Nothing Special   »   [go: up one dir, main page]

ELCHAAR 2022 Archivage

Download as pdf or txt
Download as pdf or txt
You are on page 1of 137

Design Tools for Millimeter Wave Ultra Wideband

Distributed Amplifiers
Mohamad El Chaar

To cite this version:


Mohamad El Chaar. Design Tools for Millimeter Wave Ultra Wideband Distributed Amplifiers. Micro
and nanotechnologies/Microelectronics. Université Grenoble Alpes [2020-..], 2022. English. �NNT :
2022GRALT067�. �tel-03936695�

HAL Id: tel-03936695


https://theses.hal.science/tel-03936695v1
Submitted on 12 Jan 2023

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est


archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents
entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non,
lished or not. The documents may come from émanant des établissements d’enseignement et de
teaching and research institutions in France or recherche français ou étrangers, des laboratoires
abroad, or from public or private research centers. publics ou privés.
THÈSE
Pour obtenir le grade de

DOCTEUR DE L’UNIVERSITÉ GRENOBLE ALPES


École doctorale : EEATS - Electronique, Electrotechnique, Automatique, Traitement du Signal (EEATS)
Spécialité : Nano électronique et Nano technologies
Unité de recherche : Techniques de l'Informatique et de la Microélectronique pour l'Architecture des
systèmes intégrés

Outils de conception pour l'amplification distribuée ultra large bande


aux fréquences millimétriques
Design Tools for Millimeter Wave Ultra Wideband Distributed
Amplifiers
Présentée par :
Mohamad EL CHAAR
Direction de thèse :
Florence PODEVIN Directrice de thèse
Enseignante-Chercheure, Université Grenoble Alpes
Sylvain BOURDEL Co-directeur de thèse
Enseignant-chercheur Grenoble INP, Université Grenoble Alpes
Manuel José BARRAGAN ASIAN Co-directeur de thèse
Chargé de Recherche, CNRS
Rapporteurs :
FRANK ELLINGER
PROFESSEUR DES UNIVERSITES, Technische Universität Dresden, Faculty of Electrical and
Computer Engineering
Thierry PARRA
PROFESSEUR DES UNIVERSITES, Université de Toulouse III Paul Sabatier
Thèse soutenue publiquement le 29 septembre 2022, devant le jury composé de :
Florence PODEVIN Directrice de thèse
PROFESSEUR DES UNIVERSITES, Université Grenoble Alpes
Sylvain BOURDEL Co-directeur de thèse
PROFESSEUR DES UNIVERSITES, Université Grenoble Alpes
Philippe BENECH Président
PROFESSEUR DES UNIVERSITES, Université Grenoble Alpes
FRANK ELLINGER Rapporteur
PROFESSEUR DES UNIVERSITES, Technische Universität Dresden,
Faculty of Electrical and Computer Engineering
Thierry PARRA Rapporteur
PROFESSEUR DES UNIVERSITES, Université de Toulouse III Paul
Sabatier
Jean-Yves DUPUY Examinateur
DOCTEUR EN SCIENCES, THALES DMS FRANCE SAS

Invités :
Philippe CATHELIN
INGENIEUR, STMicroelectronics
Antonio A. L. DE SOUZA
PROFESSEUR ASSOCIE, Federal University of Paraíba
Pour les amis, la famille et la France… Santé!
For Friends, Family and France… Cheers!

i
ii
“Vacation?... What is this?... Never heard of it…”
A PhD candidate in mm-wave IC design

iii
iv
Abstract

Over the last few years, both data center and network traffic increased by several folds
(more than six-fold, as stated by CISCO Systems, Inc.) and are still experiencing an
ongoing growth. This demand for higher capacity while meeting with higher end-to-end
quality service projects a continuous need for improvements in spectral efficiency
enhancement techniques; however, a given spectrum band has its limit of volume it can
support before reaching saturation levels. This pushes research efforts towards finding other
solutions and a promising one is by exploiting the uncrowded millimeter-wave (mm-wave)
spectrum through arranging more bandwidth. Signal amplification is one of the most basic
and prevalent circuit function in modern RF- and microwave- systems. On the topic of
wideband operation, distributed amplifier (DA) has proved to be a suitable and promising
candidate with its ultra-wideband (UWB) performance. Meanwhile, the design of
integrated circuit (IC) in mm-wave spectrum is becoming a challenging process due to the
unavoidable parasitic that affect circuit performance and complicate design process. The
reliability of the conventional design techniques through deriving mathematical equations,
a simple and direct design process commonly used at low frequencies, starts to diminish.
Another proposed design technique is by combining advanced-scripting languages with a
proper physical model; this will lead to a smart and efficient technique of designing mm-
wave IC DAs through a process referred as computer-automated design (CAutoD) process.
In this thesis, in particular, we propose for DAs an original ABCD-parameter based chain
matrix model, a well-known formalism originating from the passive RF-circuit domain, in
addition to a CAutoD methodology built upon it. Following this design technique is
considered complete to a much greater extent, since the model gives flexibility in
controlling a wider range of correlated design parameters. It is also considered reliable, a
crucial criterion for mm-wave design, since the model does not apply simplification or
neglect parasitic effects. It is simple, since the designer does not have to deal with the tiny
effects (parasitic) that affect DA behavior and, finally, it is considered versatile, since it
does not impose restrictions on the design component topologies the designer might be
interested to use. Our end-goal is to offer the mm-wave designer a different and interesting
technique in which the outcome of the suggested CAutoD process is a set of 3D graphical
design exploration (DSE) plots. Designers then can explore different feasible solutions and
choose the best design that meet their multiple performance objectives. Its benefit here is
demonstrated by exploring the design space of DAs with BWs ≥ 80 GHz where
STMicroelectronics (ST) 55-nm CMOS technology process is used, reporting 216 feasible
DA options to explore from. Two global optimum DAs amplifying frequencies up to
100 GHz were then implemented as circuit prototypes: a single stage DA with
6.01 GHz/mW of measured gain-bandwidth product over DC consumption (GBP⁄PDC ) and
a 17.5 THz GBP cascaded DA with 71.96 GHz/mW of GBP⁄PDC , being the highest figure
of merits, to the authors’ knowledge.

v
vi
Contents

Abstract .............................................................................................................................. v
Contents ............................................................................................................................ vii
Chapter 1 General Introduction .................................................................................... 1
1.1 Analog Signal Amplification ................................................................................... 2
1.2 Move Towards Low-Cost Production: CMOS Technology as a Solution .............. 2
1.3 Move Towards High-Frequency Band: Millimeter-Wave Band as a Solution ....... 3
1.4 Move Towards High-Frequency Band: Issues Encountered ................................... 3
1.5 Move Towards High-Frequency Band: CAutoD Technique as a Solution ............. 4
1.6 Thesis Objectives..................................................................................................... 4
1.7 Organization of the Report ...................................................................................... 5
Chapter 2 Introduction to Wideband Amplifiers – Distributed Amplifier for mm-
Wave Access ....................................................................................................................... 7
2.1 Future Trends for Millimeter-Wave Applications ................................................... 7
2.2 Amplifier Circuit Topologies with Wide Bandwidth Characteristic ..................... 10
2.2.1 Small-to-Large Bandwidth Amplifiers Topologies ....................................... 10
2.2.2 Table-of-Comparison for Performances Analysis ......................................... 17
2.3 What is a Distributed Amplifier? .......................................................................... 19
2.3.1 Walk Through History ................................................................................... 19
2.3.2 Structure and Basic Principle of Operation ................................................... 20
2.3.3 Performance Demonstration through Different Technologies and Topologies
………………………………………………………………………………25
2.3.4 State-of-the-Art Table on Distributed Amplifiers ......................................... 32
2.4 Conclusion ............................................................................................................. 35
Chapter 3 Proposition of a Novel Model to Design DAs: the Four-Port Chain
(ABCD) Model ................................................................................................................. 37
3.1 Introduction ........................................................................................................... 37
3.2 Existing Distributed Amplifier Design Techniques .............................................. 39
3.2.1 Artificial-Line Based Model [40] .................................................................. 39
3.2.2 Stage-Scaling Based Model [51] ................................................................... 40
3.2.3 Distributed-Line Based Model [52]............................................................... 41
3.2.4 Artificial Neural Network Based Model ....................................................... 45
3.2.5 Widespread Design Technique: Through CAD Software ............................. 46

vii
3.3 Design Limitations when Operating in mm-Wave Band ...................................... 47
3.4 Proposed Solution: Chain Matrix (ABCD) Based Design Model [73] ................. 49
3.5 Conclusion ............................................................................................................. 53
Chapter 4 Model Application – Computer Automated Design for Loss-
Compensated Distributed Amplifier.............................................................................. 56
4.1 Loss-Compensated Distributed Amplifier ............................................................. 56
4.1.1 Overview of Publication ................................................................................ 57
4.1.2 Loss-Compensation Technique: Graphical Analysis .................................... 58
4.1.3 Loss-Compensation Technique: Analytical Analysis ................................... 61
4.1.4 Implementing Loss-Compensated Cascode Gm-Cell in a Numerical
Computing Environment ........................................................................................... 63
4.2 Loss-Compensated Distributed Amplifier CAutoD Process ................................. 65
4.2.1 Distributed Amplifier Bandwidth Extension and Flattening Concept .......... 65
4.2.2 Algorithmic Design Methodology ................................................................. 67
4.3 3D Design Space Exploration Plots: Parameters and Variables ........................... 71
4.3.1 DA Design Parameters: Gain, Bandwidth and PDC ....................................... 71
4.3.2 DA Design Parameters: Synthesizing a Unit-cell ......................................... 73
4.4 Conclusion ............................................................................................................. 75
Chapter 5 100-GHz Single Stage CMOS Distributed Amplifier .............................. 77
5.1 STMicroelectronics 55-nm Node Process Technology ......................................... 78
5.2 100-GHz Single-Stage CMOS DA: Circuit Design .............................................. 78
5.3 100-GHz Single-Stage CMOS DA: Circuit Implementation ................................ 81
5.3.1 Microstrip Transmission Line ....................................................................... 81
5.3.2 High Quality and SRF Stacked Parallel Plate Shunt Capacitor [61] ............. 82
5.3.3 MOS Transistor Interconnection ................................................................... 89
5.3.4 Standalone Unit-Cell Layout and Microphotograph ..................................... 90
5.3.5 100-GHz Single-Stage DA: Layout and Microphotograph ........................... 91
5.4 100-GHz Single-Stage CMOS DA: Experimental Results ................................... 93
5.4.1 Vector Network Analyzer Description .......................................................... 93
5.4.2 S-parameters Analysis: Unit-Cell and DA Model Validation ....................... 94
5.4.3 Implemented DA Performance Analysis ....................................................... 95
5.5 Comparison with State-of-the-Art CMOS-Based DAs Performances .................. 98
5.6 Conclusion ............................................................................................................. 99
Chapter 6 THz Gain-Bandwidth Product 100-GHz Cascaded CMOS Distributed
Amplifier 101
6.1 THz GBP Cascaded CMOS DA: Circuit Design ................................................ 101

viii
6.2 THz GBP Cascaded CMOS DA: Circuit Implementation .................................. 103
6.2.1 Integrated Wideband Millimeter-Wave Bias-Tee [68]................................ 103
6.2.2 100-GHz Single-Stage DA: Layout and Microphotograph ......................... 109
6.2.3 Six-Stage Cascaded 100-GHz DA Layout and Microphotograph .............. 111
6.3 THz GBP Cascaded CMOS DA: Simulation Results ......................................... 112
6.3.1 S-parameter Analysis ................................................................................... 112
6.3.2 Group Delay Performance Analysis ............................................................ 113
6.3.3 Stability Performance Analysis ................................................................... 113
6.4 Comparison with State-of-the-Art DAs Performances........................................ 114
6.5 Conclusion ........................................................................................................... 116
Chapter 7 General Conclusion .................................................................................. 118
Publications .................................................................................................................... 120
Bibliography ................................................................................................................... 121

ix
x
Chapter 1

General Introduction

The aim of this first chapter is to overview the reader on the subject of this thesis work
that is wideband amplification in the millimeter-wave (mm-wave) frequency band
(30 GHz-300 GHz). This high frequency band attracts strong interest for its uncrowded
characteristic and opportunity to produce high date rate systems such that significant efforts
are being dedicated to overcome the challenges faced when designing high-end mm-wave
circuits. In particular, integrated circuit (IC) amplifier is a popular topic of study between
the research community that is undergoing continuous improvement to break the limits on
high frequency of operation, especially when the wideband criterion is targeted from a few
GHz to hundred GHz and above. The need for such criterion has lead researcher towards
distributed amplification as a promising candidate that offers such great capability and
hence a critical circuit to adequately design. This topic also gains widespread interest
especially when using complementary metal–oxide–semiconductor (CMOS) technology
process for its attractive low-cost, high-volume production features and its ability of high
frequency operation with continued device scaling. Nevertheless, a crucial challenge
encountered is the lack of appropriate and accurate design techniques at such band where
the simple analytical-based approaches, traditionally applied at low frequency spectrum,
begin to fail and therefore advocate for a revaluation of design techniques.
The following sections outlines our context starting from a general overview on analog
signal amplification and ending by stating our needs for an automated circuit design
approach when dealing with mm-wave operating frequencies. Distributed amplification,
which is the heart of this work, will be considered in Chapter 2, devoted specifically to
introducing wideband amplifiers.

Page | 1
1.1 Analog Signal Amplification

Signal amplification is one of the most basic and prevalent circuit functions in modern
RF and microwave systems. Amplifier block is found as a first component of any receiver
front-end circuitry, on which overall receiver performance depends, as well as the last
component of any transmitter front-end circuitry, on which a reliable transmission depends.
In addition, an amplifier can be considered for implementation as an intermediate block in
a complete system to compensate for any signal degradation encountered through intra-
stages circuitry.
Those applications share one common goal and it is magnifying a weak signal for further
RF-treatment or digital signal processing. From the literature, it is clear that the amplifier
design subject and application in fully integrated systems dates back to the early 21th
century. Demonstrative implementation inside receivers and transmitters front-ends, and
even transceivers have been reported using Si CMOS [1]-[3], SiGe BiCMOS [4]-[7] and
various III-V technologies such as GaAs and InP [8]-[10], to mention but a few. Those
show its widespread interest between RF- and microwave- research community for
performance exploration and improvement. While the demos of those systems can be
considered old dated, but their added contribution to the community and value stays the
same till this day. Their concept can be re-used in recent years for their originality and
enhanced by benefiting from the advancement of semiconductor technologies towards
smaller nodes for the purpose of operating at higher frequencies [11]-[13].

1.2 Move Towards Low-Cost Production: CMOS Technology as a


Solution

To meet today needs of the consumer marketplace and the radio chipset to be practical
for very large-scale production, the cost, size, and power consumption of any circuit
solution has to be significantly below what is being achieved today using compound
semiconductor technology. Traditionally, analog radio front-end ICs, especially the ones
operating at high frequencies, have been designed using III-V compound technologies
which have superior performance compared to SiGe BiCMOS and Si CMOS technologies
due to their higher electron mobility [14], [15].
Still, a CMOS implementation promises higher levels of integration and reduced cost.
ICs implemented using CMOS technology directly benefit from the higher speed limit of
the scaled technology [16] and with continued device scaling, enable circuit blocks to
operate at ever-increasing frequencies.
Typical telecommunication applications require not only transceiver and sensing
hardware, but also digital circuity to process the acquired data. Silicon CMOS technologies
are good candidates for the integration of such systems. These technologies are being
considered particularly attractive for their potential of integration with intermediate
frequency (IF) and baseband digital signal processing (DSP) functions, together with high
performance analog frontends, enabling true systems-on-chip (SoC).

Page | 2
1.3 Move Towards High-Frequency Band: Millimeter-Wave Band as a
Solution

Over the last few years, both data center and network traffic increased by several folds,
and are still experiencing an ongoing growth at an unsustainable rate [17]. This increasing
demand for higher capacity while meeting with the demand for high end-to-end quality
service projects a continuous need for improvements in spectral efficiency enhancement
techniques, particularly if the medium has to be shared with many users and interchanged
data.
However, a given spectrum band has a limit for the volume of data it can support.
Improving its throughput using bandwidth portioning and spectral aggregation techniques
has already been exhausted, leading to their evolvement arguably providing insignificant
enhancements to the recent spectrum bands since the latter have already reached their
saturation level due to being over-packed. This pushes us to search for other solutions and
one promising alternative is by arranging more bandwidth. The availability of uncrowded
spectrum in the millimeter band to exploit from makes it possible to accommodate future
significant increase in network traffic and producing multi-gigabit per second data accesses.
This motivated research attempts to overcome the limitation on high frequency of operation
through advancement in semiconductor technology process, such as technology scaling, for
instance, and in the continuous proposal of innovative analog circuits. For what concerns
the latter, it necessitates the development of advanced circuit topologies and suitable design
techniques capable of meeting with the marketplace requirements. Therefore, mm-wave
ICs is gaining a growing interest and receiving focus from recent activities by the research
and development (R&D) department of both private and public sectors.

1.4 Move Towards High-Frequency Band: Issues Encountered

Operating at millimeter-wave spectrum, however, is seeing continuous efforts in system


design to satisfy the demand for high-speed applications.
Generally speaking, synthesizing analog ICs in the high frequency spectrum subjects
one to performance degradation cause by reduced quality of passive components and
increase in parasitic that impact the highest operating frequency one could reach.
Technological-wise speaking, and in this particular case, the attractive low-cost CMOS
processes experience high sheet resistance, strict metal density rules and lossy silicon
substrate that makes it challenging to achieve high-end mm-wave CMOS circuits.
In terms of design process, designing at mm-wave renders the conventional analytical-
equation based design techniques difficult to be applied. The simplicity offered by them
starts to diminish since more behavior-dominant parasitic elements comes into role and
become difficult to neglect. Since simplicity is desirable in any design process,
approximation is adopted, inaccuracy in design parameter sizing increases and hence
dependency on electronic computer-aided design (CAD) softwares grows, such as Keysight
Advanced Design System (ADS) or Cadence Virtuoso software. Since the demand of
wireless and wireline communication is at high pace, the optimum design of any IC
component becomes major part.

Page | 3
1.5 Move Towards High-Frequency Band: CAutoD Technique as a
Solution

To meet the ever-growing demand of quality and competitiveness, iterative physical


prototyping is now often replaced by automated prototyping. The introduction of artificial
intelligence (AI) algorithms into the design process realizes an intelligent and automatic
design, and consequently considerable time and effort could be saved. Meanwhile,
designers will not have the concern that whether a specific topology can meet the design
specification or whether another parameters set exists that corresponds to better
performance.
However, before any code scripting can be performed or any proposal of algorithms and
design methodologies can be possible, developing a suitable physical model that can
properly simulate a given circuit behavior is considered an essential first step to this
process. This model is required to be complete and accurate, besides offering the designer
the possibility to have control over all design variables. Having control over all analog-
circuit components equates to having the option to propose several design methodologies
meeting multiple design objectives such as maximized output, highest speed, and cost-
effectiveness, and identifying multiple figures of merit. It also equates to obtaining a design
solution considered more of a global optimum solution than a local optimum solution. In
addition, an important metric when proposing a design model is simplicity. When speaking
of mm-wave IC design, especially, the less the designer has to deal with the plethora of tiny
parasitic effects, the easier and time-efficient his design task is considered.
Combining both circuit design-model and the widespread of advance scripting languages
will lead to a smart and efficient way of designing mm-wave ICs, which is commonly
referred to CAutoD process. Such process will give the designer the ability to virtually
explore a complete circuit before it is implemented for production.

1.6 Thesis Objectives

After demonstrating the big-picture of this thesis work, one can conclude that the theme
of this manuscript will be about IC amplifier design, implemented using low-cost CMOS
technology, with amplification bandwidth extending into the mm-wave frequency range. In
addition, it emphasizes on the need for a simple design process based on computer
automation, from proposing a versatile and complete physical model, and appropriate
methodology to achieve a desired performance criterion by the analog designer. Combining
both is required to provide optimum sizing-values for the circuit design parameters.
Among wide bandwidth amplifiers, our focus will be laid specifically on distributed
amplifiers with the following report organization.

Page | 4
1.7 Organization of the Report

The remaining of this manuscript is organized as follows.


In Chapter 2, a specific introduction on wide bandwidth amplifiers is provided. Here,
insights on the future trend towards operating at millimeter-wave are first discussed
followed by a more particular discussion on distributed amplifier device as a promising
solution to overcome the upper limit on frequency.
In Chapter 3, an overview of existing design techniques for distributed amplifiers and
their limitations for mm-wave usage are described. An original matrix-based design model,
based on four-port chain 𝐴𝐵𝐶𝐷-parameter, is then proposed as a reliable and accurate
model for this specific active circuit design to reach full CAutoD process.
In Chapter 4, a loss-compensated distributed amplifier, a topology for mitigating the
high losses issue of mm-wave frequencies, is described. Based on its concept and using the
proposed model of Chapter 3, an algorithmic design methodology is presented that
maximizes bandwidth for a given passband flatness. Its outcome is a set of 3D design space
exploration (DSE) plots that enables to explore a wide range of possible distributed
amplifier solutions.
In Chapter 5, from the resulting DSE plots of Chapter 4, a global optimum 100-GHz
bandwidth CMOS-based distributed amplifier is chosen and implemented for on-wafer
characterization.
In Chapter 6, a THz gain-bandwidth cascaded CMOS-based distributed amplifier, with
the same upper operating frequency of 100 GHz, is presented.
In the previous three chapters, the 55-nm CMOS process provided by
STMicroelectronics (ST) was adopted for both illustration, and design exploration and
implementations.
Finally, Chapter 7 provides a summary and general conclusion for the work presented
in this manuscript.

Page | 5
Page | 6
Chapter 2

Introduction to Wideband Amplifiers –


Distributed Amplifier for mm-Wave
Access

In this chapter, a particular discussion is devoted on amplifiers capable of extending their


operating frequency deep in the mm-wave band and proving distributed amplification as a
unique solution with ultra-wide bandwidth performance from a few gigahertz until
hundredth of gigahertz.
This chapter is structured as follows. Section 2.1 explains the motivation behind recent
trend followed by analog designers towards realizing ICs in mm-wave band. It also sheds
light on broadband amplification as an indispensable component of high demand in the near
future. Section 2.2 examines several amplifier circuit topologies proposed in the literature
to overcome the limitation on bandwidth, and demonstrates distributed amplification as a
promising candidate for ultra-wide bandwidth. Section 2.3 focuses on the history, principle
of operation, and technology versatility of distributed amplifiers. Here, a state-of-the-art
table in different technology processes is provided for performance tradeoff comparison.
Finally, Section 2.4 concludes this chapter and sets the tasks that will be covered in the
following chapters.

2.1 Future Trends for Millimeter-Wave Applications

The RF spectrum is a finite resource to be shared by all network devices in the world.
As years progress, the number of connected devices and data traffic is increasing at a
significant rate, as illustrated in Figure 2.1. Those statistical graphics are provided by Cisco
corporation [17] and demonstrate, via Figure 2.1(a), that by the year 2022 the overall
amount of connected devices will reach 12.3 billion units, a 43% growth from 8.6 billion
devices in 2017. In addition, via Figure 2.1(b), the overall mobile data traffic is expected to

Page | 7
grow to 77 exabytes (109 gigabytes) per month by 2022, a seven-fold increase over 2017.
Both statistics point to a no nearby deceleration in the amount of circulating information,
on the contrary, they indicate a continuous surge in telecommunication in the upcoming
years.

(a) (b)
Figure 2.1: Global (a) growth of connected devices and (b) mobile data traffic in exabyte (109
gigabytes) per month, from 2017 to 2022. [17]

Usually, the frequency spectrum is divided into several differently sized small-portions
in order to accommodate the multi RF-applications. A demonstrative example of such radio
spectrum partitioning, adopted by the United States [18], is illustrated in Figure 2.2.
Oversharing of the spectrum, however, can lead to unintentional interference. This can
happen between signals through overlapping of a given carrier sideband with upper and/or
lower sidebands of adjacent carriers, and interference that occurs when frequencies shift
slightly. In either case, they make it difficult or near impossible for receivers to pick out the
correct signal. This unintentional interference mostly occurs because generating and
treating an RF-signal is an imperfect process. Figure 2.2 can also serve as a visual proof for
demonstrating the congestion we reached in recent frequency spectrum due to oversharing.
Indeed, the common range spanning from few KHz to 30 GHz is over-populated and it is
becoming progressively difficult to open room for future devices and applications, and a
bottleneck for the support of the many attractive but bandwidth-demanding services of the
coming multimedia society.

Figure 2.2: Radio spectrum demonstrating frequency allocations adopted by the United States.
[18]

Page | 8
One common way used to alleviate this issue is by increasing the spectral efficiency,
unit of measurement defined by (bits⁄sec)⁄Hz. This can be achieved through signal
multiplexing techniques; for instance, through orthogonal frequency-division multiplexing
(OFDM). Such ways, however, are starting to reach saturation levels in recent years [19],
as shown in Figure 2.3, where capacity growth is slowing down and unable to keep up with
the continuous increase in services demand. Due to occurring congestion in the frequency
spectrum below 30 GHz, we reached a period in time where trying to push few bits of extra
information in just a single Hertz of frequency component is becoming significantly
challenging. Figure 2.3 also gives an early intuition that recent years advancement in radio
resources management techniques is slowing down in demonstrating a notable
improvement compared to the previous years; therefore, advocating the need to look for
other solutions for the years to come.

Figure 2.3: Comparing data capacity growth to demand growth. [19]

By referring back to Figure 2.2, one could perceive another promising solution and it is
exploring the mm-wave band, i.e., beyond 30 GHz, to arrange more bandwidth. As
formalized in the Shannon-Hartley theorem, enlarging bandwidths does indeed comply
with obtaining higher channel capacity. In theory, there is an unlimited amount of
unexploited bandwidth available in the 300 GHz frequency spectrum. It has motivated
many research contributions on mm-wave electronics for wireless communication in
general and broadband implementation in particular. Indeed, the sufficient broadband
spectrum that exists there has sparked interest in a variety of potential applications, such
short-range high data-rate wireless personal area networks (WPANs), wireless local area
networks (WLANs) with extraordinary capacity, and wireless uncompressed digital-video
transmission. This also allows for a variety of applications including gigabit/s point-to-
point links, wireless docking for portable devices, and vehicular radar. Millimeter-wave
systems find also place in the field of bio-medical devices. Today, imaging devices are
largely confined to health-centers and large hospitals. The dynamics of patient-health
provider interaction will radically change if these diagnostic tools are available by patient's
side rather than at long distances. Examples include large-array low-power ultrasonic
imagers and lab-on-chip devices for rapid detection of pathogens or sequencing of DNA
[20]-[22]. Diagnostic medical imaging is another application that can greatly benefit from
large capacity offered by such spectrum. Even in wireline communication, wireline
backbones, such as fiber optic communication systems, with data rates of over 100 Gb/s
per channel will be required in the near future and for those ultra-broadband systems with

Page | 9
a minimum required bandwidth of 50 GHz are key components in such systems [23].
Therefore, the 30 GHz-300 GHz spectrum spanning about 55 GHz of contiguous
bandwidth is identified as the most suitable candidate to accommodate the ever increase of
circulating data deduced. It is also appearing as a promising solution for the support of all
kinds of bandwidth-demanding services and applications that run at fiber speeds and
beyond.
In fact, significant efforts in the development of ICs with frequencies extending into the
mm-wave spectrum is already being observed. In parallel, silicon technology is being
considered as an attractive design tool for the cost-reduction and miniaturization offered by
it. Those efforts can be noticed from the available state-of-the-art research work [24]-[29],
[30]-[36], to give example of few, which highlights the trend being adopted in pushing the
limit of high frequency operation and in enlarging bandwidth performance. Through those
available researches, diverse circuit techniques are being proposed besides their
corresponding design theory and while offering the design tools and methodologies
necessary to replicate such performance characteristics.
In this thesis, efforts were concentrated on broadband mm-wave amplification, in
particular on a minimum of 50 GHz bandwidth to support future 100 Gb/s capacity and
higher. This is elaborated more in the next section (Section 2.2) where several amplifier
topologies, each with its own bandwidth limit, are listed and briefly described, where the
most promising one is a distributed structure.

2.2 Amplifier Circuit Topologies with Wide Bandwidth Characteristic

As telecommunication technology proceeds, expectations from an amplifier increase


with it. Today’s telecommunication systems require wideband, efficient and high gain
amplifiers. Depending on the application, requirements on an amplifier change. Obtaining
high-gain and wideband systems has been one of the main areas of focus in wireless and
wireline technologies. In RF and microwave amplifiers, transistors are employed as the
main component; however, physical constraints of transistors put an upper limit to the gain
and bandwidth of a practical amplifier. This fact is observed as gain-bandwidth product
(GBP) which stays constant for an amplifier design. For these reasons, special consideration
must be given to the problem of designing wide bandwidth amplifiers. Researchers have
come up with some topologies to overcome the limitation on bandwidth. Some of those
common approaches are listed below; note in each case that an improvement in BW is
achieved at the expense of complexity, i.e., more correlated design variables to take into
consideration, an added challenge when searching for optimum performance.

2.2.1 Small-to-Large Bandwidth Amplifiers Topologies


The topic of integrated amplifiers is not considered as a newborn research topic. For
that, the literature demonstrates plenty of circuit prototypes implemented in several process
nodes and each with different performance improvement objectives. In this section,
concentration is put solely on the ones with bandwidth improvement as objective. Those
topologies are listed starting by the one that offers the smallest bandwidth and ending by
the topology that offers the largest one. For each case, a single optimum design example is

Page | 10
given. A short description for each design is given accompanied by frequency behavior
examination through 𝑆-parameter analysis.

2.2.1.1 Single-Transistor Amplifier Topology [24]


We begin this listing by what could be arguably considered as the simplest topology one
could synthesize and it involves using a single transistor as Gm-cell. It is considered of
importance to demonstrate what one could obtain as frequency behavior when using just a
simple transistor biased in the active region in order to support the efforts towards more
complex topologies. It will also serve as a performance benchmark for the topologies
coming after. [24] demonstrates such amplifier through a prototype implemented in 90-nm
CMOS technology. Figure 2.4(a) and Figure 2.4(b) illustrate the circuit schematic and
microphotograph, respectively. This amplifier is synthesized as a three-stage amplifier
connected in series in order to increase its gain in the given bandwidth: a sign of weak
amplification provided by just a single transistor. From Figure 2.5(a), the fabricated
amplifier exhibits 10.8 dB of gain with 5 GHz of 3-dB bandwidth around 60 GHz, i.e.,
8.3% of fractional bandwidth (FBW). The total power consumption is only 5.5 mW from a
1-V power supply. From Figure 2.5(b), both input and output return losses, 𝑆11 and 𝑆22 ,
respectively, exhibit narrowband behavior with values better than -15 dB at 60-GHz center
of frequency.

(a) (b)
Figure 2.4: Single-transistor amplifier topology. (a) Schematic circuit diagram.
(b) Microphotograph. [24]

(a) (b)
Figure 2.5: Simulated and measured 𝑆-parameters versus frequency. (a) Insertion gain.
(b) Input and output return losses. [24]

Page | 11
2.2.1.2 Resistive-Feedback Amplifier Topology [25]
The first suggested topology for obtaining wide bandwidth amplification is through
resistive-feedback configuration [25]. Figure 2.6(a) illustrates the schematic view of this
topology. The resistive shunt-feedback provides wideband input matching. Its bandwidth
is dominated by the RC time constant at the input/output node. Figure 2.6(b) shows the
implemented microphotograph using 0.18-µm SiGe BiCMOS process. The measured 𝑆-
parameters are plotted in Figure 2.7. From Figure 2.7(a), a gain of 13 dB with 3-dB
bandwidth of 10 GHz is obtained, i.e., more than 100% of FBW. While FBW here does
seem interesting at first; however, one should note that as we increase in frequency, the
inherent parasitic of the transistor start to affect the shunt matching provided by the resistive
feedback. This makes it challenging to maintain wideband matching at mm-wave spectrum
and support the efforts for more complex topologies. The fabricated amplifier consumes
4 mA from a 2.4-V power supply. From Figure 2.7(b), the input and output return losses
are below -5 dB and -10 dB, respectively, within the declared passband. Through this
topology, bandwidth enlargement is possible compared to the single-transistor amplifier
topology [24] discussed at the beginning of this section.

(a) (b)
Figure 2.6: Resistive-feedback amplifier topology. (a) Schematic circuit diagram.
(b) Microphotograph.[25]

(a) (b)
Figure 2.7: Measured 𝑆-parameters versus frequency. (a) Insertion gain and reverse isolation.
(b) Input and output return losses. [25]

Page | 12
2.2.1.3 Cascode Amplifier Topology [26]
Since design interest is focused on the mm-wave frequency region, parasitic capacitance
becomes difficult to neglect. The most unfavorable capacitance that causes gain and
bandwidth degradation a single transistor experiences is the input-to-output overlap
parasitic, leading to a parasitic Miller capacitance. Arguably, the most widely used
technique to mitigate it is by using a conventional cascode topology, where two transistors
configuration is used at its simplest form. [26] demonstrates such cascode cell designed at
60 GHz and implemented with a 90-nm CMOS process. Its circuit topology and
microphotograph are shown in Figure 2.8(a) and Figure 2.8(b), respectively. Its frequency
response through 𝑆-parameters characterization is demonstrated in Figure 2.9. The
amplifier exhibits 14 dB gain with a 3-dB bandwidth of 23 GHz around 60 GHz at 32-mW
of total power consumption, i.e. 38.3% of FBW. The input and output return losses are
below -8 dB within the declared passband region. Through this topology, a near five-fold
fractional bandwidth enlargement is possible compared to the single-transistor amplifier
topology [24] discussed at the beginning of this section.

(a) (b)
Figure 2.8: Cascode amplifier topology. (a) Schematic circuit diagram. (b) Microphotograph. [26]

Figure 2.9: Simulated and measured 𝑆-parameters versus frequency. [26]

Page | 13
2.2.1.4 Balanced Amplifier Topology [27]
A common issue between microwave transistors is that they typically are not well
matched to 50 Ω (a common standard reference-impedance). In the case of previous
amplifiers topologies, matching networks where used. While matching network provides
good input and output matching over the desired frequency bandwidth; however, it gives
maximum gain only over a relatively narrow bandwidth. The balanced amplifier circuit
solves this problem by using two 90-degree couplers to cancel input and output reflections
from two identical amplifiers. [27] demonstrates such topology through a family of
balanced amplifiers developed to cover an operating range of 4-40 GHz and implemented
in a 0.1-µm pseudomorphic InGaAs/AlGaAs/GaAs HEMT production process. Schematic
circuit diagram of only one design is presented here in Figure 2.10(a) with its respective
microphotograph in Figure 2.10(b). The first 90° hybrid coupler divides the input signal
into two equal-amplitude components, with a 90◦ phase difference, which drive the two
amplifiers. The second coupler recombines the amplifier outputs. Because of the phasing
properties of the hybrid coupler, reflections from the amplifier inputs cancel at the input to
the hybrid, resulting in an improved impedance match; a similar effect occurs at the output
of the balanced amplifier. [27] targeted three designs with different frequency bands
including 4-9 GHz, 9-20 GHz, and 20-40 GHz, as illustrated in Figure 2.11. The two first
designs demonstrate the excellent qualities of balanced amplifiers in terms of matching
characteristics, while proving higher FBW than previously discussed designs, 77% and
76% for the 4-9 GH and 9-20 GHz cases, respectively, with a maximum of -15 dB of return
loss for the 4-9 GHz and -17 dB for the 9-20 GHz. The 20-40 GHz also proves an extended
bandwidth but return loss is not provided; nonetheless, its FBW extends up to 66.7%.
Through the two first topologies, a nine-fold fractional bandwidth enlargement is possible
compared to the single-transistor amplifier topology [24] discussed at the beginning of this
section. The third topology shows an eight-fold improvement.

