You are on page 1of 174

STUDY OF MICROWAVE METASURFACE

ANTENNAS USING CHARACTERISTIC MODE

ANALYSIS

FENG HAN LIN

NATIONAL UNIVERSITY OF SINGAPORE

2018
STUDY OF MICROWAVE METASURFACE

ANTENNAS USING CHARACTERISTIC MODE

ANALYSIS

by

FENG HAN LIN

B.Eng. (Hons.), Xidian University, China


M. Eng. (Hons.), Xidian University, China

A THESIS SUBMITTED
FOR THE DEGREE OF DOCTOR OF PHILOSOPHY
DEPARTMENT OF ELECTRICAL AND COMPUTER ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE

2018

Supervisor:
Professor Zhi Ning Chen

Examiners:
Professor Minghui Hong
Associate Professor Chengwei Qiu
Professor Buon Kiong Lau, Lund University
Declaration

I hereby declare that this thesis is my original work and it has


been written by me in its entirety. I have duly
acknowledged all the sources of information which
have been used in the thesis.

This thesis has also not been submitted for any degree in
any university previously.

Feng Han Lin

10 December 2018
Acknowledgements

In the first place, I would like to express my sincere gratitude to my

advisor, Prof. Chen Zhi Ning, for his continuous support, guidance,

encouragement, and patience throughout my journey of Ph.D studies, without

which this dissertation would not have been possible.

I would like to thank Prof. Qiu Chengwei and Prof. Hong Minghui for

being my doctoral committee and for their valuable suggestions and comments.

My sincere thanks also go to my lecturers of the electromagnetic courses,

specially, Prof. Chen Xudong, Dr. Wang Chao Fu, Prof. Guo Yongxin, and

also Prof. Qiu Chengwei, for passing me knowledge and sharpening my

thinking. Also, I would like to thank Dr. Qing Xianming, Dr. Terence See, and

Dr. Nasimuddin from ASTAR for helping me with some of my measurements,

and also Prof. Raj Mittra for fruitful discussions on the CM theory and

insightful comments on my papers.

I would like to thank all my friends in MMIC lab, Dr. Wang Zhengbing,

Dr. Zhou Yihong, Dr. Li Shunli, Dr. Liu Wei, Dr. Li Teng, Mr. Tay Chaiyan,

Mr. Siegfred Balon, Ms. Su Yuanyan, Ms. Sheng Huiwen, Dr. Deng Tianwei,

Dr. Srien Sithara, Dr. Amin Kianinejad, Dr. Liu Zhongtao, Dr. Teniou Mounir,

and many others for all the selfless help, discussions and joyful times through

the years. Outside the lab, I would like to thank my friends, Ms. Melissa Chan

for being my amazing dance partner, Ms. Sarah Kang for shaping my arts

journey, and Dr. Wang Tao, Mr. Luo Yuxuan, Mr. Jovin Hong, Mr. Shaun

Seah, Ms. Silvia Tieri, Dr. Jiang Wen, Ms. Chen Xuan and many others for

their support, companionship, and encouragement. My special thanks go to Ms.

i
Ng Chin Hui, for her supporting me through all the difficult moments with

endless love, patience, and smiles.

Mere words would fail to express my overwhelming gratitude to my family,

especially my mother, Mdm. Chen Hongwei, and my grandmother, Mdm.

Song Tianzhen, who have taught me kindness and bravery. They love me

unconditionally and support me with unwavering faith. I would like to thank

them for sacrificing everything so that I have a chance to know, to grow, to

dream, and to reach.

ii
Table of Contents

Declaration......................................................................................................... i

Acknowledgements ........................................................................................... i

Table of Contents ........................................................................................... iii

Summary .......................................................................................................... vi

List of Tables ................................................................................................... ix

List of Figures ................................................................................................... x

List of Abbreviations .................................................................................... xiv

List of Symbols .............................................................................................. xvi

List of Publications ......................................................................................xvii

Chapter 1 Introduction.................................................................................... 1

1.1 Metamaterials and metasurfaces for antennas ...................................... 4

1.2 Theory of characteristic modes........................................................... 12

1.3 Motivations and challenges ................................................................ 14

1.4 Organization of the thesis ................................................................... 16

Chapter 2 Modal Analysis of local-resonant MTSs .................................... 17

2.1 Unit cells in MTSs .............................................................................. 17

2.2 Spacing between MTS and ground ..................................................... 24

2.3 Dielectric loaded MTSs ...................................................................... 30

2.4 Summary............................................................................................. 32

iii
Chapter 3 Design of low-profile wideband local-resonant MTS antennas

.......................................................................................................................... 34

3.1 Design I: Slot-fed linearly polarized MTS antennas .......................... 34

3.1.1 Antenna design............................................................................. 34

3.1.2 Experimental results..................................................................... 52

3.2. Design II: Slot-fed dual-linearly polarized MTS antenna ................. 55

3.2.1 Antenna design............................................................................. 55

3.2.2 Experimental results..................................................................... 56

3.3 Design III: Dipole-fed MTS antenna with high FBR ......................... 61

3.3.1 Antenna design............................................................................. 61

3.3.2 Simulation results......................................................................... 65

3.4 Design IV: Probe-fed MTS antenna with optimal feeding placement 67

3.4.1 Antenna design............................................................................. 68

3.4.2 Experimental results..................................................................... 70

3.5 Summary............................................................................................. 73

Chapter 4 Modal synthesis of MTS antennas.............................................. 75

4.1 Mutual coupling in multi-antenna systems ......................................... 76

4.2 Methods of suppressing higher-order modes...................................... 83

4.2.1 Slot loading for shifting resonant frequencies ............................. 84

4.2.2 Via loading for cross-polarizing unwanted modes ...................... 86

4.3 Improvement of antenna radiation performance ................................ 90

4.3.1 Loaded MTS antenna with integrated balun ................................ 90

iv
4.3.2 Improved radiation performance.................................................. 94

4.4 Summary........................................................................................... 100

Chapter 5 Modal synthesis of MTS antennas............................................ 101

5.1. Truncated impedance-sheet model of global-resonant MTS antennas

................................................................................................................ 103

5.1.1 Control of fundamental mode by sheet impedance.................... 103

5.1.2 Analogy to waveguide modes .................................................... 111

5.1.3 Control of higher-order modes by frequency dispersion ........... 112

5.1.4 Realization of predicted modes .................................................. 114

5.2 Wideband triple-mode global-resonant MTS antennas .................... 118

5.3. Summary.......................................................................................... 129

Chapter 6 Conclusions and future work .................................................... 130

6.1 Conclusions ...................................................................................... 130

6.2 Future research ................................................................................. 133

6.2.1 From antenna radiation to scattering and absorption problems . 133

6.2.2 Optimization .............................................................................. 133

6.2.3 Non-broadside modes ................................................................ 133

6.2.4 From microwave to higher frequencies ..................................... 133

Bibliography ................................................................................................. 134

v
Summary

1. Motivation and challenges

Antennas-on-surface is the initial motivation of this work to facilitate the

development of a networked society. Realizing low-profile antennas with good

performance, especially in terms of large gain-bandwidth product, is a

classical problem in antenna design. Metasurfaces (MTSs), with rich

potentials in manipulating electromagnetic waves, have been shown promising

in improving the performance of low-profile antennas. However, most of

previous studies focus on infinite-sized and electrically-large MTSs under far-

field excitation, offering limited physical insights into the resonant behavior of

finite-sized MTSs under near-field excitation. The lack of a systematic

approach to take into consideration of the edge truncation, near-field

excitation and local field differences of finite-sized MTSs hinders further

developments of this field.

2. Objective and scope

With the primary objective of improving the performance of low-profile

antennas of broadside radiation and large bandwidth, this thesis aims to:

 introduce a new perspective to the modeling, analysis and design of

resonant MTSs;

 provide a fundamental understanding of the modal behaviors of finite-

sized MTSs consisting of a finite number of unit cells at different

resonant states;

 develop methods of selective modal excitation and modal manipulation

for high-performance antenna elements, antenna arrays, and multi-

vi
antenna systems;

 formulate a general framework for systematic design of resonant MTS

antennas.

Chapter 1 highlights the challenges of existing analysis techniques for

characterizing finite-sized resonant MTSs, followed by a detailed study of the

first class of MTSs (local-resonant MTSs) using characteristic mode analysis

(CMA) in Chapter 2. The revealed insights are then utilized to design four

low-profile wideband antennas of different features with special concerns of

exciting desired modes and minimizing the adverse effects of feeding

structures. Targeting MTS antenna arrays and multi-antenna systems, Chapter

4 introduces two approaches of mode manipulation in order to mitigate the

pattern distortion caused by near-field scattering between closely-spaced

antennas. Chapter 5 introduces a more powerful concept of global-resonant

MTS for fast synthesis of electrically-large resonant MTSs by leveraging the

concept of grid impedance. Chapter 6 concludes the thesis with several future

research directions outlined.

3. Novelty

Different from conventional excitation-dependent analysis in conjunction

with periodic boundary condition, characteristic mode analysis (CMA) is used

in an innovative manner to gain novel physical insights into the inherent

modal behaviors and radiation mechanisms of finite-sized resonant MTSs,

where the edge truncation and multi-mode near-field excitation, and local field

differences of MTSs are inherently considered.

4. Impacts

 Using source-independent characteristic mode analysis, the concepts of

vii
local-resonant and global-resonant MTSs are developed, offering novel

physical insights into the multi-mode resonant nature of finite-sized MTSs.

 A systematic analysis and design framework is formulated for resonant

MTS antennas.

 Four innovative approaches for practical feeding of resonant MTSs are

developed to meet different requirements.

 Two methods of mode manipulation, the mode shifting and mode cross-

polarizing, are proposed, offering a new wideband solution to significantly

mitigating the pattern distortion in closely-spaced multi-antenna systems.

 The method of modal synthesis using frequency-dispersive sheet

impedance is proposed for significant bandwidth enhancement of low-

profile MTS antennas.

The research work presented in Chapters 3, 4, and 5 have been published in

IEEE Transactions on Antennas and Propagation and IEEE conference

proceedings.

viii
List of Tables

Table 2.1 Summary of Resonant Frequency, MTS Size, and Unit-Cell


Size of Local-Resonant MTSs ..................................................... 19

Table 3.1 Summary of Antenna Geometry of the Slot-Fed Local-


Resonant MTS Antenna ............................................................... 36

Table 3.2 Summary of Antenna Performance for the Local-Resonant


MTS Antennas ............................................................................. 73

Table 4.1 Comparison of Antenna Performance for the Dipole-Fed MTS


Antenna with and without Suppressing HOMs ........................... 96

Table 5.1 Comparison of Antenna Performance for the Global-Resonant


MTS Antennas, Conventional MTS Antennas, and Wideband
Microstrip Patch Antennas ......................................................... 128

ix
List of Figures

Fig. 1.1 Conceptual illustration of near-field MTS with infinite size. ........ 7

Fig. 1.2 Conceptual illustration of far-field MTS with infinite size. .......... 9

Fig. 2.1 MTS of evolution and boundary setup. ....................................... 18

Fig. 2.2 Characteristic modal currents and field. ...................................... 21

Fig. 2.3 Field distribution of the odd and conventional TM03 modes.. ..... 21

Fig. 2.4 Effects of MTS-ground spacing on resonant frequency and


modal bandwidth.......................................................................... 24

Fig. 2.5 Equivalent circuit model of ground-backed MTS. ...................... 26

Fig. 2.6 Diagrams of input impedance and reflection coefficients. .......... 26

Fig. 2.7 Effects of dielectric loading on resonant frequency and modal


bandwidth..................................................................................... 29

Fig. 2.8 Effects of suspended dielectric layer on resonant frequency. ...... 30

Fig. 3.1 Configuration of the slot-fed local-resonant MTS antenna. ........ 35

Fig. 3.2 Simulation results. ....................................................................... 37

Fig. 3.3 Illustration of boundary setups for CMA. .................................... 38

Fig. 3.4 Modal analysis of ground-backed MTS without feeding slot...... 40

Fig. 3.5 Modal currents of the first four modes at 6 GHz. ........................ 40

Fig. 3.6 Modal radiation patterns of the first four modes at 6 GHz. ......... 40

Fig. 3.7 Modal currents and E-field distribution of the Q-TM03 mode at
6 GHz. .......................................................................................... 41

Fig. 3.8 Modal analysis of ground-backed MTS with top slot. ................ 43

Fig. 3.9 Modal analysis of metasurface with feeding slot......................... 44

Fig. 3.10 Illustration of potential antenna bandwidth. ................................ 46

Fig. 3.11 Parametric study based on CMA.. ............................................... 48

Fig. 3.12 Parametric study based on time-domain analysis. ....................... 51

x
Fig. 3.13 Configuration of the 2×2 MTS antenna array. ............................. 52

Fig. 3.14 Simulated and measured results of the MTS antenna array......... 53

Fig. 3.15 Simulated and measured radiation patterns of the MTS


antenna array (normalized). ......................................................... 54

Fig. 3.16 Configuration of the dual-linearly polarized MTS antenna ......... 55

Fig. 3.17 Simulated and measured results of the dual-linearly polarized


MTS antenna. ............................................................................... 58

Fig. 3.18 Normalized radiation patterns of Port 1. ...................................... 59

Fig. 3.19 Normalized radiation patterns of Port 2. ...................................... 60

Fig. 3.20 Configuration of the dipole-fed MTS antenna and boundary


setup for CMA. ............................................................................ 61

Fig. 3.21 Modal analysis of the MTS without dipole. ................................ 62

Fig. 3.22 Modal analysis of the MTS with dipole....................................... 63

Fig. 3.23 Magnitude of the modal weighting coefficients for the dipole-
fed MTS antenna. ......................................................................... 64

Fig. 3.24 Simulated results of the dipole-fed MTS antenna. ...................... 65

Fig. 3.25 Simulated radiation patterns of the dipole-fed MTS antenna. ..... 66

Fig. 3.26 Geometry of the probe-fed MTS antenna. ................................... 67

Fig. 3.27 Modal significances of the probe-fed MTS antenna. ................... 69

Fig. 3.28 Modal currents and E-field of the probe-fed MTS antenna.. ....... 69

Fig. 3.29 Fabricated prototype of the probe-fed MTS antenna. .................. 71

Fig. 3.30 Simulated and measured |S11| and boresight gain of the probe-
fed MTS antenna. ......................................................................... 71

Fig. 3.31 Simulated and measured radiation patterns of the probe-fed


MTS antenna. ............................................................................... 72

Fig. 4.1 Conceptual illustration of mutual coupling in multi-antenna


systems. ........................................................................................ 77

Fig. 4.2 Configuration of the four-port MTS antenna system................... 78

Fig. 4.3 Simulation results for the four-port MTS antenna system.. ......... 79

xi
Fig. 4.4 Distorted radiation patterns of the system. .................................. 80

Fig. 4.5 Simulated current distribution at 6 GHz on the shared ground. .. 81

Fig. 4.6 Modal significances of the first eight modes of the unloaded
MTS. ............................................................................................ 81

Fig. 4.7 Modal electrical currents and radiation patterns of the


unloaded MTS.............................................................................. 82

Fig. 4.8 Modal significances of the slots-loaded MTS. ............................ 84

Fig. 4.9 Modal currents and E-field distribution of the slot-loaded


MTS.. ........................................................................................... 85

Fig. 4.10 Modal significances of the slot-via loaded MTS. ........................ 88

Fig. 4.11 Modal currents of the slot-via loaded MTS without dipole. ........ 89

Fig. 4.12 Configuration of the MTS antenna with integrated balun. .......... 91

Fig. 4.13 Effects of integrated balun on the MTS antenna. ........................ 92

Fig. 4.14 Configuration of the slot-via loaded MTS antenna system.. ....... 93

Fig. 4.15 Comparison of current distribution with and without


suppressing HOMs. ...................................................................... 93

Fig. 4.16 Comparison of |S11| and boresight directivities of the three


cases. ............................................................................................ 95

Fig. 4.17 Simulated gain patterns of the three cases. .................................. 95

Fig. 4.18 Fabricated prototype of the slot-via loaded MTS. ....................... 97

Fig. 4.19 Simulated and measured results with HOMs suppressed. ........... 98

Fig. 4.20 Simulated and measured radiation patterns of the loaded MTS
antenna in the system (normalized). ............................................ 99

Fig. 5.1 Impedance-sheet model for global-resonant MTS antennas...... 104

Fig. 5.2 Modal manipulation by sheet-impedance. ................................. 106

Fig. 5.3 The scalar single-layered and bilateral sheet-impedance model.


................................................................................................... 108

Fig. 5.4 Modal synthesis results using the impedance-sheet model. ...... 109

Fig. 5.5 Modal electric currents for the BISM at 6 GHz......................... 110

xii
Fig. 5.6 Analogy to waveguide modes. ................................................... 111

Fig. 5.7 Effects of frequency dispersion. ................................................ 113

Fig. 5.8 Simulation setup for grid impedance extraction. ....................... 115

Fig. 5.9 Extracted grid impedance. ......................................................... 115

Fig. 5.10 Modal significances of the synthesized MTS. edge effects. ...... 117

Fig. 5.11 Modal currents and radiation patterns of the synthesized MTS. 118

Fig. 5.12 Combined currents of J1 and J6 for multi-mode excitation. ...... 119

Fig. 5.13 Geometries of the MTS antenna fed by dipoles. ....................... 120

Fig. 5.14 Effects of ground size on impedance matching and directivity. 121

Fig. 5.15 Illustration of the triple-mode operation with high excitation


purity. ......................................................................................... 122

Fig. 5.16 Magnitude of current distributions of the three dominant


modes. ........................................................................................ 123

Fig. 5.17 Geometry of the feeding network. ............................................. 124

Fig. 5.18 Photographs of the prototyped global-resonant MTS antenna. . 125

Fig. 5.19 Simulated and measured |S11| and boresight gain of the global-
resonant MTS antenna. .............................................................. 126

Fig. 5.20 Normalized radiation patterns of the global-resonant MTS


antenna. ...................................................................................... 127

xiii
List of Abbreviations

3D three-dimensional

BISM bi-layered impedance-sheet model

BOI band of interest

BW bandwidth

CM characteristic mode

CMA characteristic mode analysis

CPS coplanar-stripline

DPMA dual-linearly polarized metasurface antenna

EM electromagnetic

EMI electromagnetic interference

FBR front-to-back ratio

FSS frequency selective surface

GND ground

HIS high-impedance surface

HOM higher-order mode

IBC impedance boundary condition

ISM impedance-sheet model

LPMA linearly polarized metasurface antenna

MA metasurface antenna

MEC modal excitation coefficients

MoM method of moments

MS modal significance

MTM metamaterial

xiv
MTS metasurface

MWC modal weighting coefficients

PBC periodic boundary condition

PCB printed circuit board

PEC perfect electrical conducting

PMC perfect magnetic conducting

RF radio frequency

SC short-circuit

SISM single-layered impedance-sheet model

SMA subminiature version a

TCM theory of characteristic modes

TE transverse electric

TEM transverse electromagnetic

TLM transmission-line model

TM transverse magnetic

TO transformation optics

TSA tapered slot antenna

VSWR voltage standing wave ratio

xv
List of Symbols

β propagation constant rad/m

k wave number rad/m

λ wavelength m

ω angular frequency rad/s

f frequency 1/s

c0 speed of light m/s

ε permittivity F/m

μ permeability H/m

σ conductivity S/m

R resistance Ω

L inductance H

C capacitance F

V electric potential V

I electric current A

J surface electric current density A/m2

E electric field V/m

H magnetic field A/m

xvi
List of Publications

1. F. H. Lin and Z. N. Chen, “Low-profile wideband metasurface antennas

using characteristic mode analysis,” IEEE Trans. Antennas. Propag., vol.

65, no. 4, pp. 1706–1713, Feb. 2017. (CST UNIVERSITY

PUBLICATION AWARD 2017)

2. F. H. Lin and Z. N. Chen, “A method of suppressing higher-order modes

for improving radiation performance of metasurface multiport antennas

using characteristic mode analysis,” IEEE Trans. Antennas. Propag., vol.

66, no. 4, pp. 1894–1902, Apr. 2018.

3. F. H. Lin and Z. N. Chen, “Truncated impedance sheet model for low

profile broadband metasurface antennas using characteristic mode

analysis,” IEEE Trans. Antennas. Propag., vol. 66, no. 10, pp. 5043–5051,

Oct. 2018.

4. F. H. Lin and Z. N. Chen, “Probe-fed broadband low-profile metasurface

antennas using characteristic mode analysis,” Proc. 2017 IEEE 6th Asia-

Pacific Conf. on Antennas Propag. (APCAP 2017), Xi’an, China, Oct.

2017, pp. 664–666.

5. F. H. Lin and Z. N. Chen, “A metamaterial-based broadband circularly

polarized aperture-fed grid-slotted patch antenna,” Proc. 2015 IEEE 4th

Asia-Pacific Conf. on Antennas Propag. (APCAP 2015), Kuta, Indonesia,

Jul. 2015, pp. 353–354. (STUDENT PAPER AWARD)

xvii
Chapter 1

Introduction

The use of electromagnetic (EM) waves for communications can be dated

long back to thousands of years ago, when people lit up a torch to inform the

intrusion of enemies. It is only since two centuries ago, when Maxwell first

predicted (1865) and then Hertz observed (1888) the EM wave at the

microwave frequencies, have human’ lives been overwhelmed by the modern

inventions of wireless communication, radars, imaging, remote sensing, and

recently the internet of things. In all microwave wireless systems, antenna is a

vital component which effectively converts the EM waves from a bounded

mode in the waveguide to a desired propagating mode in space, and vice versa.

