 116(11), 1863–1868 (November 1969) 46. A.J. Scholten, H.J. Tromp, L.F. Tiemeijer, R. van Langevelde, R.J. Havens, P.W.H. de Vreede, R.F.M. Roes, P.H. Woerlee, A.H. Montree, D.B.M. Klaassen, Accurate thermal noise model for deep-submicron CMOS. International Electron Devices Meeting. IEEE Electron Devices Society, pp. 155–158, December 1999 47. C.H. Chen, M.J. Deen, Channel noise modeling of deep submicron MOSFETs. IEEE Trans. Electron Devices 49(8), 1484–1487 (August 2002) 48. A. Scholten, L. Tiemeijer, R. van Langevelde, R. Havens, A. Van Duijnhoven, R. de Kort, D. Klaassen, Compact modeling of noise for RF CMOS circuit design. IEE Proc. Circuits Devices Syst. 151(2), 167–174 (April 2004) 49. P. Jindal, Compact noise models for MOSFETs. IEEE J. Solid State Circuits 53(9), 2051–2061 (September 2006) 50. P. Kelin, An analytical thermal model of deep submicron MOSFETs. IEEE Electron Device Lett. 20(8), 339–401 (August 1999) 51. G. Knoblinger, P. Kelin, M. Tibout, A new model for thermal channel noise of deep-submicron MOSFETS and its application in RF CMOS design. IEEE J. Solid State Circuits 36(5), 831–837 (May 2001)
148
References
52. M. Obrecht, T. Manku, M. Elmasry, Simulation of temperature dependence of microwave noise in metal-oxide-semiconductor-field-effect transistors. J. Appl. Phys. 39, no. 4A, 1690– 1693 (April 2000) 53. M.S. Obrecht, E. Abou-Allam, T. Manku, Diffusion current and its effect on noise in submicron MOSFETs. IEEE Trans. Electron Devices 49(3), 524–526 (March 2002) 54. W. Liu, M.C. Chang, Transistor transient studies including transcapacitive current and distributive gate resistance for inverter circuits. IEEE Trans. Circuits Syst. I: Fundam. Theory Appl. 45(4), 416–422 (April 1998) 55. B. Razavi, R.H. Yan, K.F. Lee, Impact of distributed gate resistance on the performance of MOS devices. IEEE Trans. Circuits Syst. I: Fundam. Theory Appl. 41(11), 750–754 (November 1994) 56. J.T. Colvin, S.S. Bhatia, K.K.O, Effects of substrate resistance on LNA performance and a bondpad structure for reducing the effects in a silicon bipolar technology. IEEE J. Solid State Circuits 34(9), 1339–1344 (September 1999) 57. E.P. Vandamme, L.K. Vandamme, Critical discussion on unified 1/f noise models for MOSFETs. IEEE Trans. Electron Devices 47(11), 2146–2152 (November 2000) 58. T.H. Lee, The Design of CMOS Radio-Frequency Integrated Circuits, 2nd edn. (Cambridge University Press, New York, NY, 2004) 59. P. Wambacq, W. Sansen, Distortion Analysis of Analog Integrated Circuits, 1st edn. (Kluwer Academic, Norwell, MA, 1998) 60. V. Aparin, G. Brown, L. Larson, Linearization of CMOS LNA’S via optimum gate biasing, in IEEE International Symposium on Circuits and Systems, Vancouver, BC, May 2004, pp. 748–751 61. W. Sansen, Analog Design Essentials, 1st edn. (Springer, 2008) 62. A. Abidi, G. Pottie, W. Kaiser, Power-conscious design of wireless circuits and systems. Proc. IEEE 88(10), 1528–1545 (October 2000) 63. J.Y. Chang, A. Abidi, M. Gaitan, Large suspended inductors on silicon and their use in a 2 mm CMOS RF amplifier. IEEE Electron Device Lett. 14(10), 246–248 (May 1993) 64. S. Voinigescu, T. Dickson, T. Chalvatzis, A. Hazneci, E. Laskin, R. Beerkens, I. Khalid, Algorithmic design methodologies and design porting of wireline transceiver IC building blocks between technology nodes, in IEEE Custom Integrated Circuits Conference, San Jose, CA, September 2005, pp. 111–118 65. A. Rofougaran, J.Y. Chang, M. Rofougaran, A. Abidi, A 1 GHz CMOS RF