(a) (b)
Figure 2.10: Balanced amplifier topology. (a) Schematic circuit diagram. (b) Microphotograph.
[27]

Page | 14
(a) (b)

(c)
Figure 2.11: Measured insertion gain responses of (a) 4-9 GHz, (b) 9-20 GHz and (c) 20-
40 GHz balance-topology amplifier. [27]

2.2.1.5 Transformer-Coupled Amplifier Topology [28]


Similar to the balanced amplifier concept, the transformer-coupled topology aims at
providing a wide bandwidth input- and output- matched impedances. Instead of using a 90-
degree coupler, the following topology proposes the usage of transformers for input and
output matching. Through them, they can simultaneously perform impedance
transformation and differential-to-single-ended conversion. [28] demonstrates such circuit
realized in a 0.13-µm SiGe BiCMOS process at 60 GHz with two coupled loop inductors
as transformers. Circuit schematic is shown in Figure 2.12(a) with its microphotograph
shown in Figure 2.12(b). Performance characterization of the amplifier against frequency
is illustrated in Figure 2.13. A 3-dB bandwidth of 56 GHz, from 34 to 90 GHz, i.e. 90% of
FBW, is obtained with gain of 12.25 dB while consuming a total 𝑃𝐷𝐶 of 270 mW. Matching
is poorer than 6 dB in the lower band from 34 to 50 GHz. By reducing requirements to a
reasonable matching better than 6 dB, the passband reduces within 50 to 90 GHz, i.e. still
57% of FBW that is almost seven-fold the one of the reference single transistor.
Nonetheless, by relaxing the constraints on matching characteristic and on the basis of
solely the stated amplification-passband 3-dB bandwidth (56 GHz), through this topology
almost eleven-fold bandwidth enlargement is possible compared to the single-transistor
amplifier topology [24] discussed at the beginning of this section.

Page | 15
(a) (b)
Figure 2.12: Transformer-coupled amplifier topology. (a) Schematic circuit diagram.
(b) Microphotograph. [28]

Figure 2.13: Measured 𝑆-parameters versus frequency. [28]

2.2.1.6 Distributed Amplifier Topology [29]


We end this listing by mentioning the topology that offers the widest bandwidth. Ultra-
wide bandwidth characteristic is possible with distributed amplifier topology. As a short
overview, with more in-depth details on its principle of operation provided in Section 2.3,
the DA achieves its large bandwidth trough connecting pair of inductive-type segments at
both the input and output ports of Gm-cell. Since the Gm-cell experiences input and output
parasitic capacitance at high frequencies, the latter combined with inductive-type segments
create what is called artificial transmission line (TL). Those artificial lines provide wide
bandwidth matching which makes broadband amplification possible. Meanwhile the high
gain of the DA results from the cumulative gains of those small Gm-cells. As a
demonstrative example, [29] illustrated the performance capability of a 100-GHz DA,
centered around 50 GHz from 4 to 104 GHz, i.e. 200% of FBW, implemented in a 45-nm
CMOS SOI process. The meaning of center operating frequency loses somewhat sense with
this architecture that can be seen as an accumulation of small sections of Gm-cells, each
one showing an extended cut-off frequency due to inductive-type segments. In spite of that,
its cut-off frequency can be adjusted through its design parameters, such as transistor size,

Page | 16
and thus justifying the possibility to mention here the 50 GHz center frequency.
Figure 2.14(a) and Figure 2.14(b) represent both its schematic diagram and fabricated
circuit, respectively. From Figure 2.15, the amplifier attains 16 dB of gain while being
powered from a 5-V supply. Through this topology, a twenty-two-fold bandwidth
enlargement is possible compared to the single-transistor amplifier topology [24] discussed
at the beginning of this section.

(a) (b)
Figure 2.14: Distributed amplifier topology. (a) Schematic circuit diagram. (b)
Microphotograph. [29]

Figure 2.15: Measured 𝑆-parameter versus frequency. [29]

2.2.2 Table-of-Comparison for Performances Analysis


The six amplifier topologies that were presented in subsection 2.2.1 are summarized in
Table 2.1. The purpose of this table is to review their performance tradeoffs and highlight,
as a matter of course, why, in the context of RF spectrum enlargement towards mm-waves,
the suitable topology is distributed amplification that will be the subject of further study in
this manuscript. Since overcoming the upper limit on the gain and bandwidth of a practical
amplifier is the motivation behind the proposal of such designs, the accent is put here on
Gain-Bandwidth product (GBP) criterion. Special consideration on minimum bandwidth of
50 GHz is taken into account as it is important for future 100 Gb/s and beyond applications,
as discussed previously in Section 2.1.

Page | 17
TABLE 2.1: PERFORMANCE COMPARISON BETWEEN WIDE BANDWIDTH AMPLIFIER
TOPOLOGIES.
Amplifier Gain Bandwidth 𝐏𝐃𝐂 Chip Area GBP
Ref.
Topology (dB) (GHz) (mW) (mm2) (GHz)
Single Transistor
[24] 10.8 5 5.5 0.39 17
(Cascaded)
[25] Resistive-feedback 13 10 9.6 0.88 45
[26] Cascode 14 23 32 0.22 115
20 5 187.5 12.16 50
[27] Balanced 17 11 187.5 12.16 78
22 20 187.5 8.6 252
Transformer-
[28] 12.25 56 270 0.15 229
Coupled
[29] Distributed 16 100 N/A 0.73 640

The first three topologies: single-transistor [24], resistive-feedback [25] and cascode
[26] amplifiers are the one which present relatively simple design process compared to the
remaining topologies, since they involve few components to deal with. They even show
attractive low 𝑃𝐷𝐶 suited for low power consumption applications. However, they do not
meet the required minimum bandwidth of 50 GHz and provide the lowest GBP merit, and
hence were disregarded. Balanced amplifier topology [27] could demonstrate possible
potentials in reaching this bandwidth requirement when pushed towards higher frequencies
and it also shows flexibility in its GBP where one can obtain different BWs using the same
topology. However, its design necessitates the implementation of two Gm-cells devices, as
illustrated in Figure 2.10(a), hence twice the DC power, but with an overall delivered gain
equal to that of a single Gm-cell device.
Both transformer-coupled amplifier [28] and distributed amplifier [29] topologies fulfill
the minimum bandwidth requirement. However, as explained in Section 2.1 and illustrated
by Figure 2.1, the number of connected devices and data traffic, in other words, data
network, is constantly increasing and with no sign of stopping in the future. At one moment
in the future, even larger bandwidths will be of need. Hence, a good research practice would
be to anticipate the future and be prepared for the system demand. For that reason,
distributed amplifier is chosen as topic of further discussion since it is the only topology
that presents potential ultra-wide bandwidth performance reaching deep into the millimeter-
wave spectrum. It also coincides with continuous advancement in technology nodes where
smaller nodes are reaching higher frequency limits. It is worth also adding from a practical
point of view, that the transformer-coupled amplifier presents inability to start amplifying
from very low frequencies, if needed, compared to distributed amplifier. A requirement for
fiber optic communications. This lack of ability to amplify starting from DC-level is
because in the former topology, the input signal is being fed through a transformer to the
input of the transistor, as shown in Figure 2.12(a), compared to being feed through a
transmission line segment as in the case of a DA, as shown in Figure 2.14(a).
As a conclusion and on those bases, the remaining of this chapter is dedicated into
describing the DA circuit topology and analyzing its basic principle of operation and
performances.

Page | 18
2.3 What is a Distributed Amplifier?

In Section 2.1, a serious issue of congestion in the lower side of the radio spectrum due
to over-sharing was identified. It highlighted the efforts taken by analog designers in
moving towards the unexplored millimeter band as a promising solution to respond to the
consumer marketplace for high-speed applications and possibility, by expanding the
allocation bandwidth, to accommodate further increase in data traffic. Several proposed
amplifier topologies for enlarging bandwidth were discussed in Section 2.2 and
demonstrated the trend adopted in continuously trying to push the upper frequency limit.
Since DA presented the potentials of achieving the largest bandwidth, suitable for today
and future needs, this section is dedicated to elaborate more on the subject. For a systematic
presentation, it is divided into four subsections. It starts by presenting the history and
interest of DA design since 1936. Then, it introduces its structure and principle of operation.
The intention here is to give the reader a first step towards understanding how a DA usually
functions and what controls its behavior. Several DAs implemented in different IC
technology processes are also provided that demonstrate its widespread popularity between
research community for performance exploration. Finally, a state-of-the-art table
summarizing all IC DAs in different technology processes is presented and discussed.

2.3.1 Walk Through History


The potential of distributed amplification for obtaining gains over wide frequency bands
has long been recognized. Historically, distributed amplifier is an ancient concept appearing
in the 20th century. It was first explored by British inventor William S. Percival with
vacuum tubes to compensate for the parasitic capacitance of electrodes. He later was the
first to deposit a single patent on this topic in 1936 [37], shown in Figure 2.16, but without
any exhaustive explanation or any trace of further discussion of this idea in the literature.

Figure 2.16: Thermionic valve circuit - William S. Percival’s 1936 patent. [37]

The work became renown with the publication of [38] in 1948, where both Ginzton,
Hewlett, Jasberg and Noe, based on the concept disclosed by William Percival, discovered
that by appropriate distribution of vacuum tubes along artificial transmission lines, it is
possible to obtain amplification over much greater bandwidths than would be possible with
ordinary circuits. Their objective at that time was to find a practical solution for extremely
broadband "video" amplifiers which have flat frequency response from low audio
frequencies to frequencies as high as several hundred Megahertz. Due to technology

Page | 19
limitation at their generation, practical limitations of electron tubes prevented them to reach
frequencies up to 1000 MHz. Nevertheless, [38] was considered the first published research
work to present the principle in wide-band amplifier design and hence the term "distributed
amplifier" was born. Figure. 2.17 illustrates the general principle of distributed
amplification using vacuum tubes as amplification elements presented by [38], which later
was adopted and still being adopted in recent research work.

Figure 2.17: First schematic diagram of distributed amplifier using vacuum tubes from 1948.
[38]

Implementation using solid-state devices followed with early demonstration presented


by [39] in 1969, where an attempt was made in implementing a distributed amplifier with
2 GHz bandwidth using MESFET technology.
In 1982, [40] was published introducing a new concept of treating the DA as a
continuous-structure model. It provided the required analytical approach to represent a DA
gain behavior with respect to frequency and the design tools necessary to size one. In [40],
the design concept was enhanced and made suitable for implementing amplifiers with
several gigahertz of bandwidth. They demonstrated their idea through a practical example
where they realized a DA that covers 1-13 GHz using 1-µm GaAs FET technology.
Even in the early 21st century, this concept is still being widely adopted between authors,
such as [30]-[36], while synthesizing DAs at multi-gigahertz frequencies.

2.3.2 Structure and Basic Principle of Operation


Figure 2.18 illustrates the configuration of a conventional distributed amplifier. The core
of a DA design consists of what is commonly referred to as a unit cell. A unit cell is
composed of a Gm-cell connected to a set of four inductive-type segments, depicted here
as having an inductance value of 𝐿𝐼 ⁄2 and 𝐿𝑂 ⁄2, on both its respective input and output
sides. A set of 𝑁 unit cells are connected together until a desired overall level of
amplification is reached within a given passband. Since every Gm-cell is loading each
inductive side by its own inherent output and input complex impedances, the overall upper
and lower sides of DA are referred to, in this manuscript, as output- and input- loaded lines,
respectively.

Page | 20
Figure 2.18: Conventional distributed amplifier made of 𝑁 unit cells.

An RF signal (𝑉𝑆 ) applied at the input end of the loaded input line travels towards the
right side, where it is absorbed by the terminating load impedance 𝑍4 at the opposite end.
As it travels down, each Gm-cell samples a portion of the signal at different phase and
amplitude, and then transfer it to the output line through its own transconductance. Due to
the non-ideality of those Gm-cells, a small portion of both signals is wasted through
dissipation caused by the cells resistive loads distributed along the input and output lines.
If the phase delay of the signal at the output-loaded line is identical to the phase delay
of the input-loaded line, then the signals on the former line add constructively. The addition
will be in phase only for the forward-traveling signal. There is also a backward traveling
wave component on the output line, but the individual contributions to this wave will not
be in phase, and therefore they at least partially cancel. Any signal which travels backward
and is not completely cancelled by the out-of-phase additions, will be absorbed by the
termination output impedance, represented by 𝑍3 in Figure 2.18.
The basic principle behind distributed amplification is the absorption of the bandwidth
limiting parasitic capacitances of Gm-cell into artificial transmission lines, which allows
high bandwidths. In this approach, the added inductive-type segments combined with the
parasitic capacitances are forming artificial TLs on each sides of the DA. The artificial lines
will result in high cut-off frequencies, which will determine the new gain-bandwidth
behavior of the desired amplifier. For a more in-depth explanation of this statement, the
next discussion is dedicated to presenting the mathematical aspect behind the DA concept.
However, beforehand, for a clearer and more realistic analysis, the following
modifications are carried out. First, the DA is split into two halves, as shown in Figure 2.19.
The lower side represents the input loaded line and the upper side represents the output-
loaded line. Second, since by the end of this manuscript a FET-based DA will be
implemented as a circuit prototype for on-wafer characterization, it is considered suitable
to adopt here, in the theoretical part, a FET-based Gm-cell and represent its behavior by its
AC-equivalent lumped model. A unilateral assumption is used for its electrical model for
the purpose of simplifying the derivation for a better interpretation of the concept. Finally,
since in practice the ideal lumped inductors, 𝐿𝐼 ⁄2 and 𝐿𝑂 ⁄2, are typically implemented
each as physical microstrip, they are replaced here by microstrip transmission (µ-TL)
segments of length 𝑙𝑔 ⁄2 and 𝑙𝑑 ⁄2, respectively.

Page | 21
Figure 2.19: Simplified equivalent-circuit representation of a FET-based distributed amplifier.

Referring to Figure. 2.19, both 𝑅𝑔𝑠 , 𝐶𝑔𝑠 , 𝑅𝑑𝑠 , 𝐶𝑑𝑠 and 𝐼𝐷 are the small-signal intrinsic
elements of the FET transistor, and 𝛿𝑅, 𝛿𝐿, 𝛿𝐶 and 𝛿𝐺 are the per-unit-length series
resistance (Ω/m), series inductance (H/m), shunt capacitance (C/m) and shunt conductance
(S/m) of the lumped-element equivalent circuit of the µ-TL segment.
The combination of the distributed FET parasitic with the segments creates the artificial
transmission lines. The input (gate) and output (drain) loaded-lines are represented by their
effective series inductance-per-unit-length 𝛿𝐿𝑔 and 𝛿𝐿𝑑 , respectively, of their own
segments. Their effective shunt capacitance-per-unit-length is a result of a merge between
the parasitic capacitances from the FETs and segments, i.e., (𝛿𝐶𝑔 + 𝐶𝑔𝑠 ⁄𝑙𝑔 ) and (𝛿𝐶𝑑 +
𝐶𝑑𝑠 ⁄𝑙𝑑 ), respectively. What remains as resistive lumped elements act as per-unit-length
series and shunt losses for those equivalent transmission lines.
The previous assumption of equivalent lines is only valid when the spacing between the
FETs, i.e., 𝑙𝑔 and 𝑙𝑑 , is very small compared to the wavelength of operation (2.1).
𝑐
𝑙𝑔 | 𝑎𝑛𝑑 𝑙𝑑 |𝑚𝑎𝑥 << 𝜆𝑔𝑢𝑖𝑑𝑒𝑑 =
𝑚𝑎𝑥
√𝜀𝑒𝑓𝑓 ∙ 𝑓𝑜𝑝𝑒𝑟𝑎𝑡𝑖𝑜𝑛 (2.1)

If this condition is satisfied, both lower and upper lines can be approximated each as
continuous structure, referred to as artificial gate and drain lines, since FETs are used. Their
characteristic impedances and propagation constants are deduced to be (2.2), (2.3) and
(2.4), (2.5), respectively, [56].

𝛿𝐿𝑔
𝑍𝑐𝑔 ≈
√ 𝐶
𝛿𝐶𝑔 + 𝑔𝑠⁄𝑙 (2.2)
𝑔

𝛿𝐿𝑑
𝑍𝑐𝑑 ≈ √
𝐶 (2.3)
𝛿𝐶𝑑 + 𝑑𝑠⁄𝑙
𝑑

Page | 22
𝛿𝑅𝑔 𝛿𝐶𝑔 𝜔2 𝑅𝑔𝑠 𝐶𝑔𝑠
2
𝑍𝑐𝑔
𝛾𝑔 = 𝛼𝑔 + 𝑗𝛽𝑔 ≈ ( √ + )
2 𝛿𝐿𝑔 2𝑙𝑔
(2.4)
𝐶
+ 𝑗𝜔√𝛿𝐿𝑔 (𝛿𝐶𝑔 + 𝑔𝑠⁄𝑙 )
𝑔

𝛾𝑑 = 𝛼𝑑 + 𝑗𝛽𝑑
𝛿𝑅𝑑 𝛿𝐶𝑑 𝑍𝑐𝑑
≈( √ + )
2 𝛿𝐿𝑑 2𝑅𝑑𝑠 𝑙𝑑
(2.5)
𝐶𝑑𝑠
+ 𝑗𝜔√𝛿𝐶𝑑 (𝛿𝐶𝑑 + ⁄𝑙 )
𝑑

By preserving what was declared as initial condition in (2.1) for short length segments
between Gm-cells, the gain expression for an 𝑁-cell DA circuit can be derived as follows.
For an incident source voltage, 𝑉𝑆 , the voltage on the gate-to-source capacitance of the
th
n FET, 𝑉𝑐𝑛 , can be written as (2.6) for a phase reference from the first transistor.

1
𝑉𝑐𝑛 = 𝑉𝑆 𝑒 −(𝑛−1)𝛾𝑔 𝑙𝑔 ( )
1 + 𝑗𝜔𝑅𝑔𝑠 𝐶𝑔𝑠 (2.6)

The factor in parentheses in (2.6) accounts for the voltage division between 𝑅𝑔𝑠 and 𝐶𝑔𝑠 .
For small node FET parameters, (ω𝑅𝑔𝑠 𝐶𝑔𝑠 ) << 1; therefore, this factor can be
approximated as unity over the passband of the amplifier.
The output current on the drain line can be found by recognizing that each voltage-
controlled current-source contributes to waves of the form (−1⁄2)𝐼𝐷𝑛 𝑒 ±𝛾𝑑 𝑥 in each
direction when perfect matching is assumed from both sides of each node. Since the
transconductance of a given Gm-cell is given, by definition, as 𝐼𝐷𝑛 = 𝑔𝑚𝑛 𝑉𝐶𝑛 , the total
output current at the 𝑁 th terminal of the drain line is found to be (2.7), [56].
𝑁
1
𝐼𝑜 = − ∑ 𝐼𝑑𝑛 𝑒 −(𝑁−𝑛)𝛾𝑑 𝑙𝑑
2
𝑛=1
𝑁
1 (2.7)
= − 𝑔𝑚 𝑉𝑆 𝑒 −𝑁𝛾𝑑 𝑙𝑑 𝑒 +𝛾𝑔𝑙𝑔 ∑ 𝑒 −𝑛(𝛾𝑔𝑙𝑔 −𝛾𝑑 𝑙𝑑 )
2
𝑛=1

Using the geometric series summation formula given by (2.8), allows (2.7) to be
simplified as (2.9).
𝑁
𝑥 𝑁+1 − 𝑥
∑ 𝑥𝑛 =
𝑥−1 (2.8)
𝑛=1

Page | 23
𝑔𝑚 𝑉𝑆 𝑒 𝛾𝑑 𝑙𝑑 (𝑒 −𝑁𝛾𝑔 𝑙𝑔 − 𝑒 −𝑁𝛾𝑑 𝑙𝑑 )
𝐼𝑜 = −
2 𝑒 −(𝛾𝑔𝑙𝑔−𝛾𝑑 𝑙𝑑 ) − 1
𝑔𝑚 𝑉𝑆 𝑒 −𝑁𝛾𝑔𝑙𝑔 − 𝑒 −𝑁𝛾𝑑 𝑙𝑑 (2.9)
=−
2 𝑒 −𝛾𝑔𝑙𝑔 − 𝑒 −𝛾𝑑 𝑙𝑑

For a matched input and output ports, i.e., 𝑍𝑐𝑔 = 𝑍𝑖𝑛 = Z4 and 𝑍𝑐𝑑 = 𝑍𝑜𝑢𝑡 = Z3 , the
amplifier gain can be calculated as (2.10), [56].
1
𝑃𝑜𝑢𝑡 |𝐼𝑜 |2 𝑍𝑐𝑑
𝐺𝑎𝑖𝑛 = = 2
𝑃𝑖𝑛 1
|𝑉 |2 ⁄
2 𝑆 𝑍𝑐𝑔 (2.10)
2 2
𝑔𝑚 𝑍𝑐𝑔 𝑍𝑐𝑑 𝑒 −𝑁𝛾𝑔 𝑙𝑔 − 𝑒 −𝑁𝛾𝑑 𝑙𝑑
= | −𝛾𝑔𝑙𝑔 |
4 𝑒 − 𝑒 −𝛾𝑑 𝑙𝑑

As mentioned previously, if the phase velocity of the signal at the drain line is identical
to the one of the gate line, then the output signals on the drain line add constructively. The
addition will be in-phase only for the forward traveling signal. Referring to equation (2.7),
the terms in the summation will add only when 𝛽𝑔 𝑙𝑔 = 𝛽𝑑 𝑙𝑑 , i.e., the phase delays on the
gate and drain lines are synchronized.
Applying the synchronization condition of 𝛽𝑔 𝑙𝑔 = 𝛽𝑑 𝑙𝑑 allows equation (2.10) to be
further simplified to (2.11), [56].
2 2
𝑔𝑚 𝑍𝑐𝑔 𝑍𝑐𝑑 (𝑒 −𝑁𝛼𝑔𝑙𝑔 − 𝑒 −𝑁𝛼𝑑 𝑙𝑑 )
𝐺𝑎𝑖𝑛 =
4 (𝑒 −𝛼𝑔𝑙𝑔 − 𝑒 −𝛼𝑑 𝑙𝑑 )
2 (2.11)

For an ideal case where the DA losses are neglected, the gain equation (2.11) can be
reduced to (2.12)
2
𝑔𝑚 𝑍𝑐𝑔 𝑍𝑐𝑑 𝑁 2
𝐺𝑎𝑖𝑛 = (2.12)
4

Several interesting aspects of the distributed amplifier can be deduced from the gain
expression of (2.11) and (2.12), as follows. From Equation (2.11), for a lossless distributed
amplifier, the gain increase as 𝑁 2 . When loss is present, Equation (2.11) specify that gain
approaches zero as 𝑁 tends to infinity. This behavior is explained by the fact that the input
voltage on the input-loaded line decays exponentially. The same applies for the output-
loaded line. Thus, the FET at the end of the amplifier receive no input signal.
Those imply that, for a given set of transistor parameters, there will be an optimum value
of 𝑁 that maximizes the gain of the amplifier. This can be found by differentiating (2.11)
with respect to 𝑁 and setting the result to zero to obtain (2.13).

𝑙𝑛(𝛼𝑔 𝑙𝑔 ⁄𝛼𝑑 𝑙𝑑 )
𝑁𝑜𝑝𝑡 =
𝛼𝑔 𝑙𝑔 − 𝛼𝑑 𝑙𝑑 (2.13)

Page | 24
This result (2.13) depends on the frequency, the Gm-cell parameters and the line lengths
through the attenuation constants given in (2.4) and (2.5).

2.3.3 Performance Demonstration through Different Technologies and


Topologies
Distributed amplification topic received popularity between amplifier-research
community as being a promising candidate for breaking restrictions on upper frequency
limit. Owing to that, a noticeable amount of research work was made public demonstrating
the versatility of DA in different IC technology processes from Si, SiGe and compound III-
V technologies, such as InP and InGaAs. The common objective between those works, as
was also stated by the authors, was presenting the possible performances from gain,
bandwidth and 𝑃𝐷𝐶 one is capable of reaching through each technology. In this section, a
handful of design examples that illustrates optimum performance in each technology is
provided. Those designs are sorted out on the basis of their GBP achievements before their
electrical performance are finally summarized in a state-of-the-art table for tradeoffs
comparison in subsection 2.3.4. Only single stage DAs are duly presented herein for state-
of-the-art relevance, excluding any cascaded DA configurations where achievable gains are
inherently higher. Nonetheless, the later are introduced in Table 2.2 and their interest is
justified with the discussion that follows it.

2.3.3.1 A 4-91 GHz Bandwidth Distributed Amplifier in 90-nm Si CMOS [30]


This article presents the design of a CMOS-based DA implemented in silicon 90-nm
process node. This technology is benchmarked at 𝑓𝑡 = 160 GHz and 𝑓𝑚𝑎𝑥 = 142 GHz. The
DA is synthesized with six unit-cells. Figure 2.20(a) illustrates its circuit diagram and
Figure 2.20(b) its microphotograph for on-wafer characterization. The fabricated circuit
occupies a chip area of 0.6 ⨉ 1.2 mm2 including pads. The conventional cascode topology
with a two-transistor configuration was adopted for the Gm-cell and coplanar waveguide
(CPW) was used for realizing inductive-type segments instead of conventional microstrip
lines. Authors explain that with their back-end technology, they could guarantee lower
losses at high frequency using CPW topology.

(a) (b)
Figure 2.20: Six unit-cells 90-nm Si CMOS-based distributed amplifier. (a) Schematic
diagram. (b) Microphotograph. [30]

Page | 25
Referring to Figure 2.21, the measured 𝑆-parameters performance against frequency is
shown. Their six-unit cell DA achieves a peak gain of 7.4 dB with a 3-dB bandwidth of
80 GHz. This results in a GBP of 190 GHz. The amplifier consumes 120 mW from a 2.4-
V power supply. The measured output and input return losses across the 3-dB bandwidth
reaches a maximum of -10 dB. The reverse isolation is better than -20 dB for the whole
passband region.

Figure 2.21: Measured 𝑆-parameters of the 90-nm Si CMOS-based distributed amplifier. [30]

2.3.3.2 A 92-GHz Bandwidth Distributed Amplifier in 45-nm SOI CMOS [31]


This article presents the design of a CMOS-based DA implemented in silicon-on-
insulator (SOI) 45-nm process node. This technology is benchmarked at 𝑓𝑡 = 350 GHz and
𝑓𝑚𝑎𝑥 = 300 GHz. The DA is synthesized with three unit cells. Figure 2.22(a) shows the
schematic diagram of the device and Figure 2.22(b) shows its microphotograph for on-
wafer characterization. The fabricated circuit occupies a chip area of 0.5 ⨉ 0.8 mm2
including pads. The Gm-cell adopted is based on a cascode topology with a two-transistor
configuration and the inductive-type segments are conventional microstrip lines.

(a) (b)
Figure 2.22: Three unit-cells 45-nm SOI-based distributed amplifier. (a) Schematic diagram.
(b) Microphotograph. [31]

Referring to Figure 2.23, the measured S-parameters against frequency are plotted along
with the simulated results. The amplifier achieves a 3-dB bandwidth of 92 GHz and a gain
of 9 dB with a gain-ripple of ±1.5 dB inside the passband region. This offers an overall
gain-bandwidth of 259 GHz. The amplifier consumes 73.5 mW from a 1.2 V supply. The
measured input return loss across the entire 3-dB bandwidth is better than -10 dB, but
remains better than -15 dB below 70 GHz. The measured output return losses across the 3-

Page | 26
dB bandwidth is less than -6 dB but remains better than -10 dB below 84 GHz. The reverse
isolation is -15 dB at 88 GHz and better than -20 dB below 70 GHz.

Figure 2.23: Measured (solid lines) and simulated (dotted lines) 𝑆-parameters of the 90-nm Si
CMOS-based distributed amplifier. [31]

2.3.3.3 A 110-GHz Bandwidth Distributed Amplifier in 22-nm FD-SOI CMOS [32]


This article presents the design of a CMOS-based DA implemented in fully-depleted
silicon-on-insulator (FD-SOI) 22-nm process node. This technology is benchmarked at
𝑓𝑚𝑎𝑥 = 396 GHz (authors did not provide value for 𝑓𝑡 ). The DA is synthesized using eleven
unit cells. Figure 2.24(a) shows the schematic diagram of the device and Figure 2.24(b)
shows its microphotograph for on-wafer characterization. The chip occupies an area of
0.38 ⨉ 1 mm2 including the pads. The Gm-cell adopted is based on a cascode topology with
a two-transistor configuration and the inductive-type segments are conventional microstrip
lines.

(a)

(b)
Figure 2.24: Eleven unit-cells 22-nm FD-SOI-based distributed amplifier. (a) Schematic
diagram. (b) Microphotograph. [32]

Page | 27
Referring to Figure 2.25, the measured S-parameters against frequency are plotted along
with the simulated results. The amplifier achieves a 3-dB bandwidth of 110 GHz and a gain
of 8.5 dB. This offers an overall gain-bandwidth of 292 GHz. The amplifier consumes
80 mW from a 2-V supply. The magnitude of the input and output return losses are both
better than -15 dB for the whole amplifier bandwidth.

Figure 2.25: Simulated and measured 𝑆-parameters of 22-nm FD-SOI-based distributed


amplifier. [32]

2.3.3.4 A 112-GHz, 157-GHz, and 180-GHz Bandwidth Distributed Amplifier in 100-


nm InP HEMT [33]
This article presents the design of a high-electron-mobility transistor (HEMT)-based DA
implemented in 100-nm gate length InGaAs/InAlAs technology. This technology is
benchmarked at 𝑓𝑡 = 160 GHz and 𝑓𝑚𝑎𝑥 = 300 GHz. It demonstrates the capability of such
technology by synthesizing three DAs with different bandwidths: A 1-112 GHz, 1-
157 GHz, and a third amplifier with 5-dB gain and a 180-GHz high-frequency cutoff.
Figure 2.26(a) shows the schematic circuit diagram of the DA. The cascode topology with
a two-transistor configuration and conventional µ-TL line for the inductive segments were
adopted as unit-cell Gm-cell. Figure 2.26(b) shows the microphotograph of the fabricated
chip of 1-157 GHz DA. It consists of eleven unit cells and the die size is about 1 ⨉ 2.2 mm2
including pads.

(a) (b)
Figure 2.26: Eleven unit-cells 100-nm InP-based distributed amplifier. (a) Schematic diagram.
(b) Microphotograph of the 1-157-GHz version. [33]

Referring to Figure 2.27, measured insertion gain (𝑆21 ) for all three DAs against
frequency are plotted. From Figure 2.27(a), the 1-112 GHz and 1-157 GHz DAs provide

Page | 28
7 dB and 5 dB of gain, respectively. The third DA, with 180 GHz cutoff frequency provides
5 dB of gain as shown in Figure 2.27(b). Those offer an overall gain-bandwidth of
248 GHz, 277 GHz and 320 GHz, respectively. The authors did not provide data for power
consumptions.

(a) (b)
Figure 2.27: Measured insertion gain of (a) 1-112 GHz and 1-157 GHz, and (b) 180 GHz 100-
nm InP-based distributed amplifiers.[33]

2.3.3.5 A 105-GHz Bandwidth Distributed Amplifier in 50-nm InGaAs mHEMT [34]


This article presents the design of a metamorphic HEMT-based DA implemented in 50-
nm InGaAs technology. This technology is benchmarked at 𝑓𝑡 = 380 GHz and 𝑓𝑚𝑎𝑥 =
500 GHz. The DA fabricated consists of eight unit cells. Figure 2.28(a) illustrates the
schematic circuit diagram. The Gm-cell adopted was based on cascode topology with a
two-transistor configuration and conventional µ-TL line was used for the inductive-type
segments. Figure 2.28(b) illustrates the microphotograph of the implemented DA. The die
size is 0.75 ⨉ 2.25 mm2.

(a) (b)
Figure 2.28: Eight unit-cells 50-nm InGaAs-based distributed amplifier. (a) Schematic
diagram. (b) Microphotograph. [34]

Referring to Figure 2.29, the measured S-parameters against frequency are plotted. The
DA covers a 3-dB bandwidth of 110 GHz with 11 dB of gain while consuming a total of
450 mW from a 6-V supply voltage. This leads to a GBP of 373 GHz. The passband region
experiences a gain-ripple of around ±1 dB from 5 GHz to 110 GHz. The gain peak, in the
frequency range from 2 GHz to 5 GHz, is caused by the inductive feed of the DC bias
voltages used for on-wafer measurements. The input return loss is below -10 dB from 10

Page | 29
GHz up to 110 GHz with a maximum of -5 dB for frequencies below 10 GHz. The output
return loss is below - 7.5 dB from 5 GHz up to 100 GHz.

Figure 2.29: Measured 𝑆-parameters of 50-nm InGaAs-based distributed amplifier. [34]

2.3.3.6 A 170 GHz Bandwidth Distributed Amplifier in 130-nm SiGe HBT [35]
This article presents the design of a HBT-based DA implemented in 130-nm SiGe-
BiCMOS technology. This technology is benchmarked at 𝑓𝑡 = 300 GHz and 𝑓𝑚𝑎𝑥 =
500 GHz. The DA fabricated consists of five unit cells and the Gm-cell adopted is based
on cascode topology with a three-transistor configuration to increase voltage swing and
hence output power. Slow-wave µ-TL line topology was used for the inductive-type
segments to reduce lines length. The cascode is designed to compensate the transmission-
line-losses at high frequencies in order to extend the bandwidth as well as the GBP.
Figure 2.30(a) and Figure 2.30(b) illustrate the schematic circuit diagram and the
microphotograph of the implemented DA. The chip occupies an area of 0.85 ⨉ 0.45 mm2
including pads.

(a) (b)
Figure 2.30: Five unit-cells 130-nm SiGe-based distributed amplifier. (a) Schematic diagram.
(b) Microphotograph. [35]

Referring to Figure 2.31, the simulated and measured 𝑆-parameters against frequency
are plotted. The amplifier provides a gain of 10 dB up to a 3-dB bandwidth of 170 GHz. It
consumes a total of 108 mW FoM a 3.6-V supply. This leads to a GBP of 537 GHz. The
passband region experiences a gain-ripple of ±2 dB. The input return loss is below -20 dB

Page | 30
up to 100 GHz and then it increases with frequency reaching a maximum of -5 dB at
170 GHz. The output return loss is below -20 dB up to a higher frequency of around
140 GHz and then it starts to increase reaching a maximum of -5 dB at 170 GHz.

Figure 2.31: Simulated and measured 𝑆-parameters of 130-nm SiGe-based distributed


amplifier. [35]

2.3.3.7 A 40-222 GHz Bandwidth Distributed Amplifier in 250-nm InP HBT [36]
This article presents the design of an HBT-based DA implemented in 250-nm indium
phosphide (InP) technology. This technology is benchmarked at 𝑓𝑡 ≥ 375 GHz and 𝑓𝑚𝑎𝑥 ≥
650 GHz. The fabricated DA consists of four unit cells and the Gm-cell adopted is based
on cascode topology with a two-transistor configuration with conventional µ-TL topology
for the inductive segments. Figure 2.32(a) and Figure 2.32(b) illustrate the schematic
circuit diagram and the microphotograph of the implemented DA. The chip occupies an
area of 0.546 ⨉ 0.602 mm2 including pads.

(a) (b)
Figure 2.32: Four unit-cells 250-nm InP-based distributed amplifier (a) schematic diagram and
(b) microphotograph. [36]

Referring to Figure 2.33, the measured 𝑆-parameters against frequency with simulation
results superimposed are plotted. The amplifier provides a nominal gain of 10 dB with 3-
dB bandwidth extending from 40 GHz to 222 GHz, resulting in the highest 3-dB bandwidth
of 182 GHz reported thus far. It should be reminded that bandwidth is considered here and
not upper operating cutoff-frequency. The passband region experiences a gain-ripple of ±2
dB. The circuit consumes a total of 105 mW FoM a 4-V supply. This leads to a GBP of 575

Page | 31
GHz. The input return loss is below -10 dB from 40 GHz up to 22 GHz. The output return
loss is below -10 dB from 90 GHz to 210 GHz. The authors state that the degradation in
output return loss below 90 GHz is attributed to the on-chip dc bias network.