Physically, the mode conversion is realized by implementing a desired

distribution of surface currents over the antenna in terms of amplitude, phase,

and direction, with each being a function of frequency and space. According to

the Maxwell’s equations, antenna design is to design the boundary conditions

(geometry and materials) and the excitations (feeding structures and input

signals) such that the solution of surface currents or aperture fields realizes the

desired distributions for far-field radiation. The challenge of antenna design

comes when there are strict constraints on the boundary conditions, the

excitations, or both.

1
One of the constraints is low antenna profile for surface integration, a

growing trend in a vast number of applications, involving antenna integration

in aircrafts, ships, base stations, radars, and consumer electronics. An

ambitious goal is to integrate everything, including radiators, feeding systems,

and beamforming networks on a thin surface of thickness smaller than a tenth

even a hundredth of the operating wavelength. However, the limitation of

antenna height greatly challenges antenna design. First of all, it is difficult to

achieve wideband broadside radiation and low profile simultaneously [1].

Furthermore, if multiple antennas are closely packed for small footprint and

high integration, the strong mutual coupling and electromagnetic interference

(EMI) between antennas are critical for the performance of multi-antenna

systems, be it a multi-port antenna or an antenna array [2]–[8]. As a result,

achieving low-profile wideband antennas with low mutual coupling becomes

the main challenge that hampers the surface integration.

Many low-profile wideband antennas have been developed over the years.

The antennas work in either traveling-wave modes or standing-wave modes.

Tapered slot antennas (TSA) are popular planar traveling-wave antennas with

wideband radiation in the end-fire direction [9]. Given unlimited antenna

length, the impedance bandwidth is theoretically infinite; while in practice, the

impedance bandwidth is limited by the finite tapering length [9]–[12] and also

the bandwidth of the feeding transition [13]. The total antenna height (in the

direction of maximum radiation) is usually larger than half of the wavelength

at the lowest operating frequency. With the evolution of TSAs, the height can

be reduced to a quarter wavelength by using tightly-coupled antenna arrays or

long-slot antennas [14]–[18]. Planar leaky-wave antennas are wideband

2
antennas with a profile much less than a quarter wavelength in the direction of

maximum radiation [19]–[23]. However, the frequency scanning behavior of

such antennas is not acceptable when a wideband beam pointing at a certain

direction is required for high data rate and targeting. To avoid the frequency-

scanning behavior, antennas working in standing-wave modes are preferred.

Resonant wire antennas and resonant slot antennas are two classical types

of standing-wave antennas. Wideband resonant wire antennas, such as ground-

backed dipoles, have been widely used in communication systems [23]–[24].

The parallel spacing between the dipole and the ground is usually a quarter

wavelength for high gain in the broadside direction. As the dipole approaches

the conducting ground, the antenna bandwidth decreases dramatically. In

extreme cases, the radiation from the wire and its antiphase image cancels

each other in the broadside direction, leading to a null at the broadside and an

extremely low radiation resistance. On the other hand, microstrip patch

antenna is an important type of resonant slot antennas with broadside radiation

[25]–[27]. By virtue of the in-phase image of parallel magnetic currents over

electric conductors, the antenna height can be extremely small, typically less

than one tenth of the operating wavelength. The major drawback of the low

profile is the inherent limitation of narrowband operation due to the single-

mode operation with a high quality factor. Conventional techniques to enhance

the bandwidth of microstrip patch antenna have been well summarized in [1],

including the use of capacitive coupling feed, stacked patches, thick air

substrates, parasitic resonators, and slot loaded patches. Most of these

techniques result in a considerable increase of the antenna height or a distorted

radiation pattern, such as beam split or beam squinting.

3
With limited antenna height, the key to improving the bandwidth of

resonant patch antennas is to increase the number of operating modes and the

individual modal bandwidth with wide enough bandwidth of the feeding

system. Given single-mode operation, the modal bandwidth is inversely

related to the modal quality factor. According to the Chu’s limit of the quality

factor [28]–[29], improving the modal bandwidth of thin planar resonant

antennas requires an expanded aperture for the surface currents (either electric

currents or magnetic currents) to distribute. For patch antennas, the technique

of aperture expansion is not well developed because increasing the patch size

leads to a redshift of the resonances when the current path is lengthened but

the wave propagation constant remains the same. In this sense, methods of

modifying the propagation constant are demanded to compensate for the

frequency shift. On the other hand, the smaller the frequency dispersion of the

wave propagation constant, the larger the modal bandwidth results. Therefore,

modifying the wave propagation constant is essential for achieving a low-

profile wideband resonant antenna.

Physically, the propagation constant is directly related to the characteristics

of both the medium where wave propagates and the surface where the currents

distribute. The question is how to artificially engineer the properties of the

medium and the surface, and link them to the antenna performance.

1.1 Metamaterials and metasurfaces for antennas

Electromagnetic metamaterials (MTMs) are composite materials whose

volume-averaged permittivity ε, permeability μ, and refraction index n can be

artificially engineered by modifying the composing unit cells to manipulate

4
EM waves [30]–[34]. With only conductors and non-magnetic dielectrics, the

retrieved effective μ, ε, and n can be arbitrary in principle, even taking

negative or zero values (for a specific mode of incidence in terms of frequency,

polarization and incident angle). Numerous breakthroughs and advancements

in science and technologies have been reported, such as MTM lens for super-

resolution imaging, invisibility cloaking, tunneling of EM waves, wave

bending, and negative refraction [30], [35]–[39]. However, most of the

exciting results rely on the extreme values of the retrieved effective

parameters resulted from strong resonances, where the resultant narrow

bandwidth and high loss may have become two of the biggest obstacles for

applying MTMs in antennas. Without pursuing the extreme values, the

transformation optics (TO) derived from coordinate transformation has been

proposed to generate the desired material profile for manipulating EM waves

[40]. Though perfect in theory and simulation, realistic application of TO is

usually hampered by the difficulties of fabricating the unit cells with

extremely fine features [41]–[42]. Apart from the fabrication issue, the bulky

size of three-dimensional (3D) MTMs also makes it less attractive for low-

profile antennas at microwave frequencies. Most importantly, the effective

medium properties of MTMs are mode-specific. The parameters retrieved by

far-field excitation are different from those in the reactive near-field region of

an antenna. Nevertheless, the concept of microscopic engineering for artificial

macroscopic properties has inspired and led to the birth of metasurfaces

(MTSs), which is two-dimensional and promising for low-profile surface-

integrated antennas.

Similar to the concept of MTMs, an MTS is formed by an electrically thin

5
layer of scatters or apertures arranged on the surface of a host medium, which

exhibits extraordinary EM properties not readily available in nature, such as

high surface impedance and in-phase reflection, band gap of wave propagation,

strong anisotropy or chirality, gradient surface-wave refraction index, and

dispersive scattering coefficients and propagation constants [43]–[46].

Applications of MTSs in the microwave regime include antenna gain

enhancement [47]–[48], antenna beam steering [49], multi-beam generation

[50], polarization conversion [51], antenna scattering control [52]–[53],

antenna miniaturization [54]–[55], impedance surface waveguides [56],

antenna cloakings [57], antenna bandwidth enhancement [58], radar-cross-

section reduction [59], and near-field beam shaping [60]–[62], to name a few.

The conventional analysis methods of MTSs are based on the assumption

of infinite periods of the unit cells. It allows the effective parameters, such as

dispersion diagram, effective surface-wave refraction index, surface

impedance tensor, and reflection diagrams, to be retrieved by studying a single

unit cell [43]–[62]. However, the retrieved effective properties are generally

mode-specific or spatially dispersive. For example, the reflection phase for a

normally incident transverse electric (TE) wave usually differs from that for

an obliquely incident transverse magnetic (TM) wave. In addition, these

methods neglect the finite size hence the edge diffraction of the MTS.

MTSs can be categorized by the field region of operation and the resonant

state of the composing unit cells.

6
Fig. 1.1 Conceptual illustration of near-field MTS with infinite size.

Firstly, an MTS can be a near-field MTS or a far-field MTS. The

conceptual illustrations of the near-field and far-field MTSs are shown in Figs.

1.1 and 1.2, respectively. Both types of MTSs are assumed to be of infinite

size.

The near-field MTS is an MTS placed close to the primary source in

Region 1, as shown in Fig. 1.1, at a distance h which is usually smaller than a

one-tenth of the operating wavelength. The ground plane is optional.

Depending on the surface properties of the MTS, for example the surface

̿ (x,y), different propagating-wave modes on the


impedance tensor distribution 𝒁

surface can be supported, such as the TE waves for capacitive surfaces and

TM waves for inductive surfaces [56]–[57]. The propagating modes can be

either bounded modes (also known as surface-wave modes) or radiative modes

(also known as leaky-wave modes). By exciting the leaky-wave modes with

7
designed weightings, a tailored near-field profile (such as the Bessel beam) or

shaped far-field radiation (such as beam-shaped holographic leaky-wave

antennas) can be achieved in Region 2 [59]–[50], [53]–[55], [59], [61]–[63].

On the other hand, by exciting the surface-wave modes (or the slow-wave

modes), surface waveguides can be achieved [56]–[57].

The far-field MTS is an MTS located in the far-field region of the primary

source to produce the desired secondary wavefront, as shown in Fig. 1.2. In

general, the MTS is illuminated by a far-field source which can be analytically

treated as a combination of a series of TE and TM plane waves of different

frequencies and incident angles. In practice, a TEM plane wave of normal

incidence is the most commonly used. The MTS is designed to manipulate the

scattering response to individual incident modes so that the vector summation

of the scattering field altogether generate the desired transmitted beam in

Region 2 or reflected beam in Region 1 or both in the far-field. For absorbers,

a null of scattering field in both regions is desired. Classical applications of the

far-field MTSs are frequency-selective surfaces (FSSs), transmitarrays,

reflectarrays, and absorbers, depending on the objective of manipulation. The

far-field MTSs can be designed to manipulate the frequency (f), amplitude (T

and R), phase (φ), or polarization of EM waves. For example, the FSSs can be

regarded as frequency filters [64]–[65], the amplitude of the scattering

coefficients can be manipulated for absorbers and matching layers [66]–[67],

the phase distribution can be tailored for beamforming [48] or scattering

reduction [60], and the polarization-dependent surface properties can be

utilized for beam splitter [47] and scattering reduction [52].

8
Fig. 1.2 Conceptual illustration of far-field MTS with infinite size.

9
Secondly, an MTS can be non-resonant, local-resonant or global-resonant

at the frequency of interest.

The non-resonant MTS, with an infinite size, is an MTS composed of

electrically small and non-resonant unit cells. It can be understood as a special

boundary condition that modifies the wave propagation over the surface. Most

of the near-field MTSs shown conceptually in Fig. 1.1 are non-resonant MTSs.

Since there is no reflection from the edges due to the infinite size, the ideal

non-resonant MTSs operate with travelling-wave modes, which are complex

modes with phase variation upon propagating. In practice, an electrically large

MTS is commonly used to mimic an infinitely large MTS and there is a small

portion of unwanted standing wave caused by the edge reflection, so the

operating mode is a travelling-standing-wave mode indeed.

The local-resonant MTS, regardless of its overall size, is an MTS

composed of relatively large and resonant unit cells. Most of the far-field

MTSs shown conceptually in Fig. 1.2 are local-resonant MTSs for high

transmission or reflection, such as the FSS and reflectarray. Under a wave

incident, each resonant unit cell of the MTSs operates like a resonant antenna

that produces a local difference of phase and magnitude between the received

and the scattered signals.

The global-resonant MTS, with a finite size comparable to the operating

wavelength, is an MTS composed of electrically small and non-resonant unit

cells. Different from the non-resonant MTS of infinite size, the resonant

modes of the global-resonant MTS are real standing-wave modes caused by

the finite edge truncation. This phenomenon has been briefly touched by Pozar

in 2001 and used for the design of a perforated patch antenna with

10
experimental trial [68]. In 2011, the analysis of finite MTS using the

approximate dispersion diagram of infinite MTSs is reported based on a

transmission-line model [59].

The MTSs studied in this thesis are the local-resonant MTSs and global-

resonant MTSs, both are near-field MTSs with finite size comparable to the

wavelength of interest and they operate with real standing-wave modes.

Frequently, an MTS is used as an auxiliary component, rather than the

main radiator, to transform the wavefront of a primary wave to a secondary

wave by phase compensation, polarization conversion, and frequency selection

in the far-field. The MTS is usually electrically large to mimic an infinite MTS

and to minimize the edge diffraction.

Differently, an MTS antenna uses a finite-sized MTS as the main radiator

located in the reactive near-field of the feeding structure. The edge truncation

significantly affects the antenna performance. Due to the complicated near-

field distribution, the attempts to characterize the MTS with the effective

medium parameters (ε-μ-n) and scattering coefficients (Sij) retrieved from a

plane-wave excitation may not be able to accurately predict the behavior of the

MTSs. The dispersion diagram and surface impedance tensor have been

shown valid for leaky-wave radiation and surface-wave guides provided that

the MTSs are electrically large so the edge truncation effects are minor [43]–

[46], [49]–[51]. [53]–[56], [61]–[63], [69]–[76].

For standing-wave MTS antennas, efforts to predict and design the

resonances of the MTSs with the dispersion diagram are less successful. The

major difficulty is the characterization of a finite-sized MTS in the presence of

11
feeding structures and near-field excitation. Firstly, the edge truncation effects

of the MTS are not counted in the dispersion diagram extracted from an

infinite-sized MTS. Secondly, the realistic near-field excitation is complicated

and usually a combination of multiple modes, whereas the dispersion diagram

is mode specific. Thirdly, the assumption of periodic boundary condition

(PBC), though greatly simplifies the analysis of uniformly periodic MTSs with

electrically large sizes, fails to take into account the local field difference. For

example, the local field difference can be caused either by partially loading the

unit cells intentionally for improving antenna performance or by the feeding

structures (such as a slot, a microstrip stub, a patch, or a cable) for exciting the

MTS resonator.

It is desired to directly study the standing-wave modes of the MTS antenna

as an entity and to find the relationship between the modes of the entity and

those of the standalone MTS to reveal the roles of an MTS in generating and

manipulating the modes.

1.2 Theory of characteristic modes

The sourcing of modal solution to arbitrary structures has led us to the

theory of characteristic modes (TCM). The TCM was proposed by Garbacz in

1965 and reformulated by Harrington and Mautz [77]–[79]. The TCM studies

the integral operator that relates the surface currents to the scattered field in

the method of moments (MoM) [80]. The operator is discretized and presented

in the form of a matrix. By solving the n-th order eigencurrent Jn and the

associated eigenvalue λn from a generalized eigenvalue equation, the total

current J on a perfect electrical conducting (PEC) body under the impressed

12
electric field (E-field) Ei can be decomposed as J = ∑αn Jn, where αn is the

complex modal weighting coefficients (MWC). The TCM is then extended to

dielectric/magnetic materials [81]–[82] and composite bodies of PEC

conductors and dielectric materials [83]–[85]. The contribution of each Jn to

the total radiated power Prad in the far-field is measured by Prad = ∑|αn|2. Due

to the special formulation, all the surface currents are real or equiphasal,

meaning that the current modes studied by the TCM are standing-wave modes.

The eigenvalue λn measures the ratio of the net stored power to the radiated

power, therefore a mode resonates when λn is zero or when the modal

significance MS equals to unity, where MS = 1/|1+jλn|. The modal excitation

coefficient (MEC) Vi is defined as the inner product of the Jn and Ei. Therefore,

αn = MS ∙ Vi.

Mathematically rigorous and physically insightful, the TCM provides an

elegant modal approach to understand the operating mechanism of arbitrary

resonant scatterers [86]–[90] and antennas [91]–[118]. In antenna engineering,

the TCM has been widely employed for analysis and synthesis of antennas

[90]–[94], manipulation of antenna modes [95]–[96], design of wide-/multi-

band multi-antenna systems [97]–[100], analysis of antenna coupling [101]–

[102], design of null-steering antennas [103]–[105], design of platform-

mounted antennas [107]–[109], bandwidth enhancement of electrically small

antennas [110], and optimization of antenna shape [111].

A patch antenna loaded with a reactive-impedance substrate, which is

originally proposed in [54], is also analyzed using the TCM with the primary

interest of calculating the quality factor of the patch antenna mode [112].

However, the associated modal currents and radiation patterns are not

13
investigated. Also, the characteristic modes (CMs) of the finite-sized reactive-

impedance substrate are not addressed.

Recently in 2018, there has been increasing interests in utilizing TCM for

the analysis and design of various metasurface/metamaterial antennas,

including a metamaterial lens antenna composed of metallic rings for gain

enhancement [113], a metasurface antenna using H-shaped unit cells for

circular polarization [114], an metasurface antenna using a higher-order mode

for omnidirectional radiation [115], a patch-fed low-profile metasurface

antenna for wideband application [116], and metasurface antennas for dual-

band applications [117]–[118].

Two key features of the TCM are important to this thesis. First, the modes

are fully defined once the material and geometry of the antenna are specified.

Second, characteristic mode analysis (CMA) is independent of any excitation

so no source or port is required. Therefore the results are also independent of

any excitation, hence valid for arbitrary excitations.

1.3 Motivations and challenges

Antennas-on-surface is the initial motivation and also the ultimate goal of

this work. It is desired that the radiator and the feeding system can be

integrated onto a thin surface. With surface integration, the size, weight, cost,

and heat management of the antenna can be greatly reduced and simplified.

Other than the benefits, the challenges are also multifold. The first bottleneck

is the limited gain-bandwidth product of low-profile resonant antennas. Also,

innovative antenna feeding is demanded for good impedance matching and

selective mode excitation. For the co-existence of multiple antennas, the

14
mutual coupling between distinct antennas affects the performances of the

multi-antenna system.

MTSs have shown rich potential in manipulating EM waves not only in the

far-field but also in the near-field; however, the theory of resonant MTS

antennas is not well established due to the difficulties of dealing with the finite

edge truncation, the near-field coupling between MTSs and feeding structures,

and complicated near-field excitation source. The combination of TCM and

MTS provides a new perspective in the study of MTS antennas and a unique

opportunity to manipulate and synthesize the modes with artificially

controllable surface properties. On one hand, the TCM provides a source-free

tool to analyze the modal behaviors of an MTS with the finite truncation and

the structural coupling both inherently included. On the other hand, MTS

serves as a tool to generate and manipulate the modes. With the combination

of the analysis and synthesis tools, an arbitrary surface current distribution

may be realized on a thin surface. Note that this concept is different from the

classical antenna array theory in that the mutual coupling has been inherently

considered.

Combining the concepts of MTS and CMA, this thesis aims to:

 Introduce a new perspective to the modeling, analysis and design of MTSs

 Provide a fundamental understanding of the modal behaviors of finite-

sized MTSs

 Develop the methods of modal excitation and modal manipulation for

high-performance MTS antennas

 Build a generalized model for systematic design of MTS antennas

15
1.4 Organization of the thesis

Chapter 1 briefly introduces the background of the MTMs, MTSs, MTS

antennas, and the TCM. The motivation, challenges and the unique

opportunities of combining the TCM and MTS are outlined. Chapter 2

presents the CMA of finite-sized MTS composed of resonant unit cells. The

emphasis is on the concept of local-resonant MTSs. The effects of different

unit cells, MTS-ground spacing, and dielectric loadings on the CMs of MTSs

are studied. In Chapter 3, four excitation schemes are proposed for selective

excitation of the desired modes for low-profile wideband MTS antennas. Each

design has a specific focus and objective. Chapter 4 studies the mutual

coupling and near-field scattering between distinct MTS antennas in a multi-

antenna system. The effects of the higher-order modes of the MTSs are found

to be the main cause of pattern distortion. A method of suppressing the higher-

order modes is proposed and validated for controlling the near-field scattering

and mitigating the pattern distortion. Chapter 5 presents the advancement of

the theory and modeling of global-resonant MTS antennas. Independent of any

specific unit cells, a systematic method of controlling the modal resonances

for synthesizing an MTS is introduced. The emphasis is on the concept of

global-resonant MTSs. The thesis is concluded in Chapter 6, followed by a

brief discussion of possible future work.

16
Chapter 2

Modal Analysis of local-resonant MTSs

An MTS is uniquely defined once the sizes, shapes, lattices, and periods of

the unit cells are determined, and so are its CMs [77]–[79]. This chapter

presents a quantitative study of the effects of various unit cells and

environments on the CMS of MTSs using CMA. The study provides clear

physical insights of MTS resonances and useful guidelines for controlling

modal resonance for MTS antenna designs.

2.1 Unit cells in MTSs

Three groups of case studies are performed for illustrating the concept of

local-resonant MTSs and the effects of varying unit cells on MTS modes,

where the MTS modes stand for the CMs due to the MTS resonator. The

emphasis is on the fundamental mode with broadside radiation (broadside

mode in short in the following content).