front-end IC for a direct-conversion wireless receiver. IEEE J. Solid State Circuits 31(7), 880–889 (July 1996) 66. E.H. Nordholt, The Design of High-Performance Negative-Feedback Amplifiers (Elsevier, Amsterdam, 1983) 67. E. Duvivier, G. Puccio, S. Cipriani, L. Carpineto, P. Cusinato, B. Bisanti, F. Galant, F. Chalet, F. Coppola, S. Cercelaru, N. Vallespin, J.C. Jiguet, G. Sirna, A fully integrated zero-IF transceiver for GSM-GPRS quad-band application. IEEE J. Solid State Circuits 38(12), 2249–2257 (December 2003) 68. H. Adiseno, H. Olsson, A 1.8-V wide-band CMOS LNA for multiband multistandard front-end receiver, in IEEE European Solid-State Circuits Conference, Estoril, Portugal, September 2003pp. 141–144 69. J. Janssens, M. Steyaert, H. Miyakawa, A 2.7 volt CMOS broadband low-noise amplifier, in IEEE Symposium on VLSI Circuits, Kyoto, Japan, June 1997, pp. 87–88 70. A. van der Ziel, Noise in solid-state devices and lasers. Proc. IEEE 58(8), 1178–1206 (August 1970) 71. D.K. Shaeffer, T.H. Lee, A 1.5-V, 1.5-GHz CMOS low noise amplifier. IEEE J. Solid State Circuits 32(5), 745–759 (May 1997) 72. D.K. Shaeffer, T.H. Lee, Corrections to A 1.5-V, 1.5-GHz CMOS low noise amplifier. IEEE J. Solid State Circuits 40(6), 1397–1398 (June 2005) 73. A. Youssef, Design guidelines for the noise optimization of a 0.18 Micron CMOS low-noise amplifier. Intl. J. Analog Integr. Circuits Signal Process. 46(3), 193–201 (March 2006)
References
149
74. T.K. Nguyen, C.H. Kim, M.S. Yang, S.G. Lee, CMOS low-noise amplifier design optimization techniques. IEEE Trans. Microw. Theory Tech. 52(5), 1433–1442 (May 2004) 75. J.S. Goo, H.T. Ahn, D.J. Ladwig, Z. Yu, T.H. Lee, R.W. Dutton, A noise optimization technique for integrated low-noise amplifiers. IEEE J. Solid State Circuits 37(8), 994–1002 (August 2002) 76. L. Belostotski, J. Haslett, Noise figure optimization of inductively degenerated CMOS LNAs with integrated gate inductors. IEEE Trans. Circuits Syst. Part I: Regular Papers 53(7), 1409–1422 (July 2006) 77. F. Gatta, E. Sacchi, F. Svelto, P. Vilmercati, R. Castello, A 2-dB noise figure 900-MHz differential CMOS LNA. IEEE J. Solid State Circuits 36(10), 1444–1452 (October 2001) 78. J. Janssens, J. Crols, M. Steyaert, A 10 mW inductorless, broadband CMOS low noise amplifier for 900 MHz Wireless Communications, in IEEE Custom Integrated Circuits Conference, San Jose, CA, September 1998, pp. 75–78 79. J. Zhou, D. Allstot, A fully integrated CMOS 900MHz LNA utilizing monolithic transformers, in IEEE International Solid-State Circuits Conference, San Francisco, February 1998, pp. 132–133 80. A. Shahani, D. Shaeffer, T.H. Lee, A 12mW wide dynamic range CMOS front-end for a portable GPS receiver. IEEE J. Solid State Circuits 32(12), 2061–2070 (December 1997) 81. A. Karanicolas, A 2.7-V 900-MHz CMOS LNA and mixer. IEEE J. Solid State Circuits 31(12), 1939–1944 (December 1996) 82. A. Youssef, J. Haslett, Low power interference-robust UWB low noise amplifier in 0. 