Figure 2.33: Simulated and measured 𝑆-parameters of 250-nm InP-based distributed amplifier.
[36]

2.3.4 State-of-the-Art Table on Distributed Amplifiers


Previously described distributed amplifiers, with extra ones reported for their relevance
with this thesis work, are listed in the state-of-the-art table, Table 2.2. The purpose of this
table is to summarize the design tradeoffs between integrated technologies when addressing
the subject of UWB amplification using DAs. In addition, it will serve us as a reference for
later benchmarking when characterizing our own fabricated circuit prototypes.
It is true that the GBP is an appropriate figure-of-merit (FoM) to evaluate the level of
improvement offered by a given proposal when pushing the limitation on gain and
bandwidth, as demonstrated in Section 2.2; however, the amount of delivered gain per
milliwatt of consumption is also an important merit to consider for low-power applications.
To account for this design tradeoff and ease of comparison, we define the following new
figure of merit (2.14):
𝐺𝐵𝑃 𝐺𝑎𝑖𝑛 · 𝐵𝑎𝑛𝑑𝑤𝑖𝑑𝑡ℎ
𝐹𝑜𝑀1 = = (𝐺𝐻𝑧⁄𝑚𝑊 )
𝑃𝐷𝐶 𝑃𝐷𝐶 (2.14)

Surface area could also be of particular interest for some analog designers and
distributed amplifier are not far from reserving relatively larger on-chip area compared to
other amplifier topologies as an expense for its wider bandwidth. To highlight the area
efficiency, we also define the following figure of merit (2.15):
𝐺𝐵𝑃
𝐹𝑜𝑀2 =
𝑃𝐷𝐶 · 𝐴𝑟𝑒𝑎
𝐺𝑎𝑖𝑛 · 𝐵𝑎𝑛𝑑𝑤𝑖𝑑𝑡ℎ (2.15)
= (𝐺𝐻𝑧⁄(𝑚𝑊 · 𝑚𝑚2 ))
𝑃𝐷𝐶 · 𝐴𝑟𝑒𝑎

In this thesis, our emphasis is placed on (2.14). We believe that area reserved by an
amplifier, or any integrated circuit in a general context, is highly dependent on the process

Page | 32
technology node, where smaller nodes tend to be more efficient in occupied area than larger
nodes. In addition, it depends on the design rule restrictions that accompanies a given
process, where nodes having stringent density rules can lead to large area than less
restrictive nodes. We also believe that the amount of surface area occupied is ultimately
influenced by the end-objective of the designer circuit prototype; it depends on what design
characteristic the designer opted his efforts on and providing improving for the research
community.
Table 2.2 compares results from state-of-the-art FET- and HBT- based DAs. HBT-based
DAs, [35], [36], [48], [49], have been reported to provide larger bandwidths, beyond
100 GHz, thereby contributing to a much higher GBP merit. This is due to HBTs impressive
cutoff frequencies [14], [15] compared to their cheaper CMOS counterparts. In addition,
due to the bipolar transistor capability in providing much higher transconductance, bipolar-
based DAs provide much higher gain for relatively similar 𝑃𝐷𝐶 , in the hundreds milliwatt,
to the other technology processes. This can explain their higher 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit. In terms
of surface area reservation, they also present higher 𝐺𝐵𝑃⁄(𝑃𝐷𝐶 · 𝐴𝑟𝑒𝑎) merit. This is also
due to its advantageous transconductance. For instance, for a similar gain of around 10 dB,
[35] implemented a five unit-cells HBT-based DA while [44] implemented a nine unit-cells
CMOS-based DA.
Analyzing the design tradeoffs between FET-based DAs, III-V technologies [33], [34],
[46], [47] offer superior GBP merit than silicon semiconductor technologies. This is due to
their high electron mobility, but at the expense of much larger 𝑃𝐷𝐶 , which is why they
present low 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit and, hence, could be considered as not appropriate for
applications where low power consumption is critical.
Analyzing the performances of the CMOS-based DAs, using the cheap silicon
technology, such as in [30], [31], [44], [45], they do offer comparable gain with bandwidths
reaching above 70 GHz. In order to boost their relative low gain and obtain much higher
𝐺𝐵𝑃⁄𝑃𝐷𝐶 performance, [41] and [42] demonstrate the technique of cascading multiple
single-stage DAs. This technique helps in rising the GBP merit; for example, for the same
90-nm process node, [42] demonstrates an excess of two-fold improvement in their
cascaded DA circuit compared to[30] that implemented just a single-stage DA circuit.
A final observation worth highlighting is that all of the mentioned state-of-the-art DAs
here have adopted a cascode topology for their Gm-cell when synthesizing their unit cell.
This is even perceptible through the schematic diagrams of each research work discussed
in previous subsection 2.3.3, where the most common configuration is a cascode made of
two transistors. This could advocate the easiness that follows from implementing a cascode
topology where the purpose of adding the upper transistor is to mitigate the Miller effect
that results from the input-to-output overlap capacitance of the lower transistor. It also
advocates for its capability in being a suitable solution to provide gain for mm-wave signals.
For this reason, cascode Gm-cell topology will be adopted later when implementing our
proper DA circuits, which also would give a fair comparison to the already published state-
of-the-art ones.

Page | 33
TABLE 2.2: STATE-OF-THE-ART PERFORMANCES COMPARISON FOR DISTRIBUTED AMPLIFIERS
𝑮𝑩𝑷 GBP
Total
Process Gain Ripple£ BW& Supply PDC GBP 𝑷𝑫𝑪 PDC · Area
Ref. Area
Techno. (dB) (dB) (GHz) (V) (mW) (GHz) 𝑮𝑯𝒛 GHz
( ) (mm2) ( )
𝒎𝑾 mW · mm2
90-nm
[30] Si 7.4 -3% 80 2.4* 120 188 1.57 0.72 2.18
CMOS
90-nm
[41] Si 7@ +9%,-3 70 N/A 122 157 1.29 0.72 1.8
CMOS
90-nm
[42] digital 14@ ±1.5% 73.5 1.2* 84 368 4.38 1.72 2.54
CMOS
40-nm
(1,
[43] digital 15$ +5%,-3 80 90 450 5 0.31 16.13
1.7)**
CMOS
120-nm
[44] SOI 11 ±1.2 85 2.5* 210 302 1.44 1.28 1.13
CMOS
120-nm
[45] SOI 7.8 ±1.3 82 2.6* 130 201 1.55 1.05 1.47
CMOS
45-nm
[31] SOI 9 ±1.5% 92 1.2* 73.5 259 3.52 0.45 7.82
CMOS
22-nm
[32] FD-SOI 8.5 -3% 110 2* 80 292 3.66 0.38 9.63
CMOS
130-nm
[46] InP 14.5# ±0.7 94 3* N/A 499 N/A 2.75 N/A
HEMT
100-nm
[33] InP 7 +2.5% 111 N/A N/A 248 N/A 2.2 N/A
HEMT
70-nm
[47] InP 21@ ±3% 80 2** 300 898 3 2.47 1.21
HEMT
50-nm
[34] InGaAs 11 ±1 105 6* 450 373 0.83 1.69 0.5
mHEMT
250-nm
[48] InP 12.8 +7%,-2% 180 N/A 110 724 6.58 0.32 20.57
HBT
250-nm
[36] InP 10 -3 182 4* 105 575 5.47 0.33 16.57
HBT
130-nm
[35] SiGe 10 ±2% 170 3.6* 108 537 5 0.38 13.16
HBT
130-nm
[49] SiGe 24@ +2.5%,-1% 95 N/A 247 1506 6.09 0.65 9.37
HBT
N/A = not available.
&BWs are reported as declared by authors. £Ripple limits is determined within the authors declared BW for their reported gain.

%Estimated from plots @Cascaded Multi-stage DA (CMSDA) $Cascaded single-stage DA (CSSDA) #Flip-Chip DA
*Biased using external bias-tee **Biased using integrated bias-tee

Page | 34
2.4 Conclusion

This chapter stressed on the importance and the necessity of providing large bandwidth
amplifiers in present time due to its rise in value and need in the upcoming future for higher
spectrum capacity and data rates reasons. Several amplifier topologies from small to large
bandwidth performance were presented, but the choice of study at the end was placed on
distributed amplifier device as it showed to be the only design solution that can provide
interesting ultra-wide amplification bandwidth. This mainly due to its capability provide
impedance matching over an extremely wide frequency range by creating artificial
transmission lines through distribution of Gm-cells between inductive-type segments. Its
history, principle and technology-process versatility were also presented. From equation
(2.11), design complexity was detected where design variables from number of unit cells
and dimensions of both Gm-cell and inductive-type segments are correlated, and need to
be studied simultaneously to meet multiple design objectives to find the best design.
On that basis, the following tasks are set to be covered in the next chapters of this thesis
manuscript:
 Propose a DA model suited for mm-wave application that is considered complete
when it comes to describing frequency behaviors, such as gain response, for
instance, by taking into account the whole design parameters the conventional
DA can be made synthesized from.
 Propose a design methodology, based on our DA model, that demonstrates
optimum sizing of DA design parameters that will then assist us in choosing the
highest 𝐺𝐵𝑃⁄𝑃𝐷𝐶 metric (2.14), a desirable criterion for low-consumption
application.
 Finally, validate the proposed design process through on-wafer characterization
of both a single-stage and cascaded stages CMOS-based DAs.
In the next chapter, Chapter 3, we begin by discussing the already existing design
techniques for DAs and discussing our own newly proposed model.

Page | 35
Page | 36
Chapter 3

Proposition of a Novel Model to Design


DAs: the Four-Port Chain (ABCD)
Model

In the previous chapter, the topic of DA was introduced. Both its principle of operation
and performance abilities with different technology processes were investigated. In this
chapter, its design process is discussed. The commonly used techniques are described first,
followed by introducing a new design model for DAs that assists in successful realization
of CAutoD process and offering a different approach of designing using scripting methods.
This chapter is structured as follows. Section 3.1 starts by stating the requirements for
what a proposed physical model is believed to fulfil in order for it to be considered complete
and general, suitable for mm-wave IC design. Section 3.2 reviews already existing
techniques, from models and methods, suggested for distributed amplifier design.
Section 3.3 goes through the limitations those techniques impose, in particular when
dealing with operating frequencies reaching into the millimeter spectrum. Section 3.4
introduces a novel matrix-based design model using Chain (𝐴𝐵𝐶𝐷) parameters. Finally,
Section 3.5 summarizes and concludes this chapter.

3.1 Introduction

For the realization procedure of an IC, an initial step to consider beforehand is the design
approach; i.e., how a given circuit will be sized to reach the required performance before
being sent to the final stage of fabrication and measurement-based characterization. For a
mm-wave IC, the design process starts by having a corresponding physical model, will it
be by deriving analytical model or through graphical interface using CAD software, that
relates the circuit design variables to its performance characteristics. This aids the designer
to have a better understanding on how the circuit works and determining which parameters

Page | 37
plays what role on the circuit behavior. For instance, since the discussion is dedicate for
DA device, referring to Figure 2.18 the recurrent variables in its design are the number of
unit cells (𝑁) and Gm-cell size; by proposing a given type of physical model that relates to
its 𝑆-parameters, the performance of the DA can be analyzed and used for designing one.
The latter parameters set is highlighted since it is popularly used in microwave engineering
to describe the electrical behavior against frequency and useful in determining the electrical
properties of the circuit at hand such as gain, bandwidth and ports mismatch level.
A detailed inspection of different design techniques that can be used in DA design has
been conducted and subsequent Section 3.2 is dedicated to reviewing them. During this
inspection process, we took note of criteria we wanted to satisfy as a whole through our
own proposed design technique and we listed them here. We deduced that the degree of
usability of a design model can be determined through:
 First, how complete it is. This is defined by the amount of variables the model
takes into consideration and influences its ability in producing optimum design
end-results; when the model is constructed from the complete set of design
parameters the circuit is made from, its suggested performance starts to be
defined as a global optimum solution rather than a local one. In addition, those
amounts of variables influence the number of possible design methodologies a
designer produces from it to satisfy different desired performance merits.
 Second, its versatility. A model can be considered general when it can easily be
exchanged between technology processes, in addition to having the ability to be
adopted to several proposed topologies. That is to say, if a designer is interested
in using a different component topology from the one used when producing a
given tool, the model should be able to be adapt to the designer needs.
 Third, its complexity. Since this thesis theme is about mm-wave circuits, more
parasitic element comes into role and have a dominant impact on the circuit
frequency behavior. The attempts to take into account all tiny effects in order for
it to be considered accurate to simulate a circuit behavior lead to an excessive
complication of the model. This can be observed, for instance, when attempting
to derive an analytical-equation based model that represent the frequency
behavior for the IC under study through representing first all of its components
with their respective complete AC equivalent network.
 Finally, its reliability in simulating the circuit behavior. Again, this is crucial
since we are dealing with millimeter IC design; simplifying through neglecting
those parasitic elements degrades the overall accuracy of the proposed model,
devaluating all results of circuit analysis and thereby rendering it unsuitable for
higher frequencies usage.
The target in this thesis is set to take into account those criteria and propose a complete
model, without any approximation, for mm-wave DA design. Its main purpose is to be used
in scripting-based design technique and successful implementation of computer automation
to render the design process simple and fast. Our end-goal is to offer the designer a different
and interesting technique (other than conventional analytical approaches) in which the
outcome of the suggested CAutoD process is a set of 3D graphical design exploration plots
(DSE) plots. Designers then can explore through them different possible solutions and
choose the best design that meet their multiple performance objectives.

Page | 38
3.2 Existing Distributed Amplifier Design Techniques

Several design techniques already exists in the literature and usually each proposal
comes as an improvement for precedent works. While they are useful for preliminary
sizing, they were deduced to be not complete. Each proposed model is based on certain
assumptions that makes its outcome correct within design limitations. In addition, they
demonstrate a compromise between reliability, complexity and completeness. Simplicity
was deduced to be the most preferable through approximation and elements neglecting,
which leads to a less reliable model and, thereby, lower optimum performance distant from
being considered a global solution choice. This section starts by presenting the most
common approach in representing a DA [40], since 1982, and progress through the
available improvements until recent day’s technique is discussed, in which designers rely
more often on CAD tools. In the next section, Section 3.3, their limitations are discussed.

3.2.1 Artificial-Line Based Model [40]


The first physical model proposed for DAs dates back to 1982. It is still being referenced
by recent publications [29], [31], [33], [42], [47], [58], [59], while explaining the concept
of a DA. It is a set of mathematical equations derived from portraying the input and output
loaded lines as artificial lines. Due to this assumption, they refer to their DA as a traveling-
wave amplifier. This concept was already described in details in Chapter 2 (Section 2.3)
with equation (2.11) derived to model the DA gain behavior with respect to frequency. It
was derived while assuming the µ-TL-type segments a unit cell is made from as an
equivalent lumped-element circuit with the intrinsic per-unit-length parasitic quantities, i.e.
𝑅𝐿𝐶𝐺-lumped circuit. In addition, it considers the influence of transistor through its
unilateral AC equivalent circuit. To prove the validity of their model, the authors in [40]
implemented a 1 ̶ 13-GHz distributed amplifier using 0.1-mm GaAs process node.
Figure 3.1 illustrates the 𝑆-parameter results versus frequency through on-wafer
characterization. Indeed, their proposed mathematical gain-model is shown to fit their
measured DA gain up to 12 GHz.

Figure 3.1: Experimental performance of a 0.1-mm GaAs node distributed amplifier at 2‒


14 GHz frequency range. [40]

Page | 39
3.2.2 Stage-Scaling Based Model [51]
Due to the assumption of artificial lines and identical transistor dimension across all unit
cells, the conventional model introduced first is limited only to uniform DAs design. [51]
demonstrated a modified version of this analytical-based model where each unit cell is now
considered independent and, hence, its own µ-TL-type segments and Gm-cell can be sized
differently. Their main objective was to offer a methodology to enhance the power added
efficiency (PAE) of DAs. This was achieved through a stage-scaling technique for the
output-loaded line to utilize more voltage swing while reducing the total current
consumption. The design technique and analysis provided by [51] are derived for bipolar-
based DAs, but can as well be adapted to FET-based DAs. In a stage-scaled DA, the loaded
collector-line impedance and the transistor size are scaled simultaneously from unit cell to
another. Figure 3.2 illustrates such design example implemented by [51]. Both collector-
line TL width and cascode transistors sizes are increasing as they approach the last unit
cell. It is noticeable for this particular example that the first three stages of the amplifier are
not scaled; based on the authors, this is due to the highest T-line impedance achievable on-
chip which is restricted by the minimum metal width rule of their process technology. In
addition, the base-line TL width is kept the same since negligible current is being consumed
and hence no significant enhancement in the PAE of the DA.

Figure 3.2: Schematic diagram of the stage-scaled distributed amplifier. [51]

The wave propagation constant, derived by [51], describing the output-loaded line is
given by (3.1a). It is noticeable how the scaling process is achieved through introducing
the design term 𝜂𝑐,𝑖 (3.1b); where 𝑖 th is the position of the unit cell. It is defined as the ratio
between the intrinsic and loaded collector-line characteristic impedance, denoted by 𝑍𝐶0,𝑖
(3.1c) and 𝑍𝐶,𝑖 (3.1d), respectively. The terms 𝑅𝑐𝑙 , 𝐶𝑐𝑙 and 𝐿𝑐𝑙 represent the equivalent
𝑅𝐿𝐶𝐺-element circuit of the µ-TL-type segments, and 𝐶𝑜𝑢𝑡 and 𝐺𝑜𝑢𝑡 represent the output
complex impedance of the Gm-cell device used.
𝛾𝑐,𝑖 = 𝛼𝑐,𝑖 + 𝑗𝛽𝑐,𝑖 (3.1a)

Page | 40
1 𝜂𝑐,𝑖
= ( 2 𝜈 𝑅 𝐶 𝑘 𝑁−𝑖 + 𝑍𝑐,𝑖 𝐺𝑜𝑢𝑡 𝑘 𝑁−𝑖 )
2𝑙𝑐,𝑖 𝜂𝑐,𝑖 − 1 𝑐,𝑖 𝑐𝑙,𝑖 𝑜𝑢𝑡
𝐶𝑜𝑢𝑡,𝑖
+ 𝑗𝜔√𝐿𝑑,𝑖 (𝐶𝑐𝑙,𝑖 + )
𝑙𝑐,𝑖

𝑍𝑐0,𝑖
𝜂𝑐,𝑖 = (3.1b)
𝑍𝑐,𝑖

𝑅𝑐𝑙,𝑖 + 𝑗𝜔𝐿𝑐𝑙,𝑖 𝐿𝑐𝑙,𝑖


𝑍𝑐0,𝑖 = √ ≈√ (3.1c)
𝐺𝑐𝑙,𝑖 + 𝑗𝜔𝐶𝑐𝑙,𝑖 𝐶𝑐𝑙,𝑖

𝐿𝑐𝑙,𝑖
𝑍𝑐,𝑖 ≈
√ 𝐶 (3.1d)
𝐶𝑐𝑙,𝑖 + 𝑜𝑢𝑡,𝑖⁄𝑙
𝑐,𝑖

Assuming phase synchronization between output and input lines, i.e., 𝛽𝑐 𝑙𝑐 = 𝛽𝑏 𝑙𝑏 , the
total output voltage swing arriving at the load is found to be (3.2). The terms 𝑍𝐿 and 𝐼𝑐 are
the impedance and DC-current of the last unit cell. The term “𝑘” is the scaling factor for
both the transistor size and the dc-bias current between two adjacent unit cells in the
direction from the output to the input. The term “𝓏” is the scaling factor of the loaded
collector-line impedance seen by the Gm-cell between two adjacent unit cells in the same
direction.
𝑁 𝑁
1 𝑁−𝑛
𝑉𝑜𝑢𝑡 = ∑ 𝐼𝑐 𝑍𝐿 (𝑘 2 𝓏) 2 𝑒𝑥𝑝 (− ∑ 𝛼𝑐,𝑖 𝑙𝑐,𝑖 )
2 (3.2)
𝑛=1 𝑖=𝑛

Indeed, through the suggestion of (3.2), higher degree of design freedom is possible
since more variables are available and each unit cell is treated independently. This also
results in the ability to explore a broader range of performance solutions compared to [40].

3.2.3 Distributed-Line Based Model [52]


The physical model provided by [52] comes as an improvement for both [40] and [51]
models. Contrary to traditional ways of treating the unit-cell µ-TL-type segments as
equivalent 𝑅𝐿𝐶𝐺-lumped circuit, [52] treats each segment as a real microstrip line with its
own distributed effect. Through this change, propagation and mismatch impact between
inter-cells are now taken into considered. This provides an extra degree of accuracy when
performing preliminary sizing, since a more realistic representation of the TL losses is now
considered. In addition, [52] treats the distributed amplifier as a discrete set of cells
connected together, contrary to the continuous artificial line concept of [40] and [51]. This
treatment is considered a more realistic approach, since in practice the Gm-cells induce
amplification at specific discrete positions (nodes) between the input to output loaded-line;
therefore, it will lead to a more reliable representation of the DA amplification response.

Page | 41
Figure 3.3: Equivalent circuit of the output loaded-line side of the nth unit cell. [52]

Figure 3.3 illustrates the equivalent AC model of the output loaded-line side of a single
unit cell made from a FET-based Gm-cell. The output port of the Gm-cell is depicted by a
voltage-controlled current source with a complex parallel RC-impedance denoted by 𝑅𝑑𝑠𝑛
and 𝐶𝑑𝑠𝑛 . The Gm-cell is connected in the middle of two µ-TLs-type segments denoted by
subscripts ‘𝐿’, for left side, and ‘𝑅’, for right side. The influence of the preceding and
subsequent unit cells is taken into consideration in this illustration, as they will affect the
total output current generated by the unit cell under study. The latter is done as follows.
Equivalent impedances 𝑍𝑒𝑞𝑢𝐿𝑛 and 𝑍𝑒𝑞𝑢𝑅𝑛 model all the unit-cells seen on the left and right
sides, respectively, of the 𝑛th unit cell under study. Current 𝐼𝑜𝑢𝑡𝑛−1 , flowing from the left
side, represents the total current wave generated by the preceding cells traveling towards
the right side. Impedance 𝑍𝐿𝐿𝑛 represents the input impedance seen by the µ-TL-type
segment (𝑍𝑑𝐿𝑛 ,𝛾𝑑𝐿𝑛 ) on its left side towards its right side while being loaded on its right
side by impedance 𝑍𝐿𝑅𝑛 in parallel with output impedance of the Gm-cell (𝑅𝑑𝑠𝑛 , 𝐶𝑑𝑠𝑛 ).
Impedance 𝑍𝐿𝑅𝑛 represents the input impedance seen by the µ-TL-type segment
(𝑍𝑑𝑅𝑛 , 𝛾𝑑𝑅𝑛 ) on its left side towards its right side while being loaded on its right side by
𝑍𝑒𝑞𝑢𝑅𝑛 . Impedance 𝑍𝑅𝑅𝑛 represents the input impedance seen by the µ-TL-type segment
(𝑍𝑑𝑅𝑛 ,𝛾𝑑𝑅𝑛 ) on its right side towards its left side while being loaded on its left side by
impedance 𝑍𝑅𝐿𝑛 in parallel with output impedance of the Gm-cell (𝑅𝑑𝑠𝑛 , 𝐶𝑑𝑠𝑛 ). Impedance
𝑍𝑅𝐿𝑛 represents the input impedance seen by the µ-TL-type segment (𝑍𝑑𝐿𝑛 , 𝛾𝑑𝐿𝑛 ) on its
right side towards its left side while being loaded on its left side by 𝑍𝑒𝑞𝑢𝐿𝑛 .

Based on this configuration and applying Kirchhoff’s laws and microwave TL theory,
the total current wave traveling towards the right side of the nth Gm-cell, 𝐼𝑡𝑤𝑛 , is expressed
as (3.3);
𝐼𝑡𝑤𝑛 = −𝐼𝐷𝑛 𝟋𝑅𝑛 + 𝐼𝑜𝑢𝑡𝑛−1 𝟋𝑆𝑛 𝑇𝑑𝐿𝑛
(3.3)
where 𝟋𝑅𝑛 represents how much current generated by the nth current source, 𝐼𝐷𝑛 , is incident
on the right TL, and is calculated from (3.4);
𝑍𝑅𝐿𝑛 𝑅𝑑𝑠𝑛
𝟋𝑅𝑛 =
(1 + 𝑗𝜔𝑅𝑑𝑠𝑛 𝐶𝑑𝑠𝑛 )𝑍𝑅𝐿𝑛 𝑍𝐿𝑅𝑛 + (𝑍𝑅𝐿𝑛 + 𝑍𝐿𝑅𝑛 )𝑅𝑑𝑠𝑛 (3.4)

Page | 42
and 𝟋𝑆𝑛 represents the splitting current factor between the right part of the nth Gm-cell, just
before the right TL (𝑍𝑑𝑅𝑛 , 𝛾𝑑𝑅𝑛 ), and the dissipated part in the transistor drain impedance,
as given by (3.5);
𝑅𝑑𝑠𝑛
𝟋𝑆𝑛 =
(1 + 𝑗𝜔𝑅𝑑𝑠𝑛 𝐶𝑑𝑠𝑛 )𝑍𝐿𝑅𝑛 + 𝑅𝑑𝑠𝑛 (3.5)

Finally, the term 𝑇𝑑𝐿𝑛 is the nth left µ-TL-type segment (𝑍𝑑𝐿𝑛 ,𝛾𝑑𝐿𝑛 ) transmission coefficient
which takes into account the multiple reflections that occur at its boundaries. 𝑇𝑑𝐿𝑛 can be
calculated using (3.6a) on the basis of Figure 3.3;

(1 + Г𝑖𝑛𝑑𝐿𝑛 )(1 + Г𝑜𝑢𝑡𝑑𝐿𝑛 )𝑒 −𝛾𝑑𝐿𝑛 𝑙𝑑𝐿𝑛


𝑇𝑑𝐿𝑛 = (3.6a)
1 + Г𝑖𝑛𝑑𝐿𝑛 Г𝑜𝑢𝑡𝑑𝐿𝑛 𝑒 −2𝛾𝑑𝐿𝑛 𝑙𝑑𝐿𝑛

𝑍𝐿𝐿𝑛 − 𝑍𝑑𝐿𝑛
Г𝑖𝑛𝑑𝐿𝑛 = (3.6b)
𝑍𝐿𝐿𝑛 + 𝑍𝑑𝐿𝑛

𝑍𝑑𝐿𝑛 − 𝑍𝑅𝐿𝑛
Г𝑜𝑢𝑡𝑑𝐿𝑛 = (3.6c)
𝑍𝑑𝐿𝑛 + 𝑍𝑅𝐿𝑛

In order to determine the amount of current that a given nth unit-cell transmits to the
entire right-side unit cells, represented by an equivalent impedance 𝑍𝑒𝑞𝑢𝑅𝑛 in Figure 3.3,
the transmission coefficient of the right µ-TL-type segment (𝑍𝑑𝑅𝑛 , 𝛾𝑑𝑅𝑛 ) is introduced to
the input current expressed in (3.3), and hence the total output current comes to (3.7);
𝐼𝑜𝑢𝑡𝑛 = 𝐼𝑡𝑤𝑛 𝑇𝑑𝑅𝑛

(3.7)
= −𝐼𝐷𝑛 𝟋𝑅𝑛 𝑇𝑑𝑅𝑛 + 𝐼𝑜𝑢𝑡𝑛−1 𝟋𝑆𝑛 𝑇𝑑𝐿𝑛 𝑇𝑑𝑅𝑛

where the transmission coefficient 𝑇𝑑𝑅𝑛 stands for the 𝑛th right µ-TL-type segment (𝑍𝑑𝑅𝑛 ,
𝛾𝑑𝑅𝑛 ) and is calculated using (3.8a);

(1 + Г𝑖𝑛𝑑𝑅𝑛 )(1 + Г𝑜𝑢𝑡𝑑𝑅𝑛 )𝑒 −𝛾𝑑𝑅𝑛 𝑙𝑑𝑅𝑛


𝑇𝑑𝑅𝑛 = (3.8a)
1 + Г𝑖𝑛𝑑𝑅𝑛 Г𝑜𝑢𝑡𝑑𝑅𝑛 𝑒 −2𝛾𝑑𝑅𝑛 𝑙𝑑𝑅𝑛

𝑍𝐿𝑅𝑛 − 𝑍𝑑𝑅𝑛
Г𝑖𝑛𝑑𝑅𝑛 = (3.8b)
𝑍𝐿𝑅𝑛 + 𝑍𝑑𝑅𝑛

𝑍𝑑𝑅𝑛 − 𝑍𝑅𝑅𝑛
Г𝑜𝑢𝑡𝑑𝑅𝑛 = (3.8c)
𝑍𝑑𝑅𝑛 + 𝑍𝑅𝑅𝑛

Page | 43
Figure 3.4: Distributed amplifier output loaded-line equivalent circuit representation. [52]

Figure 3.4 demonstrates the case where 𝑁 unit cells are distributed. By developing (3.7),
a non-closed-form model for the total output current generated from the output loaded-line
of the DA is expressed in (3.9). The term 𝛿〈𝑖,𝑛〉 represents the “Kronecker” delta.
𝑁 𝑁
𝛿〈𝑖,𝑛〉 1−𝛿〈𝑖,𝑛〉
𝐼𝑜𝑢𝑡 = ∑ −𝐼𝐷𝑛 [∏(𝑇𝑑𝑅𝑖 𝟋𝑅𝑖 ) (𝟋𝑆𝑖 𝑇𝑑𝐿𝑖 𝑇𝑑𝑅𝑖 ) ]
(3.9)
𝑛=1 𝑖=𝑛

For the gate line analysis, a similar approach to that of the drain line analysis was applied.
Based on Figure 3.5, the gate side of the Gm-cell is depicted as an equivalent impedance
of series resistor and capacitor, denoted 𝑅𝑔 and 𝐶𝑔𝑠 , respectively. One can perceive that the
input loaded-line is similar to a transmission line periodically loaded by 𝑁 complex series
RC-impedances.

Figure 3.5: Distributed amplifier input loaded-line equivalent circuit representation. [52]

Based on this configuration, the approach was focused on determining the voltage 𝑉𝐶𝑛
at each gate-source capacitor 𝐶𝑔𝑠𝑛 for the 𝑛th unit-cell when a given voltage source 𝑉𝑆 is
applied (𝑉𝐶𝑛 is the control voltage of the controlled current source 𝐼𝐷𝑛 in Figure 3.3). This
resulted in the non-uniform capacitor voltage-expression (3.10).
𝑛
𝑍𝑖𝑛 1 1−𝛿
𝑉𝐶𝑛 = 𝑉𝑆 ( )( ) ∏ [𝑇𝑔𝐿𝑖 𝑇𝑔𝑅𝑖 〈𝑖,𝑛〉 ]
𝑍𝑖𝑛 + 𝑅𝑆 1 + 𝑗𝜔𝑅𝑔𝑛 𝐶𝑔𝑠𝑛 (3.10)
𝑖=1

where 𝑍𝑖𝑛 is the input impedance of the whole DA and 𝑅𝑆 is the internal resistance of the
signal source. The term 𝑇𝑔𝐿𝑖 (𝑇𝑔𝑅𝑖 ) is the transmission coefficient of the µ-TL-type segment
(𝑍𝑔𝐿𝑛 ,𝛾𝑔𝐿𝑛 ) ((𝑍𝑔𝑅𝑛 ,𝛾𝑔𝑅𝑛 )) which is determined by using the same equations as (3.6a) ((3.8a)
respectively) with the boundary reflection coefficients given in (3.11). The same
description given to the output-loaded line segments input-impedances (𝑍𝐿𝐿𝑛 , 𝑍𝑅𝐿𝑛 , 𝑍𝐿𝑅𝑛 ,

Page | 44
𝑍𝑅𝑅𝑛 ) can be applied to input-loaded line segments input-impedances, the only difference
lies in the Gm-cell inherent impedances where previously, for the output side, we were
dealing with (𝑅𝑑𝑠𝑛 , 𝐶𝑑𝑠𝑛 ) in parallel and here, for the input side, we are dealing with (𝑅𝑔𝑛 ,
𝐶𝑔𝑠𝑛 ) in series.
𝑍𝐿𝐿𝑛 − 𝑍𝑔𝐿𝑛 𝑍𝑔 − 𝑍𝑅𝐿𝑛
Г𝑖𝑛𝑔𝐿𝑛 = − ; Г𝑜𝑢𝑡𝑔𝐿𝑛 = − 𝐿𝑛 (3.11a)
𝑍𝐿𝐿𝑛 + 𝑍𝑔𝐿𝑛 𝑍𝑔𝐿𝑛 + 𝑍𝑅𝐿𝑛

𝑍𝐿𝑅𝑛 − 𝑍𝑔𝑅𝑛 𝑍𝑔 − 𝑍𝑅𝑅𝑛


Г𝑖𝑛𝑔𝑅𝑛 = − ; Г𝑜𝑢𝑡𝑔𝑅𝑛 = − 𝑅𝑛 (3.11b)
𝑍𝐿𝑅𝑛 + 𝑍𝑔𝑅𝑛 𝑍𝑔𝑅𝑛 + 𝑍𝑅𝑅𝑛

The parameter linking the output loaded-line, modeled by (3.9), and the input loaded-
line, modeled by (3.10), is the transconductance (𝑔𝑚 ) of the nth Gm-cell topology used
which relates its output current (𝐼𝐷𝑛 ) to its input control voltage (𝑉𝐶𝑛 ) as given by (3.12).

𝐼𝐷𝑛 = 𝑔𝑚𝑛 𝑉𝐶𝑛


(3.12)
Using equations (3.9), (3.10) and (3.12), the designer can now evaluate more accurately
the impact of each design parameter the distributed amplifier is made from, as well as treat
each unit cell separately and arrive at the end of the design process to an optimum and more
realistic performance.

3.2.4 Artificial Neural Network Based Model


Another modeling technique worth mentioning is through artificial neural network.
Since in mm-wave spectrum more elements for representing a certain behavior have to be
taken into consideration, the process of deriving equations is shortened through machine
learning.

Figure 3.6: Simplified representation of an artificial neural network

Artificial neural network is one topology of machine learning that identifies the
relationship (function) between a given set of data and already known outcomes. It does
this through artificial intelligence algorithms where they tune its function constants until
they find the parameter sets that best matches the training set, i.e., already observed input
and output data. Figure 3.6 illustrates an example of a simplified neural network structure.
It is made of three layers. The left most side represents its input layer, it is made of a total

Page | 45
of 𝑖 input pins where design parameter set are fed. The right most side represents the output
layer, it is made of a total of 𝑛 output pins; this depends on what the designer is interested
in providing as design characteristics. The middle layer is composed of a two-dimensional
array of neurons, which are the function constants that will be optimized. After the training
stage, the neural network can be used to predict possible characteristic outcomes from a
new different input dataset within the possibility of the pre-trained set. Such machine-
learning modeling technique has been demonstrated in [53]-[55], for instance, for modeling
power-amplifier behaviors. Compared to previously mentioned mathematical equation-
based modeling techniques, by using neural network technique there is no need for the
designer to be concerned with any tiny effects that impact circuit behavior. Thereby,
arguably it could be considered as a simple and fast technique to use during a circuit design
process, since it can be considered similar to a multi-variable curve fitting approach.
However, it should be highlighted the importance of the pre-training process step needed
to make the neural network work. For instance, when dealing with the DA circuit, as a
beforehand step the designer needs to simulate several already designed 𝑁 unit-cells DAs
and use those simulations results for the neural network model training (fitting) process.
This means that when the model becomes ready to be used again in a new design task, its
prediction relies on how adequately the pre-training circuits set were designed. For us, we
are interested in providing a design model that discovers the optimum performance (first
criterion, Section 3.1).

3.2.5 Widespread Design Technique: Through CAD Software


The final design technique that could be considered the most preferable and reliable way
for DAs, or any electronic circuit in a general context, is through a computer-aided design
(CAD) software. Traditionally, circuits are first assembled on a graphical user interface by
picking the required components from the design library. Then, a set of desired
performance goals are defined and the circuit variables intended to be tuned are
enumerated. Through the usage of iterative optimization functionality, the CAD software
configures automatically those variables through several iterations until the declared goals
are fulfilled or closely achieved. This technique is proven to be the most accurate procedure
one can have, since no assumption is made, and the closest one in simulating circuit
behavior to real-life measured behavior. Several CAD tools exists, but for RF/Microwave
integrated circuit design the most famous ones are Keysight ADS and Cadence Virtuoso
softwares. Both tools provide fully-custom integrated design environment to designers of
RF electronic products such as mobile phones, wireless networks, satellite
communications, radar systems, and high-speed data links. They support every step of the
design cycle from schematic entry, layout, design rule checking (DRC), frequency-domain
and time-domain circuit simulation, and sometimes electromagnetic field simulation,
allowing the designer to fully characterize and optimize an RF design without changing
CAD platforms. Figure 3.7 provides a demonstrative example of a DA circuit synthesized
on ADS software schematic where the transistor sizes are defined as a single variable, and
random optimization is being applied until designer-defined gain (𝑆21 ) goal is reached.

Page | 46
Figure 3.7: ADS schematic view of a distributed amplifier being optimized to satisfy a defined
gain response goal.