The evolution of the MTS of interest is shown in Fig. 2.1(a). For all the

configurations, the resonant frequency of the fundamental broadside mode is

solved through the CMA using CST 2016 with the boundary setup shown in

Fig. 2.1 (b) [119]. The spacing between the MTSs and the PEC ground of

17
(a)

(b)
Fig. 2.1 MTS of evolution and boundary setup. (a) Geometrical evolution
of the MTS from a solid square patch. (b) Boundary setup for
CMA in the CST.

infinite size is h = 2 mm. The results are summarized in Table 2.1. Group I in
Table 2.1Fig. 2.2 MTS of evolution and boundary setup. (a) Geometrical
evolution of the MTS from
Table 2.1 includes the configurations a solid
A, B, C, andsquare patch.
D in Fig. (b) Boundary
2.1(a), where all
setup for CMA in the CST.
the MTSs are with the same width of W = 25 mm, corresponding to half a

wavelength at 6 GHz in free space. The solid PEC square patch is divided into

multiple sub-patches with different periods of N (N = 1, 2, 3, 4). The spacing

18
Table 2.2
Summary of Resonant Frequency, MTS Size, and Unit-Cell Size
of Local-Resonant MTSs
Resonant Wave Width of
Config- Size of
N frequency length at Ws period
-uration MTS
(fr) fr (λ0) (Wp+Ws)
25 mm 25 mm
A 1 5.3 GHz 56.6 mm N. A.
(0.44 λ0) (0.44 λ0)
25 mm 12.25 mm
B 2 9.1 GHz 33.0 mm 1 mm
Group I

(0.76 λ0) (0.37 λ0)


25 mm 8.33 mm
C 3 12.8 GHz 23.4 mm 1 mm
(1.07 λ0) (0.36 λ0)
25 mm 6.25 mm
D 4 17.1 GHz 17.5 mm 1 mm
(1.43 λ0) (0.36 λ0)
8.33 mm 8.33 mm
C1 1 15 GHz 20 mm 1 mm
(0.42 λ0) (0.42 λ0)
16.7 mm 8.33 mm
C2 2 13.4 GHz 22.4 mm 1 mm
(0.74 λ0) (0.37 λ0)
25 mm 8.33 mm
C3 3 12.8 GHz 23.4 mm 1 mm
(1.07 λ0) (0.36 λ0)
Group II

33.7 mm 8.33 mm
C4 4 12.5 GHz 24 mm 1 mm
(1.40 λ0) (0.36 λ0)
42.3 mm 8.33 mm
C5 5 12.3 GHz 24.4 mm 1 mm
(1.73 λ0) (0.35 λ0)
51 mm 8.33 mm
C6 6 12.2 GHz 24.6 mm 1 mm
(2.07 λ0) (0.35 λ0)
68.3 mm 8.33 mm
C8 8 12.2 GHz 24.6 mm 1 mm
(2.78 λ0) (0.35 λ0)
1 mm 50 mm 16.7 mm
E 3 6.8 GHz 44.1 mm
(0.023 λ0) (1.14 λ0) (0.38 λ0)
0.5 mm 50 mm 16.7 mm
E0.5 3 6.3 GHz 47.6 mm
(0.011 λ0) (1.05 λ0) (0.35 λ0)
0.3 mm 50 mm 16.7 mm
Group III

E0.3 3 6.1 GHz 49.2 mm


(0.006 λ0) (1.02 λ0) (0.34 λ0)
0.2 mm 50 mm 16.7 mm
E0.2 3 5 GHz 60 mm
(0.003 λ0) 0.83 λ0) (0.28 λ0)
50 mm 16 mm
F 3 4.8 GHz 62.5 mm 1 mm
(0.8 λ0) (0.27 λ0)
50 mm 11.1 mm
G 3 9.1 GHz 33.0 mm 1 mm
(1.5 λ0) (0.35 λ0)

19
between adjacent patches is Ws = 1 mm and the width of the sub-patches is Wp

= W–(N–1)Ws. Group II in Table 2.1 includes variations based on

configuration C in Fig. 2.1(a). The periodicity N increases from one to eight

with the unit cell unchanged. Group III in Table 2.1 includes the

configurations E, F, and G in Fig. 2.1(a). For Group III, the overall dimension

of the MTS is increased to 2W = 50 mm. Four variations of configuration E are

involved, where Ws changes from 1 mm to 0.2 mm. From the configurations E

to F, the shape of the unit cell is changed from a square patch to a square loop.

From the configurations E to G, the shape of the lattice is changed from a

square to a parallelogram. The unit cells and the ground are both PEC and

infinitely thin. Vacuum is assumed everywhere else.

The fundamental broadside modes of the MTSs are caused by the local

resonance of the unit cells. As can be seen from Group I of Table 2.1, the

resonant frequency fr increases as N becomes larger. It means the electrical

size of the MTS at fr also increases with N. However, when the overall

electrical size of the MTS varies from 0.7 λ0 to 1.5 λ0, the resonant size of the

unit cell appears bounded in a small range from 0.26 λ0 to 0.37 λ0 in the

investigated cases, where λ0 is the wavelength at the case-specific resonant

frequency fr.

The local resonance occurs for each unit cell, where the capacitive coupling

between adjacent patches behaves as a series capacitor that lowers the

resonant frequency of the patch. Using configuration C as an example, the

vector surface current distribution and the associated vector E-field

distribution of the fundamental broadside mode are plotted in Figs. 2.2(a) and

2.2(b). The observation plane is at the lower interface between

20
(a)

(b)

Fig. 2.2 Characteristic modal currents and field. (a) Vector surface current
distribution of the first broadside mode of MTS of Configuration
C. (b) The associated vector E-field distribution over the MTS.

(a) (b) (c)

Fig. 2.3 Field distribution of the odd and conventional TM03 modes. The
MTS supports both the (a) odd and (b) conventional TM03 modes.
A solid patch only supports (c) the conventional TM03 mode.

21
the MTS and the vacuum. It is clearly seen that the MTS is essentially an array

of tightly-coupled sub-resonators, where a TM01 resonance occurs for each

unit cell. For this reason, the MTS is named as a local-resonant MTS which

resonates at an odd TM03 mode (o-TM03 in short). In this thesis, an odd mode

is a mode supported exclusively by the N×N MTS but not supported by a solid

PEC square patch, such as the o-TM03 mode. On the other hand, a TM mode

without a prefix is supported by both the patch and the MTSs studied, such as

the TM03 mode.

The o-TM03 mode is different from the conventional TM03 mode of a

conventional patch antenna. The difference is illustrated in Fig. 2.3. The TM03

mode exhibits antiphase currents while the o-TM03 mode supports all in-phase

currents. With the electric charge distribution shown in Fig. 2.3(a), E-fields

tangential to the surface are supported due to the electric potential difference

(voltage) crossing the slots, where a PEC symmetry is presented. Due to the

existence of the tangential E-fields on the surface, the ground-backed MTSs of

configurations B, C, and D are essentially capacitive surfaces where the slots

can be treated as capacitors loaded on the surface. The tangential E-fields

vanish when there is no electric potential difference across the slots, which

gives rise to the TM03 mode. Fig. 2.3(b) shows the current and E-field

distributions of the TM03 mode, where the field pattern does not require a

potential difference crossing the slots so a PMC symmetry is presented.

Therefore, the TM03 mode is also supported by the MTS. The TM03 mode for

configuration C in Group I resonates at 16.7 GHz. Since there is no potential

difference across the slots for the TM03 mode, closing the slots does not affect

the field pattern so a solid patch also supports the TM03 mode, as shown in Fig.

22
2.3(c). However, a solid patch does not support the o-TM03 mode since

tangential E-field vanishes on a solid PEC surface. Consequently, the MTS of

configuration C supports both the TM03 mode (16.7 GHz) and the o-TM03

mode (12.8 GHz), whereas a solid patch only supports the TM03 mode. Similar

analysis can be performed to configurations B and D, which support the o-

TM02 and o-TM04 modes, respectively.

With the local resonance, it is expected that the resonant frequency does not

vary noticeably with increasing number of periodicity for unchanged unit cells.

With the unchanged Wp and Ws as the configuration C, the number of

periodicity is increased gradually from 3 to 8, giving rise to Group II, where

the footnote of CN means the number of periodicity is N (N = 1, 2, 3, 4, 5, 6, 8).

The results are summarized in Table 2.1. As can be seen, both the resonant

frequency and the electrical size of the unit cell are relatively stable. It is

concluded that the resonant frequency of the fundamental broadside mode of

the local-resonant MTSs is mainly determined by the unit cells and relatively

independent of the overall size of the MTS.

Determined by the unit cells, the resonant frequency of the MTS can be

controlled by modifying the unit cells. For metallic patches, the resonant

frequency can be manipulated by tuning the slot width, namely Ws, which

changes the coupling strength between the patches. Based on configuration E,

Ws is varied from 1 mm to 0.2 mm, corresponding to the configurations E0.5,

E0.3, and E0.2 in Table 2.1 for demonstration. As Ws decreases, the resonant

frequency also decreases. Therefore, the general rule is the narrower slot, the

lower the resonant frequency.

23
Besides the coupling strength, the geometries of the unit cell and the lattice

also affect the resonant frequency. As can be seen from Group III of Table 2.1,

the resonance shifts from 6.8 GHz (E) to 4.8 GHz (F) when the shape of the

unit cell is modified to a square loop. This is equivalent to enhancing the

effective inductance of the resonant unit cell, which decreases the resonant

frequency. When the lattice shape is transformed from a square to a triangle,

the resonant frequency increases from 6.8 GHz (E) to 7.7 GHz (G). This is

equivalent to decreasing the effective inductance of the unit cell since the size

of the patch is reduced.

11 Resonant frequency
50
Resonant frequency (GHz)

Bandwidth
10 40

Bandwidth (%)
9 30
27.8 %

8 20
10.9 %
7 10
Patch (W =14.6 mm)
6 0
10 8 6 4 2 0
Height h (mm)

(a) (b)
Fig. 2.4 Effects of MTS-ground spacing on resonant frequency and modal
bandwidth. (a) Configuration of the ground-backed MTS. (b)
Resonant frequency and bandwidth against h.

2.2 Spacing between MTS and ground

The spacing between the MTS and the ground is h. The effects of varying

h on the resonant frequency and modal bandwidth of the broadside mode are

further studied. The emphasis is on the modal bandwidth, defined as the

frequency range with 3-dB drop of the modal significance from unity [91].

24
The side view of the ground-backed MTS is shown in Fig. 2.4(a), where N

= 3, Wp = 14 mm and Ws = 4 mm. Five values of h are studied and the results

are compared in Fig. 2.4(b). The resonant frequencies of the fundamental

broadside modes are 6.5 GHz (h = 10 mm), 7.3 GHz (h = 5 mm), 8.6 GHz (h =

2 mm), 9.4 GHz (h =1 mm), and 10.3 GHz (h = 0.2 mm), respectively. As can

be seen, as h becomes smaller, the modal resonant frequency increases

dramatically. In the meantime, the modal bandwidth decreases remarkably.

These effects are similar to those of a traditional microstrip patch antenna,

where the loading of ground affects the resonant frequency and limits the

bandwidth of the MTS. The modal bandwidth of a ground-backed MTS is

compared with that of a ground-backed patch of the same height and resonant

frequency. When h = 2 mm, the patch width is set to 14.6 mm, in order to

resonate at 8.6 GHz, and the modal bandwidth is 10.9 %. The modal

bandwidth is 27.8% for the ground-backed MTS with the same height and

resonant frequency. The bandwidth is approximately three times of that of the

conventional patch antenna. The enhanced bandwidth is due to the enlarged

occupying area by the currents, which in general leads to a lower quality factor.

Apparently, the MTS is a better candidate than conventional patches for low-

profile wideband antennas, providing that the occupied area is not a major

concern.

25
(a) (b) (c)
Fig. 2.5 Equivalent circuit model of ground-backed MTS. (a) Free-space
MTS, (b) half-space MTS with ground back, and (c) half- space
MTS with shunt inductor.

1600
Input Impedance (jΩ)

800

0
GND
-800
MTS+GND
MTS
-1600
0 5 10 15 20 25 30
Frequency (GHz)
(a)
180
Reflection Phase (degree)

120
60
0 GND
MTS+GND
-60
MTS
-120
-180
0 5 10 15 20 25 30
Frequency (GHz)
(b)

Fig. 2.6 Diagrams of input impedance and reflection coefficients. (a) Input
impedance and (b) reflection phase based on Configuration C of
infinite expansion, both for normal incidence.

26
When h = ∞, the ground-backed half-space MTS becomes a free-space

MTS. Qualitatively, an equivalent transmission-line model (TLM) shown in

Fig. 2.5 can be used to illustrate the difference. Note that the TLM is for an

infinite MTS with an infinite ground. Figs. 2.6(a) and 2.6(b) show the input

impedance and reflection phase at the MTS plane under a normal incidence,

respectively.

The TLM of a free-space MTS is shown in Fig. 2.5(a), where L is the self-

inductance provided by the patch, C is the mutual capacitance between

adjacent patches, ω is the angular frequency, and loss is excluded. The input

impedance Zin is the surface impedance looking toward the MTS and Zin = jωL

+ 1/ jωC for free-space MTS. When ω is small, ω2 is at a higher order smaller,

so Zin ≈ 1/ jωC. The free-space MTS is a purely capacitive surface that

supports TE waves on the surface [57]. As the frequency increases, jωL

increases and the MTS becomes less capacitive. When ω grows to the

resonant frequency ωr= 1/ (LC)1/2, the series resonance occurs and the

imaginary part of Zin approaches zero and the MTS is a low-impedance

surface like a PEC surface. After the resonance, Im(Zin) = (ωL–1/ωC) > 0, the

free-space MTS becomes inductive and starts to supports TM waves [57].

Since the L provided by the patch is very small, the series resonance happens

beyond 30 GHz as shown in Figs. 2.6(a) and 2.6(b), so the free-space MTS is

mainly capacitive in the investigated cases.

The TLM for the ground-backed MTS is shown in Fig. 2.5(b), where Z0 is

the wave impedance in free space, θ is the electrical length of the ground-MTS

spacing h, and ZG = jZ0tanθ is the equivalent impedance shunted on the free-

space MTS shown in Fig. 2.5(c). If the spacing h is electrically small, tanθ can

27
be approximated by its argument so ZG = jZ0θ. As a result, Im(ZG) = Z0θ > 0

and the ground plane is equivalent to an inductance shunted to the capacitive

MTS. Ignoring L of the free-space MTS, the input impedance for the ground-

backed MTS is Zin = jωZG /(1- ω2ZGC). At low frequencies, Zin ≈ jωZG =

jωZ0θ, so the ground-baked MTS is an inductive surface that supports TM

waves [57]. At the resonant frequency of ωr = 1/ (Z0θC)1/2, Zin = j∞,

corresponding to the well-known high-impedance surface (HIS). Therefore,

the resonant frequency here is called the HIS resonant frequency. As the

frequency further increases, Zin = jωZ0tanθ /(1–ω2Z0Ctanθ). Since ω2Z0Ctanθ–

1 >0, so Im(Zin) <0. The ground-backed MTS becomes a capacitive surface

that starts to support TE waves after the HIS resonance [57]. Here tanθ =

tan(βh) controls the first derivative ∂Zin/∂ω thus introducing the frequency

dispersion which qualitatively explains the dependence of the modal

bandwidth on h shown in Fig. 2.4(b).

Note that the HIS resonance of the infinite configuration is fundamentally

different from the CMs of the finite configuration. Intuitively, it is seen from

the reflection-phase and the input-impedance diagrams, that the CM resonance

occurs at 12.8 GHz whereas the HIS resonates at 10.4 GHz. Physically, the

HIS resonance is measured by the total scattering response due to a far-field

excitation of the entire system. The total scattering response may result from a

combination of all possible modes, including the leaky-wave modes, surface-

wave modes, and the local-resonant standing-wave modes, where the leaky-

wave and surface-wave modes are complex modes. On the contrary, the CMs

of the finite case are real modes, which are the uniquely defined intrinsic

modes of the system independent of any excitation. The HIS resonance is

28
excitation dependent and it shifts with varying incidence. Therefore, it is

challenging to identify the standing-wave modes from the excitation-

dependent total response. In this sense, the input-impedance and reflection-

phase diagrams fail to predict the standing-wave modes of the finite MTSs,

revealing the unique advantage of the combination of MTS and CMA.

However, the reflection-phase diagram (or the input-impedance diagram) is

technically the foundation of the standing-wave modes. It is known from Fig.

2.3 that the currents across the slots or the TE waves on the surface are

essential for the local-resonant TM modes. The magnetic currents are

supported when the input impedance of the ground-backed MTS is capacitive,

which happens after the HIS resonance for the investigated cases.

Consequently, the HIS resonant frequency obtained from normal incidence

can be used to estimate the initial modal sourcing frequency in the CMA.

8 40

7 Frequency
Frequency (GHz)

30
Bandwidth (%)

Bandwidth
6
20
5
10
4

3 0
0 2 4 6 8 10
Permittivity

(a) (b)
Fig. 2.7 Effects of dielectric loading on resonant frequency and modal
bandwidth. (a) Configuration of the ground-backed MTS with
dielectrics and (b) resonant frequency and bandwidth against εr.

29
1.0
6.15 6.10
6.30
0.8

Thickness (mm)
6.20
6.50 6.35 6.30
0.6 6.25

6.45
0.4
6.40

0.2
1 2 3 4 5 6
Permittivity

(a) (b)
Fig. 2.8 Effects of suspended dielectric layer on resonant frequency. (a)
Configuration of the MTS loaded with suspended dielectrics and
(b) resonant frequency against t and 𝜀c.

2.3 Dielectric loaded MTSs

Dielectric layers are usually used to support the metallic claddings. Fig.

2.7(a) shows the configuration of the ground-backed MTS with dielectric

loading, where h = 2 mm, Wp = 14 mm, and Ws = 4 mm. Lossless dielectric is

assumed. The resonant frequencies against the relative permittivity εr are

shown in Fig. 2.7(b). As can be seen, when the relative permittivity increases,

the resonant frequency and the modal bandwidth both decrease. Again, the

modal bandwidths are compared. The results suggest that air substrate

provides the optimum bandwidth while dielectrics with low 𝜀r is generally

preferred for wideband applications. Both with h = 2 mm, 𝜀r = 2 and a resonant

frequency of 6.5 GHz, the modal bandwidths are 12.4% for the ground-backed

MTS and 7.7% a ground-backed square patch (with a width of 14.5 mm). The

MTS exhibits a larger bandwidth and occupies a larger area than a patch does.

The effects of an infinitely-expanded thin dielectric slab suspended above

30
the ground-backed MTS are also investigated. The configuration is shown in

Fig. 2.8(a). The suspended dielectric layer may be used as an antenna radome

or to support other printed circuits. Therefore, it is important to know the

effects on the MTS modes. The spacing between the MTS and the slab is g.

The thickness and relative permittivity of the slab are t and εc, where the

subscript “c” stands for covering.

Fig. 2.8(b) shows the fitted curves of resonant frequencies of the

fundamental broadside mode with varying parameters, where t ∈ [0.2 mm, 1

mm] and εc ∈ [1, 6], εr = 2, and g = 1 mm. The resonant frequency is

calculated for different combinations of t and εc. The boundary line connects

all the points with the same resonant frequency. The colored area indicates the

combination of εc and t that gives rise to a resonant frequency in between two

boundaries. The frequency difference between two boundary lines is 0.05 GHz.

For example, when εc = 3 and t =0.3 mm, the resonant frequency is almost the

same as that when εc = 2 and t =0.2 mm. In general, the resonant frequency

experiences a relatively small variation from 6.1 GHz to 6.5 GHz

(approximately 6% of variation). The small variation can be explained from a

transmission-line point of view. The effective wave propagation constant of a

microstrip line, which affects the resonant frequency, is dominated by two

capacitances, one determined by the patch and the ground in the lower space

and the other one between the patch and a ground at infinity in the upper space.

The latter capacitance is much smaller than the former one in magnitude thus

has limited effect on the effective propagation constant and the resonant

frequency [23]. Similar results with g reduced to 0.2 mm suggest that the

suspended dielectric layer has limited effects on the resonant frequency of the

31
broadside mode. It is seen from Fig. 2.8(b) that when the suspended slab is as

thin as 0.2 mm, the resonant frequency is almost independent of εc in the range

studied. Similarly, when εc is low, the resonant frequency is insensitive to t

within the studied range. Consequently, the effects of a thin suspended

dielectric layer of relatively low dielectric constant on the broadside MTS

mode can be ignored in the CMA of the MTS for reduced complexity and

computational cost. On the other hand, it suggests that the thickness and

dielectric constant of the suspended slab could be freely chosen in a certain

range to facilitate the feeder design and impedance matching.

2.4 Summary

The concept of local-resonant MTS has been illustrated with the CMA.

Different from a solid patch that supports the conventional TM0N mode with

antiphase currents only, the MTSs support both the odd TM0N mode with in-

phase currents and the conventional TM0N mode, when N > 1. The CMA

results have revealed the concept of local resonance, so the overall electrical

size of the MTS at the fundamental resonant frequency can be much larger

than half a wavelength as demonstrated. In this sense, the dependence of

resonant frequency on the overall electrical size of the radiator has been

broken since the conventional higher-order mode now becomes the

fundamental mode of the MTS.

The difference between the free-space and half-space MTS has been

qualitatively revealed with a simplified transmission-line model. The effects of

varying unit cells, different MTS-ground spacing, and various dielectric

loading (including a suspended dielectric layer) have been quantitatively

32
studied with the CMA, showing that the fundamental resonant frequency of

the MTS can be effectively manipulated in a large dynamic range by

modifying the unit cells and that the environment must be taken into account

for accurately capturing the modal behaviors of an MTS.

In summary, with the same overall thickness, the MTS exhibits a much

larger modal bandwidth than a solid patch does for the fundamental broadside

mode, nearly tripled bandwidth for the chosen 3×3 configuration. In addition,

the modal resonant frequencies have been shown to be controllable by

optimizing the unit cells with the environment such as the inclusion of

dielectric loadings. The two advantages are particularly useful for low-profile

wideband antennas as they allow the modal quality factor being reduced by

expanding the antenna area without increasing the antenna height or being

concern about size-dependent resonance shift.