18-mm CMOS technology, in IEEE International Midwest Symposium on Circuits and Systems, Montreal, Canada, August 2007, pp. 1006–1009 83. S. Chehrazi, A. Mirzaei, R. Bagheri, A. Abidi, A 6.5 GHz wideband CMOS low noise amplifier for multi-band use, in IEEE Custom Integrated Circuits Conference, San Jose, CA, September 2005, pp. 801–804 84. H.J. Lee, D.S. Ha, S.S. Choi, A 3 to 5GHz CMOS UWB LNA with input matching using miller effect, in IEEE International Solid-State Circuits Conference, San Francisco, February 2006, pp. 731–740 85. R. Hu, An 8-20-GHz wide-band LNA design and the analysis of its input matching mechanism. IEEE Microw. Wireless Compon. Lett. 14(11), 528–530 (November 2004) 86. R. Hu, Wide-band matched LNA design using transistor’s intrinsic gate-drain capacitor. IEEE Trans. Microw. Theory Tech. 54(3), 1277–1286 (March 2006) 87. L. Belostotski, J. Haslett, Wide band room temperature 0.35-dB noise figure LNA in 90-nm Bulk CMOS, in IEEE Radio and Wireless Symposium, Long Beach, CA, January 2007, pp. 221–224 88. T. Cho, E. Dukatr, M. Mack, D. MacNally, M. Marringa, A single-chip CMOS direct-conversion transceiver for 900 MHz spread-spectrum digital cordless phones, in IEEE International Solid-State Circuits Conference, San Francisco, February 1999, pp. 228–229 89. J. Zhou, S. Embabi, J.P. de Gyvez, E. Sanchez-Sinencio, Using capacitive cross-coupling technique in RF low noise amplifiers and down-conversion mixer design, in IEEE European Solid-State Circuits Conference, Stockholm, Sweden, September 2000, pp. 77–80 90. S.B.T. Wang, A.M. Niknejad, R.W. Brodersen, A sub-mW 960-MHz ultra wideband CMOS LNA, in IEEE Radio Frequency Integrated Circuits Symposium, Long Beach, CA, June 2005, pp. 35–38 91. X. Li, S. Shekhar, D. Allstot, Gm boosted common-gate LNA and differential colpitts VCO/ QVCO in 0.18 µm CMOS. IEEE J. Solid State Circuits 40(12), 2609–2619 (December 2005) 92. F. Bruccoleri, E.A.M. Klumperink, B. Nauta, Noise cancelling in wideband CMOS LNAs, in IEEE International Solid-State Circuits Conference, San Francisco, February 2002, pp. 407–407 93. J. Borremans, P. Wambacq, D. Linten, An ESD-protected DC-to-6 GHz 9.7 mW LNA in 90 nm digital CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 422–423
150
References
94. R. Ramzan, S. Andersson, J. Dabrowski, C. Svensson, A 1.4 V 25 mW inductorless wideband LNA in 0.l3 mm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 424–425 95. Y. Ding and R Harjani, A +18 dBm IIP3 LNA in 0.35mm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2001, pp. 162–163 96. Y.S. Youn, J.H. Chang, K.J. Koh, Y.J. Lee, H.K. Yu, A 2 GHz 16 dBm IIP3 low noise amplifier in 0.25 mm CMOS technology, in IEEE International Solid-State Circuits Conference, San Francisco, February 2003, pp. 452–453 97. V. Aparin, N. Kim, G. Brown, Y. Wu, A. Cicalini, S. Kwok, C. Persico, A fully-integrated highly linear zero-IF CMOS cellular CDMA receiver, in IEEE International Solid-State Circuits Conference, San Francisco, February 2005, pp. 324–325 98. L. Belostotski, Wide-Band CMOS Low Noise Amplifier for the Square Kilometer Array Radio Telescope. Ph.D. Thesis, University of Calgary, 2007 99. P. van Zeijl, J.W. Eikenbroek, P.P. Vervoort, S.S. Tangenberg, G. Shipton, E. Kooistra, I. Keekstra, D. Belot, A bluetooth radio in 0.18 mm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2002, pp. 1679–1687 100. J. Rogin, I. Kouchev, Q. Huang, A 1.5 V 45 mW direct conversion WCDMA receiver IC in 0.13 mm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2003, pp. 268–269 101. J.H. Zhan, S. Stewart, A 5 GHz resistive-feedback CMOS LNA for low-cost multi-standard applications, in IEEE International Solid-State Circuits Conference, San Francisco, February 2006, pp. 721–730 102. I. Nam, B. Kim, K. Lee, CMOS RF amplifier and mixer circuits utilizing complementary characteristic of parallel combined NMOS and PMOS devices. IEEE Trans. Microw. Theory Tech. 53(5), 1662–1671 (May 2005) 103. R.P. Jindal, Noise associated with distributed