3.3 Design Limitations when Operating in mm-Wave Band

Five different techniques where mentioned for DA design. From mathematical-based


approaches, to artificial intelligence-based techniques and through schematic-level design
environments using CAD tools. However, a set of limitations have been detected for each
technique and are described as follows.
First, neglecting the inter-stage propagation, where spacing between individual Gm-cells
is assumed small compared with the wavelength of operation (𝜆𝑜 ). This is observed in the
first model (subsection 3.2.1) and second model (subsection 3.2.2) through their artificial
line assumption. While valid for low frequency of operation, it neglects the transmission
line theory for the implemented physical TL spacing that takes effect especially when
dealing with DAs extending into the mm-wave region.
Second, neglecting the fact that the Gm-cells embedded between the two lines induce
amplification at discrete positions rather than the assumed continuous manner. A limitation
also observed in the first two models discussed in Section 3.2.
This pair of limits lead to constraining design formulas usability, and hence the DSE
reliability, to only synthesizing DAs that require TL spacing between consecutive Gm-cells
with electrical length (𝛽𝑙) ≪ 𝜋⁄2, or synthesizing DAs that amplify signals with long
wavelengths, in other words DA with smaller bandwidth [56].
Both assumptions were fixed through (3.9) model (Subsection 3.2.3) where [52] re-
derived DA physical equation while keeping each TL segment as a real physical µ-TL with
its distributed effect. However, even with this improvement, until now both [40], [51], [52],
suggested models are still considered as approximate representation of the DA behavior
since they are produced based on the assumption of a unilateral two-port AC equivalent

Page | 47
model for transistor. This leads to the third design limitation caused by assuming a simple
Gm-cell model. The central element of a DA is its Gm-cell; as operating frequencies
increase, amplifier designers can no longer benefit from transistor unilateral approximation
where Miller effect starts to have a dominant impact. This assumption leads to discrepancy
in predicting its behavior, and hence that of the DA, and results in suboptimal Gm-cell
sizing. The arguably widely adopted way for improving reverse isolation, as observed in
Chapter 2 (subsection 2.3.3) where DAs in different technology processes are presented, is
by implementing a conventional cascode configuration. Even then, modeling the DA
equivalent circuit becomes complicated since the high-frequency equivalent circuit for a
single transistor biased in active region becomes large with a plethora of parameters [57].
It becomes more complicated for a cascode Gm-cell since it involves stacking a minimum
of two transistors, rendering analytical expressions based design processes for mm-wave
DAs quite cumbersome. This is without considering transistor non-quasi-static behavior
that appears when operating deep into mm-wave spectrum. Furthermore, in bulk
technology, the transistor is a four-terminal device in which the substrate influence can no
longer be disregarded.
Fourth limit, and this is more specific for neural-network based modeling approach, is
using a design model constructed from a beforehand step of function fitting to a series of
already observed data. While this approach does simplify the derivation process, where one
is no longer subjected to the hustle to derive equations and taking into account all tiny
effects; however, the degree of optimum performance it predicts for a design characteristic
is bounded within how well designed the pre-training circuit set were when used during the
fitting pre-process. Another limit for the same modeling approach is its reliability being
bounded within the same training data set used in the fitting pre-process. Beyond this set,
its predicted outcome becomes more of an extrapolation with a greater degree of
uncertainty. For example, when speaking about DA design, if the pre-training data included
a set of DAs designed up to 𝑁𝑚𝑎𝑥 unit cells, using the model for 𝑁 > 𝑁𝑚𝑎𝑥 has a higher
risk of producing meaningless results. Therefore, making the neural network technique
more of a statistical modeling technique than previously mentioned ones.
Fifth limit would be neglecting losses that become important at high frequencies and the
need for loss-compensated DA topologies. Even if distributed amplification is a promising
technique to push the limit of gain-bandwidth-product (GBP) [47], [58], [59], however,
with conventional cascode as Gm-cell, it is difficult to achieve these potentially wideband
characteristics because the resistive losses in the loaded lines are large at high frequencies.
Different techniques of loss-compensation cascodes were suggested, as in [50], [60]. This
highlights the various topologies that can be adopted and the approach considered in
enhancing the cascode topology, in which extra components are being added to enable the
DA to reach wider BWs thereby maximizing its GBP. This undoubtedly adds a new
dimension of complexity for mm-wave DA design, pushing DA designers even more
towards CAD software.
The sixth and final limit, and it is specific to using graphical schematic environment of
CAD tools, concerns the fact that some design variables are not allowed to be automated
by their optimization algorithm and hence it relies on the designer to manually do it. For
the case of DA design, semiconductor foundries provide for a given process technology a
broad range of transistor sizes to explore from. Determining the optimal transistor size to

Page | 48
use with the optimal number of Gm-cells (𝑁𝑜𝑝𝑡 ) in order to maximize DA gain for a desired
BW could be considered tricky. For instance, for CAD tools where circuit design is possible
only via drawing on schematic diagram, the designer will be obligated to manually
manipulate the number of unit cells for each component modification performed on the
DA. Designing through CAD software schematic environment is therefore the most
accurate technique available, but a time-cost consuming one indeed. It could be argued that
with recent progress of CAD tools in extending the capability of their design environment,
where they start to offer scripting environment possibilities such as application extension
language (AEL) in Keysight ADS and OCEAN language in Cadence Virtuoso, that the
design of the DA can be automated. While similar can be said when using, for instance,
MathWorks MATLAB computing environment, a scripting language will not be able to
achieve DA design automation without having first at-hand a proper physical model. A
model that takes into account the complete set of DA variables, including and most
importantly its variable 𝑁 as a free variable.
All of those limits, therefore, highlight that the success in the automated design of an IC
DA depends crucially on first the appropriate choice of the device model. They emphasize
on the need for proposing a new complete and reliable DA model that accommodates any
suggested Gm-cell topology and that provides systematic design approaches with less
dependence on CAD proprietary optimization tools. This is covered in next Section 3.4.

3.4 Proposed Solution: Chain Matrix (ABCD) Based Design Model [73]

After listing both the milestones and limitations in available distributed amplifier design
techniques, in this section an original matrix-based technique is introduced. This model is
based on 𝐴𝐵𝐶𝐷-parameters and it was proposed for the purpose of providing a more direct
way of designing mm-wave distributed amplifiers without imposing restrictions on the
designer search for optimum design solution. Compared to the previous models, we
concentrate on providing a model that deals with the criteria listed in Section 3.1: Optimum
performance by taking into account the complete set of design parameters the conventional
DA can be made from, and versatility, complexity and reliability with focus on accurate
representation of the DA mm-wave behavior.

Figure 3.8: Four-port equivalent network for a unit cell.

Page | 49
Referring to Figure 3.8, the Gm-cell can be represented symbolically by its equivalent
two-port admittance network. Its self-admittance impacts on the input and output loaded-
lines are represented by 𝑌𝐼𝐼 + 𝑌𝐼𝑂 and 𝑌𝑂𝑂 + 𝑌𝐼𝑂 , respectively, where (𝐼) subfix refers to
input of Gm-Cell and where (𝑂) subfix refers to input of Gm-Cell. Because any Gm-cell
also introduces two different kinds of coupling, leakage from output to input node and
amplification from its input to output node, their effect are represented here as two different
admittances where 𝑌𝐼𝑂 admittance is for the former type of coupling and 𝑌𝑂𝐼 admittance is
for the latter type. The choice for such a Gm-cell representation was made to keep the
proposed unit cell model (and hence the DA model) suitable for any two-port Gm-cell
topology the designer decides to use. In addition, all inductors in Figure 2.18 were replaced
by a general two-port series impedance networks, depicted as 𝑍𝑆𝐼 and 𝑍𝑆𝑂 for input and
output side, respectively, for the intention of obtaining a general unit cell model and for the
ease it provides when interchanging DA inductive-type segments topology, as will be
presented shortly.
From Figure 3.8, one can now observe that when designing a DA, one is dealing at its
core with a four-port network unit cell. Because a DA is defined by cascading this four-
port network 𝑁 times, an 𝐴𝐵𝐶𝐷-type matrix (3.13), adapted to a four-port representation,
is therefore a suitable choice to characterize a single unit cell.
𝑉1 𝐴 𝐵14 𝐴 𝐵12 𝑉4
[ 14 ] [ 12 ]
𝐼 𝐶 𝐷14 𝐶12 𝐷12 −𝐼4
[ 1 ] = [ 14 ][ ]
𝑉3 𝐴 𝐵34 𝐴 𝐵32 𝑉2 (3.13)
[ 34 ] [ 32 ]
𝐼3 𝐶34 𝐷34 𝐶32 𝐷32 −𝐼2

The left side of the equality sign represents the input side of the four-port unit cell and
the right-most matrix represents its output side. The negative sign of both 𝐼2 and 𝐼4 arises
to make the output current of one cascaded unit cell equal to the input current of the adjacent
one. When speaking of the whole DA, illustrated in Figure 2.18, its 𝐴𝐵𝐶𝐷-matrix model is
represented by (3.14); thereby, it becomes a matter of multiplying the matrix of each unit
cell together in the same order the DA network is drawn, i.e., left to right.
𝑉1 𝐴 𝐵14 𝐴12 𝐵12 𝑉4
𝑁 [ 14 ] [ ]
𝐼1 𝐶 𝐷14 𝐶12 𝐷12 −𝐼4
[ ] = ∏ [ 14 ] [ ]
𝑉3 𝐴34 𝐵34 𝐴 𝐵32 𝑉2 (3.14)
𝑛=1 [ ] [ 32 ]
𝐼3 𝐶34 𝐷34 𝐶32 𝐷32 𝑛 −𝐼2 𝑛

In order to solve either equation (3.13) or (3.14), it is necessary to identify the boundary
conditions at each terminal side of the four-port network. By proper application of
Kirchhoff’s voltage law at each port, they are found to be (3.15):
𝑉𝑆 = 𝑍1 𝐼1 + 𝑉1
0 = 𝑍2 𝐼2 + 𝑉2
0 = 𝑍3 𝐼3 + 𝑉3 (3.15)
0 = 𝑍4 𝐼4 + 𝑉4

Here it is assumed the case where only port 1 is stimulated by a microwave signal
(Figure 2.18) and all other ports are terminated by a shunted load: 𝑍2 , 𝑍3 and 𝑍4 . Depending
on the choice of Gm-cell topology and type of inductive-type segments used, it will be a

Page | 50
lumped impedance or a physical TL. Both equations (3.13) and (3.14) can be used to design
any DA that follows the conventional cascading topology of Figure 2.18.
By taking the case of series lumped impedance as a first approach for the inductive-type
segments, with admittance representation for Gm-cell, the factorized 𝐴𝐵𝐶𝐷-matrix
describing the network of Figure 3.8 is found to be the one of (3.16).
𝑉1 1 𝑍𝑆𝐼 0 0 1 0 0 0 1 𝑍𝑆𝐼 0 0 𝑉4
𝐼 0 1 0 0 𝑌 1 𝑌𝐼𝑂 0 0 1 0 0 −𝐼4
[ 1] = [ ] [ 𝐼𝐼 ][ ][ ]
𝑉3 0 0 1 𝑍𝑆𝑂 0 0 1 0 0 0 1 𝑍𝑆𝑂 𝑉2 (3.16)
𝐼3 0 0 0 1 𝑌𝑂𝐼 0 𝑌𝑂𝑂 1 0 0 0 1 −𝐼2

By further inspecting matrix (3.16), one can note the similarity between its structure and
that of Figure 3.8 from which it can be concluded that a DA unit-cell is constructed from
three four-port networks cascaded together. At the right side of the equal sign, the leftmost
and rightmost 4x4 matrices represent each the 𝐴𝐵𝐶𝐷-matrix of the parallel series-lumped
impedances network on the left and right side of the Gm-cell, respectively, and the middle
4x4 matrix, containing the 𝑌-parameters, represents the four-port 𝐴𝐵𝐶𝐷-matrix of the Gm-
cell topology used. Further inspection of (3.16) also reveals the value of such a
representation, both leftmost and rightmost 4x4 matrices are made from smaller
submatrices that represents the 𝐴𝐵𝐶𝐷-parameters of the individual input and output
segments the unit cell is made from. This demonstrates the simplicity and flexibility in
using the proposed model: DA designers now only have to replace those 𝐴𝐵𝐶𝐷-
submatrices with any other 𝐴𝐵𝐶𝐷-submatrices representing a unit-cell input and output
segments topology of their choice. For demonstration, replacing the lumped inductors used
in Figure 3.8 with physical TLs transforms equation (3.16) into equation (3.17); where the
input-side series µ-TL segments are represented by (𝑍𝑐,𝑆𝐼 , 𝛾𝑆𝐼 ) with length (𝑙𝑆𝐼 ⁄2) each and
the output-side series µ-TL segments are represented by (𝑍𝑐,𝑆𝑂 , 𝛾𝑆𝑂 ) with length (𝑙𝑆𝑂 ⁄2)
each. This also applies for the middle Gm-cell matrix; the designer can choose any Gm-
cell topology he desires and apply it to the DA model since it is represented in its general
form (𝑌-parameter). This is contrary to the conventional ways where they are restricted to
a unilateral assumption of a single transistor Gm-cell. A demonstrative example will be
provided in Chapter 4 where a method for accurately representing the frequency behavior
of a cascode topology as a scalable parameter with respect to transistor size and integrating
it with (3.16) or (3.17) is given.

Page | 51
𝑉1
𝐼
[ 1]
𝑉3
𝐼3
𝑙𝑆𝐼 𝑙𝑆𝐼
𝑐𝑜𝑠ℎ (𝛾𝑆𝐼 2
) 𝑍𝑐,𝑆𝐼 𝑠𝑖𝑛ℎ (𝛾𝑆𝐼 2
) 0 0
𝑙𝑆𝐼 𝑙𝑆𝐼
𝑠𝑖𝑛ℎ (𝛾𝑆𝐼 2
)⁄𝑍𝑐,𝑆𝐼 𝑐𝑜𝑠ℎ (𝛾𝑆𝐼 2
) 0 0
= 𝑙𝑆𝑂 𝑙𝑆𝑂
0 0 𝑐𝑜𝑠ℎ (𝛾𝑆𝑂 2
) 𝑍𝑐,𝑆𝑂 𝑠𝑖𝑛ℎ (𝛾𝑆𝑂 2
)
𝑙𝑆𝑂 𝑙𝑆𝑂
[ 0 0 𝑠𝑖𝑛ℎ (𝛾𝑆𝑂 2
)⁄𝑍𝑐,𝑆𝑂 𝑐𝑜𝑠ℎ (𝛾𝑆𝑂 2
) ]

1 0 0 0 (3.17)
𝑌 1 𝑌𝐼𝑂 0
• [ 𝐼𝐼 ]•
0 0 1 0
𝑌𝑂𝐼 0 𝑌𝑂𝑂 1

𝑙𝑆𝐼 𝑙𝑆𝐼
𝑐𝑜𝑠ℎ (𝛾𝑆𝐼 2
) 𝑍𝑐,𝑆𝐼 𝑠𝑖𝑛ℎ (𝛾𝑆𝐼 2
) 0 0
𝑙𝑆𝐼 𝑙𝑆𝐼 𝑉4
𝑠𝑖𝑛ℎ (𝛾𝑆𝐼 2
)⁄𝑍𝑐,𝑆𝐼 𝑐𝑜𝑠ℎ (𝛾𝑆𝐼 2
) 0 0 −𝐼4
[ ]
𝑍𝑐,𝑆𝑂 𝑠𝑖𝑛ℎ (𝛾𝑆𝑂 2𝑂 ) 𝑉2
𝑙𝑆𝑂 𝑙𝑆
0 0 𝑐𝑜𝑠ℎ (𝛾𝑆𝑂 2
)
−𝐼2
𝑙𝑆𝑂 𝑙𝑆𝑂
[ 0 0 𝑠𝑖𝑛ℎ (𝛾𝑆𝑂 2
)⁄𝑍𝑐,𝑆𝑂 𝑐𝑜𝑠ℎ (𝛾𝑆𝑂 2 ) ]

In order to obtain the terminal behavior of the DA, the proposed 4x4 𝐴𝐵𝐶𝐷-matrix
model is first converted into a 2x2 𝐴𝐵𝐶𝐷-matrix characterizing a two-port network in terms
of input and output ports. In the proposed conventional DA design topology, shown in
Figure 2.18, both ports 3 and 4 are terminated by matching impedances 𝑍3 and 𝑍4 ,
respectively, to absorb backward-signals and prevent reflections back to the main loaded
output and input lines, respectively, as early stated in Chapter 2 (subsection 2.3.2). Interest,
thus, falls on both ports 1 and 2, as input and output ports, respectively, leading to the
required design matrix (3.18).
𝑉 𝐴 𝐵 𝑉2
[ 1] = [ ][ ]
𝐼1 𝐶 𝐷 −𝐼2 (3.18)

Applying the previously listed boundary conditions of (3.15), the required 𝐴𝐵𝐶𝐷-
parameters were deduced to be the ones of (3.19):
𝐵
−(𝑍3 𝐶32 + 𝐴32 ) (𝐴14 + 𝑍14 )
4
𝐴= + 𝐴12 (3.19a)
𝛥

𝐵
−(𝑍3 𝐷32 + 𝐵32 ) (𝐴14 + 𝑍14 )
4
𝐵= + 𝐵12 (3.19b)
𝛥

𝐷14
−(𝑍3 𝐶32 + 𝐴32 ) (𝐶14 + )
𝑍4
𝐶= + 𝐶12 (3.19c)
𝛥

Page | 52
𝐷
−(𝑍3 𝐷32 + 𝐵32 ) (𝐶14 + 𝑍14 )
4
𝐷= + 𝐷12 (3.19d)
𝛥

where
𝐵34 𝐷34
𝛥 = (𝐴34 + ) + 𝑍3 (𝐶34 + )
𝑍4 𝑍4 (3.19e)

The two-port network 𝐴𝐵𝐶𝐷-parameters is now easily convertible into any other two-
port network parameters, such as impedance, admittance or power scattering matrix. For
the case at hand, where characterizing the behavior of the DA at terminals 1 and 2 is desired,
𝑆-parameter conversion [56] is used. For ease of referencing when implementing the design
methodology, they are rewritten hereafter (3.20). 𝑆21 represents DA insertion gain, 𝑆12 its
reverse isolation, and 𝑆11 and 𝑆22 represent reflection coefficients seen at the input and
output ports, respectively. 𝑍0 denotes the characteristic impedance reference (usually 50 Ω)
for ports 1 and 2.
𝐵
𝐴 + 𝑍 − 𝐶𝑍0 − 𝐷
0
𝑆11 =
𝐵
𝐴 + 𝑍 + 𝐶𝑍0 + 𝐷 (3.20a)
0

2
𝑆21 =
𝐵
𝐴 + 𝑍 + 𝐶𝑍0 + 𝐷 (3.20b)
0

2(𝐴𝐷 − 𝐵𝐶)
𝑆12 =
𝐵
𝐴 + 𝑍 + 𝐶𝑍0 + 𝐷 (3.20c)
0

𝐵
−𝐴 + 𝑍 − 𝐶𝑍0 + 𝐷
0
𝑆22 =
𝐵
𝐴 + 𝑍 + 𝐶𝑍0 + 𝐷 (3.20d)
0

3.5 Conclusion

This chapter discussed the available models in the literature and techniques used to
design a DA. Each approach presented its own improvement in terms of added parameters
for a more flexible design and targeting to represent the DA behavior more accurately. The
most complete and accurate design technique of course being through CAD software, such
as ADS or Cadence tools, in which the full technology design kit is available from the
manufacturing company. In spite of those, limitations exist and were discussed in this
chapter, and they can be summarized as follows. When it comes to designing DAs, the
designer has one of three options:

Page | 53
1) Proceeding with the mathematical approaches where the designer can directly
see the relation between parameters and hence determine directly the optimal set
for a design goal. However, this is at the expense of being restricted towards low
operating frequencies, in other words designing DA with smaller bandwidths,
since assumptions and simplification through neglecting parameters are more
often performed to derive them.
2) Using machine learning, such as artificial neural network, or curve fitting
techniques where the designer does not have to deal with the tiny effect (plethora
of parasitic) that affect the DA behavior or even how it is made, i.e., the DA will
be treated as a black box. However, its output prediction during the design
process relies on a large set of pre-designed DAs used in the function fitting (pre-
training) process and, also, its optimum performance relies on how adequately
they were designed, as already described in Section 3.3 (fourth limit).
3) Relying on CAD tool, either on its schematic diagram environment or scripting
environment. For the former case, some parameters such as 𝑁 are not readily
accessible to optimize and consequently the design process becomes a manual
and iterative process that relies on intuition. It is also a tedious one if elements
to enhance the DA response are being considered to be modified each time
because this pushes the designer to redo the performance investigation manually
over and over again from scratch, i.e., starting from 𝑁 = 1, per element
modification. Furthermore, when the goal changes for the same topology, even
with the experience from the previous design, relatively large simulation works
are still not avoidable. For the latter case of exploiting CAD scripting capability,
if available by the CAD tool, while it could become possible in this scenario to
automate and hence made easier. But, first and most importantly is having at
disposal a proper physical model: a (simple) model that considers the entire set
of design parameters (complete), gives flexibility in choosing DA components
topologies (versatile) and accurate in the simulation outcome it provides the
designer (reliable).
As a proposed solution, a novel matrix-based model was proposed for both unit cell and
DA based on Chain (𝐴𝐵𝐶𝐷) parameters. No approximation was performed since both the
µ-TL-type segments and Gm-cell were represented by their complete 𝐴𝐵𝐶𝐷 parameters. It
also presents versatility where the model was provided with no assumption of any kind on
neither the topology of Gm-cell nor segments used. This is contrary to what is done in the
conventional approaches where a simple unilateral AC-equivalent model was always
assumed for the transistor. The latter is considered to be not the case in practice since the
designer could be interested in adopting other Gm-cell topologies, such as a cascode Gm-
cell as was concluded in Chapter 2 (subsection 2.3.3).
In the next chapter, the potential of this model is presented through a scripting a CAutoD
methodology example where it is used to explore the different possible CMOS DAs the
55-nm ST technology can provide.

Page | 54
Page | 55
Chapter 4

Model Application – Computer


Automated Design for Loss-
Compensated Distributed Amplifier

In the last chapter, a four-port matrix-based model using 𝐴𝐵𝐶𝐷-parameters was


proposed for DA design. It was considered to be complete since no assumption was made,
and versatile since no limitations was imposed on it, neither on the topology of Gm-cell nor
on the µ-TL segments. In this chapter, the proposed model is used to explore different
CMOS-based DA solutions the 55-nm ST technology could produce.
This chapter is structured as follows. Section 4.1 starts by introducing the concept of
loss-compensated DA. A topology proposed in the nineties [50] and still commonly used
for its simplicity in reducing the high-losses issue in millimeter frequencies and expanding
the passband. Based on it and with the use of the proposed DA model (Eq. 3.17), Section 4.2
suggests a systematic algorithmic methodology for realizing a CAutoD process that
maximizes the DA bandwidth for a given passband-gain-ripple specification. Section 4.3
discusses this process outcome where a set of 3D DSE plots are obtained that enable one
to examine the tradeoffs between gain, bandwidth and power consumption, and the size
and number of Gm-cells, and arrive at the optimum desired design. Finally, Section 4.4
summarizes and concludes this chapter.

4.1 Loss-Compensated Distributed Amplifier

While the conventional DA is capable of providing us wideband amplification through


wideband matching, by dealing with the input- and output- parasitic capacitance of the Gm-
cell through connecting inductive-type segments, as already analytically analyzed in
Chapter 2 (Section 2.3); however, resistive-type losses remain unmanageable by it and
become important at high frequencies that they can no longer be neglected. Those latter

Page | 56
losses impact on the gain become more evident at mm-wave frequency that they influence
the flatness of the DA passband.
Loss-compensated DA topologies are thus proposed, and even a necessity, when it
comes to mm-wave application. Loss-compensation is achieved by adding extra
components to the DA device, such as inductors [50] or extra transistors [60], which
provides the designer with additional tuning knobs to control the resistive-type losses of
the DA. In this work, interest was set on [50] for their interesting idea of making use of the
already implemented unit cell Gm-cell contrary to [60] which requires to distribute
additional 𝑁 transistors along each loaded-lines of the DA.
A graphical study of [50] technique influence on both input- and output- loaded lines is
carried out first in subsection 4.1.2. Then, an analytical analysis is provided in
subsection 4.13. Finally, in subsection 4.1.4, a method for modeling the loss-compensated
cascode Gm-cell topology and implementing it in a numerical computing environment,
such as MathWorks MATLAB, is demonstrated.

4.1.1 Overview of Publication


The idea of a loss-compensated DA dates back to year 1996. It was first proposed by
[50] as a technique to mitigate losses for both the gate and drain loaded-lines. Indeed, from
(2.4) and (2.5), the gate-line attenuation constant (𝛼𝑔 ) and drain-line attenuation constant
(𝛼𝑑 ) are dominated by the Gm-cell input and output resistance, 𝑅𝑖 and 𝑅𝑑𝑠 , respectively,
as well as microstrip segments resistive part, represented by 𝛿𝑅𝑔 and 𝛿𝑅𝑑 , respectively.
The proposed technique comes as an improvement for the conventional cascode device, in
which two additional inductors, 𝐿𝑔 and 𝐿𝑑𝑠 , are added, as illustrated in Figure 4.1. The new
loss-compensation circuit improves the high frequency performance of the amplifier device
and, when speaking of the whole DA, it makes the gain-bandwidth product larger than that
of conventional ones.

Figure 4.1: Schematic circuit of the proposed loss-compensated distributed amplifier. [50]

To prove their suggested enhancement, [50] realized two DAs using 0.1-µm
InAlAs/InGaAs/InP HEMT technology with 50-GHz and 90-GHz bandwidth each. The 50-
GHz DA is made from eight unit cells and delivers 16 dB of gain, as shown in Figure 4.2,
while the 90-GHz DA is made from six unit cells and delivers a gain of 10 dB, as shown in
Figure 4.3. Indeed, this technique is showing its capability in granting the possibility to

Page | 57
amplify frequencies above 30 GHz, i.e. compensating losses inside the mm-wave spectrum
and producing high gain level similar to the lower side of the frequency spectrum.

Figure 4.2: Measured 𝑆-parameters of the loss-compensated DA IC with eight unit-cells. [50]

Figure 4.3: Measured 𝑆-parameters of the loss-compensated DA IC with six unit-cells. [50]

4.1.2 Loss-Compensation Technique: Graphical Analysis


Due to the increasing losses on both gate and drain lines with increase of frequency, the
DA does not experience a flat gain response inside its passband. This is demonstrated in
Figure 4.4, where attenuation constants 𝛼𝑔 and 𝛼𝑑 are plotted. For relevant insight, the
losses plotted here represent the ones of the 100-GHz DA designed in Chapter 5 which is
synthesized using the conventional two-transistor cascode for Gm-cell topology without
the compensation inductors 𝐿𝑔 and 𝐿𝑑𝑠 . One can note that the total losses (𝛼𝑡𝑜𝑡𝑎𝑙 = 𝛼𝑔 +
𝛼𝑑 ) at 100 GHz is about 14-fold higher than the total losses at the lower end of the spectrum.
This emphasizes that when speaking about the overall gain behavior with respect to
frequency, the upper-end of the conventional (non-compensated) DA passband tends to be
lower than the lower-end of its passband.

Page | 58
Figure 4.4: Attenuation constants of the conventional DA without loss-compensation
technique.

To comprehend the impact of the compensation inductors, 𝐿𝑔 and 𝐿𝑑𝑠 , on both


attenuation constants 𝛼𝑔 and 𝛼𝑑 , they are first studied here visually by increasing each
independently and plotting the latter constants behaviors with respect to frequency.

(a) (b)
Figure 4.5: Influence of loss-compensating inductor 𝐿𝑔 on (a) gate-line attenuation constant
(𝛼𝑔 ) and (b) drain-line attenuation constant (𝛼𝑑 ).

Referring to Figure 4.5, the inductor 𝐿𝑔 has a noticeable influence on the mm-wave
frequencies above 60 GHz with a dominant role on the attenuation constant 𝛼𝑑 . As 𝐿𝑔
increases, the losses on both the gate and drain loaded-lines decrease up to a certain
frequency limit as attenuation constants above this limit begin to experience a rapid
increase. Thereby, one can deduce that an excessive 𝐿𝑔 value can reduce the cut-off
frequency of both lines. For the attenuation constant 𝛼𝑑 , inductor 𝐿𝑔 is also observed to
cause negative value if excessively increased. The reason for this will be demonstrated in
the upcoming subsection 4.1.3, which deals with the analytical side of this analysis.
Nonetheless, as a short explanation here, this is the result behind [50] proposed idea where
they regulate the real part of the cascode Gm-cell output impedance through introducing a
tunable negative term in it. Doing so provides the designer with the ability to have control
over the magnitude of the DA attenuation constant with respect to frequency. A negative
attenuation constant, thereby, represents an over-compensation of the overall losses.
However, deliberately compensating losses more than what is needed to be removed while
one end-objective is to maintain a flat gain across a given BW is usually something not

Page | 59
performed when designing wideband devices. In other words, one would not remove
resistive losses excessively more than would be required to be removed at a given frequency
with respect to another, if by the end the intention is to have a proper flat-wideband
response for the amplifier device-under-test. This will be our case here, for instance, where
the DA will be designed in a way that the magnitude of the DA total losses at high frequency
will be made on the same level as the losses at low frequencies, and not over-compensated,
in order to have a proper flat passband response. This will be made clearer when discussing
Figure 4.7 next and will be further elaborated on when speaking of the loss-compensated
DA design concept in Section 4.2.

(a) (b)
Figure 4.6: Influence of loss-compensating inductor 𝐿𝑑𝑠 on (a) gate-line attenuation constant
(𝛼𝑔 ) and drain-line attenuation constant (𝛼𝑑 ).

From Figure 4.6, inductor 𝐿𝑑𝑠 allows the reduction of the losses on the gate loaded line.
It is also observed to have an impact on its cut-off frequency where 𝛼𝑔 is decreasing up to
a certain frequency limit and above this limit the loss starts to increase again rapidly.
Thereby, one can deduce that an excessive 𝐿𝑑𝑠 value can reduce the cut-off frequency of
the gate line. For the drain line, inductor 𝐿𝑑𝑠 is observed to have relatively weak influence
on 𝛼𝑑 compared to what was observed in the other plots.

Figure 4.7: Total attenuation constant (𝛼 𝑇𝑜𝑡𝑎𝑙 ) behavior with and without loss-compensation
technique.

When combing both compensation inductors, the overall losses (𝛼𝑡𝑜𝑡𝑎𝑙 ) needs to be
adjusted such that the losses at the (targeted) upper-end of the passband is on the same level
as the losses on its lower-end, as illustrated in Figure 4.7. This will result in the DA

Page | 60
experiencing a relatively flat wide amplification passband. It is being considered as
relatively flat since at the midband around 50 GHz, peak loss is observed to occur. This
will translate into a drop of gain in the middle of the DA passband. Increasing either 𝐿𝑔 or
𝐿𝑑𝑠 to flatten it even more would be difficult, since, as discussed previously, they have an
influence on the cut-off frequency of both gate and drain line, i.e., increasing 𝐿𝑔 or 𝐿𝑑𝑠
excessively will lead to shrinkage of the DA bandwidth. On that basis, one can conclude
that by using this loss-compensation topology a trade-off exists between the bandwidth and
midband gain-drop level, in other words gain-ripple. It is up to the designer to choose
between either having a DA that demonstrates wide bandwidth but with a notable midband
gain-drop or having a DA with a flat response but at the expense of a smaller bandwidth.
In this work, for instance, flatness was given priority when designing single stage DAs
where its midband was restricted within 1 dB of gain-drop. A common passband
characteristic adopted by similar work when referring to the single stage-stage DAs
mentioned in the state-of-the-art Table 2.2 (Chapter 2). .

4.1.3 Loss-Compensation Technique: Analytical Analysis


Previously, the impact of loss-compensation inductors is investigated graphically. Here,
mathematically speaking, inductors 𝐿𝑔 and 𝐿𝑑𝑠 are found to compensate the losses through
resonance with the intrinsic capacitance of the cascode cell: Inductor 𝐿𝑔 with the gate-to-
source capacitance (𝐶𝑔𝑠2 ) coming from transistor M2 and inductor 𝐿𝑑𝑠 with the drain-to-
source capacitance (𝐶𝑑𝑠1 ) coming from transistor M1. The following are proven as follows.

(a) (b)
Figure 4.8: Non-loss compensated cascode gm-cell: (a) symbolic representation and (b)
simplified AC-equivalent circuit.

From the non-compensated cascode equivalent intrinsic circuit of Figure 4.8, its output
impedance is represented by (4.1). Subscripts 1 and 2, respectively, indicate the commons-
source (CS) transistor M1 and common-gate (CG) transistor M2.
𝑍𝑑𝑠1 𝑍𝑑𝑠2 · 𝑔𝑚2
𝑍𝑜𝑢𝑡 = 𝑍𝑑𝑠2 + (𝑍𝑔𝑠2 + )
𝑍𝑑𝑠1 + 𝑍𝑔𝑠2 𝑗𝜔𝐶𝑔𝑠2 (4.1)

Since transistor M2 drain-source intrinsic resistor (𝑅𝑑𝑠2 ) is larger than its gate-source
intrinsic resistor (𝑅𝑔𝑠2 ), this leads to 𝑀𝑎𝑔(𝑍𝑑𝑠 ) ≫ 𝑀𝑎𝑔(𝑍𝑔𝑠 ); thereby, the impedance
coefficient ratio 𝑍𝑑𝑠1 ⁄(𝑍𝑑𝑠1 + 𝑍𝑔𝑠2 ) ≈ 1. Hence, (4.1) can be reduced to (4.2).

Page | 61
𝑍𝑑𝑠2 · 𝑔𝑚2
𝑍𝑜𝑢𝑡 = 𝑍𝑑𝑠2 + 𝑍𝑔𝑠2 +
𝑗𝜔𝐶𝑔𝑠2 (4.2)

Since the real part of impedances 𝑍𝑑𝑠2 , 𝑍𝑔𝑠2 and (𝑍𝑑𝑠2 · 𝑔𝑚2 )⁄(𝑗 · 𝜔 · 𝐶𝑔𝑠2 ) are (4.3a),
(4.3b) and (4.3c), respectively, the real part of the output impedance of the cascode, seen
by the drain loaded-line, is deduced to be (4.3d).
𝑅𝑑𝑠2 𝑅𝑑𝑠2
𝑅𝑒(𝑍𝑑𝑠2 ) = 𝑅𝑒 ( )= 2 2 (4.3a)
1 + 𝑗𝜔𝑅𝑑𝑠2 𝐶𝑑𝑠2 1 + 𝜔 2 𝑅𝑑𝑠2 𝐶𝑑𝑠2

𝑗
𝑅𝑒(𝑍𝑔𝑠2 ) = 𝑅𝑒 (𝑅𝑔𝑠2 − ) = 𝑅𝑔𝑠2 (4.3b)
𝜔𝐶𝑔𝑠2

2
𝑍𝑑𝑠2 · 𝑔𝑚2 𝑔𝑚2 · 𝑅𝑑𝑠2 · 𝐶𝑑𝑠2
𝑅𝑒 (−𝑗 )=− 2 2 ) (4.3c)
𝜔𝐶𝑔𝑠2 𝐶𝑔𝑠2 (1 + 𝜔 2 𝑅𝑑𝑠2 𝐶𝑑𝑠2

𝑅𝑑𝑠2 𝑔𝑚2 · 𝐶𝑑𝑠2 · 𝑅𝑑𝑠2


𝑅𝑒(𝑍𝑜𝑢𝑡 ) = 𝑅𝑔𝑠2 + 2 2 (1 − ) (4.3d)
1 + 𝜔 2 𝑅𝑑𝑠2 𝐶𝑑𝑠2 𝐶𝑔𝑠2

Indeed, from (4.3d) one can note that by using just the simple conventional cascode Gm-
cell the resistive losses cannot be reduced as they only depend on the intrinsic parameters
of the cascode. It is also clear that the second term between parenthesis
((𝑔𝑚2 · 𝐶𝑑𝑠2 · 𝑅𝑑𝑠2 )⁄𝐶𝑔𝑠2 ) operates as a negative term and contributes in decreasing the
real part of the output impedance that causes the loss on the drain line. However, circuit
designers cannot control this negative term effect freely because it depends solely on the
parameters of the transistors which can have unfavorable degradation on the overall gain
of the DA. That is why external components are proposed by [50] to be added, in this case
extra inductors 𝐿𝑔 and 𝐿𝑑𝑠 , in order to have more flexible control over Re(𝑍𝑜𝑢𝑡 ) with
significant change on its value, as demonstrated hereafter.

(a) (b)
Figure 4.9: Loss compensated cascode gm-cell: (a) symbolic representation and (b) simplified
AC-equivalent model.