33
Chapter 3

Design of low-profile wideband local-


resonant MTS antennas

Taking the advantages of larger modal bandwidth and controllable

resonant frequency of ground-backed MTSs over a solid PEC patch, this

chapter demonstrates four designs of low-profile broadband metasurface

antennas (MAs) with different modal excitation schemes. All the MAs are

exemplified at the 5-GHz WiFi bands covering 4.9–5.9 GHz. The advantages

and disadvantages of different designs are summarized.

3.1 Design I: Slot-fed linearly polarized MTS antennas

3.1.1 Antenna design

A. Antenna geometry

The geometry of the proposed linearly-polarized MTS antenna (LPMA) is

shown in Fig. 3.1. The MTS is based on configuration G shown previously in

Fig. 2.1(a). The irregular shape of the MTS, as compared to the rest in Fig. 2.1

(a), helps demonstrate the advantage of CMA in analyzing MTS of arbitrary

shapes over conventional methods.

34
Fig. 3.1 Configuration of the slot-fed local-resonant MTS antenna.

The antenna consists of three metallic layers, the MTS (P-P'), the ground

plane (G-G'), and the microstrip line (F-F') from top to bottom. The

thicknesses of substrates are 3.454 mm and 0.508 mm for the P-G layer and G-

F layer, respectively. For both substrate layers, the Rogers RO4003C with a

relative permittivity of 3.55 and loss tangent of 0.0027 is used. The MTS is

composed of an array of square patches with a width of Wp and edge-to-edge

spacing of Ws. The center patch is cut through by a slot with a width of Wc.

The width of the square ground plane is Wg. A feeding slot with a length Lf

and width Wf is located at the center of the ground plane. On the bottom layer,

the microstrip line is terminated by a virtual shot-circuit radial stub with a

radius of Rf and opening angle θf for wideband impedance matching of the

microstrip-to-slotline transition. The optimized geometrical dimensions are

summarized in Table 3.1. These parameters are kept unchanged unless

otherwise specified.

35
Table 3.1
Summary of Antenna Geometry
of the Slot-Fed Local-Resonant MTS Antenna
Symbol Quantity Value

Wp width of sub-patches 8.5 mm

Ws width of slots on metasurface 0.8 mm

Wg width of ground plane and substrate 55 mm

Wf width of feeding slot on ground 1.5 mm

Wc width of center slot on metasurface 0.7 mm

Wm width of microstrip line 1.1 mm

Lf length of feeding slot on ground 25.5 mm

ha height of substrate between patch and ground 3.454 mm

hf height of substrate between microstrip and ground 0.508 mm

rf radius of microstrip stub 5.4 mm

θf opening angle of microstrip stub 120 degree

36
(a)

(b)
Fig. 3.2 Simulation results. (a) Simulated reflection coefficient and
boresight directivity. (b) Reflection coefficient in impedance
Smith Chart.

The S-parameters and the boresight directivity are shown in Fig. 3.2(a),

calculated with the time-domain full-wave analysis using commercial

simulator CST MWS 2016. The impedance bandwidth for |S11| = 15 dB is

25.4% (4.8–6.2 GHz), within which the directivity ranges from 7.7 to 10.3

dBi. Fig 3.2 (b) shows the reflection coefficient of the antenna in the

37
impedance Smith Chart in the frequency of 4.8–6.2 GHz. As can be seen, the

top slot improves the wideband impedance matching of the antenna by

balancing the input reactance, i.e., compensating for the input inductance at

the lower frequencies and the input capacitance simultaneously.

(a)

(b)
Fig. 3.3 Illustration of boundary setups for CMA. (a) Gorund-backed MTS
without feeding slot. (b) Ground-backed MTS with feeding slot.

38
B. CMA of MTS without Feeding Slot

A two-step CMA is performed to the MTS with and without the feeding slot

to reveal the operating mechanism of the antenna. The results are calculated

with the CMA tool in commercial simulation software CST 2016, where the

ground plane and dielectric layer are infinitely extended in x and y directions.

Fig. 3.3 shows the boundary setups for the CMA of the two different

configurations in CST. In Fig. 3.3(a), the MTS is backed by an infinite ground

with a dielectric layer filled in between. Open boundaries are set in all the

directions except for the –z direction, where a PEC boundary is assumed as an

infinite ground plane. With the feeding slot, the PEC boundary in the –z

direction is replaced by an open boundary with an added space of air from the

bottom. The added space is not important as long as it is greater than zero. For

the ground-backed MTS in Fig. 3.3(a), only the sub-patches of the MTS are

meshed over which the electrical currents are solved. In Fig. 3.3(b), the slot on

the ground is also meshed, over which the magnetic currents are solved and

expressed in terms of electric currents. Loss is excluded. In all the CMA

results, J1 represents the mode of the lowest resonant frequency in the band of

observation.

39
1.0
0.9

Modal Significance
0.8
0.7
0.6
0.5
0.4 Mode 1
0.3 Mode 2
0.2 Mode 3
0.1 Mode 4
0.0
4 5 6 7 8
Frequency (GHz)
(a) (b)

Fig. 3.4 Modal analysis of ground-backed MTS without feeding slot. (a)
Top view and (b) modal significance of the ground-backed
metasurface.

(a) (b)

(c) (d)
Fig. 3.5 Modal currents of the first four modes at 6 GHz. (a) J1, (b) J2, (c)
J3, and (d) J4.

(a) (b) (c) (d)

Fig. 3.6 Modal radiation patterns of the first four modes at 6 GHz. (a) J1,
(b) J2, (c) J3, and (d) J4

40
The ground-backed MTS without the feeding slot is shown in Fig. 3.4(a).

The modal significances of the first four modes are shown in Fig. 3.4(b). As

can be seen, J1 and J2 resonate at 5.9 GHz as designed. They exhibit the same

modal significances. The next two modes J3 and J4 have the same resonant

frequencies but different modal significances. The associated modal electric

currents at 6 GHz are plotted in Fig. 3.5, where the arrows indicate the

directions of the currents. As can be seen, the y-directed J1 and the x-directed

J2 are a pair of orthogonal modes both with in-phase currents and broadside

radiation patterns, as shown in Figs. 3.6(a) and 3.6(b). For both J3 and J4,

radiation nulls appear at the boresight due to the 180°out-of-phase currents.

Consequently, J1 and J2 are the desired modes for broadside radiation.

Effective excitation of either of the two modes leads to a linearly polarized

antenna and J1 is chosen for demonstration.

(a) Ju (b) Jv (c) Eu (d) Ev

(e) side view of Eu from the v-axis direction

Fig. 3.7 Modal currents and E-field distribution of the Q-TM03 mode at 6
GHz.

41
The transformation of J1 and J2 from the x-y to a u-v coordinate system

facilitates the understanding of the modal behaviors, where the u-v coordinate

is formed by a rotation of the x-y coordinate by 45 degrees counterclockwise,

as shown in Fig. 3.4(a). By doing so, we have Ju = 0.5 (J1 + J2), Jv = 0.5 (J1 –

J2), J1 = (Ju + Jv), and J2 = (Ju – Jv). The transformed Ju, Jv, Eu, and Ev are

plotted in Figs. 3.7(a)–(d), respectively. A side view of the E-field distribution

for the center 3×3 sub-patches is shown in Fig. 3.7(e). It is seen that Ju and Jv

are directed along the u and v directions, respectively. The currents are mostly

concentrated over the centered 3×3 patches of the MTS. From an E-field point

of view, a TM01 resonance is formed for each patch, giving rise to the o-TM03

resonance for the overall MTS if ignoring the corner patches due to weak

energy distribution. The o-TM03 mode is similar to that of the configurations C

in Fig. 2.2(b). Therefore, J1 or J2 is considered as a linear combination of the

two orthogonal o-TM03 modes. However, since J1 or J2 are not the standard o-

TM03 mode, the two modes are named as the quasi-TM03 (Q-TM03 in short)

modes for ease of reference in the following discussion. It is worth pointing

out that the four corner patches play a role of enhancing the antenna gain by

0.5 dBi and decreasing the resonant frequency by 0.1 GHz due to its larger

aperture size, therefore they are maintained in the subsequent content.

C. CMA of MTS with Feeding Structure

Magnetic current represented by a narrow slot is used to excite J1 for the

sake of low profile. The magnetic current should be oriented in line with the

H-field of J1 and located where the H-field is maximum for efficient energy

coupling.

42
(a) (b)

Fig. 3.8 Modal analysis of ground-backed MTS with top slot. (a)
Configuration and (b) modal significance of the metasurface with
top slot.

It is clear from Fig. 3.5(a) that the maximum H-field of J1 and the minima

of J3 and J4 both occur at the center sub-patch. A short slot representing the

magnetic current is integrated into the center patch for selectively exciting J1.

With the center slot, the modified geometry and modal significances are

shown in Figs. 3.8(a) and 3.8(b), respectively. As can be seen, J1 and J2

diverge due to the asymmetrical perturbation introduced by the slot. Note that

J2 is y-directed and it experiences a slight blueshift compared to Fig. 3.4(b).

43
(a) (b)

(c) (d) (e) (f)

Fig. 3.9 Modal analysis of metasurface with feeding slot. (a)


Configuration, (b) modal significance, and (c)-(f) modal currents
of the metasurface with top and bottom slots. Black arrows
indicate the current direction.

To excite the small slot, a long slot representing another magnetic current

parallel to the shorter one is etched on the ground plane, as shown in Fig.

3.9(a). It is important to evaluate the effects of the feeding slot on the CMs

due to its electrical size being large enough to bring in possible new modes.

The geometrical modeling and boundary setup for CMA are shown in Fig.

3.3(b), where the lossless additional substrate with relative permittivity of 3.55

for the microstrip line is included. The modal significances of the first four

modes are plotted in Fig. 3.9(b). Here the first four modes are named as J1f, J2f,

J3f, and J4f, respectively, where the footnote stands for feeding. The resonant

frequencies of J1f, J2f, J3f and J4f are 3.5GHz, 6.0 GHz, 5.85GHz, and 6.4 GHz,

44
respectively. J1f is the new slot mode since no mode is found below 5 GHz

previously in Fig. 3.8(b).

The electric currents of the four modes at 6 GHz are plotted in Figs. 3.9

(c)-(f). J1f is in phase over the MTS and y-directed. J2f is associated with the

original MTS Q-TM03 mode of J1 in Fig. 3.5(a). J1f and J2f are excited

simultaneously by the feeding slot since they are both in phase and y-directed.

J3f is the x-directed mode, which is not excited due to orthogonality between

the modal H-field of J3f (along y-axis) and that of the feeding slot (along x-

axis). J4f is the next higher-order mode similar to J4 in Fig. 3.5(d). This mode

is not excited due to the same reason. Therefore, only the desired J1f and J2f

can be effectively excited while the unwanted mode J3f and J4f are sufficiently

suppressed. As a result, the strong coupling between the slot and the MTS

only happens between the slot mode J1f and the MTS mode J2f. This agrees

with the variation of the modal significances from Fig. 3.8(b) to Fig. 3.9(b),

where only J2f experiences a significant enhancement in modal significance in

the band of 4–5 GHz as compared with the mode 2 in Fig. 3.8 (b). The other

modes are hardly affected.

45
(a)

(b) (c)

(d)

Fig. 3.10 Illustration of potential antenna bandwidth. (a) conceptual


illustration of the potential antenna bandwidth, (b) isolated slot
without MTS, (c) modal significances of the first two slot modes,
(d) E-field distribution across slot against increasing slot length.

46
Next we discuss the potential radiation bandwidth of the antenna. Since the

proposed antenna operates with the half-wave slot mode and the fundamental

mode of the MTS, the potential operation bandwidth is windowed by the

resonant frequencies of the two modes, as qualitatively illustrated in Fig.

3.10(a). The lower frequency edge is determined by the half-wave resonance

of the slot. The upper frequency edge of the operation band is determined by

fQ03, the resonant frequency of the Q-TM03 mode, to avoid higher-order-mode

operation. The frequency ratio of the two resonances can be tuned to modify

the bandwidth by changing the length of the slot and the MTS. As the

frequency ratio becomes smaller, the bandwidth reduces with a better in-band

impedance matching. As the frequency ratio increases, the bandwidth

increases but the in-band impedance matching becomes worse because the

operating modes may not be wideband enough to guarantee good impedance

matching between the resonances. The impedance matching becomes even

worse when the frequency ratio of the two modes keeps increasing and a dual-

band antenna may be expected.

The MTS is removed to find the resonant frequency of the slot mode, as

shown in Fig. 3.10(b). The modal significances and modal E-field across the

slot for the first two modes are presented in Figs. 3.10(c) and 3.10(d),

respectively. As can be seen, the in-phase half-wave mode resonates at 3.5

GHz. Since fQ03 is 6 GHz from Fig. 3.9(b), the potential bandwidth of the

antenna is 3.5–6 GHz. The antiphase full-wave slot mode resonant at 7 GHz is

not a concern since it cannot be excited in the BOI when the microstrip line

feeds the slot at the center where a null of E-field is present for the full-wave

mode.

47
1.0 1.0

Modal Significance

Modal Significance
0.8 0.8
0.6 0.6
Wp = 8 Ws= 0.2
0.4 0.4
Wp = 8.5 Ws= 0.6
0.2 0.2
Wp = 9 Ws= 0.9
0.0 0.0
4 5 6 7 8 4 5 6 7 8
Frequency (GHz) Frequency (GHz)

(a) Wp (without slot) (b) Ws (without slot)

1.0 1.0

Modal Significance
Modal Significance

0.8 0.8
0.6 0.6
Mode 1 Mode 2 Mode 1 Mode 2
0.4 Wp= 8.0 Wp= 8.0
0.4 Ws= 0.2 Ws= 0.2
0.2 Wp= 8.5 Wp= 8.5 0.2 Ws= 0.6 Ws= 0.6
Wp= 9.0 Wp= 9.0 Ws= 0.9 Ws= 0.9
0.0 0.0
3 4 5 6 7 8 3 4 5 6 7 8
Frequency (GHz) Frequency (GHz)

(c) Wp (with slot) (d) Ws (with slot)

1.0 1.0
Modal Significance
Modal Significance

0.8 0.8
0.6 0.6
Mode 1 Mode 2
0.4 Lf= 20 Lf= 20
0.4
Mode 1 Mode 2
0.2 Lf= 24 Lf= 24 0.2 Wf= 1.0 Wf= 1.0
Lf= 28 Lf= 28 Wf= 1.8 Wf= 1.8
0.0 0.0
3 4 5 6 7 8 3 4 5 6 7 8
Frequency (GHz) Frequency (GHz)

(e) Lf (with slot) (f) Wf (with slot)

Fig. 3.11 Parametric study based on CMA. Variation of modal significances


against different dimensions: (a)-(b) without bottom slot (c)-(f)
with bottom slot. (unit: mm)

D. Parametric study and sensitivity analysis

Conventional parametric study is performed for understanding of the

operating principle and for the optimum design, where an excitation source

48
has to be predefined. Differently, here the parametric study is carried out by

the source-free CMA first to identify the most critical parameters that

qualitatively affect the behaviors of the MTS. Based on the sensitivity of

parameters, time-domain analysis is performed for optimized performance.

This avoids the mutual dependence of parameters in conventional optimization.

The effects of Wp and Ws on the resonant frequencies and modal

significances of the MTS without and with the feeding slot are shown in Figs.

3.11(a)–(b) and (c)–(d), respectively. As can be seen in both cases, the

resonant frequency of the MTS mode decreases with increasing Wp and

reducing Ws, the same trend as that is concluded in Chapter 2. This

demonstrates again the frequency tunability of the MTS mode. Since each

patch contributes a local TM01 resonance, the shift of resonant frequency

follows that of a conventional rectangular microstrip patch antenna [23]. The

increasing Wp leads to a larger length along the resonant direction while the

decreasing Ws leads to a larger mutual-capacitance between the patches, both

leading to a lower resonant frequency of the patch.

The effects of Lf and Wf on the modal significance with the feeding slot are

also studied. As shown in Figs. 3.11(e) and (f), Lf determines the resonant

frequency of the slot mode. The resonant frequency of the MTS mode is

hardly affected. The in-band modal significance also varies against Lf. The

resonant frequency and in-band modal significance are both much less

sensitive to varying Wf.

With decreasing Wp, increasing Ws, and increasing Lf, the modal

significances of the two modes both decreases at frequencies between the

resonant frequencies of the slot mode and the MTS modes. Therefore, the

49
potential gain bandwidth increases but wideband impedance matching may

become more difficult because the modal significance indicates the potential

radiating capability of the mode.

The CMA-based parametric study helps identify the most critical

parameters for fast optimization of antenna performance. The four parameters

studied can be sequenced in a descending order with respect to their

contribution to manipulating the modal significances, i.e., Lf, Ws, Wf and Wp,

which indicates the weight of each parameter in the subsequent time-domain

optimization. In other words, Lf should be optimized first whereas Wp the last.

The parametric study performed with time-domain analysis agrees well

with the CMA results. As shown in Fig. 3.12, the operating band shifts upward

with growing Ws and decreasing Lf and Wp, whereas the center frequency

remains with varying Wf. In addition, Lf and Ws mainly contribute to

impedance matching. From the radiation perspective, the directivity is mainly

determined by the metasurface design while the feeding slot size does not

affect it much.

To summarize, the upper-end frequency and directivity are dominated by

the MTS design, while the lower-end frequency and impedance matching are

mainly determined by the size of the feeding slot.

50
0 12

Directivity (dBi)
-10
|S11| (dB) 9
-20
Wp = 8.0 Wp = 8.0
6
-30 Wp = 8.5 Wp = 8.5
Wp = 9.0 Wp = 9.0
-40 3
4 5 6 7 8 4 5 6 7 8
Frequency (GHz) Frequency (GHz)
(a)
0 12

Directivity (dBi)
-10
|S11| (dB)

9
-20
Ws = 0.2 Ws = 0.2
6
-30 Ws = 0.6 Ws = 0.6
Ws = 0.9 Ws = 0.9
-40 3
4 5 6 7 8 4 5 6 7 8
Frequency (GHz) Frequency (GHz)
(b)
0 12
Directivity (dBi)

-10
9
|S11| (dB)

-20
Lf = 20
6
-30 Lf = 20 Lf = 24
Lf = 24 Lf = 28 Lf = 28
-40 3
4 5 6 7 8 4 5 6 7 8
Frequency (GHz) Frequency (GHz)
(c)
12
0
Directivity (dBi)

-10 9
|S11| (dB)

-20
Wf = 1.0 Wf = 1.0
6
-30 Wf = 1.2 Wf = 1.2
Wf = 1.8 Wf = 1.8
-40 3
4 5 6 7 8 4 5 6 7 8
Frequency (GHz) Frequency (GHz)
(d)

Fig. 3.12 Parametric study based on time-domain analysis. Effects of (a)Wp,


(b) Ws, (c) Lf, and (d) Wf on |S11| and directivity. (unit: mm)

51
3.1.2 Experimental results

Based on the proposed MTS antenna element, a 2×2 antenna array is

investigated experimentally. Fig. 3.13 (a) shows the configuration of the array.

The optimized distance between each element is D = 42 mm and the ground

width is 97 mm (1.78 λ0 at 5.5 GHz). Shown in Fig. 3.13(b) is the prototype

fabricated to experimentally verify the validity of the design concept. The

simulated and measured S-parameters and boresight gain are compared in Figs.

3.14(a) and (b), both with good agreement. The slight discrepancy may be

attributed to the non-ideal dielectric thickness after multilayer-PCB processing.

The measured impedance bandwidth for |S11|= –15 dB is 24.6 % (4.85–6.21

GHz), within which the boresight gain varies from 12.2 dBi to 14.1 dBi.

Compared with a single element with a fractional bandwidth of 25.4 %, the

broadband property is maintained.

(a) (b)
(a) (b)
Fig. 13 Configuration of the 2×2antenna array and the fabricated prototype.
Fig. 3.13 Configuration of the 2×2 MTS antenna array. (a) Bottom view,
and (b) top view of the fabricated prototype.

52
0
-10

|S11| (dB)
-20
-30
-40 Simulation
Measurement
-50
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(a)

15 15

14 14

13
Gain (dBi)

13

12 12

11 11

10 Simulation
10

9 Measurement 9

8 8

4.5 5.0 5.5 6.0 6.5


Frequency (GHz)
(b)

Fig. 3.14 Simulated and measured results of the MTS antenna array. (a) S-
parameters and (b) gain.

53
x-z plane y-z plane
Simu_Co Simu_Co
5.0 GHz  = 00° Simu_X 5.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co Simu_Co
5.5 GHz
 = 00° Simu_X 5.5 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10

0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co Simu_Co
6.0 GHz  = 00° Simu_X 6.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Fig. 3.15 Simulated and measured radiation patterns of the MTS antenna
array (normalized).

Fig. 3.15 shows the excellent agreement between the simulated and

measured radiation patterns at 5.0 GHz, 5.5 GHz, and 6.0 GHz. Due to the

symmetrical feeding design and high mode purity revealed in CMA, the

measured cross-polarization at boresight is below -30 dB across the operating

band.

54
3.2. Design II: Slot-fed dual-linearly polarized MTS antenna

Wideband dual-polarized antennas are demanded in wireless

communications for enhanced communication capacity by polarization

diversity. The design methodology for the linearly-polarized MTS antenna

(LPMA) is extended to the design of a dual-linearly polarized MTS antenna

(DPMA).