resistance of MOSFET gate structures in integrated circuits. IEEE Trans. Electron Devices ED-31(10), 1505–1509 (October 1984) 104. B. Heydari, M. Bohsali, E. Adabi, E. A. Niknejad, Low-power mm-wave components up to 104 GHz in 90 nm CMOS, in IEEE International Solid-State Circuits Conference, San Francisco, February 2007, pp. 200–201 105. C.H. Chen, C.S. Chang, C.P. Chao, J.F. Kuan, C.L. Chang, S.H. Wang, H.M. Hsu, W.Y. Lien, Y.C. Tsai, H.C. Lin, C.C. Wu, C.F. Huang, S.M. Chen, P.M. Tseng, C.W. Chen, C.C. Ku, T.Y. Lin, C.F. Chang, H.J. Lin, M.R. Tsai, S. Chen, C.F. Chen, M.Y. Wei, Y.J. Wang, J.C.H. Lin, W.M. Chen, C.C. Chang, M.C. King, C.M. Huang, C.T. Lin, J.C. Guo, G.J. Chern, D.D. Tang, J.Y.C. Sun, 90 nm CMOS MS/RF based foundry SOC technology comprising superb 185 GHz fT RFMOS and versatile, high-Q passive components for cost/performance optimization, in IEEE International Electron Devices Meeting, Washington, pp. 2.5.1–2.5.4, December 2003 106. B. Razavi, Design of Analog CMOS Integrated Circuits (McGraw-Hill, 2001) 107. Y. Li, S.M. Yu, J.R. Hwang, F.L. Yang, Discrete dopant fluctuation 20 nm/15 nm-gate planer CMOS. IEEE Trans. Electron Devices 55(6), 1449–1455 (June 2008) 108. T. Hook, J. Brown, P. Cottrell, E. Adler, D. Hoyniak, J. Johnson, R. Mann, Lateral ion implant straggle and mask proximity effects. IEEE Trans. Electron Devices 50(3), 1946– 1951 (September 2003) 109. Balun Transformer B4F Type #617 Data Sheet, Toko America Inc. 110. M. Koutani, H. Kawamura, S. Toyoyama, K. Iizuka, A digitally controlled variable-gain low-noise amplifier with strong immunity to interferers. IEEE J. Solid State Circuits 42(11), 2395–2403 (November 2007) 111. T. Yamaji, N. Kanou, T. Itakura, A temperature stable CMOS variable-gain amplifier with 80-dB linearly controlled gain range. IEEE J. Solid State Circuits 37, 553–558 (May 2002) 112. Q.H. Doung, L. Quan, S.G. Lee, An all CMOS 84 dB-linear low-power variable gain amplifier, in IEEE Symposium on VLSI Circuits, Kyoto, Japan, June 2005, pp. 114–117
References
151
113. S.C. Tsou, C.F. Li, P.C. Huang, A low-power CMOS linear-in-decibel variable gain amplifier with programmable bandwidth and stable group delay. IEEE Trans. Circuits Syst. II 53(12), 1436–1440 (December 2006) 114. K. Iizuka, Integrated RF-ICs for mobile TV applications, in IEEE International Workshop on Radio-Frequency Integration Technology, Singapore, December 2007, pp. 116–121 115. B. Gilbert, A low noise wideband variable-gain amplifier using an interpolated ladder attenuator, in IEEE International Solid-State Circuits Conference, San Francisco, February 1991, pp. 280–281 116. R. van de Plassche, Integrated Analog-To-Digital and Digital-To-Analog Converters (Kluwer Academic, 1994) 117. G.L. Radulov, P.J. Quinn, P.C.W. van Bee, J.A. Hegt, A.H.M. van Roermund, A binaryto-thermometer decoder with built-in redundancy for improved DAC yield, in IEEE International Symposium on Circuits and Systems, Kos, Greece, May 2006, pp. 1414–1418 118. H. Chang, K. Hyun, A 4 GHz direct digital frequency synthesizer utilizing a nonlinear sineweighted DAC in 90 nm CMOS, in IEEE Asia Pacific Conference on Circuits and Systems, Macao, China, November 2008, pp. 1700–1703 119. A. Youssef, E. Youssoufian, and J. Haslett, Digitally-controlled RF passive attenuator in 65 nm CMOS for mobile TV tuner ICs, in IEEE International Symposium on Circuits and Systems, Paris, 2010 120. L. Larson, D. Jessie, Advances in RF packaging technologies for next-generation