Referring to Figure 4.9, 𝐿𝑔 and 𝐿𝑑𝑠 are added to see how they influence the behavior of
𝑅𝑒(𝑍𝑜𝑢𝑡 ). The former inductor is considered through its impedance 𝑍𝐿𝑔 = 𝑗𝜔𝐿𝑔 in series

Page | 62
with 𝑍𝑔𝑠2 = 𝑅𝑔𝑠2 + 1⁄𝑗𝜔𝐶𝑔𝑠2 and the latter inductor is considered through its impedance
𝑍𝐿𝑑𝑠 = 𝑗𝜔𝐿𝑑𝑠 in series with 𝑍𝑑𝑠1 (𝐶𝑑𝑠1 in parallel with 𝑅𝑑𝑠1 ). This results in (4.1) becoming
(4.4).
𝑍𝑜𝑢𝑡 =

(𝑍𝑑𝑠1 + 𝑗𝜔𝐿𝑑𝑠 ) 𝑍𝑑𝑠2 · 𝑔𝑚2 (4.4)


𝑍𝑑𝑠2 + {(𝑍𝑔𝑠2 + 𝑗𝜔𝐿𝑔 ) + }
(𝑍𝑑𝑠1 + 𝑗𝜔𝐿𝑑𝑠 ) + (𝑍𝑔𝑠2 + 𝑗𝜔𝐿𝑔 ) 𝑗𝜔𝐶𝑔𝑠2

Now it is clear how the compensation takes place through resonances: 𝑍𝑑𝑠 interacts with
𝑍𝑑𝑠1 and 𝑍𝑔 interacts with 𝑍𝑔𝑠2 , where both together contribute into reducing the frequency
impact on the degrading part of 𝑍𝑜𝑢𝑡 . Through (4.5a) it is found that inductor 𝐿𝑑𝑠 resonates
specifically with the parasitic capacitor 𝐶𝑑𝑠1 and through (4.5b) inductor 𝐿𝑔 resonates
specifically with the parasitic capacitor 𝐶𝑔𝑠2 .
𝑅𝑑𝑠1
𝑍𝑑𝑠1 + 𝑗𝜔𝐿𝑑𝑠 = + 𝑗𝜔𝐿𝑑𝑠 (4.5a)
1 + 𝑗𝜔𝑅𝑑𝑠1 𝐶𝑑𝑠1

1
𝑍𝑔𝑠2 + 𝑗𝜔𝐿𝑔 = 𝑅𝑔𝑠2 + + 𝑗𝜔𝐿𝑔 (4.5b)
1 + 𝑗𝜔𝐶𝑔𝑠2

The analysis demonstrated up to now, while insightful for explaining how the
compensation inductors influence the losses on gate and drain loaded-lines, it is considered
approximate and not fully suited for practical usage when designing in the mm-wave
spectrum. Adding more elements for more accurate depiction will cause the information to
be lost, since more parasitic will be intertwined, and will result in a more complicated
design process since each parasitic needs to be extracted for different transistor sizes. For
that reason, the next subsection deals with the modeling aspect of loss-compensated
cascode by introducing a simple matrix-based method for implementing the complete Gm-
cell topology on a coding software such as MathWorks MATLAB.

4.1.4 Implementing Loss-Compensated Cascode Gm-Cell in a Numerical


Computing Environment
Compared to conventional ways of using AC equivalent circuit to derive parametric
expressions to replicate the cascode Gm-cell frequency behavior for use in circuit design,
the method presented hereafter treats the adopted loss-compensated cascode as a matrix-
based model.
This method is considered an important step since in either DA model (3.14) or (3.17),
the Gm-cell is represented in its general form through the middle 𝐴𝐵𝐶𝐷-matrix. In addition,
this method will allow the inductors to be implemented as tuning knobs, a necessary step if
full computer automation is needed. Since no approximation for simplification purposes is
carried out, this method benefits from the complete and accurate representation of cascode
behavior, hence of DA behavior. It also benefits from ease of scalability where a collection
of matrices can be extracted for several transistor sizes and then depending on the desired
cascode size the corresponding transistors matrices are used.

Page | 63
Figure 4.10: Flow diagram for converting the loss-compensated cascode Gm-cell into a 𝑌-
matrix representation.

This implementation method is illustrated in Figure 4.10 as a flow diagram. When the
designer has biased his cascode cell to the desired operating point, a collection of both CS
and CG transistor 𝑆-parameter matrices are extracted first for the envisaged sizes and
frequency range during DA design process. Since CS transistor in standalone has its gate
pin as input port, drain pin as output port and source pin grounded, its corresponding 𝑆-
parameters will be a two-port matrix extraction. Since CG transistor in standalone has its
source pin as input port, drain pin as output port and gate pin as a third port eventually
connected to a shunt inductor 𝐿𝑔 , its corresponding 𝑆-parameters will be a three-port matrix
extraction.
Next step involves converting CG transistor three port 𝑆-matrix into a two-port matrix
form. This can be done by using (4.6a), where Г represents the reflection coefficient
between the added inductor 𝐿𝑔 and the reference characteristic impedance (𝑍0 ) the CG
three-port 𝑆-parameters was extracted with (4.6b). In case µ-TL topology is used to
implement inductor 𝐿𝑔 , (4.6c) is a suitable option where 𝑍C , 𝛾𝑇𝐿𝐿𝑔 and 𝑙 𝑇𝐿𝐿𝑔 are the
characteristic impedance, propagation constant and length of this inductor.
𝑆𝑖3 𝑆𝑖3 Г
𝑆𝑖𝑗′ = 𝑆𝑖𝑗 + ; (𝑖, 𝑗) ∈ (1,2)
1 − 𝑆33 Г (4.6a)

where
𝑗𝜔𝐿𝑔 − 𝑍0
𝛤=
𝑗𝜔𝐿𝑔 + 𝑍0 (4.6b)

𝑍𝑐 𝑡𝑎𝑛ℎ(𝛾𝑇𝐿𝐿𝑔 𝑙 𝑇𝐿𝐿𝑔 ) − 𝑍0
𝛤=
𝑍𝑐 𝑡𝑎𝑛ℎ(𝛾𝑇𝐿𝐿𝑔 𝑙 𝑇𝐿𝐿𝑔 ) + 𝑍0 (4.6c)

Page | 64
Afterwards, both CS and CG two-port 𝑆-parameters matrices are converted into an
𝐴𝐵𝐶𝐷-matrix format using the conversion provided by [2] and multiplied with 𝐴𝐵𝐶𝐷-
matrix of the series inductor 𝐿𝑑𝑠 , as shown in (4.7a). In case µ-TL topology is used to
implement inductor 𝐿𝑑𝑠 , (4.7b) is a suitable option where 𝑍C , 𝛾𝑇𝐿𝐿𝑑𝑠 and 𝑙 𝑇𝐿𝐿𝑑𝑠 are the
characteristic impedance, propagation constant and length of this µ-TL (𝑇𝐿𝐿𝑑𝑠 ).
𝐴 𝐵 𝐴 𝐵 1 𝑗𝜔𝐿𝑑𝑠 𝐴 𝐵
[ ] =[ ] [ ] [ ]
𝐶 𝐷 𝐶𝑎𝑠𝑐𝑜𝑑𝑒 𝐶 𝐷 𝐶𝑆 0 1 𝐿𝑑𝑠 𝐶 𝐷 𝐶𝐺 (4.7a)

𝐴 𝐵
[ ]
𝐶 𝐷 𝐶𝑎𝑠𝑐𝑜𝑑𝑒
𝑙𝑇𝐿 𝑙𝑇𝐿𝐿𝑑𝑠
𝐴 𝐵 𝑐𝑜𝑠ℎ (𝛾𝑇𝐿𝐿𝑑𝑠 2𝐿𝑑𝑠 ) 𝑍𝑐 𝑠𝑖𝑛ℎ (𝛾𝑇𝐿𝐿𝑑𝑠 ) 𝐴 𝐵
=[ ] [
2
] [ ] (4.7b)
𝐶 𝐷 𝐶𝑆 𝑠𝑖𝑛ℎ (𝛾 𝑙𝑇𝐿𝐿𝑑𝑠 𝑙𝑇𝐿𝐿𝑑𝑠 𝐶 𝐷 𝐶𝐺
𝑇𝐿𝐿𝑑𝑠 2
)⁄𝑍 𝑐 𝑐𝑜𝑠ℎ (𝛾𝑇𝐿𝐿𝑑𝑠 2
)
𝑇𝐿𝐿𝑑𝑠

Finally, the obtained two-port 𝐴𝐵𝐶𝐷-matrix describing the proposed Gm-cell amplifier
is converted to a two-port 𝑌-matrix from using [56] and now can be joined with proposed
DA 𝐴𝐵𝐶𝐷-models, (3.14) or (3.17), by filling up their middle 𝐴𝐵𝐶𝐷-matrix with the
adequate 𝑌𝑖𝑗 parameters. Due to no simplification performed on the transistors and the
possibility to represent the complete cascode Gm-cell through (4.7b), DA designers can
now benefit from high reliability in accurate behavior representation, one of the desired
criterion highlighted in Chapter 3 (Section 3.1). In addition, designers can benefit by using
to proposed design methodologies that produce accurate DA sizing outcome, as will be
presented later on when speaking of our own algorithmic DA design methodology.

4.2 Loss-Compensated Distributed Amplifier CAutoD Process

Before discussing the offered loss-compensated DA algorithmic design methodology,


the design concept of the loss-compensated cascode Gm-cell [50] adopted as a technique
that can be used in extending the DA bandwidth while providing control over its mid-band
flatness is first examined. Using the proposed cascode 𝑌-matrix based model, (4.7a) or
(4.7b), in conjunction with the DA 𝐴𝐵𝐶𝐷-model (3.14), the influence of inductors 𝐿𝑔 and
𝐿𝑑𝑠 on DA 𝑆21 response is first investigated.

4.2.1 Distributed Amplifier Bandwidth Extension and Flattening Concept


The DA under study is the one implemented and measured in Chapter 5, whose design
parameters were obtained from the design methodology of the upcoming subsection 4.2.2.
Those parameters are only summarized here as follows. The DA consists of five unit cells
(𝑁 = 5), where the unit cell is modeled using (3.17) and its inductive-type segments are
implemented as µ-TLs using the thickest copper metal available in the BEOL of ST 55-nm
used process with 131 µm in length. The middle 𝐴𝐵𝐶𝐷-matrix of (3.17) is filled up with
the 𝑌-parameters of the two-port admittance of the loss-compensated cascode consisting
made from 31-µm CS nMOS, M1, and 62-µm CG nMOS, M2, biased at a current density
of 0.16 mA per CS micrometer width (0.16 mA/µm). During this concept analysis, both 𝐿𝑔

Page | 65
and 𝐿𝑑𝑠 were left as ideal components, (4.6b) and (4.7a), and used as tuning knobs. The
DA is intended to cover a bandwidth of 100 GHz.

Figure 4.11: 𝑆21 response versus frequency when 𝐿𝑑𝑠 = 0 pH and 𝐿𝑔 increases.

Referring to Figure 4.11, for the given 𝑁 and cascode size, the DA gain is determined
by its gain level at lower of spectrum – here, it is 7.1 dB. When 𝐿𝑔 and 𝐿𝑑𝑠 are set to zero,
one can observe the gradual decrease in gain with respect to frequency. The cause of this
is related to the high losses both physical µ-TLs and cascode Gm-cell itself experience at
mm-wave frequencies, as illustrated in Figure 4.4. By shorting 𝐿𝑑𝑠 , i.e., 𝐿𝑑𝑠 = 0 pH,
increasing 𝐿𝑔 rises gain at a certain high frequency component. Indeed, inductor 𝐿𝑔 is
extending the DA BW, enabling it to reach higher frequencies―in this case 100 GHz―by
compensating for losses at those frequencies, as demonstrated in Figure 4.5.

Figure 4.12: 𝑆21 response versus frequency when 𝐿𝑑𝑠 increases and 𝐿𝑔 =40 pH.

Regardless of the achievable extension, however, Figure 4.11 highlights an undesirable


behavior between lower and upper passband limits: the mid-band exhibits a gradual gain
roll-off. This is due to 𝐿𝑔 being a loss-compensating technique with a major impact on the
magnitude of the attenuation constant 𝛼𝑔 at frequencies above 60 GHz, as was deduced
from Figure 4.5, and insignificant impact on the lower spectrum. To solve this issue causing
midband gain drop, the addition of 𝐿𝑑𝑠 is required. By maintaining 𝐿𝑔 at the value that
peaks DA gain at the higher-end of the spectrum to the same level of gain as the lower-end
of the spectrum, in this case 40 pH, increasing 𝐿𝑑𝑠 leads to lifting the mid-band gain drop,

Page | 66
as illustrated in Figure 4.12―here, it is desired to have 1-dB mid-band gain drop.
Therefore, to summarize up to now, one can conclude the following: Inductor 𝐿𝑔 mainly
influences upper-end peaking while inductor 𝐿𝑑𝑠 plays its main role on mid-band, but
unfortunately not exclusively.

Figure 4.13: 𝑆21 response versus frequency when 𝐿𝑑𝑠 and 𝐿𝑔 are tuned to maintain
simultaneously the extended bandwidth of 100 GHz and 1-dB mid-band gain drop at 50 GHz.

Figure 4.12 also highlights an undesirable behavior: increasing 𝐿𝑑𝑠 has an influence on the
level of peaking produced from 𝐿𝑔 where it increases it even further and on the position of
its gain-peaking frequency where it shifts it towards lower spectrum. To fix it, 𝐿𝑔 and 𝐿𝑑𝑠
must simultaneously be decreased and increased, respectively, to maintain the desired mid-
band drop (1-dB) while conserving the acquired high frequency component (100 GHz), as
demonstrated in Figure 4.13. Indeed, it is observed that through the usage of 𝐿𝑔 and 𝐿𝑑𝑠 ,
one can extend the BW while flattening the passband gain-ripple

4.2.2 Algorithmic Design Methodology


A step-by-step algorithmic DA design methodology is proposed based on the design
concept discussed previously. This methodology uses both DA 𝐴𝐵𝐶𝐷-model (3.14) with
the unit cell model (3.17), where the physical μ-TL topology is used for the inductive-type
segments, and the presented loss-compensated cascode Gm-cell with its matrix
implementation method presented in subsection 4.1.4. For the latter, its inductors 𝐿𝑔 and
𝐿𝑑𝑠 are implemented as physical μ-TL and thereby (4.6c) and (4.7b), respectively, will be
used during the design process.
The purpose of this methodology is to explore all loss-compensated DAs a given
technology process can provide at a given biasing point, by maximizing bandwidth for
various combinations of transistor widths and number of unit cells while satisfying a
specified passband flatness. It is therefore noted that the initiating point for such method is
the designer’s choice of bias point beforehand, what follows will be automated and the end
outcome will be 3D DSE plots to pick DA solutions from.
The design methodology is summarized as a flowchart in Figure 4.14 and its systematic
steps are presented hereafter. The DSE plots outcome of this methodology are discussed in
the next Section 4.3. Since 55-nm ST process node is used at the end for fabrication, its

Page | 67
process design kit (PDK) is used throughout the design steps for illustration. For the
numerical computing environment, in our case MathWorks MATLAB software is used. It
should be noted that technology components, such as μ-TL and transistors, were first
simulated on Keysight ADS software, since PDK is available on it by semiconductor
foundry and, as a second step, their 𝑆-parameters were exported into MATLAB to perform
the required conversions and matrix manipulations.

Figure 4.14: Flowchart presenting our DA CAutoD procedure using proposed DA and loss-
compensated cascode models.

Step 1) Extract the 𝑆-parameters for a range of PDK μ-TL lengths at a given thickness
and width. They will be used to model the unit-cell segments as well as 𝑇𝐿𝐿𝑔 (4.6c) and
𝑇𝐿𝐿𝑑𝑠 (4.7b). Choosing the thickest μ-TL in technology BEOL offers the lowest losses
versus frequency, since it will have the lowest sheet resistance, and hence aids DA gain
response. For eight-metal version of 55-nm CMOS process, it is the ultra-thick metal M8.
For what concerns the width, in this metal layer, the minimum width allowed is 0.6 µm,
but to comply with the DC-current handling limits expressed on the Design Rule Manual
(DRM), we decided to adopt a width of 1 µm. We also decided to apply the same 1 µm
width for all µ-TLs with the aim of having later on a simpler design implementation process
where all the lines will share the same width. After that, the two-port 𝑆-parameters are
extracted for length ranging from 10 µm up to 400 µm from ADS. Since chain (𝐴𝐵𝐶𝐷)

Page | 68
model is used for both unit-cell and cascode models, the μ-TL characteristic impedance
(𝑍𝐶 ) and propagation constant (𝛾) can be obtained using (4.8) when needed, where
conversion between two-port 𝑆-parameters to 𝐴𝐵𝐶𝐷-parameters is done on MATLAB.

𝐴·𝐵
𝑍𝐶 = √ (4.8a)
𝐶·𝐷

1 𝐴+𝐷 𝐴·𝐵
𝛾= 𝑐𝑜𝑠ℎ−1 ( )√ (4.8b)
𝑙 2 𝐶·𝐷

Step 2) At a desired drain current per unit gate width (𝐼𝑑 (𝐴⁄𝑤𝑖𝑑𝑡ℎ)), extract the 𝑆-
parameter matrices of both nMOS CS and CG FETs separately for a range of different
channel widths, as explained in Figure 4.10, in order to link the loss-compensated cascode
as a two-port 𝑌-matrix with the DA model (3.17). Here, the CG FET, M2, channel width is
set to be twice the size of the CS FET, M1, in order to have similar magnitude for the
cascode Gm-cell input and output capacitive-type reactance. From that, the unit-cell μ-TL
segments can be set equal in length, i.e., 𝑙𝑆𝐼 ⁄2 = 𝑙𝑆𝐼 ⁄2 = 𝑙 ⁄2, inside (3.17) which will aid
us later in having a simpler layout implementation process. Those two design choices were
taken in order to equate the loaded input- and output-lines phase delays; a necessity for
producing constructive interference on the output line, as was discussed in the basic
operation section of a DA in Chapter 2.
In this thesis, the cascode bias was desired to be biased at 0.16 mA per CS micrometer
width (0.16 mA/𝜇m) at 𝑉𝐷𝐷 =1.2 V (nominal value of technology). While this choice of
current density does not represent the full capability of ST 55-nm CMOS
(𝑓𝑡 = 160 GHz⁄𝑓𝑚𝑎𝑥 = 246 GHz at 𝐼𝑑 (𝐴⁄𝑤𝑖𝑑𝑡ℎ) = 0.22 mA⁄µm): our choice of current
density is to nearly maximize 𝑔𝑚 while being consumption-efficient since our end goal is
concerned about a metric of GBP over 𝑃𝐷𝐶 . On that basis and as a first step, the CS and CG
FETs were inserted as separate standalone devices on ADS schematics from the technology
PDK library and their drain current per unit gate width was set to our 0.16 mA/𝜇m choice.
As a second step, a sweep for a desired range of widths is performed; in this work, their
width was varied for a range of 14 𝜇m≤ 𝑊𝐶𝑆 ≤ 40 𝜇m and, hence, 28 𝜇m≤ 𝑊𝐶𝐺 ≤ 80 𝜇m,
and their respective 𝑆-parameters were extracted. It should be reminded that for the CS
FET it will be a two-port matrix extraction and for the CG FET it will be a three-port matrix
extraction, as already described in Figure 4.10. Wider widths are possible, but were
disregarded in being mentioned in this discussion. This is because during the time of
proposal and testing of this methodology, 𝑊𝐶𝑆 > 40 𝜇m were observed to offer DAs with
BWs ≤ 80 GHz, as will be proven in Section 4.3 (Figure 4.16); thereby, of no interest for
us to highlight them here and in the remaining of this manuscript. In addition, DAs with
𝑊𝐶𝑆 < 14 𝜇m required 𝑇𝐿𝐿𝑑𝑠 μ-TLs with lengths above our set design limit (400 𝜇m)
extracted in Step 1) and hence did not converge, and, for us, were disregarded since they
are solutions with excessively large on-wafer areas.
After this step, the DA model (3.17) is now ready with its tunable μ-TLs lengths (unit-cells
segments, 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 ), transistors width and 𝑁 knobs to perform automated design for
a defined matching and passband flatness levels.

Page | 69
Step 3) By fixing the number of unit cells of DA to one in (3.14), i.e. 𝑁 = 1, iterate
through each 𝑊𝐶𝑆 and increase the μ-TL segments length (𝑙 ⁄2) until the input- and output-
ports reflection coefficient bandwidth stops expanding at a defined targeted maximum
level. In (3.20), 𝑍0 is set to 50 Ω that represents the impedance of both RF probes for the
measurement setup. In our case, the maximum acceptable mismatch level defined for a
single unit cell input and output ports is set at 𝑆11𝑚𝑎𝑥 = 𝑆22𝑚𝑎𝑥 = −10 dB. The scripted
methodology at each transistor width is set to increase 𝑙 ⁄2 until the far-end frequency of
𝑆11 and 𝑆22 bandwidth where mismatch level is at −10 dB stops expanding.

Step 4) For a given value of 𝑊𝐶𝑆 , iterate through 𝑁 and for each combination of (𝑊𝐶𝑆 , 𝑁)
detect the length of 𝑇𝐿𝐿𝑔 that increases gain at the rightmost side of the frequency spectrum
to attain same level of gain of that at the lower-end of the DA passband. The lower-end
passband frequency is usually determined by the AC-grounded capacitor used in series with
termination resistors 𝑅3 and 𝑅4 when implementing the loaded-lines termination loads 𝑍3
and 𝑍4 . Both 𝑅3 and 𝑅4 were set as ideal 50-Ω resistors since it is desired here to design
DAs at 𝑍0 = 50 Ω, as stated in Step 3. As a first approach during the design process and
before starting the layouting process of making the choice of AC-ground topology to
synthesis, the capacitors were considered here as a perfect GND connection (i.e. 𝑍3 = 𝑅3
and 𝑍4 = 𝑅4 ). This is observed to provide equal gain over all frequencies ≤ 10 GHz and
with a flat shape, as visible from Figure 4.11. Since it is commonly known that moving
higher in frequency benefits IC designers from implementing components of reduced on-
chip area, we decided to go with 10 GHz as the lower-corner frequency (𝐹𝑙𝑜𝑤𝑒𝑟 ) of the DA
passband. The latter choice will benefit us later on during layout implementation with
synthesizing small AC-grounded capacitor for 𝑅3 and 𝑅4 compared to if 𝐹𝑙𝑜𝑤𝑒𝑟 of 1 GHz
was to be picked, for instance. Accordingly, for a given (𝑊𝐶𝑆 , 𝑁), the µ-TL 𝑇𝐿𝐿𝑔 length is
increased until a certain high frequency component gain becomes on the same gain level as
at 10 GHz (see Figure 4.11). The frequency at which this peaking first occurs is considered
the DA upper-corner frequency (𝐹𝑢𝑝𝑝𝑒𝑟 ).

Step 5) For a given value of 𝑊𝐶𝑆 , iterate through 𝑁 and for each combination of (𝑊𝐶𝑆 , 𝑁)
detect the length of 𝑇𝐿𝐿𝑑𝑠 in order to flatten passband for a desired mid-band gain drop. In
this manuscript, the desired mid-band gain drop variable was assigned to a value of 1 dB
and the mid-band frequency is determined by (𝐹𝑢𝑝𝑝𝑒𝑟 + 𝐹𝑙𝑜𝑤𝑒𝑟 )⁄2; accordingly, for a given
𝑊𝐶𝑆 and 𝑁, the length of 𝑇𝐿𝐿𝑑𝑠 is increased until the gain at mid-band is made 1 dB less
than gain at 𝐹𝑙𝑜𝑤𝑒𝑟 (see Figure 4.12). As mentioned in subsection 4.2.1, 𝑇𝐿𝐿𝑑𝑠 has an
influence on gain peaking established by 𝑇𝐿𝐿𝑔 and vice-versa. The proposed way to
counteract it is just to script the algorithm to reduce 𝑇𝐿𝐿𝑔 in order to maintain gain at 𝐹𝑢𝑝𝑝𝑒𝑟
on the same level as gain at 𝐹𝑙𝑜𝑤𝑒𝑟 while increasing 𝑇𝐿𝐿𝑑𝑠 to maintain mid-band gain drop
limited to 1 dB with respect to gain at 𝐹𝑙𝑜𝑤𝑒𝑟 (see Figure 4.13).

Step 6) Repeat Steps 4) to 5) for all combinations of 𝑊𝐶𝑆 extracted in Step 2) with the
desired range of unit cells. In this manuscript, we extracted 27 transistor widths (14 𝜇m≤
𝑊𝐶𝑆 ≤ 40 𝜇m) and explored DA configurations made of 2 ≤ 𝑁 ≤ 9. For 𝑁 > 9, BWs ≤
80 GHz were obtained and hence were disregarded. This results in the automated DA
design program in producing 216 feasible loss-compensated DAs solutions.

Page | 70
Step 7) Final step includes evaluating for each design option, in this case 216 solutions,
its power consumption (𝑃𝐷𝐶 ). This can be accounted for by using (4.9) where
𝐼𝑑 (𝐴⁄𝑤𝑖𝑑𝑡ℎ) = 0.16 𝑚𝐴⁄µ𝑚 at 𝑉𝐷𝐷 = 1.2 𝑉 (defined in Step 2).
𝑃𝐷𝐶 ≅ 𝑁 · 𝑉𝐷𝐷 · 𝐼𝑑 · 𝑊𝐶𝑆
(4.9)

4.3 3D Design Space Exploration Plots: Parameters and Variables

When successfully implementing the proposed design methodology, the DA designer


will end up with an outcome of six DSE plots. Each 3D plot illustrates a given characteristic
against transistor size on x-axis and number of unit cells (𝑁) on the y-axis. Subsection 4.3.1
covers the three DSE plots that represent parameter of interest from gain, bandwidth and
𝑃𝐷𝐶 . Subsection 4.3.2 covers the remaining DSE plots that represent the required design-
variables from µ-TL segments length and compensation µ-TL-type inductors 𝑇𝐿𝐿𝑔 and
𝑇𝐿𝐿𝑑𝑠 lengths.

4.3.1 DA Design Parameters: Gain, Bandwidth and 𝐏𝐃𝐂


Figure 4.15 illustrates the variation of DA gain against 𝑊𝐶𝑆 and 𝑁. One can observe that
when either of those two parameters increase, the overall DA gain increases.

Figure 4.15: 3D DSE plot showing possible DA gain with respect to 𝑊𝐶𝑆 and 𝑁.

This is in agreement with is in agreement with (2.11) where the gain expression is
proportional to transconductance term (𝑔𝑚 ). The transconductance parameter is defined by
the transistor width. Having larger width equates to having higher 𝑔𝑚 , hence higher gain.
This plot also is in agreement with (2.12) where DA gain is clearly defined with the number
of unit cells.
Thereby, one can conclude that the small transconductance provided by narrow transistor
can be compensated by adding more unit cells to achieve higher gain and this is also
equivalent to realizing a DA with big transistors and fewer number of unit cells.

Page | 71
Figure 4.16: 3D DSE plot showing possible DA 1-dB bandwidths with respect to 𝑊𝐶𝑆 and 𝑁.

Figure 4.16 illustrates the variation of DA bandwidth against 𝑊𝐶𝑆 and 𝑁. It highlights
that for each 𝑁 there is a maximum BW attainable that is wider for smaller 𝑊𝐶𝑆 sizes;
however, until a certain limit where even smaller sizes lead to smaller BWs. This could be
explained by the inability of smaller transistors from providing enough gain to compensate
for important losses at high mm-wave regions. Therefore, this pushed both 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠
to be increased even more to elevate the gain at 𝐹𝑢𝑝𝑝𝑒𝑟 towards the same level of gain at
𝐹𝑙𝑜𝑤𝑒𝑟 while maintaining the 1-dB mid-band gain drop constraint. As already discussed in
Section 4.1 (subsection 4.1.2), this leads to shrinkage of BW where excessive increase of
either compensation inductors leads to shift of loaded-lines cut-off frequencies towards
lower frequencies.
Decreasing 𝑁 while also decreasing 𝑊𝐶𝑆 can permit operation at wider BWs but at the
expense of lower gain, as shown by Figure 4.15. Therefore, the first parameter to choose
from would be the bandwidth required for the DA to cover and, then, as a second step, to
choose the desired gain level from the available options obtained for that specific
bandwidth.

Figure 4.17: 3D DSE plot showing possible DA PDC with respect to 𝑊𝐶𝑆 and 𝑁.

Page | 72
Figure 4.17 illustrates the variation of DA power consumption against 𝑊𝐶𝑆 and 𝑁. It
shows that a DA constituting of more Gm-cells, i.e., larger 𝑁, leads to higher 𝑃𝐷𝐶 . The
same applies for having a DA made from larger transistors. The highest consumption,
hence, would be the design in which the DA is synthesized with both big transistors and
large 𝑁 at the same time. One reason for this design point could be for delivering larger
gain. For instance, from Figure 4.17, choosing a DA made from 𝑁 = 9 with 𝑊𝐶𝑆 =
40 µm leads to 70 mW of power consumption but with a high gain close to 12 dB;
however, from Figure 4.16, this choice would restrict the design to a passband of only
80 GHz.
As a conclusion, from those three figures, one can conclude that a compromise between
those three parameters (gain, bandwidth and 𝑃𝐷𝐶 ) exists and the choice would depend on
the designer application orientation. Having an amplifier that delivers high gain will
equates to having a smaller bandwidth and higher 𝑃𝐷𝐶 and vice-versa. In this work, major
interest is laid on the bandwidth as a first choice. This will be further elaborated on in
Chapter 5 where a 100-GHz BW DA is needed for on-wafer characterization, and then, as
a second step, the choice was made to have high gain with low 𝑃𝐷𝐶 .

4.3.2 DA Design Parameters: Synthesizing a Unit-cell


In this section, the design parameters needed to obtain the performance characteristics
covered previously are presented. The loss-compensated unit cell in discussion is
synthesized from µ-TL segments and a cascode Gm-cell with loss compensation inductors
implemented as µ-TL (𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 ). Its electrical schematic is presented in Figure 4.18.
Those parameters (length of segments, 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 ) are plotted independently with
respect to 𝑊𝐶𝑆 and 𝑁, as shown in Figure 4.19 and Figure 4.20.

Figure 4.18: Schematics of DA unit cell with loss-compensated cascode Gm-cell.

Figure 4.19 presents the length needed for µ-TL segments (𝑙 ⁄2) needed for obtaining a
DA with reference impedance (𝑍0 ) around 50 Ohms. It is noticeable that when the size of
transistor increases, the length of segments increases. It is interesting to note the linear
dependence of segments length against 𝑊𝐶𝑆 . This is consistent with loaded lines
characteristic impedance equation (2.2) and (2.4) discussed in Chapter 2, which states that
inductance of segments must increase when the Gm-cell input and output capacitances are
increased in order to maintain a required DA characteristic impedance. We remind here
that the length of the µ-TL segments is determined by setting 𝑁 =1, what explains a 2D
plot.

Page | 73
Figure 4.19: DA 2D DSE plot showing μ-TL segments lengths to achieve widest matching BW
for ‖𝑆11𝑚𝑎𝑥 ‖ = ‖𝑆22𝑚𝑎𝑥 ‖ = −10 dB

For the compensation inductors 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 , referring to Figure 4.20, one can note
that larger inductances are needed for DAs built using either smaller transistor sizes or
larger number of unit cells. This is due to the small transistor incapability in providing
sufficient gain, in other words transconductance, to mitigate the losses at high frequency.
This becomes even more drastic when N increases, since total losses of gate and drain are
defined by 𝑁(𝛼𝑔 𝑙𝑔 ) and 𝑁(𝛼𝑑 𝑙𝑑 ), as shown in (2.11).

(a) (b)
Figure 4.20: 3D DSE plot showing possible lengths for (a) shunt µ-TL (𝑇𝐿𝐿𝑔 ) and (b) series µ-
TL (𝑇𝐿𝐿𝑑𝑠 ), with respect to 𝑊𝐶𝑆 and 𝑁.

Page | 74
4.4 Conclusion

This chapter started by introducing the idea of loss-compensation for DAs where it
showed to be a promising topology for extending and flattening the passband response. The
necessary graphical and mathematical analysis of its function was provided. It was shown
to influence the passband response through regulating both the input (gate) and output
(drain) loaded-line losses (Figure 4.7). In addition, a technique for modeling this cascode
device in a matrix-form and use it in MATLAB was demonstrated that enables a complete
representation of the cascode device (Figure 4.10) and designing one through scripting.
Based on this concept and using the 𝐴𝐵𝐶𝐷-matrix model provided in Chapter 3, a step-
by-step algorithmic methodology was provided for the designer to implement in any
scripting language he desires and explore the possible DA design solutions a given
technology node can offer. For demonstration, in this work, the 55-nm CMOS ST process
was used and 216 possible feasible DAs with BW ≥ 80 GHz were resulted to explore from.
Owing to this design process, graphical plots were produced that presented in an
interesting 3D view the behavior of gain (Figure 4.15), bandwidth (Figure 4.16) and 𝑃𝐷𝐶
(Figure 4.17) with respect to both the transistor size and number of unit-cells used. Using
them, the designers can now study and analyze the correlation between each design
parameter and arrive at the optimum design point, as also demonstrated through Figure 4.19
for choice of µ-TL segments length and Figure 4.20 for choice of compensating inductor
𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 .

In the next chapter, this DSE plots were used to choose a 100-GHz CMOS DA that offers
the highest FoM of 𝐺𝐵𝑃⁄𝑃𝐷𝐶 for implementation and on-wafer characterization.

Page | 75
Page | 76
Chapter 5

100-GHz Single Stage CMOS


Distributed Amplifier

In the last chapter, using both the DA (3.14) matrix-based model with its unit-cell (3.17)
representation, a systematic algorithmic methodology was proposed. Its aim is to explore,
through scripting computer-automated algorithms, the possible amplifier gains and
bandwidths obtainable from a given process node for any desired range of 𝑁 and transistor
size (𝑊𝐶𝑆 ). A design process was determined to be challenging to perform especially when
using CAD tools in which 𝑁 has to be manipulated manually. For demonstration, the 55-
nm ST technology was used and the methodology was coded to automatically size CMOS-
based DAs for a range of 2 ≤ 𝑁 ≤ 9 and 14 𝜇m≤ 𝑊𝐶𝑆 ≤ 40 𝜇m. A total of 216 feasible
solutions were obtained with design space solutions covering 80 GHz ≤ BW ≤ 130 GHz
and ≈0 dB ≤ Gain ≤ 12 dB. In this chapter, focus is put on DA solutions that amplify up
to 100 GHz and choosing the one that demonstrates global optimum 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit for
on-wafer implementation and characterization.
This chapter is structured as follows. Section 5.1 starts by describing the 55-nm ST
technology process used. Section 5.2 covers the circuit design part. It demonstrates how the
obtained 3D DSE plots from Chapter 4 were used to pinpoint the global optimum 100-GHz
DA that satisfies our defined FoM criterion. Section 5.3 covers the circuit implementation
part of the latter DA. Here, a new stacked parallel plate shunt capacitor [61] with high
quality factor at mm-waves and high self-resonance frequency (𝑆𝑅𝐹) reaching 368 GHz is
introduced. Section 5.4 discusses the experimental results of the fabricated DA circuit. A
comparison with the performance of the state-of-the-art CMOS DAs, from Chapter 2, is
presented in Section 5.5. Finally, Section 5.6 concludes and summarizes this chapter.

Page | 77
5.1 STMicroelectronics 55-nm Node Process Technology

The technology process adopted in this work is introduced first since it will aid in
providing a better visualization for when discussing integrated components and circuit
realization. In this work, the eight-metal version of the 55-nm ST process is used.

(a) (b)
Figure 5.1: (a) Simplified cross-section of metal stack and (b) transistor 𝐹𝑡 and 𝐹𝑚𝑎𝑥 as a
function of drain current per unit gate width for nMOSFET available with ST eight-metal
version 55-nm process.

As illustrated in Figure 5.1, this process uses copper metallization with eight metal layer
backend and aluminum metallization with the top-most metal later (AP) used for pads
probing. Its back end of line (BEOL) is composed of one ultra-thick top metal layer (M8)
on top of two thick metals (M7 and M6), in order to sustain high current densities and to
reduce sheet resistance for high-quality passive circuit design in Analog/RF applications.
Beneath them, five thin lower layers (M5-M2 and M1) are available for dense
interconnections and contact with silicon devices. Under optimal bias conditions, nMOS
transistors offer a unity-gain cutoff frequency (𝑓𝑡 ) and maximum power-gain cutoff
frequency (𝑓𝑚𝑎𝑥 ) of 160 GHz and 246 GHz, respectively, suited for mm-wave applications.

5.2 100-GHz Single-Stage CMOS DA: Circuit Design

This section concentrates on explaining how we chose the optimum values for 𝑁 and
𝑊𝐶𝑆 to design a CMOS-based DA with high 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit. For this optimum design
point, it also demonstrates how we determined the required dimensions for the inductive-
type segments and the loss-compensation inductors, 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 , to synthesize its unit
cell.
At the beginning of this manuscript, we stated our interest for DAs that cover a minimum
BW of 50 GHz for the opportunity to offer futuristic 100 Gb/s and above data-rates
systems. In order to benefit from meaningful results for performance analysis with already
available works and gain better perspective on the DA model capabilities, our choice is
designated on bandwidth performance similar to the evaluated DAs from the literature
presented in Chapter 2. It is noticeable that related works on silicon-based DA provided
amplification up to around 100 GHz. Therefore, for a fair comparison, we decided to
implement an amplifier device with similar passband. It should be noted that even higher

Page | 78
BWs are possible for realization using 55-nm process, reaching as high as 130 GHz as
indicated by Figure 4.16, but were ignored in the scope of model validation.
To determine the possible options for such BW criterion, the first step is to sort the
available solutions from Figure 4.16 for an upper-corner frequency of 100 GHz. Since the
available solutions do not present exact values of 100 GHz but around it, we decided to
relax our search margin with a spread of ±2 GHz, i.e. we picked DA solutions that present
98 GHz ≤ BW ≤ 102 GHz. This led us to obtain a narrower DSE range of 35 possible
options to choose from instead of the 216 options, as shown in Figure 5.2, with their
respective gain and the required values of 𝑁 and 𝑊𝐶𝑆 .