Fig. 3.16 Configuration of the dual-linearly polarized MTS antenna

3.2.1 Antenna design

The geometry of the DPMA is shown in Fig. 3.16. The three metallic

layers are the MTS (M-M’), the ground plane (G-G’), and the feeding

microstrip lines (F-F’). The substrate is Rogers RT5880 (εr = 2.2 and tanσ =

0.0009). The thicknesses of the substrates are 3.14 mm and 0.508 mm for the

MTS and the microstrip line, respectively. The MTS is based on configuration

E, where the edges of the square patch and the square ground are in parallel, in

order to facilitate the alignment of the circuit boards in experiments. The patch

55
width is Wp =12 mm and the slot width is Ws = 1.4 mm to place the broadside

MTS mode at 5.9 GHz. For each polarization, a two-element H-plane slot

array is proposed to excite the o-TM03 MTS mode through the microstrip

power divider on the bottom layer. Each slot contains two parts. One part with

a length of Lf = 14 mm is centered underneath the square patch. The second

part, as an extension for introducing the slot mode, has a length of Le = 7 mm.

The total length of each slot is Lf + Le = 21 mm and the slot width is Wf =1.5

mm. The microstrip line crosses the center of the first part of the slot and

forms a microstrip-to-slotline transition [13]. The extension of the microstrip

stub has an optimized length of 12 mm for impedance matching. The width of

the microstrip line is Wm = 1.4 mm, corresponding to a characteristic

impedance of 50 Ω. Two sets of T-junction power dividers are designed to

feed the horizontal and vertical pairs of slots for vertical and horizontal

polarizations, separately. The width of the square ground is Wg = 66 mm,

optimized for high gain.

3.2.2 Experimental results

The simulated and measured S-parameters are compared in Fig. 3.17(a).

Good agreement is observed. The measured impedance bandwidth with return

loss greater than 10 dB is 4.8–6.2 GHz (or 25%) for both polarizations, over

which the port isolation is higher than 35 dB. Fig. 3.17(b) shows the simulated

and measured boresight gains. The achieved boresight gain is 8.7–10.3 dBi for

both polarizations over the 10-dB impedance bandwidth. The discrepancy

between the simulated and measured gains is less than 0.5 dB. The simulated

and measured radiation patterns under the separated excitations of Port 1 and

Port 2 are compared in Figs. 3.18 and 3.19, respectively. Good agreement is

56
achieved. While the measured cross-polarization level is better than -28 dB

across the band, the simulated cross-polarization level in the operating band is

below -60 dB at the broadside. The discrepency is attributed to the imperfect

testing environment where the testing cable, antenna supporters, and the

polarization purity of the horn antenna (source) affect the measurement of the

cross-polarization level.

Consequently, fed by two slot arrays, each polarization of the antenna

operates at two modes simultaneously, including the o-TM03 mode and the

slot-array mode. With an overall size of λ0×λ0×0.06λ0 (λ0 is the wavelength in

free space at the lowest operating frequency), the proposed antenna achieves

an impedance bandwidth of 25% with return loss better than 10 dB and a

boresight gain of 8–10 dBi. The in-band port isolation and cross-polarization

level are better than 35 dB and 28 dB, respectively.

57
0
Meas. Simu.

S-Parameters (dB)
-10 |S11| |S11|
-20 |S22| |S22|
-30 |S21| |S21|
-40
-50
-60
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(a)

11
P1_Simu
10 P1_Meas
Gain (dBi)

P2_Simu
9
P2_Meas
8
7
6
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(b)

Fig. 3.17 Simulated and measured results of the dual-linearly polarized


MTS antenna. (a) S-parameters and (b) gains.

58
Simu_Co Simu_Co
5.0 GHz  = 00° Simu_X 5.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co Simu_Co
5.5 GHz  = 00° Simu_X 5.5 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co Simu_Co
6.0 GHz  = 00° Simu_X 6.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Fig. 3.18 Normalized radiation patterns of Port 1. The left and right columns
are for the E- and H-planes, respectively.

59
Simu_Co Simu_Co
5.0 GHz  = 00° Simu_X 5.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 150°
210 150
150° 0 150°
210 150
150°
(dB) 180
180° (dB) 180
180°

Simu_Co Simu_Co
5.5 GHz  = 00° Simu_X 5.5 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 150°
210 150
150° 0 (dB) 150°
210 150
150°
(dB) 180
180° 180
180°

Simu_Co Simu_Co
6.0 GHz  = 00° Simu_X 6.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 150°
210 150
150° 0 (dB) 150°
210 150
150°
(dB) 180
180° 180
180°

Fig. 3.19 Normalized radiation patterns of Port 2. The left and right columns
are for the E- and H-planes, respectively.

60
3.3 Design III: Dipole-fed MTS antenna with high FBR

The slot-coupling fed MTS antennas have achieved wideband

performances with a low profile. The main drawback is the relatively large

back radiation from the resonant slot. The achieved front-to-back ratios (FBR)

are less than 13 dB, which may cause EMI problems to the radio-frequency

(RF) frontends and digital circuits at the back of the antenna. A top-feed MTS

antenna is proposed for broadband operation, high FBR, and low profile

simultaneously.

Fig. 3.20 Configuration of the dipole-fed MTS antenna and boundary setup
for CMA.

3.3.1 Antenna design

The configuration of the top-fed MTS antenna is shown in Fig. 3.20. Two

dielectric layers are used as the Substrates I and II with the thicknesses of d

and h, respectively, between which an air gap of thickness g is sandwiched in

between. The substrate Rogers RO4003C with a relative permittivity of 3.55

and loss tangent of 0.0027 is used. A planar dipole with a length of Ld and

width of Wf is printed on top of Substrate I. The MTS is printed on top of the

61
J4 J6
(J1, J2) J3 J5 (J7, J8)
1.0 J1
J2
Modal Significance 0.8
J3
0.6 J4
J5
0.4
J6
0.2 J7
0.0 J8
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(a)

(b)

Fig. 3.21 Modal analysis of the MTS without dipole. (a) Modal
significances and (b) associated modal electric currents of the
MTS without the dipole and Substrate I.

grounded Substrate II. The MTS is in the same 3×3 configuration as the

DPMA with the center patch removed for placing the vertical twin-wires to

feed the dipole. The width of the patch and the spacing between adjacent

patches are Wp and Ws, respectively. The length and width of the dipole are Ld

and Wf, respectively. The broadside MTS mode is designed to resonate at 6

GHz by tuning Wp and Ws. The dipole mode is designed to resonate at 4.9

GHz by tuning Ld and Wf. The dipole mode and the MTS modes are

simultaneously excited to cover the operating band, 4.9–5.9 GHz, for

62
(a)

(b)
Fig. 3.22 Modal analysis of the MTS with dipole. (a) Modal significances
and (b) associated modal electric currents of the MTS without the
dipole and Substrate I.

wideband performances.

The calculated modal significances of the first eight modes of the MTS

without and with the dipole are shown in Figs. 3.21(a) and 3.22 (a). The modal

currents and radiation patterns at 6 GHz are plotted in Figs. 3.21(b) and 3.22

(b), respectively. For ease of comparison, the modal indices are manually

tracked in an ascending order of frequencies at 0.707 of the modal

significances. The modes are assigned with the identical index if their modal

currents exhibit similar symmetries. As can be seen from Figs. 3.21(a) and

63
3.22(b), the modes J1 and J2 are the broadside MTS modes both resonating at

6 GHz. The modes (J3, J4, J5, J6) are the unwanted modes due to radiation

nulls at the broadside. Even higher-order modes are not considered due to low

MS in the band of interest. It is seen from Fig. 3.22(a) that the dipole

introduces an additional mode J0 at 4.8 GHz. The modal index “zero” is

assigned to the dipole mode. The desired y-polarized MTS mode J1' shifts

from 6 GHz to 6.2 GHz. Slight frequency shift is also observed for the rest of

the modes (J2', J3', J4', J5', J6') after introducing the dipole.

0.20
Modal weighting coefficient

J0
0.15 J1'

0.10

0.05
J7'
0.00
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 3.23 Magnitude of the modal weighting coefficients for the dipole-fed
MTS antenna.

The lumped port at the center of the dipole only effectively excites J0 and

J1' in the band of interest (BOI) since the two modes both exhibit the in-phase

currents over the dipole, as shown in Fig. 3.22(b). To validate the dual-mode

operation of the proposed MA, the |MWC| or |αn| of the first ten modes are

calculated using the commercial software Altair FEKO 2017.1.2 [120]. Fig.

3.23 shows the calculated |αn| (n = 1, 2…10). As can be seen, the dipole mode

J0 and the MTS mode J1' dominate the radiation at lower and higher

frequencies, respectively.

64
(a)

(b)

Fig. 3.24 Simulated results of the dipole-fed MTS antenna. (a) |S11|, (b)
boresight gain.

3.3.2 Simulation results

Fed by lumped ports, the simulated |S11| and boresight gain of the dipole-

feed MTS antenna are shown in Figs. 3.24(a) and 3.24(b), respectively. The

achieved operating bandwidth is 29% with an in-band gain of 7.5–8.5 dBi.

The simulated radiation patterns are shown in Fig. 3.25. The achieved FBR is

higher than 20 dB throughout the operating band. In comparison with the slot-

feed MTS antennas, approximately 10-dB higher FBR has been achieved.

Also, in comparison with traditional ground-backed dipole antennas with a

height of a quarter wavelength, the proposed antenna with an overall

dimension of 0.47 λd× 0.47 λd× 0.081 λd has achieved a height reduction of 67%

with the wideband performance maintained.

65
4.9 GHz
 =180
0° 5.5 GHz
0 30°
210 30°
150 5.8 GHz
-10 60°
120
60°
240
-20
-30 90°
270 90°
90

-20
120°
300 120°
60
-10
E-plane
0 (dB) 150°
330 30
150°
0
180°

(a) y-z plane


4.9 GHz
 =180
0° 5.5 GHz
0 30°
210 30°
150 5.8 GHz
-10 60°
120
60°
240
-20
-30 90°
270 90°
90

-20
120°
300 120°
60
-10
H-plane
0 (dB) 150°
330 30
150°
0
180°

(b) x-z plane

Fig. 3.25 Simulated radiation patterns of the dipole-fed MTS antenna


(normalized). (a) y-z plane and (b) x-z plane

66
3.4 Design IV: Probe-fed MTS antenna with optimal feeding

placement

Both the dipole-fed and slot-fed MTS antennas have achieved wideband

performances with low profile. A probe-fed MTS antenna is proposed.

Different from the previous designs where the MTS provides only one

broadside mode, a secondary MTS mode is introduced by the proposed slot-

embedded MTS for simple structure, low profile, and wideband performances

simultaneously. The CMA results are used to determine the optimum feeding

placement of the cable to avoid possible modal distortion which may

deteriorate the radiation pattern and impedance matching of the antenna.

Fig. 3.26 Geometry of the probe-fed MTS antenna.

67
3.4.1 Antenna design

The geometry of the proposed MTS antenna is shown in Fig. 3.26. The

MTS is composed of three rectangular patches each with a width of Wp and

length of 3Wp+2Ws, where Ws is the spacing between the adjacent patches, as

shown in Fig. 3.26(a). Since only y-polarization is needed, the patches in the

x-axis direction are connected for the convenience of slot integration. The

MTS is separated from the ground plane at a height of h = 3.048 mm, where

the dielectric filled in between has a relative permittivity 𝜀r = 3.55 and loss

tangent of 0.0027, as shown in the side view in Fig. 3.26(b). The feeding cable

is placed at the top of the MTS, as shown in Fig. 3.26(c). The cross-section is

shown in Fig. 3.26(d), where the spacing between the center of the vertical

cable and the center of the MTS is d. The coaxial cable penetrates the

substrate with its outer conductor soldered to the center patch of the MTS. The

inner probe is electrically connected across the center slot. The optimized

dimensions are Wp = 10 mm, Ws = 1.5 mm, Lf = 20 mm, Wf = 0.8 mm, and d =

13.8 mm, respectively.

In the CMA, the cable is not included. Instead, a thin conducting planar

strip of 1 mm in width is used to connect the opposite sides of the slot at the

center. The in-phase electric current on the short strip is a qualitative indicator

of the modes excited when a voltage source is applied across the short strip.

This avoids the time-consuming calculation of the Prad or the |αn| for the

individual modes for identifying the excited modes.

68
1.0 J1
J2

Modal Significance
0.8 J3
J4
0.6
J5
0.4 J6
J7
0.2 J8
J9
0.0 J10
4.0 4.5 5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

Fig. 3.27 Modal significances of the probe-fed MTS antenna.

Fig. 3.28 Modal currents and E-field of the probe-fed MTS antenna. (a)–(c)
modal electric currents and (d)–(f) modal E-fields of J9, J3 and J7
at resonant frequencies.

69
With the top slot and the feeding strip integrated, the modal significances

of the first ten modes are shown in Fig. 3.27. By checking the modal currents,

the modes with in-phase currents on the feeding strip are J3, J7, and J9, whose

modal electric currents and associated modal E-field distribution are shown in

Fig. 3.28. For the rest of modes, the currents are either toward the x-direction,

generating a curl, or vanishing over the short strip, which are all nonphysical

under a real probe excitation. As a result, these modes are not excited by the

probe. As can be seen from Figs 3.27 and 3.28, J3 is the o-TM03 mode

resonating at 5.3 GHz. The J7 is the TM03 mode resonating at 6.5 GHz. J9 is

the TM22 mode resonating at 4.9 GHz due to the embedded slot. All the three

modes exhibit broadside radiation of y-polarization.

To implement the excitation scheme, a coaxial cable through the substrate

from the bottom is needed. It plays the same role as a metallic via that short-

circuits the patch to the ground. The original modes (thus the antenna radiation

patterns) are distorted when the cable is placed at a point where a potential

difference is present between the patch and the ground. In other words, there

are currents flowing over the vias. To avoid the modal distortion, it is

proposed to locate the cable at the null of the normal E-field of the desired

modes. Based on the E-field distribution of J3, J7, and J9 shown in Figs. 3.28

(d)–(f), the optimal position is the margin of the center patch where the normal

E-field vanishes. The optimized distance is d = 13.8 mm.

3.4.2 Experimental results

The structure of the proposed antenna with the integrated coaxial cable is

shown in Fig. 3.29. The simulated and measured |S11| and boresight gain are

70
Fig. 3.29 Fabricated prototype of the probe-fed MTS antenna.

0 12 12

Simu
-10 Meas Gain (dBi)
|S11| (dB)

9 9

-20
6 6

-30 TM22 Simu


TM03
o-TM03 Meas
-40 3 3

4.0 4.5 5.0 5.5 6.0 6.5 7.0 4.5 5.0 5.5 6.0 6.5
Frequency (GHz) Frequency (GHz)

Fig. 3.30 Simulated and measured |S11| and boresight gain of the probe-fed
MTS antenna.

shown in Fig. 3.30. As can be seen, the experimental results agree well with

the simulation results. The measured 10-dB impedance bandwidth is 4.5–6.5

GHz (36%) over which 6–10 dBi measured boresight gain is achieved. The

measured cross-polarization level at the broadside direction is below -21 dB.

The discrepancy shown in the reflection coefficients is mainly caused by the

fabrication errors since the antenna prototype is completely hand made.

As can be seen by comparing Figs. 3.27 and 3.30, the TM22 MTS mode is

dominant near 5 GHz. From 5 to 6 GHz, the o-TM03 mode takes over. From

approximately 6 to 6.5 GHz, the operating mode is a combination of the o-

TM03 and the TM03 modes. Due to the antiphase currents of the TM03

71
E-plane H-plane

Simu_Co Simu_Co
5.0 GHz  = 00° Simu_X 5.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co
Simu_Co
5.5 GHz
5.5 GHz
 = 00° Simu_X  = 00° Simu_X
30° 0 30°
330 30°
30 Meas_Co
0 30°
330 30 Meas_Co
Meas_X Meas_X
-10 -10 60°
60
60°
300 60°
60 60°
300

-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10

0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Simu_Co Simu_Co
6.0 GHz  = 00° Simu_X 6.0 GHz  = 00° Simu_X
0 30°
330 30°
30 Meas_Co 0 30°
330 30°
30 Meas_Co
Meas_X Meas_X
-10 60°
60 -10 60°
60
60°
300 60°
300
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
240 120°
120 120°
240 120°
120
-10 -10
0 (dB) 150°
210 150
150° 0 (dB) 150°
210 150
150°
180
180° 180
180°

Fig. 3.31 Simulated and measured radiation patterns of the probe-fed MTS
antenna (normalized).

mode, the boresight gain decreases near 6.5 GHz, where the TM03 mode

dominates. The simulated and measured radiation patterns at 5.5 GHz are

compared in Fig. 3.31. Excellent agreement has been achieved.

72
Table 3.2
Summary of Antenna Performance for the Local Resonant MTS Antennas
Size of the MTS Excitation
Designs Bandwidth FBR
Antenna Schemes
38×38×3.96 mm3 Slot
I 31 % <13 dB
0.61×0.61×0.064 λL3 Coupling
39×39×3.65 mm3 Slot-array
II 25 % <13 dB
0.62×0.62×0.058 λL3 Coupling
30×30×5.06 mm3 Suspended
III 21 % >20 dB
0.47×0.47×0.081 λL3 Dipole
33× 33×5.15 mm3 Coaxial
IV 29 % >20 dB
0.49×0.49×0.047 λL3 Cable

To summarize, a probe-feed low-profile broadband MTS antenna has been

proposed, analyzed and experimentally validated. Simple structure, wideband

operation, high FBR, and reduced cost have been achieved simultaneously.

Three MTS modes, the o-TM03, TM03, and TM22 MTS modes are generated

and combined for wideband operation. The optimum feeding placement has

been identified for effective excitation of the desired modes while avoiding

possible modal distortion.

3.5 Summary

Four designs of wideband MTS antennas have been proposed and

experimentally validated at the 5-GHz WiFi bands under the constraint of low

profile, i.e., antenna height smaller than one tenth of the wavelength at the

lowest operating frequency. The key features have been summarized in Table

3.2. In comparison with a conventional patch antenna with low profile and a

typical bandwidth of 10% [23], the proposed MTS antennas have achieved up

to tripled bandwidth with low profile.

The modal current and field distributions have been studied in detail for

73
both the desired and unwanted modes, providing insightful guidance for

selectively exciting the desired modes. While conventional characterization

methods of MTSs are strictly restricted to infinite-sized periodic structures and

fail to count the effects of local loadings, such as the slot-loaded ground, the

proposed combination of CMA and MTS allows a finite-sized non-periodic

MTS being fully characterized and locally modified for optimized

performance with the effects of feeding structure inherently included.

74
Chapter 4

Modal manipulation of local-resonant MTS


antennas

Low-profile wideband MTS antennas have been realized in Chapter 3,

where the attention has mainly been paid to the fundamental broadside mode

of the MTS. Multiple higher-order modes (HOMs) also exhibit large modal

significances near the fundamental modes. However, these HOMs have been

sufficiently suppressed for an antenna element by the feeding designs, as

demonstrated in Chapter 3. However, for a multi-antenna system with the

feeding scheme predefined for each element, the HOMs of one antenna may

be effectively excited by the wave radiated from another antenna, especially

when all the antenna elements are closely spaced. The anomalous radiation

from the excited HOMs can cause unexpected EMI problems that degrade the

performances of the multi-antenna system, such as severe radiation-pattern

distortion, poor port isolation, and significant drop of antenna efficiency. To

mitigate the EMI problems, methods of suppressing the unwanted HOMs are

needed. One of the most powerful aspects of MTSs is that their modes can be

artificially engineered by modifying the unit cell. A simple example of this

concept has been demonstrated in Chapter 2, where the modal resonance shifts

when the unit cell is changed.

75
In this chapter, the effects of the HOMs on the performances of a multi-

antenna system are studied. It is found that the HOMs are the main cause of

the severe pattern distortion. A method of suppressing the HOMs by unit-cell

loading is proposed for mitigating the pattern distortion, where the optimum

loading positions are determined with the modal E-field pattern. With the

suppression of the HOMs, the pattern distortion is significantly alleviated. The

concept is experimentally validated for potential surface-integrated compact

multiport antenna applications.

4.1 Mutual coupling in multi-antenna systems

It is a challenge to achieve stable broadside radiation patterns against

frequency for distinct antennas in broadband multiport antenna systems. The

antennas may not necessarily work at the same frequency bands [68], [121]. In

addition to poor port isolation, the strong mutual coupling between the closely

spaced antennas also contributes to the distortion of radiation patterns [2]–[8].

Theoretical analysis of this phenomenon by treating antennas as loaded

scatterers using a network description are reported [122]–[126], where it is

concluded that the radiation pattern of one excited antenna is contributed by its

direct radiation and the multipath scattering from its ambient antennas (loaded

scatterers).

A conceptual illustration of the pattern distortion caused by mutual

coupling is shown in Fig. 4.1. Suppose JA is the primary currents excited on

Antenna A with the radiated E-field of EA. As the incident field, EA induces

the currents JsB on Antenna B. Part of the JsB enters into the Port B and

76
Fig. 4.1 Conceptual illustration of mutual coupling in multi-antenna
systems.

contributes to the transmission coefficients SBA from Port A to Port B. The

other part of JsB adds up with the portion reflected from Port B due to

impedance mismatching at Port B and re-radiates the scattering field EsB, part

of which further induces JsA on Antenna A to produces EsA, and so on so forth.

The scattering fields EsA and EsB add up in space with the primary radiation

EA, which altogether result in the total radiated fields different from the

desired EA in general, hence the pattern distortion.

The mutual coupling becomes much more complicated as the number of

antennas increases. The analysis above only qualitatively helps with the

general understanding of the pattern distortion. The port network parameters,

such as the scattering matrix, only count the portion of energy entering into the

ports thus limited in characterizing the scattering waves radiated from the

reflected currents.