wireless communications applications, in IEEE Custom Integrated Circuit Conference, San Jose, CA, September 2003, pp. 323–330 121. S. Blaakmeer, E. Klumperink, D. Leenaerts, B. Nauta, An inductorless wideband BalunLNA in 65nm CMOS with balanced output, in IEEE European Solid-State Circuits Conference, Munich, Germany, September 2007, pp. 364–367 122. A. Youssef, A. Ismail, and J. Haslett, A sub-2dB noise figure LNA in 65 nm CMOS for mobile TV applications, in IEEE Radio and Wireless Symposium, Louisiana, January 2010 123. A. Youssef, A. Ismail, and J. Haslett, A sub-2 dB noise figure LNA in 65 nm CMOS for mobile TV application, in IEEE Topical Meeting on Silicon Monolithic Integrated Circuits in RF Systems, Louisiana, January 2010
Index
A Aardvark host adaptor, 83, 93 Active feedback, 34, 42, 45, 47, 50–52, 65 Active gain control, 13, 101 Active mixer, 61, 140 Adaptive bias current control, 98 AGC test, 136–138 Analog-to-digital converter (ADC), 5, 12, 59, 96, 133 Analog TV, 6 B Bake-gate resistance, 18 Balun, 78, 84, 85, 140 Bandgap reference, 72–74, 76, 83, 92 Bias generator circuit, 63, 65, 71–77, 81, 92, 97 Binary-weighted voltage, 74 Bluetooth radio, 44 Broadband L-degenerate, 36–38 C Cadence, 135 Calibration kit, 84 Capacitance density, 70, 92, 119, 120 Capacitive cross-coupling technique, 38, 44 CG amplifier, 39, 42, 52 Channel length modulation, 17, 29, 68 Channel thermal noise, 16, 30 China multimedia mobile broadcasting (CMMB), 2 Chip scale package (CSP), 144 Class AB amplifier, 76 CMMB. See China multimedia mobile broadcasting (CMMB) Coax cable, 85, 109, 124 Code rate, 98 Coupling capacitors, 34, 65, 70, 76, 80, 92, 118
Cross-modulation, 9, 54 CSP. See Chip scale package (CSP) Current mirrors, 72–75 Current reuse amplifiers, 34–36, 55, 56 Current steering technique, 98 D DC stability analysis, 76 Deep N-well, 70, 92 Demodulator, 6, 131, 132, 140 Depletion capacitance, 20, 24 Diffusion current, 17 Digital CMOS technology, 59 Digital comparators, 133 Digital signal processing, 6 Digital TV (DTV), 3, 6, 8–10 Doppler, 98 Drain noise current, 16, 17 Dummy devices, 75 DVB-H LNA, 59, 61–63, 66–69, 83, 85, 89, 92, 105, 113, 117, 142 DVB-H receiver, 9, 10, 59–63, 66, 69, 92, 113, 140, 142–144 E Equivalent input noise voltage, 18 F Flicker noise, 6, 16, 18, 19 Front-end specification, 6, 60–61 G GaAs, 139 Gain compression, 9
153
154 Gain step, 69, 85–88, 98–100, 103–105, 108, 113, 123, 135–137, 140, 142 GSM interference, 12 H Harmonic mixing, 3–5, 9, 143 I I2C chips, 81, 83 Image rejection (IR), 3, 5–7 IM3 products, 61 Inductorless topologies, 45 Inductorless-type amplifier, 43 InP, 139 Integrated Services Digital BroadcastingTerrestrial (ISDB-T), 2, 6, 103 Integrated TV tuner, 5 Interferer blockers, 19, 95 Interferers, 9, 10, 15, 60–62, 69, 131, 132, 136–138, 140, 142, 143 Interferer signal pattern, 61 Inter-integrated circuit, 81 ISDB-T. See Integrated Services Digital Broadcasting-Terrestrial (ISDB-T) L Layout parasitics, 63 LC tracking filter, 3, 5 L-degenerate amplifier, 13, 26, 31–34, 36–38, 52 Linear distortion, 19 Linearity test pattern, 9, 10, 105 LNA common-mode voltage, 72, 137 LNA physical implementation, 69–71 LNA shielding, 70 LO. See Local oscillator (LO) Local oscillator (LO), 4, 5, 9, 142, 143 LO harmonics, 5 Look-up table, 113, 116, 117 Low-pass filter (LPF), 5, 65, 76 L3 pattern reception, 11 M Maximum gain mode, 69, 85–88 Measurement equipments, 63, 76, 84, 93 Microcontroller, 81 MIM-capacitors, 70 Minimum gain mode, 69, 86 Minimum noise figure, 16, 19, 30, 31, 36, 40, 43