Figure 5.2: Insertion gain of thirty-five DA design options, with respect to 𝑁 and 𝑊𝐶𝑆 , that
offer 98 GHz ≤ 𝐵𝑊 ≤ 102 GHz.

As a second step, from this new set of 35 solutions, we analyze how DC efficient is each
DA option when delivering its own insertion gain. This can be performed through
Figure 5.3, where 𝐺𝐵𝑃⁄𝑃𝐷𝐶 performance for each of the 35 DAs is plotted with the required
values of 𝑁 and 𝑊𝐶𝑆 . It is visible, in our case, that this performance varies between
7.2 GHz/mW and 8.4 GHz/mW, with the relatively lowest values being for the DA options
delivering higher gain.

Figure 5.3: 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit of thirty-five DA design options, with respect to 𝑁 and 𝑊𝐶𝑆 , that
offer 98 GHz ≤ 𝐵𝑊 ≤ 102 GHz.

From here, the final step now depends on the application purpose of the DA; either the most
𝑃𝐷𝐶 -efficient DA with degradation in gain or the highest gain with a degradation in 𝑃𝐷𝐶 .
Since our interest by the end of this thesis work is to publish our research efforts in order
to assist in future mm-wave DA designs, our search was focused towards DAs that

Page | 79
demonstrate comparable gain to the already published single-stage CMOS DAs; this is in
order to highlight the added value from using our design technique. Doing so leads us to an
even more restricted range and from it, we chose the DA that offers high 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit.
Referring back to the sate-of-the-art Table 2.2, the minimum gain published for single-stage
CMOS DA is 7 dB; consequently, from Figure 5.2, our options were restricted to only
four options (highlighted in green): a six unit-cells DA with 26 µm ≤ 𝑊𝐶𝑆 ≤ 28 µm and a
five unit-cell DA with 𝑊𝐶𝑆 =31 µm. Since it was of our interest here to also implement a
𝑃𝐷𝐶 -efficient amplifier, choice was directed on DAs offering high 𝐺𝐵𝑃⁄𝑃𝐷𝐶 between those
four options; thereby, from Figure 5.3, our choice is now limited to either choosing a DA
with 𝑁 = 6 and 𝑊𝐶𝑆 = 26 µm or a DA with 𝑁 = 5 and 𝑊𝐶𝑆 = 31 µm, respectively, where
both offer ≈7.4 GHz/mW of 𝐺𝐵𝑃⁄𝑃𝐷𝐶 .
To choose from those two options, we opted for the DA with the smallest core area. This
requires pre-knowledge of how the DA will be implemented. In this work, all inductors are
implemented as conventional microstrip lines; by referring to Figure 4.18, it is clear that
𝑇𝐿𝐿𝑑𝑠 has a dominant role in defining the physical height of the unit cell and hence of the
DA. Thereby, from the DSE plot in Figure 4.20(b), that gives the required lengths for 𝑇𝐿𝐿𝑑𝑠 ,
it was deduced that the five unit-cells DA option offers 31% less core area than its six unit-
cells counterpart since its loss-compensation inductors are smaller in value.
As a result, the DA solution made from 𝑁 = 5 unit-cells with 𝑊𝐶𝑆 = 31 µm is
determined to be our global optimum choice for circuit implementation and on-wafer
characterization. Figure 5.4 shows the loss-compensated DA preliminary schematic
diagram with its design parameters summarized in Table 5.1 and its design characteristics
in Table 5.2. This DA will provide a simulated gain of 7 dB up to 100 GHz with a 𝑃𝐷𝐶 of
30 mW. It is composed of five loss-compensated cascode Gm-cells, placed at a distance of
2 · (𝑙 ⁄2) from each other; where the unit-cell segment length (𝑙 ⁄2) is 131 µm. Transistor
M1 is set to 31 𝜇m width and M2 twice that, i.e. 62 𝜇m. To compensate the high losses at
100 GHz while maintaining 1-dB of midband gain drop, a 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 of 43 µm and
129 µm, respectively, are needed as indicated by Figure 4.20. To prevent voltage drop
across the termination resistors 𝑅3 and 𝑅4 when biasing the DA, the latter are AC-grounded
through bypass capacitors 𝐶𝑆 . In addition, 𝑇𝐿𝐿𝑔 is AC-grounded through a capacitor 𝐶𝑔 .

Figure 5.4: Preliminary schematic diagram of the five unit-cells loss-compensated DA (biasing
not shown). Both 𝐶𝑠 and 𝐶𝑔 are shunt capacitors to insure AC ground connections.

Page | 80
TABLE 5.1: SINGLE-STAGE DA DESIGN PARAMETERS DETERMINED FROM CAUTOD PROCESS

M1 M2 TL Segments (𝒍⁄𝟐) 𝐓𝐋𝐋𝐠 𝐓𝐋𝐋𝐝𝐬


N
(µm) (µm) (µm) (µm) (µm)
5 31 62 131 43 129

TABLE 5.2: SINGLE-STAGE DA PRELIMINARY DESIGN CHARACTERISTICS

Gain Bandwidth 𝐏𝐃𝐂 Gain Ripple GBP 𝐆𝐁𝐏⁄𝐏𝐃𝐂


(dB) (GHz) (mW) (dB) (GHz) (𝐆𝐇𝐳⁄𝐦𝐖)
7 Up to 100 30 ±1 224 7.46

It is worth highlighting how the circuit design process was simple, fast and direct. Owing
to our matrix-based DA model (3.14), with its unit-cell model (3.17), and the possibility it
offers for automating 𝑁 and 𝑊𝐶𝑆 through MATLAB, analog designers can just directly pick
their optimum design point in an interesting graphical way for any FoM they please. There
is no need to change 𝑁 manually and run optimization for each time a design parameter is
modified in order to search for possible improvement in performance. In the next
Section 5.3, the layout aspect is covered and Section 5.4 covers the experimental results of
the on-wafer measurements.

5.3 100-GHz Single-Stage CMOS DA: Circuit Implementation

This section discusses the layout implementation side of the previously chosen circuit.
In subsection 5.3.1, the microstrip line topology used for implementing all type of inductors
making the DA from unit-cell segments, 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 is presented. In subsection 5.3.2,
a new high quality stacked parallel plate shunt-capacitor, used to implement 𝐶𝑆 and 𝐶𝑔 , is
introduced. In subsection 5.3.3, the MOS transistor interconnection that are used when
implementing are presented. Finally, subsection 5.3.4 is devoted to presenting the final
implemented layouts of both unit-cell and DA with their own on-chip microphotographs.

5.3.1 Microstrip Transmission Line


For the TL, we have taken advantage of the accurate and flexible models that are
provided in the PDK since they are based on pre-measurements from the semiconductor
foundry. Its topology is presented in 3D view in Figure 5.5(a). Metal 1 (M1, in green) is
the ground floor, metals 2 to 7 (M2 to M7) are used for complying with the local density
rules and are connected to the ground while metal 8 (M8) is the signal strip.

Page | 81
(a) (b)
Figure 5.5: (a) Isometric 3D view of the microstrip line. (b) Its attenuation constant when
implementing a 76-pH inductor for different width and length.

Figure 5.5(b) illustrates the attenuation constant (𝛼𝑇𝐿 ) of the TL with respect to
frequency for different widths ranging from 1 µm to 18 µm (limit imposed from PDK). For
a fixed inductance of 76 pH, which represent the needed value to synthesize the unit cells
segments, it is noticeable that the smaller the width, the lower 𝛼 𝑇𝐿 . It might seem that any
µ-TL configuration has negligible loss impact; however, it must be noted that a DA is made
of two inductive segments cascaded 𝑁 times, as evident in Figure 2.18. This means that the
input microwave signal will propagate through a total of 2⨉N physical segments, in other
words it will experience 2∙N times the losses, on both the loaded input and output lines of
the DA. Therefore, having the smallest width for TL can benefit from having the smallest
length that aid in reducing surface area of the DA when implemented and it contributes in
lowering the overall losses of the DA.
Since the 5-unit cells DA that we chose consumes 30 mW from a 1.2-V supply (25 mA),
and on the bases of the DC-current handling limits expressed on the Design Rule Manual
(DRM), we adopted for a width of 1 µm. We also decided to apply the same 1 µm width
for all µ-TLs with the aim of having later on a simpler design implementation process where
all the lines will share the same width.

5.3.2 High Quality and 𝑺𝑹𝑭 Stacked Parallel Plate Shunt Capacitor [61]
A common issue when synthesizing circuits at mm-wave is having at hand suitable
passive components that demonstrate high quality (𝑄) factor. As made clear from the
capacitor 𝑄-factor formula (𝑄 = 1⁄(𝜔 · 𝑅 · 𝐶) for 𝜔 < (2 · 𝜋 · 𝑆𝑅𝐹); where ω is the
angular frequency, 𝑅 is the equivalent parasitic resistor, 𝐶 is the equivalent parasitic resistor
and 𝑆𝑅𝐹 is the self-resonance frequency), its 𝑄-value is constantly experiencing decrease
with increase of operating frequency, thereby highlighting, in a general context, that
capacitors may become a bottleneck for high performance mm-wave ICs. In this work,
referring to Figure 5.4, capacitors 𝐶𝑆 and 𝐶𝑔 are used to provide DC decoupling and AC
grounding for the termination resistors and inductor 𝑇𝐿𝐿𝑔 , respectively.

As a solution, we proposed an on-chip passive stacked parallel plate shunt-capacitor


(SPP-SC) [61], constructed without additional process steps and with focus on mm-wave
performance. To prove its mm-wave performances, the proposed SPP-SC is compared with
other state-of-the-art capacitor topologies found in the literature in addition to a woven-
type MOM capacitor already available in the 55-nm ST process design kit.

Page | 82
5.3.2.1 Stacked Parallel Plate Shunt Capacitor
Figure 5.6(a) depicts the physical structure of a 300-fF SPP-SC synthesized within
existing masks using the 8-metal version of 55-nm process and Figure 5.6(b) illustrates a
symbolic view of its concept on how the process BEOL stack is used.

(a) (b)
Figure 5.6: 300-fF SPP-SC (a) 3D isometric view, and (b) symbolic front and side views of its
topology.

The odd metal plates are linked together by a stack of via array placed on each side of the
capacitor distributed in the longitudinal direction; this is to lower the resistive and inductive
parasitic, where path towards ground for current density is shorter. It also contributes in
confining the electric field within the plates, which contributes to the increase of
capacitance and prevent coupling with adjacent components. The even metal plates are
linked together from the front by a stack of via array and the stack is considered as the
access side. The multi-plate structure offers several parallel paths, thus minimizing AC
current density and lowering even more the aforementioned parasitic. Both via and structure
configurations give rise to a capacitor with an overall high 𝑄 and 𝑆𝑅𝐹 merits, as will be
demonstrated when discussing its on-wafer measurements.
𝑆𝐹−1
ɛ𝑟𝑛
𝐶0 ≅ ɛ0 · 𝑊 · 𝐿 · ∑ = 𝑊 · 𝐿 · 𝐶𝑑
𝑡𝑛 (5.1)
𝑛=1

From Figure 5.6(b), design equation (5.1) can be deduced and used as a first approach,
before layout implementation, to size the required the design variables: physical dimensions
(𝐿 x 𝑊) and stacking factor (𝑆𝐹), to obtain a desired intrinsic capacitance (𝐶0 ) value. The
term 𝑆𝐹 is the number of metal plates desired and made available from a given technology
node. Terms ɛ𝑟𝑛 and 𝑡𝑛 are the relative permittivity and thickness of each dielectric layer,
respectively, and ɛ0 is the permittivity of free space. From (5.1), the benefit of the proposed
topology can also be perceived: as 𝑆𝐹 increases with process node advancement, higher
capacitance density per unit area (𝐶𝑑 ) is expected.
Referring to the state-of-the-art work on high-frequency passive capacitors [62]-[66],
similar works presented capacitors on the order of 300 fF at most. On that bases, and since
by the end our main objective is to compare with what already exists by fixing a parameter
and showing the added values, the capacitance 𝐶0 in (1) is set to 300 fF. To demonstrate
the proposed capacitor performance, the 300-fF SPP-SC is designed as follows. The entire
metal stack of 55-nm node, i.e. 𝑆𝐹 = 8, is used to maximize to total 𝐶𝑑 (≈0.91 fF⁄µm2 )

Page | 83
for smaller dimensions and, also, to lower parasitic. Referring to Figure 5.6, reducing 𝑊
contributes to lower capacitor self-resistance, since GND connections from each side will
be closer to the capacitor input; however, since the metal layers of our capacitor share equal
dimensions, we were restricted by our Design-Rule-Manual (DRM) rules of the stringent
thin-metal layers (M1-M5) length (𝐿). Thereby, 𝐿 is restricted up to 10.8 µm, what led to
a wider width (𝑊) of 30.6 µm, accepted by the DRM, in order to preserve 𝐶0 at 300 fF,
(1). Figure 5.7 illustrates the microphotograph of final capacitor for on-wafer
measurements.

Figure 5.7: Microphotograph of the fabricated 300-fF SPP-SC.

5.3.2.1 On-wafer Characterization and Simulation Results


The 𝑆-parameters of the capacitor was measured via on-wafer probing using Anritsu 2-
port VectorStar ME7838D VNA setup. Conventional Thru-Reflect-Line (TRL) calibration
method [67], with a 630-μm line and an open reflect, was then applied. The
microphotographs of the TRL structures are shown in Figure 5.8.

(a) (b) (c)


Figure 5.8: Microphotograph of the fabricated Thru-Line-Reflect calibration kit.

The de-embedded capacitor can be represented as the electrical equivalent model


illustrated in Figure 5.9; where 𝑅0 , 𝐿0 and 𝐶0 are the intrinsic resistance, inductance and
capacitance, respectively, and 𝐶eff is an effective capacitance that represents 𝐿0 in series
with 𝐶0 .

Figure 5.9: SPP-SC equivalent electrical model. (𝑅0 , 𝐿0 , 𝐶0 ) represents intrinsic lumped-
elements of the capacitor and (𝑅𝑖𝑛 ,𝐶𝑖𝑛 ) represents the measured input-impedance.

Page | 84
To characterize those elements, 𝑆- to 𝑍-parameter conversion is first performed [56].
Intrinsic resistor 𝑅0 is determined from the measured capacitor equivalent input-series
resistance (𝑅in ), (5.2a). Intrinsic capacitance 𝐶0 (5.2b) is determined from the measured
capacitor equivalent input series-capacitance (𝐶in ) using (5.2c) and the intrinsic inductance
𝐿0 is determined from (5.2d) where 𝜔|𝑄=0 is the angular frequency at 𝑆𝑅𝐹 since
m{𝑍11 }|𝑄=0 = 0.

𝑅𝑖𝑛 (𝜔) = ℜ𝑒{𝑍11 } (5.2a)


𝐶0 = 𝐶𝑖𝑛 |𝜔≈0 (5.2b)
1
𝐶𝑖𝑛 (𝜔) = 𝐶𝑒𝑓𝑓 (𝜔) = − (5.2c)
𝜔 · 𝑚{𝑍11 }

1
𝐿0 = 2
(𝜔|𝑄=0 ) · 𝐶0 (5.2d)

𝑚{𝑍11 }
𝑄(𝜔) = − (5.2e)
ℜ𝑒{𝑍11 }

Figure 5.10 illustrates the measured and simulated RLC-elements against frequency for
the implemented 300-fF SPP-SC. The capacitor experience a relatively constant 𝑅in against
frequency with a measured average magnitude of 0.37 Ω. The increase of resistance at
frequencies below 6 GHz is attributed to the lower limit of TRL defined by the 630-𝜇m
line. Since 𝑅in is shown here to have a flat variation in the measurements frequency range,
it can safely be considered as a lumped 𝑅0 . The slight difference in extracted resistance
between EM simulation and measurement results could be attributed to limitation in our
equipment to accurately quantify very low losses. The proposed topology provides a
relatively constant value of 𝐶𝑖𝑛 within the frequency range considered, signifying that the
𝑆𝑅𝐹 are all beyond our measurement limit (100 GHz). Measured intrinsic capacitances 𝐶0
of 285 fF is obtained for the calculated 300-fF. The slight difference between what was
calculated (5.1) and measured could be due to process variation in dielectric layers
permittivity (ɛ𝑟 ).

Figure 5.10: Measured and HFSS EM-simulated equivalent input series-resistance (𝑅𝑖𝑛 ) and
capacitance (𝐶𝑖𝑛 ) of de-embedded 300-fF SPP-SC. (Solid line: Measured. Dashed line: EM
Simulated)

Page | 85
The measured 𝑄-factor is plotted in Figure 5.11 using 𝑄(𝜔) from (5.2e). 𝑄-value of 14.7
is obtained at 100 GHz. It must be noted that the drop of 𝑄-value here is dominated by the
increase of frequency rather than resistance, since from Figure 5.10 it is clear that 𝑅𝑖𝑛 , at
any given frequency, remains almost the same. The proposed capacitor also contributes to
a high 𝑆𝑅𝐹 merit, as demonstrated in Figure 5.11, where measured 𝑄-factor is extrapolated
using MathWorks MATLAB and 𝑆𝑅𝐹 is then extracted at frequency in which 𝑄 = 0 (since
m{𝑍11 }|𝑄=0 = 0). The 300-fF capacitor is estimated to reach an 𝑆𝑅𝐹 of 368 GHz.

Figure 5.11: Estimated SRF for 300-fF SPP-SC capacitor, extracted at 𝑄 = 0 through
extrapolating the measured 𝑄-factor.

Using those values, 𝐿0 is then obtained using (5.2d) and summarized within Table 5.3
with the other intrinsic elements modeling the SPP-SC in addition to the values extracted
from EM simulation. A slight difference in the extracted values of 𝐿0 exists between
measurement and simulation. This is due to the measured 𝐿0 being determined from an
extrapolation of a measured 𝑄-factor with evident distortion caused by the measured low-
value resistance fluctuation against frequency, as illustrated in Figure 5.10.

TABLE 5.3: EXTRACTED INTRINSIC LUMPED ELEMENTS OF 300-𝑓F SPP-SC.

𝑹𝟎 (Ω) 𝑪𝟎 (𝐟𝐅) 𝑳𝟎 (𝐩𝐇)


Capacitor
Meas. EM Sim. Meas. EM Sim. Meas. EM Sim.
300-fF 0.37 0.25 285 298 0.66 0.55

Figure 5.12 (next page) compares de-embedded to both lumped modeled (Figure 5.9)
and EM simulated 𝑆11 parameters for all 300-fF SPP-SC. A very good agreement is
obtained. This indicates how accurate the proposed physical model is up to 100 GHz.

Page | 86
Figure 5.12: Equivalent electrical model validation through 𝑆11 parameter plot of 300-fF

5.3.2.2 Performance Comparison with State-Of-the-Art Capacitor Topologies


After introducing the capacitor topology with its on-wafer performance, a comparison
with state-of-the-art capacitors is performed here. As part of AC-grounded components,
radial stub [62] is used for its ease of design but reserves large chip area. Slow-wave stub
[63] is proposed that benefits from length reduction; however, wider footprint is required
resulting in comparable area. Metal-Insulator-Metal (MIM) capacitor is one topology that
can offer high capacitance density due to its special high-𝑘 dielectric layer but suffers from
high dielectric losses at mm-waves. In [64], a 120-fF MIM capacitor demonstrated a 𝑄-
value of 3.8 at 100 GHz. Also, the majority of low-cost CMOS technologies lack MIM
capacitor options. Metal-Oxide-Metal (MOM) capacitor is an alternative option that uses
existing metals in the process backend-of-line (BEOL) but experiences lower 𝑄-values due
to its structure consisted of narrow thin metal wires. For instance, a 129-fF woven-structure
MOM capacitor provides a 𝑄-value below 3.2 at 100 GHz, [65]. In [66], for a common
multi-stack interdigital capacitor, the same value of 120-fF provided a 𝑄-value of 2.5 at
100 GHz. In addition, MOM capacitors experience a rapid 𝑆𝑅𝐹 reduction when higher
capacitance value is desired which could be attributed to the substantial increase in its
parasitic-inductance caused by its wire-type topology. It is also worth mentioning the
simulation efforts where they require large processing resources since condensed meshing
process is required for their closely spaced wires when optimizing ICs in full 3D-EM
simulators. Other approaches referred to as active capacitors such as implementing gate
oxide capacitors [64] exist to improve 𝑄-factor and area-efficiency, but they need a DC
bias and are strongly process and temperature dependent.

TABLE 5.4: PERFORMANCE COMPARISON BETWEEN WIDE BANDWIDTH AMPLIFIER TOPOLOGIES.


Process Stack 𝑪𝟎 𝑺𝑹𝑭 Area 𝑪𝒅
Ref. Topology 𝑸&%
Node Layers (fF) (GHz) (mm2) (𝒇𝑭⁄µ𝒎𝟐 )
[62] Stub 65-nm 9 348 150 N/A 10053 0.035
[64] MIM 65-nm 7 120 194 3.8 N/A N/A
£ #
[65] MOM 45-nm 7 129 160 < 3.2 129 1
#
[65] MOM 45-nm 7 215 110 0.5 196 1.1
#
[66] MOM 45-nm 11 120 140 2.5 32 3.75
[66] MOM 45-nm 11 350 75 < 0* 92# 3.8
This
SPP 55-nm 8 285 368 14.7 331 0.86
Work
%Estimated £Estimated #Estimated
N/A = not available. from plots. from Q-value provided at 60 GHz. from fF/𝜇m2.
&Determined at 100 GHz. *Due to SRF being below 100 GHz.

Page | 87
Table 5.4 compares the measured capacitor performance with results from state-of-the-
art capacitors. To the authors’ knowledge, this work demonstrated a capacitor that offers
the highest 𝑆𝑅𝐹 and 𝑄 merits while providing large capacitance values for mm-wave
frequency range. In terms of occupied area, it is true that the SPP capacitors reserved larger
area between the stated ones [64]-[66], but this is attributed to the characterized SPP in this
work having larger 𝐶0 . For the case of [66] 350-fF capacitor compared to the 285-fF SPP,
the former presented compact area which is largely related to [66] using an 11-metal version
of a 45-nm node compared to the 8-metal version of a larger 55-nm node used in this work
and, also, at the expense of a lower 𝑆𝑅𝐹.
As a conclusion, measurements showed attractive performance from excellent quality
factor and high 𝑆𝑅𝐹, making it suitable for analog systems operating until 100 GHz and
possibly even up to 368 GHz, proving eligibility for both broadband and high data-rate ICs.

5.3.2.3 Performance Comparison with 55-nm Woven-Structure MOM Capacitor


In the design kit of the technology at hand, a common woven-structure MOM capacitor
is available for usage as a parameterized cell (PCell). Figure 5.13 illustrates its topology
where it makes use of existing metals (M2-M4) in the process BEOL.

Figure 5.13: Cross section of a Metal 2/Metal 4 woven-structure MOM capacitor in 55-nm
PDK

To justify even further the need for our proposed SPP-SC topology, a comparison with
the mm-wave performance of this PCell MOM capacitor is also performed. For a fair
comparison, the value of 300-fF capacitance was fixed for both topologies, and both
resistance, capacitance variation against frequency and quality factors were plotted in
Figure 5.14. The 300-fF PDK woven-structure MOM capacitor reserved area of
30.6 µm ⨉ 6.23 µm, a lower area relative to the SPP capacitor due to the woven-structure
topology providing higher capacitance density per area (1.5 fF⁄µ𝑚2)

Page | 88
(a) (b)
Figure 5.14: Measured (a) equivalent input capacitance, and (b) equivalent input resistance and
𝑄-factor for 300-fF SPP-SC and woven-structure MOM capacitor in 55-nm process.

From Figure 5.14(a), for the same capacitance value of 300-fF, the SPP-SC
demonstrated flat capacitance variation against frequency, which signify an 𝑆𝑅𝐹 well
above 100 GHz (close to 368 GHz as determined from Figure 5.11), whereas the woven-
structure PDK MOM demonstrates a rapid increase in capacitance which signify its 𝑆𝑅𝐹 is
close to 100 GHz. From Figure 5.14(b), the SPP-SC delivers a self-resistance much lower
than that of the PDK MOM capacitor and hence a much desirable quality factor at
millimeter wave regions. At 100 GHz, the former capacitor has 0.33 Ω of self-resistance
and 𝑄 = 14.7, whereas the latter capacitor has 3.1 Ω of self-resistance and 𝑄 = 0.53.
Through this analysis, we validated even more the efforts taken to propose a different
capacitor than the already available ones for the purpose of gaining better overall
performance when designing our 100-GHz DA and any mm-wave ICs in a general context.

5.3.3 MOS Transistor Interconnection


The transistor connection is crucial in mm-wave design. In order to reduce the resistive
losses that can load the DA and hence degrade its passband, we have adopted the topology
illustrated in Figure 5.15 during layout implementation. The CG-FET uses the one of
Figure 5.15(a) where gate-, drain- and source- pins are accessed through M8, and the CS-
FET uses the same one (Figure 5.15(b)), but with source-pin directly shorted to GND plane.

(a) (b)
Figure 5.15: Isometric 3D view of the interconnection used. (a) CG-FET. (b) CS-FET.

Page | 89
5.3.4 Standalone Unit-Cell Layout and Microphotograph
A single unit cell is implemented as a standalone device for the purpose of validating
the proposed model (3.17) frequency behavior with measurements through 𝑆-parameters
characterization.
Referring to the design parameters determined in Section 5.2 (Table 5.2) and the
electrical schematic provide in Figure 5.4, and using both the microstrip line topology, the
proposed SPP capacitor topology for implementing 𝐶𝑔 as a decoupling capacitor and the
transistor interconnection introduced in subsection 5.3.1, 5.3.2 and 5.3.3, respectively, the
unit cell can be synthesized as shown in Figure 5.16. Due to the high 𝑆𝑅𝐹 and quality factor
of a single SPP-SC capacitor, as was determined in Figure 5.10, we were able to provide a
strong AC-ground at the upper end of the 43-µm 𝑇𝐿𝐿𝑔 facing the DC-input side and with
𝑆𝑅𝐹 well above 100 GHz. Figure 5.17 illustrates the microphotograph of this unit cell as a
standalone circuit. Extra DC pads were attached on top for biasing gate-pin of the CG
transistor (M2) through external power supply. For the drain pin of transistor M2 and gate
pin of transistor M1, biasing is provided through the RF-pads by the probes of the VNA.

Figure 5.16: Isometric 3D representation of a single unit cell with loss-compensated cascode as
Gm-cell attached to a 600-fF AC grounded SPP capacitor 𝐶𝑔 .

Page | 90
Figure 5.17: Microphotograph of the fabricated standalone unit cell in 55-nm process for on-
wafer characterization.

5.3.5 100-GHz Single-Stage DA: Layout and Microphotograph


In order to achieve the targeted 7 dB of gain up to 100 GHz, we need to connect five
replicas of the same unit cell, Figure 5.4, together as suggested by the design method.
Figure 5.18 (next page) presents the final electrical schematic diagram of our loss-
compensated CMOS-based DA. Table 5.5 summarizes all of its design parameters values
used during layout implementation. All µ-TL were implemented using thickest metal
available in technology, M8, with same width of 1 µm, as explained in subsection 5.3.1.
The μ-TL segments are done with length 𝑙 ⁄2 of 131 μm. Both 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 are
implemented with length of 43 μm and 129 μm, respectively. Both termination impedances
𝑍3 and 𝑍4 are realized as series RC loads. Both 𝑅3 and 𝑅4 are implemented using
technology component as N+ silicided polysilicon resistors, and were AC-grounded
through a 1-pF bypass SPP capacitor (𝐶𝑆 for both 𝐶3 and 𝐶4 ) to have a DA passband starting
from 10 GHz; as was already stated in Chapter 4 design algorithm (Step 4), we decided to
go with the highest 𝐹𝑙𝑜𝑤𝑒𝑟 of 10 GHz instead of 1 GHz, for instance, since they showed
equal gain; but choosing higher frequency will benefit us from having a bypass capacitor
with a small on-chip area, hence smaller DA. After finishing the design layout
implementation and before sending to fabrication, one last simulation (as a validation step)
was performed on ADS schematic to make sure that everything is in order. 𝑅3 and 𝑅4 had
to be refine hand-tuned to 53.5 Ω to slightly readjust the passband midband gain drop to be
precisely at 1-dB limit. This step is not always necessary and can vary from one designer
to another. But, it could be considered as a possible quick solution when slight re-
adjustment is required at the end-stage before sending to fabrication instead of, for instance,
increasing 𝑇𝐿𝐿𝑑𝑠 that will affect the height of the DA (even if 𝑇𝐿𝐿𝑑𝑠 was slightly increased)
and forces the designer to redo the synthetization process once more followed by
DRC/GAG validation procedure.

Page | 91
TABLE. 5.5: DA CIRCUIT DESIGN PARAMETERS FOR LAYOUT IMPLEMENTATION

M1 M2 TL Segments (𝒍⁄𝟐) 𝐓𝐋𝐋𝐠 𝐓𝐋𝐋𝐝𝐬 𝐑𝟑 𝐑𝟒


N
(µm) (µm) (µm) (µm) (µm) (Ω) (Ω)
5 31 62 131 43 129 53.5 53.5

Supply voltages for the gate pin of each M1 FET and the drain pin of each M2 FET are
provided through RF-probes by the measurement setup. The gate pins of M2 FETs are
biased together by an external bench power supply delivered from G-DC-G pads and
through resistor-type RF-chokes (𝑅𝑔 ) implemented using high-resistivity polysilicon
resistor of 10-kΩ. This is possible since we are dealing with FET and hence there is no
impact on overall 𝑃𝐷𝐶 . A 600-fF decoupling SPP capacitor (𝐶𝑔 ) was added on 𝐿𝑔 terminal
pin facing the DC pads for AC-grounding.
Finally, GSG-type RF-pads with 50-μm pitch are connected to DA input and output ports.
Both the final DA layout view and microphotograph are shown in Figure 5.19 and
Figure 5.20, respectively. The DA presents a core are of 0.52 mm2 and total area of
0.83 mm2 including pads.

Figure 5.18: Circuit schematic of the 5 unit-cells 100-GHz CMOS-based DA.

Figure 5.19: Layout view of the implemented 5 unit-cells 100-GHz CMOS-based DA.

Page | 92
Figure 5.20: Microphotograph of the fabricated 5 unit-cells 100-GHz CMOS-based DA in 55-
nm process for on-wafer characterization.

5.4 100-GHz Single-Stage CMOS DA: Experimental Results

In this section, simulation and measurement results of the amplifier device are presented
and compared. The measurement setup is first introduced in subsection 5.4.1. The 𝑆-
parameters analysis of both the unit-cell and DA are discussed in subsection 5.4.2. In
subsection 5.4.3, a performance analysis of the implemented DA is carried out. Here,
stability and group delay analysis are presented in addition to power capabilities analysis
from 1-dB gain compression point (P1dB), saturation power (Psat ) and third order intercept
point (IP3). Finally, a comparison with state-of-the-art CMOS DAs is carried out in
Section 5.5.

5.4.1 Vector Network Analyzer Description


All 𝑆-parameters measurements are performed via on-wafer probing using the VNA
setup shown in Figure 5.21.

Figure 5.21: Bias-tee simulated and measured isolation (𝑆21) versus frequency between DC-
and RF- ports.

Page | 93
This setup consists of an Anritsu 2-port VectorStar ME7838D mm-wave system
available with the MS4640B series Vector Network Analyzer (VNA) and the 3739C test
set to control two MA25300A mm-wave modules, each connected to Cascade Microtech
Infinity GSG 50-μm pitch probes. Both MS4640B and 3739C devices cover frequencies
< 54 GHz then transition towards MA25300A modules occurs to cover ranges above 54
GHz. For biasing, two external bench power supplies are used, Keysight B2901A to bias
the gate pins of cascode CG FETs and Keysight B2902A to provide bias along the loaded
gate- and drain-line of the DA through RF probes. VNA calibration using line-reflect-
reflect-match (LRRM) technique was performed before measurements to remove probing
RF impact.

5.4.2 S-parameters Analysis: Unit-Cell and DA Model Validation


Figure 5.22(a) and Figure 5.22(b) compare the 𝑆-parameter results versus frequency
obtained from MATLAB-based 𝐴𝐵𝐶𝐷-model simulation, (3.14) with (3.17), to both ADS-
based simulation and measurement of a single unit-cell and DA, respectively. Great
agreement is obtained between ADS and 𝐴𝐵𝐶𝐷-model results. This indicates how reliable
the proposed model is when used as design tool, this support the forth criterion. Unit-cell
pads were de-embedded from measurements but the DA pads were not for a better
representation of the latter behavior when used as standalone IC in real-world applications.

(a)

(b)
Figure 5.22: 𝑆-parameters simulation and measurement results versus frequency. (a) Single
unit cell. (b) DA.

Page | 94
Biased at 𝑉DD =1.2 V, the unit cell consumed an overall measured current of 5 mA and
the DA consumed 25 mA—same consumptions when calculated during design process
using (4.9). This resulted in a measured 𝑃𝐷𝐶 of 6 mW and 30 mW, respectively. The DA
achieves an average insertion gain 𝑆21 of 6.7 dB. A peak gain of 8 dB is observed at 91 GHz
resulting in an in-band ripple of ±1.3 dB when referenced to mid-band gain and, hence, a
coverage range from 𝐹𝑙𝑜𝑤𝑒𝑟 of 16.7 GHz up to an 𝐹𝑢𝑝𝑝𝑒𝑟 of 100 GHz. This gives an
amplified passband BW of 83.3 GHz and leads to a GBP of 180 GHz. The slight difference
(3.3 GHz) at the upper side of the frequency spectrum, where the measured DA reaches
higher frequencies compared to the simulated one, could be attributed to a slight
overestimation of the parasitic the transistors used are made from. The increase of gain at
frequencies < 10 GHz is attributed to the small bypass capacitor (𝐶𝑆 ) used to cover down
to 10 GHz (a decision we made in Step 4 of our design methodology in subsection 4.2.2).
This can be improved by either using larger 𝐶𝑆 (MIM capacitors in this case could be
considered for their larger capacitance density) or shorting both 𝑅3 and 𝑅4 directly to GND
as illustrated in Figure 4.1. The latter option will offer a DA with a flat gain starting from
DC level, but at the expense of higher 𝑃𝐷𝐶 since DC-current will pass through those
resistors (caused by the DC-voltage bias along the output loaded-line for 𝑅3 and along the
input loaded-line for 𝑅4 ). The dip in gain between 𝐹𝑙𝑜𝑤𝑒𝑟 and 10 GHz is caused by an
overestimation of loaded-lines termination resistances during implementation. Both input
and output matching levels, S11 and 𝑆22 , respectively, are better than −10 dB starting from
4.35 GHz. The amplifier maintains input matching up to 100 GHz while the output
matching remains < −10 dB up to 91.6 GHz with a slight dip back at 99 GHz. The reverse
isolation S12 is less than −50 dB below 1 GHz and reaches a maximum of −10 dB at
100 GHz. The overall good agreement between simulation and measurement results
validates the design process.

5.4.3 Implemented DA Performance Analysis


After demonstrating the benefits this matrix-based model offers from ease and direct
design process, and the reliability it provides where good agreement was obtained between
simulations and measurements, focus here is placed on analyzing the performance of the
fabricated 100-GHz CMOS-based DA. Both group-delay, stability and large-signal
performance analysis is performed hereafter. It should be pointed out that those three design
characteristics were not part of the CAutoD algorithm presented in Chapter 4 and what is
demonstrated in this subsection is just an illustration of what we would obtain if we decided
to go with a DA design process that focuses solely on extending and flattening passband.
The same three analysis were not applied to the standalone unit-cell since from
Figure 5.22(a) this device does not provide amplification and hence it has no interest for
practical usage.

5.4.3.1 Group Delay Performance Analysis


Figure 5.23 shows the DA group-delay behavior besides its 𝑆21 phase. Inside its
passband ranging from 𝐹𝑙𝑜𝑤𝑒𝑟 of 16.7 GHz to an 𝐹𝑢𝑝𝑝𝑒𝑟 100 GHz, determined previously,
the group delay experiences a measured ±5 ps variation up to 80 GHz and then increases
quickly as it approaches the upper side 100 GHz. This is because amplifier S21 phase is

Page | 95
linear against frequency until 80 GHz and becomes non-linear as it approaches passband
steep roll-off at higher frequency end.