Under the framework of the TCM, the total current Jtot on an antenna,

including the induced current Js due to the mutual coupling, is a combination

of the modal currents. Each excited mode of the antenna serves as the

77
secondary source that excites the modes of the other antennas, leading to the

mutual excitation of all the current modes in the entire antenna system. For

MTS antennas with each of multiple modes (though suppressed in an element

level), the mutual excitation of the unmatched HOMs causes the unwanted

multi-path structural scattering between the antennas and leads to severer

pattern distortions, such as main beam tilt/squinting, main beam split, and

unexpected sidelobes [122]–[126]. To alleviate the pattern distortion, it is

desired to suppress the unwanted HOMs of each antenna without affecting the

functional broadside modes.

Fig. 4.2 Configuration of the four-port MTS antenna system.

A realistic four-port antenna system is investigated to illustrate the pattern

distortion caused by HOMs, based on the dipole-fed MTS antenna element

proposed in Section 3.3. The emphasis is on the radiation patterns of each

antenna. The configuration of the antenna system is shown in Fig. 4.2. All the

78
dimensions of the element are kept the same as those previously described.

The four-port MA system is formed by the four antennas in the 2×2

configuration with a rotational symmetry to enable the polarization diversity

and to mimic a complex electromagnetic environment while lowering the

analysis complexity [127]. Due to the rotational symmetry, the analysis of any

one antenna captures the characteristics of the entire system. Therefore, only

Port 1 is excited and the other three are terminated with matched loads for all

source-dependent analysis in the subsequent content. For all the MTS antennas,

the lumped ports are defined at the center of the dipoles with 50-Ω source

impedance, where the feeding gap is 0.2 mm in width. The ground size is Wg =

40 mm or 0.65 λ0 (λ0 is the operating wavelength in free space at 4.9 GHz) for

each element, optimized for impedance matching and consistent gain. The

center-to-center spacing between two horizontal and vertical MTS antennas

are also Wg and the edge-to-edge spacing is 10.5 mm or 0.17 λ0.

0 9
Boresight Directivity (dBi)

-10
S-Parameters (dB)

-20 8
-30
-40 7
-50 |S11| 2×2 |S21| 2×2
-60 |S31| 2×2 |S41| 2×2 6 isolated
-70 Antenna 1 of the system
isolated |S11|
-80 5
4.5 5.0 5.5 6.0 6.5 4.5 5.0 5.5 6.0 6.5
Frequency (GHz) Frequency (GHz)

(a) (b)
Fig. 4.3 Simulation results for the four-port MTS antenna system.
Comparison of S-parameters and boresight directivities of (a) the
isolated metasurface antenna and (b) the same antenna in the 2×2
antenna system.

The simulated S-parameters and boresight directivities of the four-port

antenna system are compared with those of an isolated MTS antenna in Fig.

79
4.3. As can be seen, the impedance matching is hardly affected by the

configuration whereas the boresight directivities differ greatly from each other.

Specifically, the directivity is consistent for the isolated antenna but

fluctuating for the same antenna in the system. The drop of the directivity

means that all the energy emitted by Antenna 1, which is originally due to the

two functional modes only, is now partly carried by other modes so the energy

goes to directions off the boresight due to antiphase currents. The increase of

the directivity at 5.7 GHz and the corresponding poorest port isolations

observed at the same frequency means that the currents are spread over all the

MTSs and result in a larger aperture size.

Fig. 4.4 Distorted radiation patterns of the system. Results are plotted at
4.9 GHz, 5.5 GHz and 5.8 GHz when Port 1 is excited. Pattern
distortion is observed. The radiation pattern at 5.5 GHz of an
isolated antenna is plotted for comparison.

The distorted radiation patterns are shown in Fig. 4.4 and compared with

that of the isolated MTS antenna (green dashed lines) at the center frequency

5.5 GHz. Symmetric and single broadside beam are observed throughout the

operating band as detailed in Section 3.3. It is seen from Fig. 4.4 that after

integrating the MTS antennas in the system, the beam in the H-plane is tilted

80
Fig. 4.5 Simulated current distribution at 6 GHz on the shared ground. The
magnitude is normalized to the maximum current over dipole 1.

J4 J6
(J1, J2) J3 J5 (J7, J8)
1.0 J1
J2
Modal Significance

0.8
J3
0.6 J4
J5
0.4
J6
0.2 J7
0.0 J8
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 4.6 Modal significances of the first eight modes of the unloaded MTS.

off from the boresight by 19°at 4.9 GHz, then the beam splits into two at 5.5

GHz, and two sidelobes occur at 5.8 GHz. Similarly in the E-plane, the levels

of the sidelobes and the tilt angle of the main beam increase versus frequency.

81
Fig. 4.7 Modal electrical currents and radiation patterns of the unloaded
MTS.

The pattern distortion is attributed to the excitation of the unwanted modes

of the three surrounding MTSs. The excited current distribution over the

shared ground plane is shown in Fig. 4.5. The modal significances and the

associated modal currents shown in Section 3.3 are reproduced in Figs. 4.6

and 4.7 for ease of reference. As can be seen from Fig. 4.5, the unwanted

currents are excited, including the out-of-phase currents under Antennas 2, 3,

and 4 and the x-directed currents under Antennas 2 and 3. According to the

TCM, any induced currents must be a linear superposition of the modal

currents. Therefore, the excited unwanted currents are from the combination of

the unwanted modes (J3, J4, J5, J6) because they exhibit similar symmetries of

the current distributions. A desired single broadside radiation beam can be

achieved by suppressing the unwanted currents.

82
4.2 Methods of suppressing higher-order modes

In principle, suppressing a mode means reducing its contribution to the

total radiated power. Mathematically, the contribution of a mode Jn to the total

radiated power Prad is measured by the magnitude of its modal weighting

coefficients (MWC), or |αn| which equals to the product of the modal

significance and |Vi| (the magnitude of the MEC). Therefore, there are two

methods for suppressing a mode, i.e., reducing the modal significance or |Vi|.

The first method proposed is the “mode shifting” for suppressing the modal

significance at the frequency of interest. A small modal significance requires a

large |λn|, meaning that the mode is relocated to resonate elsewhere far from

the original operating frequency, provided that λn is monotonically continuous

with respect to frequency so it approaches to zero at a lower or higher

frequency.

The second method proposed is the “mode cross-polarizing” for

suppressing the |Vi| that equals to the inner product of the modal currents Jn

and the impressed E-field Ei. It is proposed to mismatch the direction of Jn and

the polarization of Ei for a small |Vi|. For a given Ei, the modal currents Jn

should be cross-polarized to be orthogonal to that of Ei. While for a given Jn,

the incident field Ei should be cross-polarized, which has been applied to

determine the modal excitation schemes in the last chapter.

Both the two methods are based on the manipulation of λn and Jn, which are

completely determined by the geometry and material properties of the object.

Therefore, the proposed methods are independent of the excitation Ei once it is

impressed. In implementation, the modal suppression is implemented

83
Js1 Js3 Js4

1.0
Modal Significance

0.8

0.6

0.4

0.2

0.0
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 4.8 Modal significances of the slots-loaded MTS.

by loading the unit cells of the MTS.

4.2.1 Slot loading for shifting resonant frequencies

It is understood from Fig. 4.6 that the unwanted HOMs have smaller modal

significances before the resonances than after the resonances. It is therefore

decided to shift the HOMs to higher frequencies by shortening the current

paths of the HOMs. Since J1 is y-directed while all the HOMs (J2, J3, J4, J5, J6)

are with x-directed components, shortening the electrical lengths of x-directed

currents does not affect J1 but the unwanted modes.

Two slots, each of a width Wc = 0.3 mm, are cut into each of the unit cell

of the MTS, as shown in Fig. 4.8. The electric lengths of the x-directed

currents are reduced to approximately one third of the lengths of the desired y-

84
(a)

(b)

Fig. 4.9 Modal currents and E-field distribution of the slot-loaded MTS. (a)
Modal currents at 6 GHz (b) |Ez| distribution of the slot-loaded
MTS at 6 GHz. Note the different locations of nulls for |Ez1| and
|Ez4|.

directed currents. As the number and width of slots increase, the strips become

narrower and the potential bandwidth of J1 decreases though the modal

significances of the HOMs also decreases. Therefore, two slots are used to

shift the HOMs out of the BOI. The configuration of three strips looks similar

to that of the probe-fed MTS in Section 3.4, where the resonance along the

shorter edge is excited for wideband operation. Differently, here the resonance

along the shorter edge is the unwanted modes resonant at frequencies higher

than the BOI.

The modified modal significances are shown in Fig. 4.8. The modified

modal currents (Js1, Js3, Js4) and the associated z-components of the modal E-

85
fields are plotted at 6 GHz in Figs. 4.9(a) and (b). The subscript “s” denotes

the slot loading and i is the modal index. As can be seen from Fig. 4.8 and Fig.

4.9(a), the original modes (J2, J5, J6) have been shifted to higher frequencies

while (J1, J3, J4) are less affected. The calculated |αn| is similar to those in Fig.

4.5(c). It is desired to further suppress Js3 and Js4 as they may well be excited

in the system.

4.2.2 Via loading for cross-polarizing unwanted modes

The mode cross-polarizing method is proposed for further suppressing the

magnitude of MEC or |Vi| of the co-polarized HOMs. The idea is to change the

modal currents to a direction perpendicular to the polarization of the external

E-field. For Antenna 1, the external E-field is the superposition of the E-field

emitted from Antennas 2, 3, and 4, which are mostly φ-polarized since for the

two dominant modes the currents are mostly concentrated on the dipole that

generates φ-polarized radiating fields. Therefore, the currents should be z-

directed for reducing the portion being excited by the φ-polarized E-field. In

other words, the unwanted modal currents should be grounded using vertical

vias.

The challenge is to determine the loading positions so that the targeted

currents can be shorted without affecting the desired modes. For TM modes,

adding shorting vias at the nulls of E-field does not affect the current

distribution due to equal electric potentials at the two opposite ends of the vias.

On the contrary, it does so if potential difference is present. To minimize the

effect of the grounded vias on the desired modes while maximize it on the

HOMs, the vias should be located where the maximum E-field of the HOMs

and the minimum E-field of the desired modes are both present. As a result,

86
the |MEC| is reduced.

Now the target is to determine the optimum loading positions from the

modal E-field distribution, as shown in Fig. 4.9(b). As can be seen, the nulls of

|Ez1| are located at the strip center for the second row of unit cells. For the top

and bottom rows of the strips, the nulls of |Ez1| are slightly off center shrinking

inward by approximately 1 mm. At all these nulls, only |Ez4| exhibits large

values. Loading shorting vias at the nulls of |Ez1| significantly affects Js4 but

Js1.

87
Jsc1 Jsc2 Jsc3 Jv1 Jv3
1.0
Modal Significance 0.8

0.6
shorted
0.4

0.2

0.0
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(a)

J0 Jv3' Jv1'

1.0
Modal Significance

0.8

0.6 Jsc3'
Jsc2' shorted
0.4

0.2 Jsc1'

0.0
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)
(b)

0.20
Modal weighting coefficient

J0
0.15 Jv1'

0.10

0.05
Jsc1'
0.00
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

(c)

Fig. 4.10 Modal significances of the slot-via loaded MTS. (a) without
dipole, (b) with dipole. (c) |αn| of the slot-via loaded metasurface
antenna fed by lumped port.

88
Fig. 4.11 Modal currents of the slot-via loaded MTS without dipole.

The geometry of the via-loaded MTS is shown in Fig. 4.10(a). The

diameters of the vias are 0.6 mm. The vias are loaded at the center of the strips

for the second row. For the first and third rows, the vias are loaded with 1 mm

off from the patch centers toward the origin. The resultant modal significances

and modal currents are shown in Figs. 4.10(a) and 4.11. As can be seen, the

mode 4 disappears in the BOI but negligible influences are observed for Jv1

and Jv3. The subscript “v” denotes the modes after via loading. It is noticed

that three new modes resonating below 5 GHz are introduced, namely Jsc1, Jsc2,

and Jsc3, where the subscript “sc” stands for “short-circuit”. The currents of

Jsc2 and Jsc3 are mostly z-directed as expected regardless of their relatively

large modal significances in the operating band.

Fig. 4.10(b) shows the impact of the dipole on the modal significances of

89
the slot-via loaded MTS. Again, the dipole mode J0 is introduced at the lower-

end frequency and the desired MTS mode Jv1 is slightly shifted upward. The

calculated |αn| with a delta gap voltage is shown in Fig. 4.10(c). Similar to the

unloaded MTS antenna, J0 and Jv1 are simultaneously excited and they

dominate the radiation. The modal currents and radiation pattern of Jsc1' are

also shown in Fig. 4.10(c), where two unwanted sidelobes are observed. The

contribution of the slightly excited Jsc1 at 5.7 GHz is negligible as compared

to the two dominant modes. When the center column of the strips are also

loaded with shorting vias, the currents over the center column of strips become

stronger, which equivalently makes the system more capacitive. Therefore, the

mode significance and |αn| of Jsc1 slightly increase over the BOI since Jsc1 is

an inductive mode over the BOI. As a result, the maximum of the calculated

|αn| of Jsc1 slightly increases to 0.05 from 0.04 in the band of 5.5–6.2 GHz,

which is not desired. For the two reasons, the center column of the strips is not

loaded.

With the proposed mode-suppression techniques, the HOMs (J2, J4, J5, J6)

are well suppressed, whereas the desired dipole mode J0 and MTS mode J1 are

not affected. To further suppress J3, more advanced techniques may be

required, such as miniaturization design of the MTS in the x-direction because

this mode presents an x-directed resonance with antiphase currents, as

indicated by |Ez3| in Fig. 4.9(b). It is now expected that the pattern distortion

could be mitigated after the suppression of most of the HOMs.

4.3 Improvement of antenna radiation performance

4.3.1 Loaded MTS antenna with integrated balun

90
(a) (b)

Fig. 4.12 Configuration of the MTS antenna with integrated balun. (a)
bottom view with the dielectric removed; (b) side view at A-A’
cutting plane.

A series of transitions from the SubMiniature-version-A (SMA) connector

to the parallel twin-wires are integrated as a balun, as shown in Fig. 4.12. An

additional layer of substrate with a thickness of 0.508 mm is attached to the

bottom of the ground plane for printing the balun. The width of the feeding

gap is 0.2 mm unchanged. Two parallel vertical wires, each with a diameter of

0.55 mm and the edge-to-edge spacing of 0.4 mm, connect the dipole to the

coplanar-strip-line (CPS) on the bottom layer [128], where the SMA connector

is soldered to a 50-Ω microstrip line. Between the microstrip line and the CPS

is a 180°hybrid circulator, whose dimensions are optimized to provide equal

power and 180°phase difference for balanced feed of the dipole [129]. The

radius and width of the ring are 8.5 mm and 0.4 mm, respectively. The

characteristic impedances of the four ports of the circulator are designed at 50

. The loaded resistor at the isolation port is 52 .

The simulated S-parameters and boresight gain of the isolated MTS

antenna fed by lumped port and the integrated balun are compared in Fig. 4.13.

91
0

-10

|S11| (dB)
-20

-30
Lumped Port (Ld=18 mm)
-40
Balun (Ld=19.5 mm)
-50
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

9
Boresight Gain (dBi)

8
7
6
5
Lumped Port
4 Balun
3
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 4.13 Effects of integrated balun on the MTS antenna. Comparison of


|S11| and directivities of two isolated MTS antennas fed by lumped
port and balun, respectively.

The ground widths are 40 mm for both the cases. For impedance matching, Ld

is adjusted to 19.5 mm from 18 mm. Dielectric loss is included with a loss

tangent of 0.0027. The insertion loss of the balun is around 0.1 dB. As can be

seen, the difference between the two designs is negligible because the vertical

twin-wires have minor effects on the CMs when the modal E-field of the

dipole mode and the MTS mode are both null at the center of the MTS, similar

to the discussion in the probe-fed MTS antenna, where a cable penetrates the

substrate at null of the modal E-field without affecting the desired modes.

92
Fig. 4.14 Configuration of the slot-via loaded MTS antenna system. Left-
hand side: top view. Right-hand side: bottom view.

Fig. 4.15 Comparison of current distribution with (upper figure) and without
(lower figure) suppressing HOMs.

93
4.3.2 Improved radiation performance

The unloaded MTS antennas in Fig. 4.1 are replaced by the loaded MTS

antennas, as shown in Fig. 4.14. By exciting Port 1 and terminating the other

three ports with 50-Ω loads, the simulated surface currents on the ground

plane are plotted in Fig. 4.15. It is obvious that the currents under the three

terminated MTS antennas are significantly reduced in magnitude, indicating

that the presence of the unwanted modes has been effectively suppressed. In

addition, it is seen that the currents are mainly confined locally under the

Antenna 1, meaning that the neighboring ground and antennas both become

less visible to the active antenna.

To highlight the improvement of radiation performances, three cases are

compared in terms of the radiation patterns, 10-dB impedance bandwidth,

boresight directivities, and the S-parameters. The three cases are: (1) four

dipoles matched at 5.5 GHz (23 mm in length and 1.5 mm in width) at a height

of a quarter wavelength (10.5 mm) horizontally placed above a ground plate;

(2) four unloaded MAs shown in Fig. 4.1; and (3) four loaded MAs shown in

Fig. 4.14. For all the cases, the square ground is 80 mm in width and 50-Ω

lumped ports are used to terminate the four antennas whereas only Port 1 is

excited. The comparison is summarized in Table 4.1. Detailed S-parameters,

directivities and radiation patterns are shown in Figs. 4.16 and 4.17.

94
0

S-Parameters (dB)
-10

-20

-30 Dipole
Loaded
Unloaded
-40
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

10
9
Directivity (dBi)

8
7
6
Dipole
5 Loaded
4 Unloaded
3
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 4.16 Comparison of |S11| and boresight directivities of the three cases.

Fig. 4.17 Simulated gain patterns of the three cases. The patterns are plotted
at 5 GHz, 5.5 GHz and 5.8 GHz for (1) four dipoles with a height
of quarter wavelength over a ground, (2) four unloaded
metasurface antennas, and (3) four loaded metasurface antennas.

95
TABLE I Table 4.1
Comparison
Comparison ofofAntenna
AntennaPerformances
Performance for
withtheand
Dipole-Fed
without MTS Antennaof
Suppression
Higher-Order Modeswith and without Suppressing HOMs
Beam Beam Bandwidth*2, Boresight |S21|,
Model 1
Split Tilt* Height Directivity |S31|*3
10 %, 28 dB,
Dipoles No 3°~10° 7.5~8 dBi
20 mm 45 dB
Unloaded 24 %, 22 dB,
Yes 3°~19° 6.8~8 dBi
MAs 5 mm 33 dB
Loaded 21 %, 26 dB,
No 2°~8° 8.5~9 dBi
MAs 5 mm 36 dB

*1. Beam tilt: the tilt angle of the beam maximum in the operating band.
*2. Bandwidth: the fractional bandwidth with |S11| < -10 dB.
*3. Only |S31| is given since |S31| and |S41| are equal for the first two cases, and
they are of <1dB difference for the last case. Worst values over BOI are given.

As can be seen from Table 4.1, the conventional dipole design has a large

height of 20 mm or a quarter wavelength for high gain, and a limited

bandwidth of typically 10%. The two limitations are mitigated by the proposed

dual-mode MA with a 5-mm profile and a fractional bandwidth of 24%.

However, the application of the unloaded MA in compact multiport antenna

systems is hampered by the severe pattern distortion caused by the unwanted

HOMs. With the HOMs suppressed, the pattern distortion is significantly

alleviated, and the low-profile and wideband characteristics are maintained. In

addition, the directivities and port isolations of the loaded MAs are improved

in comparison with the unloaded MAs. It is worth pointing out that the

improved radiation patterns of the MAs, though not perfected, have been

comparable to those of the original dipole antennas. There are slight tilted

angles for the improved patterns in Fig. 4.17, which are mainly caused by the

asymmetrical ground plane but much smaller than those of the unloaded MA

and the dipole antennas (see Table 4.1).

96
Fig. 4.18 Fabricated prototype of the slot-via loaded MTS.
Fig. 18 Photograph of the fabricated sample of the metasurfaces.

To validate the concept and results, the loaded MA system is fabricated

and tested. The photograph of the fabricated sample is shown in Fig. 4.18. The

simulated and measured S-parameters, boresight gain, and radiation patterns

are in good agreement, as shown in Figs. 4.19 and 4.20. Compared with the

antennas fed by lump ports, there is no noticable difference in the bandwidth

and radiation patterns for the balun-integrated design, but the port isolation is

improved by 5 dB owing to the decoupling function of the hybrid [129]. The

achieved port isolation is better than 30 dB, as shown in Fig. 4.19.

97
Meas |S11| |S21| |S31| |S41|
Simu |S11| |S21| |S31| |S41|
0

S-Parameters (dB)
-10
-20
-30
-40
-50
-60
-70
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

9
Boresight Gain (dBi)

8
7
6
5 Meas
4 Simu
3
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

Fig. 4.19 Simulated and measured results with HOMs suppressed. The
simulated and measured S-parameters and boresight gain are
compared, respectively. For gain measurement, only Port 1 is
excited and the rest are terminated with 50-Ω loads.