Index Mobile TV standard, 1, 9, 11–13, 59, 63, 84, 106, 130, 139, 142 Modulation error rate (MER), 98 MOS-cap, 119–121 MOS capacitors, 65, 70, 71, 74 MOSFET overdrive voltage, 18, 20, 22, 24, 25 weak-inversion region, 21 Multi-mode buffers, 63, 92 Multi-standard radios, 43 N Nanometer (NM) CMOS technology, 2, 12–14, 16–18, 21, 65, 67, 70, 81, 123, 135, 139, 142 Nano-scale technologies, 144 Narrowband LNA, 15 Noise cancellation mechanisms, 42, 51 Noise-cancelled load, 51 Noise-cancelling CG amplifier, 57 Noise floor, 12, 70, 71, 80, 87, 105, 136 Nonlinear output resistance, 67 Nonlinear thermometer coding, 113 O Open loop gain, 30, 76 Operational transconductance amplifier (OTA), 65, 67, 69, 72, 75–77, 79, 84, 85, 88, 92 interdigitating layout style, 77 phase margin, 76 OTA. See Operational transconductance amplifier (OTA) Output resistance nonlinearity, 22, 66, 69, 144 Oxide stress effect, 77 P Package bond wired, 120–123 Package bondwire parasitics, 71 Package downbonds, 144 Package pins, 136 Pad capacitance, 71 Passive gain control, 13, 99–102, 130, 142 Passive mixer, 12, 61, 86 Phase noise, 12, 142, 143 Pinch-off, 67 Poly-gate resistor, 18 Polysilicon material, 13, 69 Post-mixer amplifier (PMA), 12, 59–61
Index Power supply noise, 63, 73 Process and temperature variations, 21, 37, 43, 63, 72–74 Q QAM. See Quadrature amplitude modulation (QAM) Quad-flat package, 121 Quadrature amplitude modulation (QAM), 10, 11, 60, 61 Quadrature passive mixer, 12 Quality(Q) factor, 33, 43, 119, 120 R Reliability test, 89 Resonance, 15, 31, 33, 64 Reverse isolation, 84, 85 RF front-end, 4–7, 10, 12–14, 60–62, 66, 95–101, 103–105, 118, 123, 131–133, 135–138, 143 RF gain, 63, 95–104, 106, 130, 132, 135–137, 140, 142 RF level indicator (RFLI), 131–138, 142 circuit, 131–135, 138, 142 threshold voltages RFLI. See RF level indicator circuit (RFLI) RF passive attenuator, 12, 61, 98, 106, 108, 110, 123–129, 135, 136, 140–142, 144 RF tracking filter, 5 Routing parasitics, 70, 71, 123 S Second-order nonlinearity, 55 Selectivity, 105, 113 Selectivity test pattern, 9 Sensitivity, 2, 9–11, 43, 56, 60, 61, 70, 96, 98, 103, 104, 130 SFB amplifier, 13, 26–30, 32, 34, 39–41, 56, 57 Short-channel MOSFET, 28 Shot noise, 17 Shunt-feedback amplifier, 28, 29, 34, 40 Shunt-series amplifier, 80, 81 Shunt-shunt feedback type amplifier, 52 Signal-to-noise ratio (SNR), 10, 96, 98, 103, 104, 130, 132, 137, 138 Silicide material, 71, 92 Silicon layout, 63 Silicon tuner, 3 Single-ended to differential converter, 140
155 SMA connectors, 85, 124 SOC. See System On Chip (SOC) Soft nonlinearity, 20 S1 pattern reception, 61 Specifications control factor, 6 Spectrum analyzer, 77, 80, 85, 87, 124, 136 Step-by-step design, 12 Substrate contact, 18, 70, 120 Substrate coupling, 13, 70 System On Chip (SOC), 140 T Thermometer coding, 108, 113 Thin-oxide technology, 21 Three-phase mixer, 143 Threshold voltage, 18, 72, 132, 133, 135, 137 Transistor breakdown, 72 Transistor layout, 13, 69, 70, 75, 92 Transistor matching, 75 TV-can tuner, 3, 4 U U/D signal power ratios, 132 UHF. See Ultra High Frequency (UHF) Ultra High Frequency (UHF), 1, 4, 36, 46, 61, 64, 84, 85, 87, 106, 120, 123, 124, 138, 139 Ultra-thick metal, 71 Ultra-wide band, 15 Universal Serial Bus, 83 Up-down type double-conversion architecture, 5 V Variable gain, 59, 61–63, 66–69, 103, 117, 142 Variable-gain LNA, 12, 13, 99, 103, 104, 106, 108, 139–140 Variable resistor, 73–76, 123 Very high frequency (VHF), 4, 17, 36, 46, 124 VHF. See Very high frequency (VHF) Video reception, 1 Voltage headroom, 53, 65 W Well proximity effect, 75, 92 Wideband CMOS LNA, 12, 15–57, 139, 140 Wireless Local Area Network (WLAN), 3 WLAN. See Wireless Local Area Network (WLAN)