Figure 5.23: Group delay simulation and measurement results versus frequency. Simulations
were performed on ADS software

5.4.3.2 Stability Performance Analysis


The DA 𝑘- and |𝛥|-factor derived from measurements and compared to simulation can
be seen in Figure 5.24. The DA is unconditionally stable for the whole measured passband
except at frequency of 100 GHz where it becomes conditionally stable with 𝑘 = 0.85. This
decline is attributed to a higher measured S12 reaching an utmost of −10 dB compared to
−17 dB in simulation and a wider S21 -passband. This led to a higher |S12 S21 | product which
causes 𝑘-factor to degrade since 𝑘 = (1 − |S11 |2 − |S22 |2 + |𝛥|2 )⁄(2|S12 S21 |) [20].

Figure 5.24: DA 𝑘- and |𝛥|-factor versus frequency. Simulations were performed on ADS
software.

5.4.3.3 Large-Signal Performance Analysis


While the main scope of this thesis is to come up with a different and new design
technique for mm-wave DAs, the large signal performance are worth mentioning to gasp
the performance of the UWB amplifier device. The following analysis is presented hereafter
where the power delivery capability of the former device is analyzed. Due to limitation in

Page | 96
our measurement equipment, the provided data in Figure 5.25 and Figure 5.26 are
simulation-based results.

Figure 5.25: Simulated output power and gain at 10 GHz, 60 GHz and 90 GHz

Figure 5.26: Simulated P1dB , Psat and OIP3 as a function of frequency

Figure 5.25 shows the simulated output power and gain as a function of input power at
10 GHz, 60 GHz and 90 GHz. The maximum output power at the three frequencies is
9.5 dBm, 9.1 dBm and 6.3 dBm, respectively. At 10 GHz and 60 GHz, the gain exhibits
similar behavior while at 90 GHz, the gain is compressed slightly earlier. Figure 5.26 shows
the simulated OP1dB, Psat and IP3 as a function of frequency. The average P1dB , Psat and
OIP3 over the entire frequency band are 2.4 dBm, 8.6 dBm and 12.5 dBm, respectively.
The maximum Psat of 9.5 dBm and OIP3 of 14.3 dBm were achieved at frequencies below
40 GHz.

TABLE 5.6: DA SIMULATED AND MEASURED 𝑂𝑃1𝑑𝐵 ESULTS


Frequency (GHz)
𝐎𝐏𝟏𝐝𝐁 10 30 60 90
Simulation (dBm) 4.3 4 2.4 -1.1
Measurement (dBm) 4.6 3.4 2 -1.6

Table 5.6 compares the simulated P1dB results to the ones we were able to measure at
10 GHz, 30 GHz, 60 GHz and 90 GHz. Both OIP3 and Psat measurements were not possible
to provide due to lack of coupler equipment to send two tones to perform the former
measurement and the incapability of our VNA in providing more than 0 dBm for the latter

Page | 97
measurement. One can note that measured results are well correlated with simulation and
are considered acceptable performance for such a CMOS-based small-signal distributed
amplifier.

5.5 Comparison with State-of-the-Art CMOS-Based DAs


Performances

In Chapter 2, state-of-the-art DA performances in different process nodes and


technologies was introduced. It was concluded that indeed HBT-based DA can outperform
FET-based DAs due to the HBT superiority in providing higher transconductance. Since
this is commonly known and for a fair comparison, here our characterized DA is compared
to only CMOS-based DAs. Table 5.7 summarizes our measured results to the other start-
of-the-art silicon DAs.

TABLE 5.7: STATE-OF-THE-ART PERFORMANCE COMPARISON FOR CMOS-BASED DISTRIBUTED


AMPLIFIERS
𝑮𝑩𝑷 GBP
Total
Process Gain Ripple£ BW& Supply PDC GBP 𝑷𝑫𝑪 PDC · Area
Ref. Area
Techno. (dB) (dB) (GHz) (V) (mW) (GHz) 𝑮𝑯𝒛 GHz
( ) (mm2) ( )
𝒎𝑾 mW · mm2
90-nm
[30] Si 7.4 -3% 80 2.4* 120 188 1.57 0.72 2.18
CMOS
90-nm
[41] Si 7@ +9%,-3 70 N/A 122 157 1.29 0.72 1.8
CMOS
90-nm
[42] digital 14@ ±1.5% 73.5 1.2* 84 368 4.38 1.72 2.54
CMOS
40-nm
(1,
[43] digital 15$ +5%,-3 80 90 450 5 0.31 16.13
1.7)**
CMOS
120-nm
[44] SOI 11 ±1.2 85 2.5* 210 302 1.44 1.28 1.13
CMOS
120-nm
[45] SOI 7.8 ±1.3 82 2.6* 130 201 1.55 1.05 1.47
CMOS
45-nm
[31] SOI 9 ±1.5% 92 1.2* 73.5 259 3.52 0.45 7.82
CMOS
22-nm
[32] FD-SOI 8.5 -3% 110 2* 80 292 3.66 0.38 9.63
CMOS
55-nm
This
Si 6.7 ±1.3 83.32 1.2* 30 180 6.01 0.83 7.24
Work
CMOS
N/A = not available.
&BWs are reported as declared by authors. £Ripple limits is determined within the authors declared BW for their reported gain.
%Estimated from plots @Cascaded Multi-stage DA (CMSDA) $Cascaded single-stage DA (CSSDA)
*Biased using external bias-tee **Biased using integrated bias-tee

Page | 98
This work demonstrates comparable gain while being the most power-efficient DA. As
a result, this work reports the highest GBP⁄PDC figure of merit for single-stage CMOS DA
of 6.01 GHz/mW, to the authors’ knowledge. It is true that our presented DA does not
provide the highest GBP between the stated CMOS DAs; for instance, [43] provides a GBP
of 450 GHz. However, one can benefit from our low 𝑃𝐷𝐶 to cascade three of our DA stage
until [43] PDC of 90 mW is reached. This would lead to a higher GBP of 843 GHz and,
hence, a GBP⁄PDC of 9.37 GHz/mW; highlighting once more, that thanks to our model and
algorithmic scripted method, we were able to squeeze out the last drop of possible
performance improvement from each design parameter when design the single stage DA.
It is true that cascading leads to increase in total area; however, in this case, layout
improvements such as pads rearrangement would have been considered. In addition, it is
worth mentioning the contribution a smaller technology node (e.g. 40-nm [43] and 22-nm
[32]) can bring in area reduction compared to the 55-nm adopted in this work, which could
explain their higher GBP⁄(PDC · Area) merit.

5.6 Conclusion

In this chapter, the 𝐴𝐵𝐶𝐷-model was put to practice where it was used to realize a 100-
GHz single-stage loss-compensated CMOS DA using ST 55-nm CMOS technology. Both
the circuit design and layout implementation aspects of the design process were covered. It
showed how we were able to directly determine the optimum number of unit cells and
transistor size, besides other design parameters, needed to satisfy our desired FoM of high
𝐺𝐵𝑃⁄𝑃𝐷𝐶 with reasonable gain. It demonstrated how the design process can be rendered
simple and fast by using the proposed methodology and scripting, and interesting through
3D-DSE plots.
During the design process, a lack of high-quality capacitor suitable for mm-wave design
was detected. As a solution, a viable passive capacitor topology is demonstrated and
characterized. An accurate design equation (5.1) is also provided with at most 5% of
discrepancy with the characterization performed. Measurements showed attractive
performance from excellent quality factor and high 𝑆𝑅𝐹, making it suitable for analog
systems operating until 100 GHz and possibly even up to 368 GHz, proving eligibility for
both broadband and high data-rate ICs. Furthermore, it is fabricated without extra mask
costs (unlike MIM capacitor) and easy to simulate in a time-efficient manner in full 3D-
EM simulators if compared to MOM capacitor, since it does not involve a mesh of closely
spaced thin metal wires. Finally, it is a promising topology in advanced technology nodes
since higher capacitance density is expected with metal stack increment.
When comparing overall behavior of the unit-cell and DA, Figure 5.21, to both ADS-
based simulation and measurements, good agreement was obtained validating therefore the
accuracy of the model. In addition, when compared to the state-of-the-art CMOS DAs
(Table 5.7), optimum 𝐺𝐵𝑃⁄𝑃𝐷𝐶 performance was obtained owing to the capability of the
proposed model in presenting the complete DA and taking all of its design parameters into
consideration during the automated sizing. Joining the latter two criteria contributes to
maximizing the performance that could possibly come from each design parameter.

Page | 99
Page | 100
Chapter 6

THz Gain-Bandwidth Product 100-


GHz Cascaded CMOS Distributed
Amplifier

In the last chapter, a demonstration on how we can use the proposed model to realize a
single-stage DA with high 𝐺𝐵𝑃⁄𝑃𝐷𝐶 characteristic was presented. There, both circuit
design and implementation sides of the design process were covered. In addition,
performance analysis and comparison with state-of-the-art CMOS-based DAs was
performed where ours demonstrated better overall performance. In this chapter, the model
is reused in combination with device cascading technique to demonstrate its potentiality in
realizing a CMOS-based DA with the highest THz GBP characteristic possible when
compared to the literature.
This chapter is structured as follows. Section 6.1 covers the circuit design part of the
design process from determining the new set of required design variables (𝑙 ⁄2, 𝑁, 𝑊𝐶𝑆 ,
𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 ) in addition to a new design variable (𝑀) which defines the cascading
factor. Section 6.2 covers the design implementation part of the design process. Here, a new
integrated inductive bias-tee [68] is presented that offers wideband characteristic.
Section 6.3 discusses the simulated results of the implemented cascaded DA. A comparison
with the performances of state-of-the-art DAs in different technologies and process nodes
is presented in Section 6.4. Finally, Section 6.5 concludes and summarizes this chapter.

6.1 THz GBP Cascaded CMOS DA: Circuit Design

From the state-of-the-art presented in Table 2.2 (Chapter 2), and reminded in Section 6.4
of this chapter, the idea of cascading DA was introduced by [41], [42], [47], [49], as a
technique to obtain higher overall gain covering the large passband where the gain of a
single-stage 𝑁-unit cells DA is multiplied by the number of cascaded stages (𝑀). In [69],

Page | 101
an analytical-based study was provided that demonstrates a trade-off between 𝑁 and 𝑀 to
reach a certain gain while being 𝑃𝐷𝐶 -efficient. Herein, we decide to make use from the
outcome of our previously described algorithmic methodology to choose new set of
adequate design-variables sizes in order to have the optimum GBP per single stage. This
proposition will aid in maximizing the cascaded DA overall GBP, which in return will have
advantageous influence on having a much higher overall 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit.
The circuit design part for a 100-GHz bandwidth criterion was already covered in
Chapter 5 (Section 5.2). It was determined that 35 DA options were only able to satisfy
such criterion, from the whole 216 DAs design space of Figure 4.11, with their respective
optimum values for 𝑁 and 𝑊𝐶𝑆 .
Since our aim here is to show that maximizing GBP is possible through our model, we
decided to pick the optimum design option that offers the highest gain. Referring to
Figure 5.2, the DA with 𝑁 = 6 and 𝑊𝐶𝑆 = 28 µm offers gain of 7.7 dB over 100 GHz of
passband, higher compared to the previous option of 𝑁 = 5 and 𝑊𝐶𝑆 = 31µm (7 dB). Of
course, from Figure 5.3, it is visible that the former option offers relatively lower 𝐺𝐵𝑃⁄𝑃𝐷𝐶
than the latter due to its higher 𝑃𝐷𝐶 (32.3 mW instead of 30 mW, determined from
Figure 4.17). However, when cascading the DAs, the higher gain will benefit us from
having a much higher overall GBP and therefore will lead to an even greater 𝐺𝐵𝑃⁄𝑃𝐷𝐶
performance, as will be proven here shortly after.
For the chosen design option of 𝑁 = 6 and 𝑊𝐶𝑆 = 28 µm, the segments length (𝑙 ⁄2)
was determined to be 125 µm from Figure 4.19 and the values of its loss-compensating µ-
TL-type inductors, 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 , are found to be 54 µm and 155 µm from Figure 4.20(a)
and Figure 4.20(b), respectively.
It should be highlighted how the re-design process was easy and determining the
required design variables was immediate without any re-calculation of any kind. This would
have not been the case if the previously methods mentioned in Chapter 3 (Section 3.2) were
used nor would it have been the case if CAD tools that only offers schematic design
environment where change of FoM is followed by re-doing the manual exchange of 𝑁 from
the scratch and consequently followed by a re-optimization process per exchange.
What remains is determining the number of cascaded stages (𝑀) needed. From state-of-
the-art Table 2.2, it is noticeable between cascaded DAs that the commonly adopted gain
ripple is ±3 dB over the passband. For this reason, we can benefit from our 1-dB restriction
imposed during our design process, as was already specified in the algorithmic design
methodology (Step 5), and cascade the single-stage DA six times, i.e., set 𝑀 = 6. This will
lead to a preliminary simulated gain of 46.2 dB ±3 dB. For a 100-GHz passband, this will
result in a GBP of 20.42 THz. Of course, 𝑃𝐷𝐶 will also be 6-fold higher, but when
calculating the overall 𝐺𝐵𝑃⁄𝑃𝐷𝐶 of the cascaded DA, one will obtain 105.5 GHz/mW;
proving the previous statement, that cascading does indeed increase 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit. Such
value is determined to be the highest between the recorded values in the literature, based
on state-of-the-art Table 2.2. Those design parameters in order to implement such cascaded
DA are summarized in Table 6.1 with the design characteristics in Table 6.2.
Next Section 6.2 covers the layout implementation of the cascaded DA.

Page | 102
TABLE 6.1: CASCADED DA DESIGN PARAMETERS DETERMINED FROM CAUTOD PROCESS

M1 M2 TL Segments (𝒍⁄𝟐) 𝐓𝐋𝐋𝐠 𝐓𝐋𝐋𝐝𝐬


M N
(µm) (µm) (µm) (µm) (µm)
6 6 28 56 125 54 155

TABLE 6.2: CASCADED DA PRELIMINARY DESIGN CHARACTERISTICS

Gain Bandwidth 𝐏𝐃𝐂 Gain Ripple GBP 𝐆𝐁𝐏⁄𝐏𝐃𝐂


(dB) (GHz) (mW) (dB) (THz) (𝐆𝐇𝐳⁄𝐦𝐖)
46.2 Up to 100 193.54 ±3 20.42 105.5

6.2 THz GBP Cascaded CMOS DA: Circuit Implementation

This section discusses the layout implementation part of the previously chosen circuit.
Since we will implement a DA circuit similar to the one implemented in Chapter 5, but just
with different values for 𝑁 and 𝑊𝐶𝑆 , both discussions on microstrip line topology and shunt
SPP capacitor topology were already covered in Section 5.3 and are not re-written here.
Only the extra difference is highlighted where the output loaded-line of the DA here is
biased through a newly proposed integrated wideband bias-tee whereas the former DA of
Chapter 5 was biased from the external VNA through its RF-probes.

6.2.1 Integrated Wideband Millimeter-Wave Bias-Tee [68]


In a general context, realizing wideband active ICs is accompanied with additional
design efforts to produce suitable components used to synthesize them, mainly ones capable
of providing wideband characteristic. For instance, in Chapter 5 an issue with having high
𝑄-factor and 𝑆𝑅𝐹 shunt capacitor components, suited for the mm-wave band, was detected
and was resolved by proposing our own SPP topology. Another challenge encountered is
when the wideband IC is required to be synthesized with its own bias tee. This requires the
implementation of a compact, low power-overhead and wideband RF-choke. For the case
at hands, for instance, the bias-tee has to be able to provide RF-choke capability covering
frequencies reaching the highest operating frequency of the DA, i.e., 100 GHz. This is
considered challenging since parasitic at high frequency become dominant and can disrupt
the performance of the bias tee and, consequently, have negative influence on the behavior
of the device itself.

6.2.1.1 Already Existing Techniques for Realizing Wideband Bias-Tees


There already exists traditional techniques to realize a bias-tee and they are summarized
as follows. The most common technique to realize a wideband and compact biasing
network is by using integrated resistive elements as RF-chokes [50]. With this approach, a
series resistance is used between power-supply equipment and on-wafer device to inhibit
RF signals from leaking outwards and any possible interaction between both two. This

Page | 103
approach also benefits from allowing the lower-corner frequency of the IC to extend
downwards to DC levels (0 Hz). However, this approach is not suitable when DC power
efficiency is important since it causes a voltage drop across the resistor and thus 𝑉𝑑𝑑 must
be increased to preserve the same operating point of the biased circuit. External bias-tees
with large inductance are commonly used, such as in [31], [45] and [70], to mitigate this
issue owing to their attractive low DC-resistance while providing high reactance that
chokes RF signal. This technique, however, requires the use of a bulky and expensive off-
chip inductor not suited for fully integrated systems. Embedding a large spiral inductor as
part of the design to avoid cost penalty, such as in [71], is suited for relatively small
operational bandwidths where they can be positioned below 𝑆𝑅𝐹 [72], but this integrated
solution may become problematic when dealing with several tens of gigahertz bandwidth.
This is due to the parasitic elements having larger influence at high frequencies and hence
affecting the behavior of the wideband bias-tee. Most often, this happens to the bias-tee
𝑆𝑅𝐹 where the overall parasitic can alter it and subsequently alter the passband behavior of
the device being designed. Shifting this 𝑆𝑅𝐹 outside such wide passband by reducing those
parasitic is considered challenging especially when the intended circuit to be biased
operates up to 100 GHz.
As a solution, a compact fully integrated bias-tee with low DC-overhead and wideband
performance is proposed. Compared to conventional approach where bias-tees, the ones
that use inductor components as RF-choke, are designed to reduce the overall parasitic and
push their 𝑆𝑅𝐹 outside the device passband, i.e., above the highest operating frequency of
the biased device, here the bias-tee is designed by taking advantage of the parasitic
capacitance and pushing its 𝑆𝑅𝐹 below the passband of the biased device.
For the sake of simplicity in demonstrating the proposed bias-tee idea, its design concept
and process are first introduced followed, as a second step, by a practical example where a
100-GHz bandwidth standalone bias-tee is implemented and its simulation versus
measurement is discussed.

6.2.1.2 Proposed Wideband Bias-Tee: Design Concept


Figure 6.1 illustrates the complete biasing network configuration. The left side
represents the DC-power supply, the right side represents the intended RF-circuit to be
biased and the middle network represents the lumped-element circuit of the proposed bias-
tee. The bias-tee here is a two-port RF network, with port-1 towards DC-power supply side
and port-2 towards the RF-circuit side. When compared to the practical implementation in
Figure 6.6 where each single-stage DA is biased from its output-loaded line, it is noticeable
that the termination resistance 𝑅 and DC-decoupling capacitance 𝐶 (𝑅3 and 𝐶3 ,
respectively, on Figure 6.6) are considered as part of the bias-tee.

Figure 6.1: Lumped model representation of the proposed bias-tee topology

Page | 104
Conventionally, the bias-tee reactance-type RF-choke is designed such that its 𝑆𝑅𝐹 is
pushed above the biased circuit highest operating frequency (𝐹𝑢𝑝𝑝𝑒𝑟 ) while providing
sufficiently high inductive reactance across its passband. This is usually achieved by
optimizing the structure on the basis of 𝐿𝑆 while attempting to diminish RF-choke parasitic
𝐶𝑆 . Our proposed way, however, takes advantage of the 𝑆𝑅𝐹 instead of suffering from it.
This is done by making use of RF-choke parasitic capacitance 𝐶𝑆 and considering it now as
a design parameter while pushing bias-tee overall resonance frequency below the biased
circuit lower-corner frequency (𝐹𝑙𝑜𝑤𝑒𝑟 ). For this reason, the bias-tee of the proposed
approach could be viewed as having a “capacitive-reactance RF-choke”, i.e. the passband
of the device is operating in the capacitive region established by 𝐶𝑆 . Referring to Figure 6.1,
shunt capacitor 𝐶𝑎 is added as a filtering capacitor to protect the integrated mm-wave circuit
against any perturbation that comes from DC power supply and its wiring. In practice, off-
chip capacitors are also implemented all along the DC-path and probe coming from external
power supply. In addition, large on-wafer integrated capacitors are also added at pad-level
to enhance protection. Those have no effect on bias-tee as long as 𝐶𝑎 acts as a perfect AC-
ground in the DA frequency range of interest. Capacitors 𝐶𝑎 and 𝐶 own parasitic are not
considered herein, being negligible as compared to 𝐿𝑆 -𝐶𝑆 tank elements and would be, for
instance, if the SPP capacitor introduced in Chapter 5 (Section 5.3) is used. Finally, the
purpose of adding 𝐿𝑒 , a small value inductor, is to remove excess imaginary part of the bias
tee input impedance seen by the device being biased. This will be better explained when
discussing equation (6.5).
In a first approach, 𝐿𝑆 and 𝐶𝑆 are being considered as independent parallel components.
Equation (6.1) demonstrates the impedance of the 𝐿𝐶-tank. Emphasis was put intentionally
here on the parasitic capacitance 𝐶𝑆 as a design parameter, instead of 𝐿𝑆 , as it is noticeable
how capacitor 𝐶𝑆 can be benefited from to create high impedance RF-choke. Later on, when
dealing with the practical example of designing a 100-GHz bandwidth bias-tee for the DA,
an empirical relationship between both is provided as an example that designers can adopt
to choose adequately the RF-choke parameters.
−𝑗
𝑍𝑡𝑎𝑛𝑘 (𝜔) = 2
𝜔
𝜔𝐶𝑆 (1 − ( 𝑡𝑎𝑛𝑘 ) ) (6.1a)
𝜔

1
𝜔𝑡𝑎𝑛𝑘 =
√𝐿𝑆 𝐶𝑆 (6.1b)

By ignoring 𝐿𝑒 at first, and if 𝐶𝑎 is chosen large enough to be safely assumed as an AC-


connection to ground for the whole passband, the bias-tee equivalent input impedance is

derived first from 𝑍𝑖𝑛 in (6.2).


𝐶2
𝑅𝑒{𝑍𝑖𝑛 (𝜔)} = 𝑅′𝑖𝑛 (𝜔) = 𝑅
𝛥 (6.2a)

′ ′
1 𝐶 + 𝜌𝐶𝑠 + 𝜔2 𝜌𝐶𝑠 𝐶 2 𝑅 2
𝐼𝑚{𝑍𝑖𝑛 (𝜔)} = 𝑋𝑖𝑛 (𝜔) = − ·
𝜔 𝛥 (6.2b)

Page | 105
𝜔𝑡𝑎𝑛𝑘 2
𝜌 =1−( )
𝜔 (6.2c)

𝛥 = (𝐶 + 𝜌𝐶𝑠 )2 + (𝜔𝜌𝐶𝐶𝑠 𝑅)2


(6.2d)
′ ′ ′ ′
Impedance 𝑍𝑖𝑛 is in the form of 𝑅𝑖𝑛 ± j|𝑋𝑖𝑛 |. For matching purpose, |𝑋𝑖𝑛 | should be

reduced as much as possible while 𝑅𝑖𝑛 be made as close as possible to the real part of the
output impedance of the device being biased―in this case, to the single-stage DA. Up to
this point, an additional relationship for 𝐶𝑆 and 𝐿𝑆 is necessary to pursue the proposed bias-
tee design; it can be obtained through practical implementation considerations, as done
next.

6.2.1.3 100-GHz Bandwidth Bias-Tee: Circuit Design Process


The design process of our wideband bias-tee, suitable for the DA discussed in
Section 6.1, will be covered here. Beforehand, this process requires pre-knowledge of the
bandwidth lower-limit of the device tackled. In our case, we will target by the end of this
chapter a DA that will start amplifying from an 𝐹𝑙𝑜𝑤𝑒𝑟 of 5 GHz until an 𝐹𝑢𝑝𝑝𝑒𝑟 of 100 GHz.
The RF-choke used is a microstrip spiral inductor realized using the eighth metal layer of
ST 55-nm technology. It is chosen here since it is widely known to offer an equivalent
parallel 𝐿𝑆 -𝐶𝑆 tank at high frequency without the need to add explicitly extra lumped
components. However, for this choice, 𝐶𝑆 becomes a parameter correlated to 𝐿𝑆 and,
thereby, requires as a first step to be modeled analytically. Figure 6.2 gives the simulated
𝑓𝑡𝑎𝑛𝑘 and DC self-resistance of the spiral inductor metallic winding ―this will be referred
back to when discussing measurement results―, using ANSYS’ HFSS, as a function of its
self-inductance 𝐿𝑆 .

Figure 6.2: Plot of 𝐿𝑆 -𝐶𝑆 tank 𝑓𝑡𝑎𝑛𝑘 frequency and its DC self-resistance against different
values of self-inductance (𝐿𝑆 ).

By curve fitting the collected 𝑓𝑡𝑎𝑛𝑘 to 𝐿𝑆 , an empirical expression is formulated in (6.3).


𝜔𝑡𝑎𝑛𝑘
𝑓𝑡𝑎𝑛𝑘 = = 3.778 · 104 · 𝐿−0.6479
𝑆(𝐻) + 4.355 · 109
2𝜋 (6.3)

The design parameter 𝐶𝑆 can be represented analytically as (6.4).


1
𝐶𝑆 = (6.4)
((2𝜋 · 𝑓𝑡𝑎𝑛𝑘 )2 · 𝐿𝑆 )

Page | 106
Equation (6.3) is optional, but simplifies the design task by relating 𝐶𝑆 to any
implemented spiral design through its 𝑓𝑡𝑎𝑛𝑘 . Another possible way could be to execute
multiple 3D EM simulations and extract 𝐶𝑆 for each spiral inductor winding turns. This will
result in discrete values for 𝐶𝑆 instead of continuous ones as provided by the use of
equations (6.3) and (6.4).
After having modeled the 𝐿𝑆 -𝐶𝑆 tank, the bias-tee circuit design can now begin as
follows. As stated previously, the proposed idea for wideband performance is to place the
resonance frequency of the overall bias-tee (𝑓𝑟𝑒𝑠𝐵𝑖𝑎𝑠−𝑇𝑒𝑒 ) at most at the targeted 𝐹𝑙𝑜𝑤𝑒𝑟 of
the device. Since we will deal with an 𝐹𝑙𝑜𝑤𝑒𝑟 of 5 GHz, both 𝐶𝑎 and 𝐶 are set to 2 pF to
insure enough AC-ground while not reserving too much area when being implemented as
close as possible to the DA stage. Resistance 𝑅 is set to 50 Ω that represents the DA 𝑅3
termination resistor, as already mentioned at the beginning of this subsection. By sweeping
𝐿𝑆 from 1-nH up to 10-nH, the input reactance against frequency can thus be calculated
from equation (6.2b) and for each 𝐿𝑆 -value, 𝑓𝑟𝑒𝑠𝐵𝑖𝑎𝑠−𝑇𝑒𝑒 corresponds to when

𝑋𝑖𝑛 (𝜔𝑟𝑒𝑠𝐵𝑖𝑎𝑠−𝑇𝑒𝑒 ) = 0 Ω. Figure 6.3 summarizes those values of 𝑓𝑟𝑒𝑠𝐵𝑖𝑎𝑠−𝑇𝑒𝑒 with respect to
𝐿𝑆 . One can observe that a minimum of 5.6 nH is needed for a resonance frequency of the
bias-tee below 5 GHz. We decide to choose a slightly higher 6-nH inductance to be cautious
against process variation.

Figure 6.3: Plot of bias-tee overall resonance frequency against different values of self-
inductance (𝐿𝑆 ).

′ ′
Figure 6.4 shows the bias-tee 𝑋𝑖𝑛 variation with respect to frequency. Above 5 GHz, 𝑋𝑖𝑛
is acting as a negative reactance for 𝐿𝑆 ≥ 5.6 nH proving, therefore, when used with the
DA later on, that its passband will experience an equivalent capacitive behavior with small
magnitude. In order to obtain wideband low input reactance up to the intended 𝐹𝑢𝑝𝑝𝑒𝑟 , for
better port matching purposes, and since the negative part of the reactance is quite small, it
can be compensated at higher frequencies by simply adding a low-value inductor, 𝐿𝑒 . By
doing so, one arrives at the complete bias-tee of Figure 6.1, providing an input impedance
𝑍𝑖𝑛 , described by equation (6.5).
′ ′
𝑍𝑖𝑛 = 𝑍𝑖𝑛 + 𝑗𝜔𝐿𝑒 = 𝑅𝑖𝑛 ≈𝑅
(6.5)

Page | 107

Figure 6.4: Bias-tee equivalent input reactance (𝑋𝑖𝑛 ) versus frequency for different values of
𝐿𝑆 .


The effective input resistance 𝑅𝑖𝑛 is observed to decrease against frequency as a
consequence of using 𝐶𝑆 , as indicated by (6.2a) and (6.2d). A deducible solution would be
to increase 𝑅 so that the average of 𝑅𝑖𝑛 becomes 50 Ω. However, based on the refine hand-
tuning performed at the end of the design process with the previous single-stage DA
implemented in Chapter 5, the resistor 𝑅 was left here untouched until the end process
before sending the cascaded DA to fabrication. Table 6.3 summarizes the required design
parameter determined previously and needed to implement such bias-tee suitable to cover
the passband of a DA amplifying from 5-GHz up to 100 GHz. In the next part of this
subsection, simulation and measurements of the implemented bias-tee as a standalone
device are presented.

TABLE 6.3: 100-GHZ BANDWIDTH INDUCTIVE BIAS-TEE DESIGN PARAMETERS

𝐋𝐬 𝐂𝐒 𝐂𝐚 𝐂 𝐋𝐞
(nH) (𝐟F) (pF) (pF) (pH)
6 27.6 2 2 39

6.2.1.4 100-GHz Bandwidth Bias-Tee: Simulation vs Measurement Results


Figure 6.5 illustrates a 3D view of the implemented bias-tee that represents the electrical
model of Figure 6.1. The 𝐿𝑆 -𝐶𝑆 tank with 𝐿𝑆 of 6 nH and 𝐶𝑆 of 27.6 fF was implemented as
a microstrip spiral inductor covering a surface area of 82 µm x 82 µm with the minimum
spacing allowed by the DRM of 0.6 µm between metallic windings to conserve area. The
39-pH inductor 𝐿𝑒 is implemented using µ-TL of 50-µm in length. Both 2-pF shunt
capacitors, 𝐶𝑎 and 𝐶, where implemented using SPP-topology introduced in Chapter 5 [61]
since this high-quality capacitor topology demonstrated to be the most suitable for such
applications at mm-wave band

Page | 108
Figure 6.5: 3D isometric view of the implemented 100-GHz bandwidth bias-tee.

Figure 6.6 compares the simulated and de-embedded measured results. The simulation
was performed on ANSYS 3D EM simulator HFSS. The discrepancy at frequencies below
6 GHz is attributed to the lower limit of TRL defined by the 630-μm line (Figure 5.8). A
slight shift exits at lower frequencies, which could be caused by a slightly higher inductance
than the anticipated 𝐿𝑆 of 6 nH. In overall, great agreement is obtained between both results
validating the design process and concept. The proposed bias-tee is capable of delivering
an RF-choke bandwidth of around 100 GHz since isolation between DC- and RF- port is
≤ -20 dB up to 100 GHz (the intended 𝐹𝑢𝑝𝑝𝑒𝑟 of the DA). As additional note, referring back
to Figure 6.2, the price of such a bias-tee is a small DC resistance of about 6.3 Ω. It is
considered an attractive property for such a wideband performance since it will lead to a
smaller DC-power overhead compared to using a resistor-based bias-tee, for instance, that
provides similar wideband performance but for a much higher DC-power overhead.

Figure 6.6: Bias-tee simulated and measured isolation (𝑆21) versus frequency between DC- and
RF- ports.

Up to now, the new bias-tee topology was discussed and its design concept was validated
through measuring a 100-GHz standalone device version of it. In the next subsection, the
layout implementation of single stage DA, of Section 6.1, with such bias-tee is presented.

6.2.2 100-GHz Single-Stage DA: Layout and Microphotograph


All the needed components to implement the six unit-cells single-stage DA piece of the
complete cascaded DA are now available. Figure 6.7 presents the final electrical schematic
diagram of our loss-compensated single-stage 6-unit cells CMOS-based DA of 7.7-dB

Page | 109
amplification and Figure 6.8 its microphotograph. Table 6.4 summarizes all of its design
parameters. All TL are implemented using thickest metal available in technology, M8, with
same width of 1 µm, as explained in Chapter 5, subsection 5.3.1. The μ-TL segments are
done with length 𝑙 ⁄2 of 125 μm. Both 𝑇𝐿𝐿𝑔 and 𝑇𝐿𝐿𝑑𝑠 are implemented with length of
54 μm and 155 μm, respectively. Both termination impedances 𝑍3 and 𝑍4 are realized as
series RC loads. Both 𝑅3 and 𝑅4 are implemented using technology component as N+
silicided polysilicon resistors, and were AC-grounded through a 2-pF bypass SPP capacitor
(both 𝐶3 and 𝐶4 ) which will lead to a DA passband starting from 5 GHz while being
compact in area.

TABLE 6.4: SINGLE-STAGE DA CIRCUIT DESIGN PARAMETERS FOR LAYOUT IMPLEMENTATION


M1 M2 TL Segments (𝑙 ⁄2) TLLg TLLds
N
(µm) (µm) (µm) (µm) (µm)
6 28 56 125 54 155

Supply voltage for the drain pin of each M2 FET is provided through the newly proposed
integrated bias-tee topology (Figure 6.5) connected to the output loaded-line of the single-
stage DA. The biasing of all gate pins is provided through resistive RF-choke implemented
using high-resistivity polysilicon resistor (𝑅𝑔 ) of 10-kΩ. This is possible since we are
dealing with FET and hence there is no impact on overall 𝑃𝐷𝐶 . A 600-fF decoupling SPP
capacitor (𝐶𝑔 ) was added on each 𝐿𝑔 terminal pin facing the DC pads for AC-grounding.

Figure 6.7: Circuit schematic of the 6 unit-cells 100-GHz CMOS-based DA with integrated
bias-tee.

Figure 6.8: Layout view of the implemented 6 unit-cells 100-GHz CMOS-based DA with
integrated bias-tee.

Page | 110
6.2.3 Six-Stage Cascaded 100-GHz DA Layout and Microphotograph
The final design step remaining is to realize the cascaded DA. In order to obtain 46.2 dB
of gain, the single-stage DA of Figure 6.8 is cascaded six times (𝑀 = 6). Figure 6.9
demonstrates the final layout of the implemented cascaded DA.

Figure 6.9: Layout view of the implemented 6-stage cascaded 100-GHz CMOS-based DA.

The core area of the DA is 2.3 mm2 and the total area is 3.5 mm2 including pads. To
prevent biasing interference between consecutive DAs where the drain-line of a preceding
DA is connected to the gate-line of the subsequent DA, a 2-pF MOM capacitor was added
in series as a DC-decoupling capacitor. Series DC-decoupling capacitors were also added
at both the input and output sides of the cascaded DA to prevent any influence from VNA
RF-probes on the biasing point of both first and sixth single-stage DAs, respectively. Due
to high gain delivered by the integrated amplifier, any influence from the outside word ―in
this case external biasing equipment― can interfere with the behavior of the DA and could
render it unstable. For that reason, 400-pF AC-grounded capacitor was added on each side
of the DA on DC-pad that provides 𝑉𝑑𝑑 to filter any incoming low-frequency noise. Woven-
structure MOM capacitors, already available in the PDK, were used for both roles (DC-
decoupling and AC-grounding on DC-pads) due to its capability in providing high
capacitance density per unit area, hence considered advantageous for not wasting additional
silicon area. Finally, GSG-type pads with 50-μm pitch are connected to the cascaded DA
input and output RF-ports. After finishing the design layout implementation and before
sending to fabrication, one last simulation (as a validation step) was performed on ADS
schematic to make sure that everything is in order. Termination resistor 𝑅3 was hand-tuned
to 57 Ω while taking into consideration the resistive drop caused by the implemented bias-

Page | 111
tee (Figure 6.4) and fixing any slight imperfection in midband gain-drop to preserve
precisely the 3-dB limit. Concerning resistor 𝑅4 , it was left at 53.5 Ω since this part is
similar to the one used in the implementation of single-stage DA of Chapter 5 without a
bias-tee.
In the next Section 6.3, the performance analysis of the presented cascaded CMOS-
based DA is discussed.