98
Simu_Co Simu_Co
4.9 GHz  =180
0° Simu_X 4.9 GHz  =180
0° Simu_X
0 30°
210 30°
150 Meas_Co 0 30°
210 30°
150 Meas_Co
Meas_X Meas_X
-10 -10 60°
120
60°
240 60°
120 60°
240

-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 E-plane -10 H-plane
0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

5.5 GHz  =180


0° 5.5 GHz  =180

0 30°
210 30°
150 0 30°
210 30°
150

-10 60°
120 -10
60°
240 60°
240 60°
120
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 E-plane -10 H-plane
0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

5.8 GHz  =180


0° 5.8 GHz  =180

0 30°
210 30°
150 0 30°
210 30°
150

-10 60°
120 -10
60°
240 60°
240 60°
120
-20 -20
-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 E-plane -10 H-plane
0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

Fig. 4.20 Simulated and measured radiation patterns of the loaded MTS
antenna in the system (normalized). Results are plotted at three
representative frequencies. All antennas are connected to SMA
connectors. Only Port 1 is excited and the rest are terminated with
50-Ω loads.

99
4.4 Summary

With the theory of characteristic modes, the pattern distortion of multi-

mode MTS antennas in a wideband multiport antenna system has been

modeled, analyzed, and mitigated. It has been proven that the HOMs have

mainly caused the pattern distortion. The method to suppress the HOMs by

slotting and grounding the unit cells of the MTS has been proposed and

experimentally validated. As a result, the radiation patterns of the unloaded

metasurface antennas have been significantly improved. Furthermore, in

comparison with the conventional dipole antenna system, the proposed

antenna system has achieved a height reduction of 75%, and a bandwidth

enhancement of 100%.

In addition to the advantageous modal-resonance control ability and large

bandwidth of MTSs for antenna elements, we have demonstrated that a certain

mode amongst multiple modes of an MTS can be selectively controlled by

refining and optimizing the unit cells, which allows for a capability of shaping

the near-field scattering between antennas for a multi-antenna system adopting

MTSs. Beyond a single antenna element, the proposed mode-manipulation

technique paves the way for surface-integrated wideband multiport antenna

systems.

100
Chapter 5

Modal synthesis of MTS antennas

With various mode excitation and manipulation techniques introduced in

Chapters 3 and 4, respectively, the key to designing an MTS antenna is to

synthesize the broadside modes at the desired frequencies. For MTSs antennas

employing the local TM01 unit-cell resonance, the resonant frequency can be

tuned by modifying the geometry of the unit cells, such as patch width, slot

width, shape of the unit cells, and shape of the lattice, as introduced in Chapter

2. The searching method is conceptually simple but practically inefficient

since the searching space is infinite. It requires rich experience of the designer

and considerable computational resources to calculate and search the mode but

may not warrant a success of the modal synthesis.

An electrically large MTS is promising for high gain as well as wideband

operation due to an enlarged occupying area of surface currents with reduced

modal quality factor. However, the searching method is inefficient for the

design of electrically large MTSs. First, as the number of unknowns increases

with the growing electrical size of the MTS, the required computation cost

(time and memory storage) rises up dramatically fast because the computation

cost is O(N3) for the standard MoM [80] and at least O(N2) for the matrix

decomposition in solving the eigenvalue problem [79], where N is the number

101
of the unknowns. On the other hand, the number of significant modes also

increases, which requires much denser frequency samples for accurate modal

tracking. In addition to the mathematical complexity, the physics of the local

resonance also contributes to the increasing number of meshes (unknowns).

For a local TM01 resonance, the magnetic currents are mostly concentrated at

the edges of the unit cells. To represent the current variation, much denser

meshes are required at the edges than in the central area of the patch. It is

therefore expensive to use uniform meshes since the central area of the patch

needs to be meshed unnecessarily in the same density as that for the edges

without contributing to increasing the accuracy. It may be more expensive to

use non-uniform meshes due to the iterative mesh-adaptive process.

Consequently, the local-resonance concept has become the main obstacle for

the efficient and robust modal synthesis of electrically large MTSs containing

a great number of fine details, though effective for electrically small and

medium cases.

In this chapter, the concept of global-resonant MTSs is introduced. Herein

the unit cell is non-resonant with its electrical size being in the order of

subwavelength. The non-resonant unit cell exhibits slowly varied phase for the

currents so that the currents and fields can be averaged over one period of the

unit cell. In this manner, the MTS can be represented by a homogenized

impedance sheet or impedance boundary condition (IBC). As a result, the

dependence of unit cells can be eliminated and the meshes can be uniform and

relatively sparse for the efficient and robust modal analysis of MTSs. With the

homogenization, the solution to the optimum unit-cell design can be achieved.

Also, the modal analysis, modal synthesis, and further modal manipulation can

102
be performed in a macroscopic level without concerning the specific design of

the unit cell beforehand. Moreover, the global-resonant concept provides a

unique opportunity to incorporate the HOMs into operation, allows the

bandwidth of MTS antennas being significantly enhanced on the basis of the

previous works.

5.1. Truncated impedance-sheet model of global-resonant MTS

antennas

5.1.1 Control of fundamental mode by sheet impedance

With non-resonant unit cells, the analysis and synthesis of an MTS can be

greatly simplified by approximating the MTS with a homogeneous impedance

sheet or IBC that gives a similar macroscopic scattering response in far field.

The concepts of impedance sheet and IBC have been used for analyzing the

scattering coefficients of frequency selective surfaces [70]–[74], [130]–[132]

and recently for the analysis and synthesis of surface-wave and leaky-wave

devices [56], [75]–[76], [133]–[140]. On the other hand, the CMs of wire

objects are controllable with partially loaded lumped elements [86], [88]–[90],

[92], [95], [101]. However, the CMA of a truncated impedance sheet has not

been reported.

The proposed single-layer impedance-sheet model (SISM) for a class of

MTS antennas is shown in Fig. 5.1. In the SISM, an infinitely thin and purely

reactive impedance sheet is separated from an infinite PEC ground plane. A

dielectric layer with a thickness of h is filled in between. The shape of the

sheet can be arbitrary while a square sheet with a width of w is assumed for

103
Fig. 5.1 Impedance-sheet model for global-resonant MTS antennas.

proof of concept. Vacuum is assumed for the dielectric layer. The sheet is

isotropic, homogeneous, and non-dispersive in the initial study for simplicity.

The proposed idea is to simplify the modal analysis and synthesis with the

ideal model first and then approximate the ideal model by periodic structures

to reconstruct the predicted modes. The effects of frequency dispersion are

considered in Section 5.1.3.

The sheet impedance, Zs in Fig. 5.1, is the impedance of an ideal

homogeneous sheet, analogous to the permittivity of a homogeneous material

such as water and glass. In an isotropic case, the scalar sheet impedance is Zs =

Rs + jXs and Et = JsZs, so Zs links the tangential E-field and the surface current

density, where Et is the tangential E-field at a point on the sheet and Js is the

surface current density at the same point [69]. It is a scalar IBC that specifies

the ratio of the Et to the Js on the sheet. For purely reactive impedance sheets,

Rs = 0 and Zs = jXs. The unit of Zs is “ohms per square”, denoted as “Ω/sq”,

104
which is dimensionally equal to Ω but invariable under scaling therefore can

be used to compare the electrical properties of impedance sheets with different

sizes [67], [76]. It is exclusively used for sheet impedance to avoid confusing

1 Ω/sq of sheet resistance with 1 Ω of lumped resistor.

The grid impedance, Zg in Fig. 5.1, is the effective or averaged sheet

impedance of a grid (period) structures, analogy to the effective permittivity of

a metamaterial. Mathematically, Zg = Rg + jXg (Ω/sq) is defined in the same

way as Zs and can be measured in experiments or simulations using S-

parameters [76]. The difference is that the tangential E-field and the surface

current density for Zg are averaged over one period. Therefore, Zg loses its

definition when looking inside one unit cell whereas Zs is not restricted (in the

context of classical electrodynamics). Since complex numbers cannot be

directly compared, in the following content, the Rs, Xs, Rg, Xg stand for the real

values of the associated physical quantities without units for ease of

description and value comparison.

With the time convention of ejt, the sheet is capacitive for Xs < 0. When Xs

= 0, the sheet is simply a PEC patch and the SISM degenerates to a classical

microstrip patch antenna. The sheet is meshed using triangles on which

electric basis functions are defined in numerical modeling. By performing

CMA on the system, the modal resonant frequencies, the characteristic

currents Jn (n = 1, 2, 3…) and the modal far-field radiation patterns are

calculated. For proof of concept, w = 0.5 λ0 and h = 0.05 λ0, where λ0 is the

wavelength in free space at f0. The maximum mesh size is 0.01 λ0. The sheet is

modeled as the “tabulated surface impedance” in the CST with Zs = jXs (Ω/sq).

In principle, the problem is frequency independent, which is also validated in

105
(a)
3.7 % 4.4 % 6.9 %
1.0
Zs = j20
0.8 Zs = 0
Modal Significance

Zs = - j20
0.6

0.4

0.2

0.0
2.0 2.2 2.4 2.6 2.8 3.0
Frequency (GHz)

(b)
Fig. 5.2 Modal manipulation by sheet-impedance. (a) Eigenvalue and (b)
modal significance of the first broadside radiating mode against Zs.

simulations by testing f0 at 1 GHz, 2 GHz, 6 GHz, and 30 GHz, respectively.

The results are the same.

The eigenvalue can be controlled by changing the sheet impedance.

According to the TCM, a mode is inductive, resonant, or capacitive when the

associated eigenvalue is positive, zero, or negative, respectively. Fig. 5.2(a)

shows the effects of varying Xs on the eigenvalue of the first broadside mode.

106
As can be seen, when the sheet becomes capacitive (Xs < 0), the eigenvalue

decreases as compared to the PEC case, meaning that the net stored electric

energy increases. As a result, an inductive mode turns into a resonant mode at

a frequency higher than that of the PEC case. On the other hand, when the

sheet becomes inductive (Xs > 0), the eigenvalue increases and the net stored

magnetic energy increases, so the mode resonance shifts to a lower frequency.

Fig. 5.2(b) shows the associated modal significances against Xs. As can be

seen, when Xs decreases from +20 to –20, the resonant frequency, the

electrical size of the MTS at the resonance, and the modal bandwidth all

increase. Note that the physical size of the sheet does not vary. For only

capacitive sheets, it is expected that the mode resonance can be shifted to a

higher frequency as the sheet becomes less capacitive (or Xs more negative),

so the MTS becomes electrically larger for enhanced modal bandwidth. Also

note that a larger capacitance is associated with a smaller magnitude of the

reactance of a capacitor because the magnitude of the reactance of a capacitor

is inversely proportional to its capacitance.

The merits of the effect are in two folds. First, it suggests that the modal

resonant frequency is tunable without changing the size of the sheet but the

sheet impedance. Second, it suggests that the size of the sheet (thus the

antenna aperture) can be scalable without changing the modal resonant

frequency. With a predefined width of w and thickness of h, the ISM can be

used for estimating the required Xs to relocate the desired mode to the targeted

frequency. In comparison with traditional microstrip antennas using a solid

PEC sheet, a new parameter of sheet impedance has been introduced to control

the mode.

107
Fig. 5.3 The scalar single-layered and bilateral sheet-impedance model.

The single-layered model can be generalized to the bi-layered impedance

sheet model (BISM). The BISM introduces an additional shunt sheet

impedance to facilitate the PCB implementation. The dielectric-loaded SISM

and BISM are shown in Fig. 5.3. The dielectric layers (including air) are used

for implementing the sheet impedance with the PCB technologies [135]–[136].

In the SISM, a dielectric layer is coated on top of the sheet. An additional

impedance sheet is added in the BISM. For fair comparison, the dielectric

layer has the same thickness of t and relative permittivity of εr for both models.

The sheet impedances are Xs1 and Xs2 for the SISM and BISM, respectively.

All dielectrics are assumed lossless, homogeneous, isotropic, and infinitely

expanded in the transverse directions unless otherwise specified.

The objective is to control the resonant frequency of the fundamental

broadside mode by modifying the sheet impedance such that the mode is

placed at the desired frequency f0 with the sheet geometry predefined. For

108
1.0 J1
J2

Modal Significance
0.8
J3
0.6 J4
0.4 J5
J6
0.2 J7
SISM
0.0 J8
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)
(a)
1.0 J1
J2
Modal Significance

0.8
J3
0.6 J4
0.4 J5
J6
0.2
BISM J7
0.0 J8
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)
(b)

Fig. 5.4 Modal synthesis results using the impedance-sheet model. Modal
significances are calculated for (a) the SISM with Xs1= -185 and
(b) the BISM with Xs2 = -373.

proof of concept, we assume f0 = 6 GHz, corresponding to the wavelength in

free space of λ0 = 50 mm, and w = λ0, h = 0.1 λ0 for both models in Fig. 5.3,

where t = 0.813 mm and εr = 3.55. When Xs = 0 (PEC), the TM01 mode

resonates at approximately 2.4 GHz for both models. To relocate the first

broadside mode to 6 GHz, the required Xs is capacitive and solved through a

few rounds of parameter scanning in the CMA of the SISM and BISM,

respectively. The results are Xs1 = –185 and Xs2 = –373 for the SISM and

BISM, respectively.

109
Fig. 5.5 Modal electric currents for the BISM at 6 GHz.

Fig. 5.4 shows the modal significances of the SISM and BISM. As can be

seen, the results are similar so both models are capable of manipulating the

modes. It is noticed that Xs1 is approximately half of Xs2 because the two

sheets of the BISM are shunted in the transmission-line model (see Section

5.1.4), resulting in a halved shunt impedance of 0.5 Xs2. Therefore, the Xs1

obtained from the CMA of the SISM can be directly used for estimating Xs2

without further CMA of the BISM, provided that the spacing t is much smaller

than wavelength.

Fig. 5.5 shows the modal currents on the top sheet of the BISM. The SISM

has similar current distributions (not shown here for brevity). Two orthogonal

pairs of broadside radiating modes are observed, namely, the modes (J1, J2)

resonating at 6 GHz and the modes (J6, J7) at frequencies higher than 7 GHz,

respectively. For broadside radiation with y-polarization, the modes (J1, J6)

are the desired modes. The modes (J2, J7) generate the cross-polarized field.

The rest modes are the unwanted modes due to radiation nulls at the broadside.

110
Fig. 5.6 Analogy to waveguide modes. The vector field patterns are plotted
for the modes J1 and J6 on the top surface of the impedance
sheet. An analogy is made to a rectangular PMC-walled
waveguide.

5.1.2 Analogy to waveguide modes

The desired modes are physically different from those of the local-resonant

MTSs discussed in Chapter 2. For revealing the difference, an analogy is made

to the well-known waveguide modes and we take J1 and J6 as two examples

[141]. Fig. 5.6 shows the vector E-field and H-field patterns of the modes J1

111
and J6 on the upper surface of the top sheet (the same for either side of the two

sheets). As can be seen, for both modes the Eφ components dominate the

aperture and the Ez components occur mainly near the edges. For the H-fields,

there are only the Hφ components and Hz = 0. Therefore, the two modes are the

TMz modes. To find the modal indices, an analogy is made to the vector field

patterns of a perfect magnetic conducting (PMC) walled rectangular

waveguide since the H-fields of J1 and J6 are normal to the four virtual side

walls. The field patterns of the TMz01 and TMz21 modes of a PMC-walled

rectangular waveguide are inserted in Fig. 5.6 for comparison. It is clearly

seen that J1 and J6 are the TMz01 and TMz21 modes, respectively. Note that the

TMz01 and TMz21 modes are caused by the global resonances of the overall

sheet, whereas in the previous works, the operating mode is the quasi-TM0N

(N>1) MTS mode contributed by the local resonances of the unit cells.

5.1.3 Control of higher-order modes by frequency dispersion

It has been demonstrated that the resonant frequency of the TMz01 mode

can be controlled by the sheet impedance, where a constant Zs against

frequency (non-dispersive) is assumed. In realistic situations, the implemented

effective sheet impedance is usually frequency dispersive, meaning that the

impedance exhibits different values at different frequencies. Since the

eigenvalue is directly related to the sheet impedance at the same frequency,

the frequency dispersion can be used to control the resonant frequency of the

HOMs. While the suppression of selected HOMs helps reduce the mutual

coupling between antennas, an HOM can be constructive. For example, the

above mentioned TMz21 mode can be used for bandwidth enhancement of

broadside radiation when it resonates at a frequency close to the TMz01 mode.

112
-100

Sheet Reactance (jΩ/sq)


0th-order (non-dispersive)
-200
1st- order (dispersive)
-300

-400

-500

-600
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

(a)

bandwidth reduced
1.0 J1
J2
Modal Significance

0.8
J3
0.6 J4
0.4 J5
J6
0.2
Dispersive BISM J7
0.0 J8
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)
(b)

Fig. 5.7 Effects of frequency dispersion. (a) Dispersive sheet reactance Xs


versus frequency and (b) the associated modal significances.

In the non-dispersive case shown in Fig. 5.4, the TMz21 mode resonates above

7 GHz. A redshift of the TMz21 mode is desired so the value of the Xs2 at the

TMz21 resonance should be larger than -373, corresponding to a positive slope

of the sheet reactance against frequency, as shown in Fig. 5.7(a).

The effects of a linear (first-order) positive slope of Xs2(f) on the modal

resonant frequencies and modal bandwidth are studied. In most cases, the first-

order model is a good approximation if there is no strong unit-cell resonance

113
in the BOI to change the slope dramatically. The non-dispersive case can be

regarded as the zeroth-order model. Fig. 5.7(a) shows the frequency dispersion

of Xs2. Let Xs2 = 373 at 6 GHz be unchanged and act as the anchor point, the

slope of the sheet reactance is determined by setting Xs2 to 515 at 4.5 GHz

(which gives a similar slope as that implemented in the next sub-section).

The associated modal significances are shown in Fig. 5.7(b). As compared

to the non-dispersive BISM in Fig. 5.4, the first eight resonance peaks of the

dispersive BISM shrink toward 6 GHz from both sides when the slope of Xs2

increases. For example, the resonant frequency of J6 shifts down to

approximately 6.7 GHz from above 7 GHz. On the other hand, the modal

bandwidths (with modal significance greater than 0.8, for example) of both J1

and J6 decrease. Therefore, the frequency dispersion is effective in controlling

the distance between the two resonant peaks. In general, the larger the slope,

the closer spacing between the modes, and the smaller the bandwidth results.

The closely located TMz01 and TMz21 modes can be combined for bandwidth

enhancement of MAs.

5.1.4 Realization of predicted modes

To approximate the desired sheet impedance, a square patch in a square

lattice is a typical choice for isotropic grid impedance. The simulation setups

for extracting the grid impedance using numerical methods are shown in Fig.

5.8. The lattice periodicity p is 5 mm (0.1λ0) to meet the homogenization

condition. Periodic boundary conditions (PBCs) are defined at the four sides

and Floquet ports are defined on the top and bottom with the wave vectors

normal to and pointing toward the patch. With the transmission-line model

(TLM) and the PBCs, the scalar grid impedance Zg = Rg + Xg is extracted

114
Fig. 5.8 Simulation setup for grid impedance extraction. The geometry of
the unit cell and simulation setup for extracting Zg.
Grid Reactance (jΩ/sq)

-150 -183

-300 -375

-450
-450 Wp=4.61 mm
-600
Wp=4.74 mm
-750 Wp=4.98 mm
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

(a) (b)
Fig. 5.9 Extracted grid impedance. (a) Zs versus the width of square patch,
(b) the constructed MTS with ten periodicities and same patterns
on both sides.

using two y-polarized normal incidences, each from one port, respectively [76],

[135]–[136]. The Z0 in the TLM denotes the characteristic impedance of the

air, whose electrical length is λ0 in the setup.

Fig. 5.9(a) shows the extracted Xg for different Wp. The value of the grid

resistance Rg is in the order of 10-5 due to numerical errors, thus being treated

115
as zero. As can be seen, Xg is more capacitive in the observing band for a

larger Wp. To approach Xs1 = 185 for the SISM and Xs2 = 373 in the BISM,

the required Wp is approximately 4.98 mm and 4.74 mm, respectively.

Apparently, it is difficult to implement Xs1 for the SISM using conventional

PCB technologies due to the small spacing (~20 μm) between neighboring

patches. For this reason, only the BISM is discussed in the subsequent

contents.

The required sheet impedance Zs has been approximated by the grid

impedances Zg of periodic structures. Fig. 5.9(b) shows the geometry of the

constructed MTS. Ten periodicities are used in each direction and the same

patterns are printed on both sides. The overall width of the MTS is w = 50 mm.

With the ideal sheets replaced by the MTSs with Wp = 4.74 mm, the modal

significances of the first eight modes are shown in Fig. 5.10(a). As can be seen,

all the first eight modes predicted by the equivalent model have been

reproduced in terms of the similarity of modal significances. The resonant

frequencies of the realized modes are slightly lower than those in Fig. 5.7(b).

This is caused by the edge effects due to the loss of coupling environment for

the edge elements. To compensate for the edge effects, the patch size is

reduced to tune the resonant frequencies of J1 and J2 back to 6 GHz. A fast

guess is made. Since the predicted J1 resonates at 6 GHz, which is 1.2 times of

the realized resonant frequency, the estimated value of Xg is ‒373×1.2 =

447.6, corresponding to Wp = 4.61 mm, as shown in Fig. 5.9(a). The modal

significances with the compensation are shown in Fig. 5.10(b), where J1 has

been relocated at 6 GHz.

116
1.0 J1
J2

Modal Significance
0.8 5.2 GHz
J3
0.6 J4
0.4 J5
J6
0.2 4.74 mm
J7
0.0 J8
4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

(a)

1.0 J1
J2
Modal Significance

0.8
6 GHz J3
0.6 J4
0.4 J5
J6
4.61 mm
0.2 J7
0.0 J8
5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

(b)
Fig. 5.10 Modal significances of the synthesized MTS. edge effects. (a)
Without edge-effects compensation and (b) with edge-effects
compensation.