Author Biographies
AHMED YOUSSEF received his B.Sc. (Hon.) and M.Sc. degrees both in electrical engineering from Ain Shams University, Cairo, Egypt, in 1998 and 2002, respectively and a Ph.D. degree in electrical engineering from the University of Calgary, AB, Canada in 2009. Dr. Youssef worked with TRlabs, in Alberta, Canada from 2004 to 2007 as RF integrated circuits (RFICS) researcher, where he was mainly engaged in developing low-power front-end circuits for WLAN applications. From 2007 to 2009 he worked at Newport Media Inc. in Lake Forest, CA, where he was involved with the design and research of CMOS RFICs for multimedia applications. Dr. Youssef was the recipient of the Mobinil Telecommunication Inc. Pre-master Fellowship in 2000. He also received the Young Scientist Award at the Maastricht General Assembly of the International Union of Radio Science in 2002 in the Netherlands and an Honorable Mention at 2003 in the Symposium of the Microelectronics Research & Development in Montreal, Canada. In 2006 he received the Gordon Lewis Hedberg Doctoral Award at the University of Calgary and in 2007 the TRLabs Scholarship Award. Dr. Youssef has presented several short courses and tutorials in the area of high speed integrated circuit for wireless communication systems at IEEE conferences throughout the world and for various programs in industry as well.
157
158
Author Biographies
JAMES HASLETT is a Professor in the Department of Electrical and Computer Engineering at the University of Calgary. He has been an academic staff member for the past 39 years, and was the Head of the EE Department at Calgary from 1986 to 1997. Dr. Haslett was appointed “University Professor” in July of 2007, and is currently the Director of the provincial iCORE (informatics Circle of Research Excellence)-funded Advanced Technology Information Processing Systems (ATIPS) Lab at the University of Calgary. He held the TRLabs/iCORE/NSERC Senior Industrial Research Chair in Wireless Communications from 2002 to 2007, building a team of researchers specializing in radio frequency integrated circuit (RFIC) design for wireless communications applications. His current research focuses on radio frequency integrated circuit design for a variety of next-generation high-speed wireless communications applications and on the design of new integrated sensors for biomedical applications. He has published over 200 papers in peer-reviewed journals and conference proceedings, and holds 12 patents, several of which have been licensed to industry. He has graduated over 40 M.Sc. and Ph.D. students during his career. Dr. Haslett is a Life Fellow of the Institute of Electrical and Electronics Engineers (IEEE), a Fellow of the Engineering Institute of Canada, and a Fellow of the Canadian Academy of Engineering. He and his students have won numerous national and international awards for their research work. Dr. Haslett is currently a member of the Editorial Review Committees of five IEEE Transactions, is a member of several technical and executive committees of international IEEE Conferences, and is also a member of the provincial iCORE internal review committee that establishes research chair programs in Alberta.