6.3 THz GBP Cascaded CMOS DA: Simulation Results

In this section, simulation results of the amplifier device are presented. Measurements
were not possible to provide at that time due to silicon wafer still being in the fabrication
stage. The 𝑆-parameter analysis of the cascaded DA are first discussed. Then, performance
analysis is performed from stability and group delay analysis. Regarding power capability
analysis from P1dB, Psat and IP3, they are not possible to simulate due to insufficient
processing resources when performing large-signal harmonic balance simulation on ADS
of such a DA with great deal of non-linear components. A comparison with state-of-the-art
DAs is performed in the next Section 6.4.

6.3.1 𝑺-parameter Analysis


Figure 6.10 illustrates the simulated 𝑆-parameter results versus frequency for the
cascaded DA. While the design process was performed using the simulation capability of
our 𝐴𝐵𝐶𝐷-model on MATLAB, however, the final simulation of the complete cascaded
DA device is performed on ADS software. This is due to some design blocks, such as bias-
tee, being added as touchstone files (.s2p) whereas our DA model does not possess such
functionality.

Figure 6.10: 𝑆-parameter simulation results versus frequency of cascaded DA.

Biased at 𝑉dd =1.5 V instead of 1.2 V, to compensate for the bias-tee self-resistance,
each single-stage DA consumes 27 mA and the overall cascaded DA consumes a total of
162 mA. This results in a 𝑃𝐷𝐶 of 40.5 mW and 243 mW, respectively. The DA achieves an
average insertion gain 𝑆21 of 45.3 dB with in-band ripple of ±3 dB and, hence, a coverage
range from 𝐹𝑙𝑜𝑤𝑒𝑟 of 5 GHz up to an 𝐹𝑢𝑝𝑝𝑒𝑟 of 100 GHz. This gives an amplified passband
BW of 95 GHz and leads to a GBP of 17.5 THz. Due to bias-tee power overhead and the

Page | 112
DA starting from 5 GHz instead of DC-level, the new 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merit is found to be
71.96 GHz/mW. Both input and output matching levels, S11 and 𝑆22 , respectively, are better
than −10 dB starting from 5 GHz up to 100 GHz. The reverse isolation S12 reaches a
maximum of −93 dB at 100 GHz. This high isolation can be explained from using cascode
topology as Gm-cell for the unit cells and since we are cascading single-stage DAs 𝑀-
times, the overall reverse isolation is also multiplied by the number of cascaded stages the
same way as the overall gain is.

6.3.2 Group Delay Performance Analysis


Figure 6.11 shows the DA group-delay behavior besides its 𝑆21 phase. The group delay
experiences a flat variation inside passband up to 70 GHz and then increases quickly as it
approaches 𝐹𝑢𝑝𝑝𝑒𝑟 of 100 GHz (the cut-off frequency of the DA where sharp roll-off
occurs). This is because amplifier S21 phase is linear against frequency until 70 GHz and
becomes non-linear as it approaches passband steep roll-off at higher frequency end.

Figure 6.11: 𝑆-parameter simulation results versus frequency of cascaded DA.

6.3.3 Stability Performance Analysis


The cascaded DA 𝑘- and |𝛥|-factor derived from simulation can be seen in Figure 6.12.

Figure 6.12: DA 𝑘- and |𝛥|-factor versus frequency. Simulations were performed on ADS
software.

Page | 113
It is unconditionally stable for the whole passband. This DA demonstrates 𝑘-values
≥ 100 and |𝛥|-values ≤ 0.1. This high stability is attributed to the device S12 being
interestingly very low, as illustrated in Figure 6.10, compared to single-stage DA of
Chapter 5 (Figure 5.21(a)). The decline of both 𝑘- and |𝛥|-factors as frequency increases is
due to the same reason of S12 increasing until -93 dB at 100 GHz.

6.4 Comparison with State-of-the-Art DAs Performances

In Chapter 2, a comparison between state-of-the-art DA performances in different


process nodes and technologies was introduced. In order to highlight that it is possible using
CMOS technologies to obtain high gains with UWB performances reaching frequencies as
high as 100 GHz, since the former is usually considered as a weak technology due to its
low transconductance, the characterized cascaded DA is mentioned with all the different
DAs of Table 2.2. This performance comparison is summarized in Table 6.5 (next page).
It is observed that for a comparable bandwidth, our DA demonstrated to offer the highest
gain recorded while being power efficient, hence possessing the highest 𝐺𝐵𝑃⁄𝑃𝐷𝐶 between
the stated DAs. In addition, even though our cascaded DA reserves a relatively higher total
area of 3.5 mm2; however, a comparison between all 𝐺𝐵𝑃⁄(𝑃𝐷𝐶 · 𝐴𝑟𝑒𝑎) values shows that
our DA contributed to the highest one. This indicates an efficient usage of per silicon area
to deliver the stated 45.3 dB of gain and re-validates once more the ability our model can
give us in squeezing the most from each deign component the DA is made from to maximize
the latter gain performance.
Through this performance comparison, we can conclude that while it is true what is said
that CMOS technology do provide low transconductance compared to the other mentioned
process technologies; however, with adequate and proper choice of optimum design
parameters provided thanks to our model, CMOS technologies can be aided even more in
proving their abilities in mm-wave circuit design applications.

Page | 114
TABLE 6.5: STATE-OF-THE-ART PERFORMANCES COMPARISON FOR DISTRIBUTED AMPLIFIERS
𝑮𝑩𝑷 GBP
Total
Process Gain Ripple£ BW& Supply PDC GBP 𝑷𝑫𝑪 PDC · Area
Ref. Area
Techno. (dB) (dB) (GHz) (V) (mW) (GHz) 𝑮𝑯𝒛 GHz
( ) (mm2) ( )
𝒎𝑾 mW · mm2
90-nm
[30] Si 7.4 -3% 80 2.4* 120 188 1.57 0.72 2.18
CMOS
90-nm
[41] Si 7@ +9%,-3 70 N/A 122 157 1.29 0.72 1.8
CMOS
90-nm
[42] digital 14@ ±1.5% 73.5 1.2* 84 368 4.38 1.72 2.54
CMOS
40-nm
(1,
[43] digital 15$ +5%,-3 80 90 450 5 0.31 16.13
1.7)**
CMOS
120-nm
[44] SOI 11 ±1.2 85 2.5* 210 302 1.44 1.28 1.13
CMOS
120-nm
[45] SOI 7.8 ±1.3 82 2.6* 130 201 1.55 1.05 1.47
CMOS
45-nm
[31] SOI 9 ±1.5% 92 1.2* 73.5 259 3.52 0.45 7.82
CMOS
22-nm
[32] FD-SOI 8.5 -3% 110 2* 80 292 3.66 0.38 9.63
CMOS
130-nm
[46] InP 14.5# ±0.7 94 3* N/A 499 N/A 2.75 N/A
HEMT
100-nm
[33] InP 7 +2.5% 111 N/A N/A 248 N/A 2.2 N/A
HEMT
70-nm
[47] InP 21@ ±3% 80 2** 300 898 3 2.47 1.21
HEMT
50-nm
[34] InGaAs 11 ±1 105 6* 450 373 0.83 1.69 0.5
mHEMT
250-nm
[48] InP 12.8 +7%,-2% 180 N/A 110 724 6.58 0.32 20.57
HBT
250-nm
[36] InP 10 -3 182 4* 105 575 5.47 0.33 16.57
HBT
130-nm
[35] SiGe 10 ±2% 170 3.6* 108 537 5 0.38 13.16
HBT
130-nm
[49] SiGe 24@ +2.5%,-1% 95 N/A 247 1506 6.09 0.65 9.37
HBT
55-nm
This
Si 45.3 ±3 95 1.5** 243 17487 71.96 3.5 20.56
Work$
CMOS
N/A = not available.
&BWs are reported as declared by authors. £Ripple limits is determined within the authors declared BW for their reported gain
%Estimated from plots @Cascaded Multi-stage DA (CMSDA) $Cascaded single-stage DA (CSSDA) #Flip-Chip DA
$Simulated performances
*Biased using external bias-tee **Biased using integrated bias-tee

Page | 115
6.5 Conclusion

Through this chapter, a 45.3-dB cascaded DA using ST 55-nm CMOS technology was
introduced. Results demonstrated the highest 𝐺𝐵𝑃 and 𝐺𝐵𝑃⁄𝑃𝐷𝐶 characteristics of
17.5 THz and 71.96 GHz/mW, respectively. The highest recorded FoMs in the literatures.
Owing to the proposed 𝐴𝐵𝐶𝐷-matrix based model and through the provided design
methodology in Chapter 4, the optimum design parameters were able to be determined in
order to provide the highest gain per DA-stage possible while being power efficient.
Throughout this chapter, a bias-tee topology with UWB performance was also
introduced where it is based on a newly proposed technique for pushing the resonance
below the lower-end of the wide passband compared to the traditional way of pushing it
towards the higher-end of the wide passband. The latter becomes quite difficult especially
if we are speaking about futuristic BWs ≥ 100GHz. Our proposed bias-tee showed to be
efficient in power-consumption where it demonstrated lower power-overhead compared to
resistive-based bias-tee and simpler in designing compared to the conventional inductive-
based bias-tee since parasitic capacitance was part of the design instead of a component
suffering from it. As a proof-of-concept, the proposed topology was designed for the
intended DA and was characterized as a standalone device, which showed its capability in
providing RF-choke functionality reaching 100 GHz of bandwidth.

Page | 116
Page | 117
Chapter 7

General Conclusion

In this thesis, we contributed in improving the design process of mm-wave distributed


amplifier in integrated circuits. We did so by focusing on various aspects of its design
requirements as follows. 1) Proposing a reliable and versatile design model that represents
the distributed amplifier complete frequency behavior, offering the possibility for scripting-
based design technique. 2) Providing an algorithmic methodology that can be used in the
framework of automated designing to explore the wide space of possible distributed
amplifier solutions a given process technology can offer. In addition, 3) proposing
integrated passive components specifically designed to address the issue of wideband AC-
grounding and device biasing. Our efforts have been validated by using STMicroelectronics
55-nm process as a technology to explore from and discover the feasible CMOS DAs and,
by the end, arrive at implementing a single-stage and cascaded CMOS DA options that
amplify frequencies up to 100 GHz and demonstrate global optimum performances.
First, different DA design techniques were investigated from mathematical-based
approaches, to artificial intelligence-based techniques such as neural-network and through
schematic-level design environments using CAD tools; however, each technique presented
its own weakness. It ranged from the assumption of a continuous loaded-lines structure,
which restricts the highest frequency of operation where the model can still be considered
valid. The assumption of a simple unilateral Gm-cell AC equivalent circuit, where in reality
a more complex representation is needed when dealing with high frequencies since parasitic
become dominant. In addition, when speaking especially of mm-wave frequency bands,
loss-compensated Gm-cells become necessary to be adopted in order to mitigate the
parasitic losses and provide amplification at those bands, but they introduce extra design
variables. This complicates the circuit design process where more variables, besides the
one of the DA, needs to be taken into consideration. Depending on the Gm-cell topology
chosen by the designer, those variables can change between one DA to another which
highlights, therefore, the need for a general DA model. The most preferred design technique
is often determined to be through CAD environments, such as Keysight ADS or Virtuoso
Cadence software. For CAD environments where only schematics diagram simulation is

Page | 118
provided, the DA is drawn first and as a second step, optimization is applied until a specific
performance goal is achieved. This technique lacks the capability to optimize the number
of unit-cells (𝑁) which falls on the designer task to manually do it and, consequently,
perform re-optimization per 𝑁 changement. Since the DA structure has several degrees of
freedom, a large volume of simulation work is necessary for a specific topology and goal,
hence affecting the simplicity of such design process. For CAD environments where
scripting capabilities are provided, this optimization process can be made easier through
automation, however, of course, whatever the scripting language used, this requires first
having at hand a proper physical model. On those basis, a matrix-based design model based
on 𝐴𝐵𝐶𝐷 parameters, famously used in describing passive circuitry, was introduced that
does not employ restrictions on any kind of design parameter, that considers a complete
and general Gm-cell representation, and that offers the possibility to consider all design
parameters as possible tunable variables, including 𝑁. Efforts were dedicated on providing
a model for successful implementation of CAutoD process for DAs through scripting.
Secondly and as a practical example, the basic principle-of-operation of a loss-
compensated DA was presented as a topology that can aid in the possibility to extend the
amplification passband response deep into the mm-wave frequency range. Based on its
concept, a systematic algorithmic methodology, with the help of the 𝐴𝐵𝐶𝐷-based model,
was offered that maximizes passband for any combination of unit cells and transistor size.
As a result, its outcome was 3D DSE plots that offer the designer an interesting graphical
overview of the possible tradeoffs between all the design characteristics of interest (in this
case gain, bandwidth and PDC ). Our CAutoD method also provided the optimal and accurate
dimensions needed for design parameters set such as transistor width, 𝑁, TL segments
length, and, in this particular case of loss-compensated cascode, the inductive-type
elements 𝐿𝑔 and 𝐿𝑑𝑠 , to synthesize the suitable design option.

Finally, two CMOS-based DAs using ST 55-nm process were designed and
implemented for on-wafer characterization. The first amplifier was a single-stage 100-GHz
bandwidth DA that demonstrated high 𝐺𝐵𝑃⁄𝑃𝐷𝐶 that served for model validation purposes.
The second amplifier was a cascaded 100-GHz bandwidth DA that demonstrated 17.5 THz
𝐺𝐵𝑃 and 71.96 GHz/mW 𝐺𝐵𝑃⁄𝑃𝐷𝐶 merits, the highest recorded characteristics compared
to the state-of-the-art literature. During the design process, two new design components
were introduced that can contribute in synthesizing high quality and broadband mm-wave
ICs. The first component was an AC-ground capacitor that offers the highest 𝑆𝑅𝐹- and 𝑄-
merit while providing large capacitance values at high frequency range. For instance, for a
300-fF capacitance, an 𝑆𝑅𝐹 of 368 GHz and 𝑄-value of 14.7 at 100 GHz were obtained.
The second component being a new bias-tee topology based on the idea of benefiting from
the parasitic and pushing below its SRF below the lower-end operating frequency of the
DA. The latter demonstrated promising wideband performance reaching 100 GHz.
Through our work, we prevision that by the usage of our model and method we could
assist analog designers in their future design process where it facilitates the alternation
between layout- and design- stages where more focus can be dedicated on layout and the
challenges it brings to mm-wave circuitry, and proposing enhanced components, as was
done here. In addition, to give the opportunity to discover the optimality a given technology
node can yield, and explore and suggest enhancements that can break the limits and further
advance the performance of future mm-wave distributed amplifiers.

Page | 119
Publications

Journals:
[Published] M. El-Chaar, L.Vincent, J. -D. Arnould, A. A. L. de Souza, S. Bourdel and F. Podevin,
“Accurate Design Method for Millimeter Wave Distributed Amplifier based on Four-Port
Chain (ABCD) Matrix Model,” in IEEE Transactions on Circuits and Systems I: Regular
Papers, vol. 69, no. 11, pp. 4510-4523, Nov. 2022.
[Published] M. El-Chaar, F. Podevin, S. Bourdel, A. A. L. de Souza and J. -D. Arnould, "Integrated
Stacked Parallel Plate Shunt Capacitor for Millimeter-Wave Systems in Low-Cost Highly
Integrated CMOS Technologies," in IEEE Solid-State Circuits Letters, vol. 5, pp. 114-117,
2022.
Conferences with acts:
M. EL Chaar, A. L. de Souza, M. Barragan, F. Podevin and S. Bourdel, "A Non-Closed-Form
Mathematical Model for Uniform and Non-Uniform Distributed Amplifiers," 2020 IEEE
MTT-S International Conference on Microwaves for Intelligent Mobility (ICMIM), 2020, pp.
1-4,
M. E. Chaar, A. L. de Souza, M. Barragan, F. Podevin, S. Bourdel and J. -D. Arnould, "Integrated
Wideband Millimeter-Wave Bias-Tee – Application to Distributed Amplifier Biasing," 2021
19th IEEE International New Circuits and Systems Conference (NEWCAS), 2021, pp. 1-4,
I. Alaji, T. C. Mouvand, M. El-Chaar, A. Lisboa-de-Souza, F. Podevin and S. Bourdel, "Cascaded
tunable distributed amplifiers for serial optical links: Some design rules," 2020 18th IEEE
International New Circuits and Systems Conference (NEWCAS), 2020, pp. 234-237,
M. El-Chaar, F. Podevin, S. Bourdel, A. Lisboa-de-Souza, J.-D. Arnould “Méthode de conception
d’amplificateurs distribués millimétriques basée sur la matrice chaîne (ABCD) 4 ports ”, »,
22èmes Journées Nationales Microondes, Limoges, 7-10 juin 2022.
Seminar, worshops:
M. El-Chaar, A. de Souza, F. Podevin, S. Bourdel, J.-D. and M. Barragan, “Distributed Amplifier:
Application to future means of communication in the field of millimeter-wave frequencies”,
Journée Scientifique du Pole PEM, Grenoble, France, 21 octobre 2021.

Page | 120
Bibliography

[1] D. Alldred, B. Cousins, and S. P. Voinigescu, “A 1.2 V, 60 GHz radio receiver with on-chip
transformers and inductors in 90 nm CMOS,” in Proc. IEEE Compound Semiconductor
Integrated Circuits Symp., Nov. 2006, pp. 51–54.
[2] B. Razavi, "A 60GHz direct-conversion CMOS receiver," ISSCC. 2005 IEEE International
Digest of Technical Papers. Solid-State Circuits Conference, 2005., 2005, pp. 400-606 Vol. 1
[3] Razavi, "A 60-GHz CMOS receiver front-end," in IEEE Journal of Solid-State Circuits, vol.
41, no. 1, pp. 17-22, Jan. 2006,
[4] Q. Gordon, T. Yao and S. P. Voinigescu, "65-GHz receiver in SiGe BiCMOS using monolithic
inductors and transformers," Digest of Papers. 2006 Topical Meeting on Silicon Monolithic
Integrated Circuits in RF Systems, 2006, pp. 4 pp.-,
[5] B. A. Floyd, S. K. Reynolds, U. R. Pfeiffer, T. Zwick, T. Beukema and B. Gaucher, "SiGe
bipolar transceiver circuits operating at 60 GHz," in IEEE Journal of Solid-State Circuits, vol.
40, no. 1, pp. 156-167, Jan. 2005,
[6] C. H. Wang, Y. H. Cho, C. S. Lin, H. Wang, C. H. Chen, D. C. Niu, J. Yeh, C. Y. Lee, and J.
Chern, “A 60 GHz transmitter with integrated antenna in 0.18 m SiGe BiCMOS technology,”
in IEEE Int. Solid-State Circuits Conf. (ISSCC) Dig. Tech. Papers, Feb. 2006, pp. 186–187.
[7] B. Floyd, S. Reynolds, U. Pfeiffer, T. Beukema, J. Grzyb, and C. Haymes, “A silicon 60 GHz
receiver and transmitter chipset for broadband communications,” in IEEE Int. Solid-State
Circuits Conf. (ISSCC) Dig. Tech. Papers, Feb. 2006, pp. 649–658.
[8] S. Gunnarsson, C. Karnfelt, H. Zirath, R. Kozhuharov, D. Kuylenstierna, A. Alping, and C.
Fager, “Highly integrated 60 GHz transmitter and receiver MMICs in a GaAs pHEMT
technology,” in IEEE J. Solid-State Circuits, vol. 40, no. 11, pp. 2174–2186, Nov. 2005.
[9] M. Siddiqui et al., “GaAs components for 60 GHz wireless communication applications,”
presented at the GaAs Mantech Conf., San Diego,CA, Apr. 2002.
[10] H. Zirath, T. Masuda, R. Kozhuharov, and M. Ferndahl, “Development of 60 GHz front-end
circuits for a high-data-rate communication system,” IEEE J. Solid-State Circuits, vol. 39, no.
10, pp. 1640–1649, Oct. 2004.
[11] O. Memioglu, Y. Zhao and B. Razavi, "A 300GHz 52mW CMOS Receiver with On-Chip LO
Generation," 2022 IEEE International Solid- State Circuits Conference (ISSCC), 2022, pp. 1-
3,
[12] S. Hara et al., "A 416-mW 32-Gb/s 300-GHz CMOS receiver," 2017 IEEE International
Symposium on Radio-Frequency Integration Technology (RFIT), 2017, pp. 65-67.

Page | 121
[13] K. Takano et al., "17.9 A 105Gb/s 300GHz CMOS transmitter," 2017 IEEE International
Solid-State Circuits Conference (ISSCC), 2017, pp. 308-309.
[14] J. Hacker et al., "THz MMICs based on InP HBT Technology," 2010 IEEE MTT-S
International Microwave Symposium, 2010, pp. 1126-1129.
[15] J. Rieh, B. Jagannathan, H. Chen, K. T. Schonenberg, D. Angell, A. Chinthakindi, J. Florkey,
F. Golan, D. Greenberg, S.-J Jeng, M. Khater, F. Pagette, C. Schnabel, P. Smith, A. Stricker,
K. Vaed, R. Volant, D. Ahlgren, “SiGe HBTs with cut-off frequency of 350 GHz”, Digest of
2002 International Electron Devices Meeting (IEDM '02), pp. 771-774, 2002
[16] International Technology Roadmap for Semiconductors, 2007, see http://public.itrs.net/
[17] Cisco, “Cisco Visual Networking Index: Global Mobile Data Traffic Forecast Update, 2017-
2022” Cisco Visual Netw. Index, San Jose, CA, USA, White Paper, Feb. 2019.
[18] National Telecommunications and Informatio Administration, "United States Frequency
Allocations: The Radio Spectrum Chart" United States Department of Commerce, Jan. 2016.
[19] Richard N. Clarke, Expanding mobile wireless capacity: The challenges presented by
technology and economics, Telecommunications Policy, Volume 38, Issues 8–9, 2014, Pages
693-708.
[20] P. Liu, K. Skucha, M. Megens, and B. Boser, “A CMOS Hall-effect sensor for the
characterization and detection of magnetic nanoparticles for biomedical applications,” IEEE
Trans. Magn., vol. 47, no. 10, pp. 3449–3451, 2011.
[21] B. Jang, P. Cao, A. Chevalier, A. Ellington, and A. Hassibi, “A CMOS fluorescent-based
biosensor microarray,” in IEEE ISSCC Dig. Tech. Papers, 2009, pp. 436–437.
[22] D. Di Carlo, L. Y. Wu, and L. P. Lee, “Dynamic single cell culture array,” Lab Chip, vol. 6,
pp. 1445–1449, 2006.
[23] Gustav Veith, Eugen Lach, Karsten Schuh, 100 Gigabit-per-second: Ultra-high transmission
bitrate for next generation optical transport networks, Comptes Rendus Physique, Volume 9,
Issues 9–10, 2008, Pages 1002-1011, ISSN 1631-0705.
[24] Chun-Lin Ko, Chieh-Pin Chang, Chien-Nan Kuo, Da-Chiang Chang and Ying-Zong Juang, "A
1-V 60 GHz CMOS low noise amplifier with low loss microstrip lines," Proceedings of
Technical Program of 2012 VLSI Design, Automation and Test, 2012, pp. 1-4.
[25] Y. Park, C. . -H. Lee, J. D. Cressler, J. Laskar and A. Joseph, "A very low power SiGe LNA
for UWB application," IEEE MTT-S International Microwave Symposium Digest, 2005.,
2005, pp. 1041-1044.
[26] Sananes, R.; Socher, E.: '52–75 GHz wideband low-noise amplifier in 90 nm CMOS
technology', Electronics Letters, 2012, 48, (2), p. 71-72,
[27] W. R. Deal, M. Biedenbender, P. Liu, J. Uyeda, M. Siddiqui and R. Lai, "Design and Analysis
of Broadband Dual-Gate Balanced Low-Noise Amplifiers," in IEEE Journal of Solid-State
Circuits, vol. 42, no. 10, pp. 2107-2115, Oct. 2007
[28] A. Arbabian et al., "A 90GHz-carrier 30GHz-bandwidth hybrid switching transmitter with
integrated antenna," 2010 IEEE International Solid-State Circuits Conference - (ISSCC),
2010, pp. 420-421
[29] O. El-Aassar and G. M. Rebeiz, "A Compact pMOS Stacked-SOI Distributed Power Amplifier
With Over 100-GHz Bandwidth and Up to 22-dBm Saturated Output Power," in IEEE Solid-
State Circuits Letters, vol. 2, no. 2, pp. 9-12, Feb. 2019.
[30] Ren-Chieh Liu, To-Po Wang, Liang-Hung Lu, Huei Wang, Sung-Hsiung Wang and Chih-Ping
Chao, "An 80GHz travelling-wave amplifier in a 90nm CMOS technology," ISSCC. 2005
IEEE International Digest of Technical Papers. Solid-State Circuits Conference, 2005, San
Francisco, CA, USA, 2005, pp. 154-590 Vol. 1.

Page | 122
[31] J. Kim and J. F. Buckwalter, "A 92 GHz Bandwidth Distributed Amplifier in a 45 nm SOI
CMOS Technology," in IEEE Microwave and Wireless Components Letters, vol. 21, no. 6, pp.
329-331, June 2011.
[32] P. V. Testa, D. Fritsche, S. Schumann, W. Finger, C. Carta and F. Ellinger, "110 GHz
travelling-wave amplifier in 22 nm FD-SOI CMOS," 2017 IEEE Asia Pacific Microwave
Conference (APMC), Kuala Lumpur, Malaysia, 2017, pp. 406-409.
[33] B. Agarwal et al., "112-GHz, 157-GHz, and 180-GHz InP HEMT traveling-wave amplifiers,"
in IEEE Transactions on Microwave Theory and Techniques, vol. 46, no. 12, pp. 2553-2559,
Dec. 1998.
[34] C. Zech et al., "An ultra-broadband low-noise traveling-wave amplifier based on 50nm InGaAs
mHEMT technology," 2012 The 7th German Microwave Conference, Ilmenau, Germany,
2012, pp. 1-4.
[35] P. V. Testa, G. Belfiore, D. Fritsche, C. Carta and F. Ellinger, "170 GHz SiGe-BiCMOS Loss-
Compensated Distributed Amplifier," 2014 IEEE Compound Semiconductor Integrated
Circuit Symposium (CSICS), 2014, pp. 1-4.
[36] S. Yoon, I. Lee, M. Urteaga, M. Kim and S. Jeon, "A Fully-Integrated 40–222 GHz InP HBT
Distributed Amplifier," in IEEE Microwave and Wireless Components Letters, vol. 24, no. 7,
pp. 460-462, July 2014.
[37] W. S. Percival, “Thermionic Valve Circuits,” British Patent Specification no. 460,562, Jul. 24
1936.
[38] E. L. Ginzton, W. R. Hewlett, J. H. Jasberg and J. D. Noe, "Distributed Amplification," in
Proceedings of the IRE, vol. 36, no. 8, pp. 956-969, Aug. 1948.
[39] W. Jutzi, "A MESFET distributed amplifier with 2 GHz bandwidth," in Proceedings of the
IEEE, vol. 57, no. 6, pp. 1195-1196, June 1969.
[40] Y. Ayasli, R. L. Mozzi, J. L. Vorhaus, L. D. Reynolds and R. A. Pucel, "A Monolithic GaAs
1-13-GHz Traveling-Wave Amplifier," in IEEE Transactions on Microwave Theory and
Techniques, vol. 30, no. 7, pp. 976-981, Jul. 1982.
[41] Ming-Da Tsai, Huei Wang, Jui-Feng Kuan and Chih-Sheng Chang, "A 70GHz cascaded multi-
stage distributed amplifier in 90nm CMOS technology," ISSCC. 2005 IEEE International
Digest of Technical Papers. Solid-State Circuits Conference, 2005., San Francisco, CA, USA,
2005, pp. 402-606 Vol. 1.
[42] A. Arbabian and A. M. Niknejad, "A tapered cascaded multi-stage distributed amplifier with
370GHz GBW in 90nm CMOS," 2008 IEEE Radio Frequency Integrated Circuits Symposium,
Atlanta, GA, USA, 2008, pp. 57-60.
[43] Po-Han Chen, Kuang-Sheng Yeh, Jui-Chih Kao and H. Wang, "A high performance DC-80
GHz distributed amplifier in 40-nm CMOS digital process," 2014 IEEE MTT-S International
Microwave Symposium (IMS2014), Tampa, FL, USA, 2014, pp. 1-3.
[44] Jonghae Kim et al., "A 12dBm 320GHz GBW distributed amplifier in a 0.12/spl mu/m SOI
CMOS," 2004 IEEE International Solid-State Circuits Conference (IEEE Cat.
No.04CH37519), San Francisco, CA, USA, 2004, pp. 478-540 Vol.1.
[45] J. Plouchart et al., "A 4-91-GHz traveling-wave amplifier in a standard 0.12-/spl mu/m SOI
CMOS microprocessor technology," in IEEE Journal of Solid-State Circuits, vol. 39, no. 9,
pp. 1455-1461, Sept. 2004.
[46] S. Masuda, T. Takahashi and K. Joshin, "An over-110-GHz InP HEMT flip-chip distributed
baseband amplifier with inverted microstrip line structure for optical transmission system," in
IEEE Journal of Solid-State Circuits, vol. 38, no. 9, pp. 1479-1484, Sept. 2003.

Page | 123
[47] Y. Li, W. -L. Goh and Y. -Z. Xiong, "A 2 to 92 GHz distributed amplifier using 70-nm InP
HEMTs," 2015 IEEE International Wireless Symposium (IWS 2015), Shenzhen, China, 2015,
pp. 1-4.
[48] K. Eriksson, I. Darwazeh and H. Zirath, "InP DHBT Distributed Amplifiers With Up to 235-
GHz Bandwidth," in IEEE Transactions on Microwave Theory and Techniques, vol. 63, no. 4,
pp. 1334-1341, April 2015.
[49] A. Arbabian, S. Callender, S. Kang, M. Rangwala and A. M. Niknejad, "A 94 GHz mm-Wave-
to-Baseband Pulsed-Radar Transceiver with Applications in Imaging and Gesture
Recognition," in IEEE Journal of Solid-State Circuits, vol. 48, no. 4, pp. 1055-1071, April
2013
[50] S. Kimura, Y. Imai, Y. Umeda and T. Enoki, "Loss-compensated distributed baseband
amplifier IC's for optical transmission systems," in IEEE Transactions on Microwave Theory
and Techniques, vol. 44, no. 10, pp. 1688-1693, Oct. 1996.
[51] J. Chen and A. M. Niknejad, "Design and Analysis of a Stage-Scaled Distributed Power
Amplifier," in IEEE Transactions on Microwave Theory and Techniques, vol. 59, no. 5, pp.
1274-1283, May 2011.
[52] M. EL Chaar, A. L. de Souza, M. Barragan, F. Podevin and S. Bourdel, "A Non-Closed-Form
Mathematical Model for Uniform and Non-Uniform Distributed Amplifiers," 2020 IEEE
MTT-S International Conference on Microwaves for Intelligent Mobility (ICMIM), 2020, pp.
1-4.
[53] Danh Luongvinh and Youngwoo Kwon, "Behavioral modeling of power amplifiers using fully
recurrent neural networks," IEEE MTT-S International Microwave Symposium Digest, 2005.,
2005, pp. 1979-1982
[54] Taijun Liu, S. Boumaiza and F. M. Ghannouchi, "Dynamic behavioral modeling of 3G power
amplifiers using real-valued time-delay neural networks," in IEEE Transactions on Microwave
Theory and Techniques, vol. 52, no. 3, pp. 1025-1033, March 2004
[55] Ali Reza Zirak and Sobhan Roshani, “Design and Modeling of RF Power Amplifiers with
Radial Basis Function Artificial Neural Networks” International Journal of Advanced
Computer Science and Applications (ijacsa), 7(6), 2016.
[56] D. M. Pozar, Microwave Engineering, 4th ed. Hoboken, NJ, USA: Wiley, 1998.
[57] Y. Tsividis, Operation and modeling of the MOS transistor, 2nd ed. New York; Oxford: Oxford
University Press, 1999.
[58] O. El-Aassar and G. M. Rebeiz, "A DC-to-108-GHz CMOS SOI Distributed Power Amplifier
and Modulator Driver Leveraging Multi-Drive Complementary Stacked Cells," in IEEE
Journal of Solid-State Circuits, vol. 54, no. 12, pp. 3437-3451, Dec. 2019.
[59] O. El-Aassar and G. M. Rebeiz, "A Cascaded Multi-Drive Stacked-SOI Distributed Power
Amplifier With 23.5 dBm Peak Output Power and Over 4.5-THz GBW," in IEEE Transactions
on Microwave Theory and Techniques, vol. 68, no. 7, pp. 3111-3119, July 2020.
[60] K. Moez and M. Elmasry, "A New Loss Compensation Technique for CMOS Distributed
Amplifiers," in IEEE Transactions on Circuits and Systems II: Express Briefs, vol. 56, no. 3,
pp. 185-189, March 2009
[61] M. El-Chaar, F. Podevin, S. Bourdel, A. A. L. de Souza and J. -D. Arnould, "Integrated Stacked
Parallel Plate Shunt Capacitor for Millimeter-Wave Systems in Low-Cost Highly Integrated
CMOS Technologies," in IEEE Solid-State Circuits Letters, vol. 5, pp. 114-117, 2022.
[62] M. Seo, B. Jagannathan, J. Pekarik, and M. J. W. Rodwell, “A 150 GHz Amplifier With 8 dB
Gain and +6 dBm Psat in Digital 65 nm CMOS Using Dummy-Prefilled Microstrip Lines,” in
IEEE Journal of Solid-State Circuits, vol. 44, no. 12, pp. 3410–3421, Dec. 2009

Page | 124
[63] A.-L. Franc, E. Pistono, D. Gloria, and P. Ferrari, "High-performance Shielded Coplanar
Waveguides for the Design of CMOS 60-GHz Band-pass Filters," in IEEE Trans. on Electron
Device, vol. 59, no.5, pp. 1219-1226, May 2012
[64] A. Ghadiri and K. Moez, "High-Quality-Factor Active Capacitors for Millimeter-Wave
Applications," in IEEE Transactions on Microwave Theory and Techniques, vol. 60, no. 12,
pp. 3710-3718, Dec. 2012
[65] T. Quémerais, L. Moquillon, P. Benech, J. Fournier and S. Pruvost, “CMOS 45-nm 3D metal-
oxide-metal capacitors for millimeter wave applications,” Microwave and Optical Technology
Letters, vol. 53, no. 7, pp. 1476-1478, Apr. 2011
[66] J. Shi, K. Kang, Y. Z. Xiong, J. Brinkhoff, F. Lin and X. Yuan, "Millimeter-Wave Passives in
45-nm Digital CMOS," in IEEE Electron Device Letters, vol. 31, no. 10, pp. 1080-1082, Oct.
2010
[67] G. F. Engen and C. A. Hoer, "Thru-Reflect-Line: An Improved Technique for Calibrating the
Dual Six-Port Automatic Network Analyzer," in IEEE Transactions on Microwave Theory and
Techniques, vol. 27, no. 12, pp. 987-993, Dec. 1979
[68] M. E. Chaar, A. L. de Souza, M. Barragan, F. Podevin, S. Bourdel and J. -D. Arnould,
"Integrated Wideband Millimeter-Wave Bias-Tee – Application to Distributed Amplifier
Biasing," 2021 19th IEEE International New Circuits and Systems Conference (NEWCAS),
2021, pp. 1-4
[69] I.Alaji, T. Capelli Mouvand, M. El-Chaar, A. Lisboa-de-Souza, S. Bourdel, F. Podevin,
“Cascaded tunable distributed amplifiers for serial optical links: some design rules," Proc. 18th
IEEE Int. New Circuits and Systems Conf. (NEWCAS), Montreal, Canada, June 16-19 2020,
remote conference
[70] F. Ellinger, "60-GHz SOI CMOS traveling-wave amplifier with NF below 3.8 dB from 0.1 to
40 GHz," in IEEE Journal of Solid-State Circuits, vol. 40, no. 2, pp. 553-558, Feb. 2005.
[71] T. Huang, Y. Lin, J. Cheng, J. Kao, T. Huang and H. Wang, "A high-gain low-noise distributed
amplifier with low DC power in 0.18-µm CMOS for vital sign detection radar," 2015 IEEE
MTT-S International Microwave Symposium, Phoenix, AZ, 2015, pp. 1-3
[72] V. Pulijala and S. Azeemuddin, "Spiral inductors for V-band wireless ICs," 2011 IEEE Applied
Electromagnetics Conference (AEMC), Kolkata, 2011, pp. 1-4.
[73] M. El-Chaar, L. Vincent, J. -D. Arnould, A. A. L. de Souza, S. Bourdel and F. Podevin,
"Accurate Design Method for Millimeter Wave Distributed Amplifier Based on Four-Port
Chain (ABCD) Matrix Model," in IEEE Transactions on Circuits and Systems I: Regular
Papers, vol. 69, no. 11, pp. 4510-4523, Nov. 2022.

Page | 125

You might also like