The predicted and the reconstructed modes are different in terms of

detailed current distributions but they give similar far-field radiation patterns.

Therefore, the realized J1 and J6 can be named as the quasi-TMz01 and quasi-

TMz21 modes (Q-TM01 and Q-TM21 modes in short), respectively.

The equivalent model is proved to be accurate and effective in predicting

the functional CMs independent of any specific unit cells. It allows the

functional MTS modes to be analyzed and engineered before it is physically

constructed, leaving the synthesis of the MTS greatly simplified into designing

a single unit cell for the desired grid impedance. The desired grid impedance is

117
Fig. 5.11 Modal currents and radiation patterns of the synthesized MTS.

aimed at the predefined frequency for the TMz01 mode. For simultaneous

control of the TMz01 mode and the TMz21 mode, establishing a relation

between the geometry of the unit cell and the frequency dispersion is required

for simultaneous control of the the grid impedances at both resonant

frequencies of the two modes.

5.2 Wideband triple-mode global-resonant MTS antennas

The MTS with edge compensation is excited as an antenna. The modal

currents are used for guiding the excitation design. Fig. 5.11 plots the modal

electric currents and the associated modal radiation patterns for the first eight

modes. In comparison with the averaged currents in Fig. 5.5, the symmetries

118
Fig. 5.12 Combined currents of J1 and J6 for multi-mode excitation.

of the current distributions are well preserved. The synthesized modes J1, J2,

J6, and J7 exhibit broadside radiation patterns as the equivalent modes do. For

broadband broadside y-polarized radiation, it is desired that both J1 and J6 can

be both effectively excited in phase. Fig. 5.12 plots the combined current

distribution Jc = J1 + J6 at 6 GHz for both the equivalent and constructed

modes. As can be seen, two regions with strong currents stand out for both

cases. Two feeders are therefore used to excite the modes J1 and J6

simultaneously. The two regions are also clearly seen in Fig. 5.11, as circled

out. Using the combination of currents is a more robust approach to find a

good placement of the feeders.

Fig. 5.13 shows the geometry of the proposed MA fed by two coplanar-

dipoles. The dipoles are placed near the two maxima of Jc with a spacing of D.

To avoid overlapping, a slot loop with a width of Ws is etched around each

dipole. Each dipole consists of two strips printed on the opposite sides of the

substrate and connected by metallic vias of 0.6 mm in diameter. The length

and width of the strips are Ld and Wd, respectively. The feeding gap is 0.2 mm

119
Fig. 5.13 Geometries of the MTS antenna fed by dipoles. Lump ports are
used.

in width and a lumped port is defined across the gap with a port impedance of

Zp = 145 Ω. To broaden the gain bandwidth, Ld is designed so that the even

mode of the dipoles (thus the sum beam) resonates at a frequency lower than

that of the Q-TM01 mode, similar to the design concept in Section 3.3. With an

in-phase feed of the two dipoles, the dimensions of Ld, Wd, Ws, and D are

optimized for impedance matching and gain bandwidth using CST 2018. The

optimized dimensions are Ld = 27 mm, Wd = 1.1 mm, Ws = 0.7 mm, and D = 25

mm.

120
0

-10

|S11| (dB)
-20
Wg = ∞
-30 Wg = 60 mm
Wg = 70 mm
-40
3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

Fig. 5.14 Effects of ground size on impedance matching and directivity.


Lumped ports are used for calculating the |S11| and boresight
directivities.

Fig. 5.14 shows the simulated |S11| and the boresight directivities with

different ground sizes. With the infinite ground, the achieved bandwidth for

|S11| < ‒10 dB is 45%, within which the directivity is 9.8–13.5 dBi. The finite

sized ground plane hardly affects the impedance matching. When Wg = 60 mm,

the boresight directivity is 0.8–1.4 dB lower than that of the infinite case. It is

concluded that the model with an infinite ground has well approximated the

real problem. Here Wg = 60 mm is chosen to host the plastic spacers shown in

Fig. 5.13. The width of the dielectric portion of the MTS extends to Wg.

121
10

Radiated Power (mW)


Total Jd
8 J1
6 J6
4
2
0
4.0 4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

(a)
1.0
Jd
Modal Significance

0.8
J1
0.6 J6
0.4
0.2
0.0
3.5 4.0 4.5 5.0 5.5 6.0 6.5
Frequency (GHz)

(b)

Fig. 5.15 Illustration of the triple-mode operation with high excitation


purity. (a) Modal far-field radiated powers and (b) associated
modal significances.

The triple-mode operation of the proposed MA is clearly revealed by

studying the modal radiated powers. Fig. 5.15(a) shows the total far-field

radiated power of the antenna and that decomposed into each mode using

FEKO 2017.2 [120]. The port impedance is Zp (unchanged) and a constant gap

voltage of 1 volt is specified for each port. The impedance mismatching and

loss are excluded in the calculation of antenna directivity. The modes with

radiated power below 0.1 mW are not shown. As can be seen from Fig. 5.15(a),

the antenna radiation is dominated by three modes, whose modal significances

122
Fig. 5.16 Magnitude of current distributions of the three dominant modes.
From left to right: Jd at 4 GHz, J1 at 5.5 GHz, and J6 at 6.2 GHz.

are shown in Fig. 5.15(b). The magnitude distributions of the modal currents

are plotted in Fig. 5.16, at the frequencies with the highest radiated power in

the observing band. The first mode is defined as Jd where the subscript “d”

stands for dipole. As can be seen from Figs. 15 and 16, Jd is the additionally

introduced dipole mode (resonating at 3.6 GHz). The next two dominant

modes are associated with the two MTS modes J1 (Q-TMz01) and J6 (Q-TMz01)

in Fig. 5.11, respectively. With the proposed excitation scheme, the unwanted

modes are effectively suppressed in the band of interest.

The significances and advantages of the proposed coplanar modal

excitation scheme are summarized. First, only the desired modes are

effectively excited for wide gain bandwith at the presence of numourous

unwanted modes, which is usually challenging. Second, the back radiation is

significantly reduced in comparison with a slot-fed MTS antenna. Third, the

antenna bandwith is improved with the additional dipole mode. Most

importantly, the proposed low-profile widband feeding technology does not

increase any height of the antenna in constrast to the conventional cavity-

backed and ground-backed slot-coupling feed [142]–[144].

123
Fig. 5.17 Geometry of the feeding network.

Fig. 5.17 shows the feeding network for a wideband balanced feed of the

two dipoles, where two cascaded 180° hybrids are connected to an SMA

connector through a T-shaped power divider [145]. The diameter of the copper

wire is 0.6 mm and the spacing between the twin wires is 0.4 mm, giving rise

to a characteristic impedance of Z0 = 145 Ω for the vertical twin-wire

transmission line, which has been used as Zp for the lumped ports. The feeding

network is optimized for differential and equal-power output in the frequency

range of 4–7 GHz. The Rogers RO4003C substrate (εr = 3.55 and tanδ =

0.0027) with a thickness of t = 0.813 mm is used for both the feeding network

and the MTS. With the dielectric and copper loss included, the simulated

insertion loss of the feeding network is lower than 0.35 dB. The return loss is

greater than 20 dB.

124
Fig. 5.18 Photographs of the prototyped global-resonant MTS antenna. The
long copper wire is chopped into the segments to connect the
dipole on the top and the balun at the bottom.

Fig. 5.18 shows the prototyped antenna. Good agreements between the

simulated and measured |S11| and boresight gain are shown in Fig. 5.19. The

achieved impedance bandwidth for |S11| < -10 dB is 3.95–6.23 GHz or 45%,

within which the simulated and measured boresight gains are 8.9~11.7 dBi

and 8.5~11.6 dBi, respectively. Fig. 5.20 shows that the simulated and

measured radiation patterns (normalized) are in good agreement in both the E-

plane (yz-plane) and H-plane (xz-plane). The simulated and measured in-band

cross-polarization levels at the boresight direction are lower than -80 dB and -

30 dB, respectively.

125
0
-5 Simulated
Measuered

|S11| (dB)
-10
-15
-20
-25
3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

14
Boresight Gain (dBi)

12
10
8 Simu. Dir.
Simu. Gain
6
Meas. Gain
4
3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0
Frequency (GHz)

Fig. 5.19 Simulated and measured |S11| and boresight gain of the global-
resonant MTS antenna.

126
E-plane H-plane Simu_Co
Simu_Co
4.0 GHz  =180
0° Simu_X 4.0 GHz  =180
0° Simu_X
0 30° 0 30°
210 30°
150 Meas_Co
30°
210 150 Meas_Co
Meas_X Meas_X
-10 -10 60°
120
60°
240 60°
120 60°
240

-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 -10

0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

5.5 GHz  =180



5.5 GHz  =180

0 30° 0 30°
210 30°
150
30°
210 150

-10 -10 60°


120
60°
240 60°
120 60°
240

-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 -10

0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

6.0 GHz  =180


0° 6.0 GHz  =180

0 30° 0 30°
210 30°
150
30°
210 150

-10 -10 60°


120
60°
240 60°
120 60°
240

-20 -20

-30 90°
270 90°
90 -30 90°
270 90°
90

-20 -20
120°
300 120°
60 120°
300 120°
60
-10 -10

0 (dB) 150°
330 30
150° 0 (dB) 150°
330 30
150°
0
180° 0
180°

Fig. 5.20 Normalized radiation patterns of the global-resonant MTS


antenna. Patterns plotted at 4 GHz, 5.5 GHz, and 6 GHz.

127
Table 5.1
Comparison of Antenna Performance for the Global-Resonant MTS Antennas,
Conventional MTS Antennas, and Wideband Microstrip Patch Antennas
No. of
Design Antenna size (λL3) Excitation scheme FBR, BW
modes

Prop. 0.8×0.8×0.09 3 Coplanar Dipoles >20 dB, 45%

[55] 0.4×0.8×0.06 2 Suspended Monopole ~14 dB, 15%

[142] 0.8×0.6×0.07 2 Cavity backed slot >20 dB, 22%

[143] 0.5×0.4×0.13 2 Ground backed slot ~20 dB, 25%

[144] 0.7×0.7×0.18 2 Cavity backed slot >14 dB, 44%

Table 5.1 compares the key features of the proposed global-resonant MA

with those of the conventional MAs and wideband microstrip patch antennas,

including the antenna size, antenna height, the number of operating modes, the

excitation schemes, the front-to-back ratio (FBR), and the simulated

bandwidth (BW) using the criteria of return loss greater than 10 dB, where λL

is the wavelength in free space at the lowest operating frequency. In

comparison with conventional low-profile MAs [142]–[144], the bandwidth of

the proposed MA has been significantly improved with high FBR and

comparable antenna height. In comparison with the conventional microstrip

patch antennas to achieve wideband performances at the cost of increasing

antenna height, the proposed MA is advantageous in low profile and large

bandwidth simultaneously [56], [142]. For further reducing the antenna height

while simultaneously enhancing the bandwidth, a larger MTS is suggested

because increasing the electrical size of the MTSs offers a more favorable

potential gain-bandwidth product by compensating for the bandwidth

128
reduction due to reduced antenna height. With the proposed impedance-sheet

model, the modes can be efficiently synthesized at a unit-cell level without the

expensive blind searching.

5.3. Summary

A truncated impedance-sheet model has been proposed and validated for

accurate modeling, fast analysis and systematic design of a new class of low-

profile broadband metasurface antennas. The proposed concepts have allowed

the modes of a complicated and mesh-dense metasurface being efficiently

analyzed and effectively engineered with a simple equivalent homogeneous

sheet before designing the unit cells. The unit-cell independent modeling has

played a significant role in simplifying the synthesis of the metasurface from

expensive blind searching into the systematic design of a single unit cell. A

new measure of engineering surface properties has been proposed to

manipulate the modal resonant frequencies, which is the key to incorporate the

higher-order metasurface mode for bandwidth enhancement and paves the way

for further height reduction and bandwidth enhancement of metasurface

antennas. The Q-TMz01 and Q-TMz21 modes caused by finite truncation have

been identified, providing clear physical insights of the broadband behavior of

the proposed antenna. A new coplanar-feeding technology has been proposed

for selective multi-mode excitation, bandwidth enhancement, and the

reduction of antenna height, weight, and cost at the same time. The concepts

can be readily extended to different bands with dual polarizations and circular

polarizations for applications, such as cellular base stations and satellite

communications.

129
Chapter 6

Conclusions and future work

6.1 Conclusions

Microwave MTS antennas have been studied in this thesis using CMA, with

the primary aim of improving the performance of low-profile antennas and

antenna arrays.

In summary, the challenges of existing methods for analyzing finite-sized

MTSs are highlighted in Chapter 1, including the edge truncation not modeled,

the complexity of multi-mode near-field excitation not considered, and the

periodic boundary condition failing to take into consideration of local field

differences (non-periodic structures). Taking advantage of the TCM in

analyzing arbitrary structures independent of any excitation, a detailed CMA

of the first class of MTSs (local-resonant MTSs) is presented in Chapter 2,

where the multi-mode resonant nature of finite-sized MTSs have been clearly

revealed and the key parameters influencing MTS modes have been identified.

The novel insights gained on the MTS modes are then utilized by different

excitation schemes in Chapter 3 to design four low-profile wideband MTS

antennas providing significantly improved performance over existing designs.

Apart from exciting the selected modes while suppressing the unwanted

higher-order modes by the feed design for a single antenna element, two

130
methods of modal manipulation by unit-cell loading are proposed in Chapter 4,

in order to circumvent the pattern distortion caused by the mutual excitation of

higher-order modes in a multi-antenna system. Chapter 5 introduces a more

powerful concept of global-resonant MTS, which simplifies the design of an

MTS from optimizing a large number of parameters into the design of a single

unit cell by treating an MTS as a homogenous impedance sheet.

The contribution of this thesis is highlighted in the following.

Firstly, the introduction of CMA offers a new analysis method of MTSs. It

provides a new perspective to gain novel physical insights into the inherent

operating mechanism of finite-sized MTSs with arbitrary shapes. In

comparison with conventional methods which offer limited physical insights,

the combination of CMA and MTS reveals clearly the multi-mode resonant

nature of finite-sized MTSs and enables accurate modeling of the edge

truncation, near-field excitation, and local-field differences. Moreover, it

allows the effects of feeding structures on the MTS modes being quantitatively

characterized, favorably employed or effectively minimized.

Secondly, a systematic analysis and design framework for MTS structures

has been formulated. Two categories of planar multi-mode resonators have

been introduced, including the local-resonant MTS and the global-resonant

MTS. For both categories, the MTSs have been shown capable of generating

multiple orthogonal modes with distinct radiation patterns, resonant

frequencies and surface currents, which offer new solutions to fulfilling

challenging antenna design requirements. The effectiveness of the framework

is demonstrated by applying it to achieve significantly better performance of

low-profile antennas and arrays than existing designs.

131
Thirdly, while the concept of local-resonant MTS represents the philosophy

of searching and optimizing based on the innovative use of an analysis tool,

the concept of global-resonant MTSs, developed by virtue of the powerful idea

of field averaging, represents a significant advance in the state of the art

concerning how to inversely synthesize an MTS with particular modes

resonant at predefined frequencies. The proposed sheet-impedance model not

only greatly simplifies the modal analysis but also offers a new method to

design MTSs by establishing a link between the grid impedance and the modal

resonant frequencies. As a result, the design of MTSs has been greatly

simplified to matching the grid impedance and the geometry of a single unit

cell, with which more imaginative designs and new physical phenomena can

be expected. It is worth pointing out that the proposed impedance-sheet model

can be viewed as a generalized model of conventional microstrip patch

antennas, where two additional degrees of freedom (the sheet impedance and

its frequency dispersion) have been introduced. In this sense, a new design

paradigm of microstrip patch antennas has been introduced.

In addition, the concept of using MTSs to manipulate characteristic modes is

proposed and validated. Although mode manipulation is not fundamentally

new, so far it has only been applied to antenna element design and scattering

control. Two methods of modal manipulation, the unit-cell loading and the

grid-impedance engineering, have been proposed for manipulating the modal

resonant frequencies, the modal bandwidths, the inter-mode frequency spacing,

and the modal current/field distributions.

Last but not least, it is believed that the concepts and the antenna designs

proposed in this thesis can be readily adapted to make useful contributions in

132
facilitating the development of a networked society and an internet of things.

6.2 Future research

It is believed that the proposed combination of MTS and CMA has opened a

new avenue for further research. Several future directions are suggested below.

6.2.1 From antenna radiation to scattering and absorption problems

This thesis has mainly focused on the far-field radiation and near-field

scattering problems for antennas, but the proposed concepts may well be

extended to far-field scattering and absorption problems.

6.2.2 Optimization

Square patch and square MTSs have been studied in the thesis for their

simplicity. Optimizing the geometries of both unit cell and MTS can be a

natural extension of this thesis toward an optimum design.

6.2.3 Non-broadside modes

The broadside modes have been the main focus. It is of interest to extend the

proposed concepts to other non-broadside modes for distinct antenna radiation

characteristics.

6.2.4 From microwave to higher frequencies

The concepts and methods developed in this thesis may have paved the way

for applications at higher frequencies up to possibly the visible lights. In

consideration of possible loss and non-linear effects, the TCM and the

proposed modeling may both need to be adapted.

133
Bibliography

[1] K. L. Wong, Compact and Broadband Microstrip Antennas. New York:

Wiley, 2006.

[2] J. L. Allen and B. L. Diamond, “Mutual coupling in array antennas,”

M.I.T. Lincoln Laboratory, Lexington, Mss., Technical Rep. 424 (ESD-

TR-66-443), Oct. 1966.

[3] J. W. Wallace and M. A. Jensen, “Mutual coupling in MIMO wireless

systems: a rigorous network theory analysis,” IEEE Trans. Wireless

Comm., vol. 3, no. 4, pp. 1317–1325, Jul. 2004.

[4] M. A. Jensen and J. W. Wallace, “A review of antennas and propagation

for MIMO wireless communications,” IEEE Trans. Antennas Propag.,

vol. 52, no. 11, pp. 2810–2824, Nov. 2004.

[5] Y. Wang and Z. Du, “Dual-polarized slot-coupled microstrip antenna

array with stable active element pattern,” IEEE Trans. Antennas

Propag., vol. 63, no. 9, pp. 4239–4244, Sep. 2015.

[6] H. Li, B. K. Lau, and S. He, “Design of closely packed pattern

reconfigurable antenna array for MIMO terminals,” IEEE Trans.

Antennas Propag., vol. 65, no. 9, pp. 4891–4896, Sep. 2017.

[7] H. M. Bernety and A. B. Yakovlev, “Reduction of mutual coupling

between neighboring strip dipole antennas using confocal elliptical

metasurface cloaks,” IEEE Trans. Antennas Propag., vol. 65, no. 4, pp.

1554–1563, Apr. 2015.

134
[8] H. M. Bernety and A. B. Yakovlev, “Decoupling antennas in printed

technology using elliptical metasurface cloaks,” J. Applied Phys., vol.

119, no. 1, pp. 1–11, Apr. 2016.

[9] P. J. Gibson, “The Vivaldi aerial,” Proc. 9th Eur. Microw. Conf.,

Brighton, U.K., 1979, pp. 101–105.

[10] L. R. Lewis, M. Fasset, and J. Hunt, “A broadband stripline array

element,” Proc. IEEE Int. Antennas Propag. Symp. Dig., vol. 12, Jun.

1974, pp. 335–337.

[11] J. Shin and D. H. Schaubert, “Toward a better understanding of

wideband Vivaldi notch antenna arrays,” Proc. Antenna Application

Symp., Allerton Park/Monticello, IL, Sep. 20–22, 1995.

[12] F. Lin, Y. Qi, and Y. C. Jiao, “A 0.7–20 GHz strip-fed bilateral tapered

slot antenna with low cross-polarization,” IEEE Antennas Wireless

Propag. Lett., vol. 12, no. 1, pp. 737–740, Jun. 2013.

[13] F. Lin, Y. C. Jiao, and Y. Qi, “A wideband microstrip to bilateral

slotline transition using constant impedance bilateral slotline and heart-

shaped irregular cavity,” IEEE Microw. Wireless Comp. Lett., vol. 23,

no. 5, 255–257, Apr. 2013.

[14] H. A. Wheeler, “Simple relations derived from a phased-array antenna

made of an infinite current sheet,” IEEE Trans. Antennas Propag., vol.

AP-13, no. 4, pp. 506–514, Jul. 1965.

[15] I. Tzanidis, K. Sertel, and J. L. Volakis, “UWB low-profile tightly

coupled dipole array with integrated balun and edge terminations,” IEEE

Trans. Antennas Propag., vol. 61, no. 6, pp. 3017–3025, Jun. 2013.

135
[16] S. Shi et al., “Ultrawideband optically fed tightly coupled phased

array,” J. Lightwave Tech., vol. 33, no. 23, pp. 4781–4790, Dec. 2015.

[17] J. J. Lee, S. Livingston, and R. Koenig, “Wide band long slot array

antennas,” Proc. IEEE Antennas Propag. Symp., vol. 2, Columbus, OH,

June 2003, pp. 452–455.

[18] A. Neto and J. J. Lee, “Ultrawide-band properties of long slot

arrays,” IEEE Trans. Antennas Propag., vol. 54, no. 2, pp. 534–543, Feb.

2006.

[19] W. W. Hansen, Radiating electromagnetic waveguide, U.S., Patent No.

2.402.622, 1940.

[20] C. H. Walter, Traveling Wave Antennas. New York: McGraw-Hill, 1965.

[21] C. Caloz, T. Itoh, and A. Rennings, “CRLH metamaterial leaky-wave

and resonant antennas,” IEEE Antennas Propag. Mag., vol. 50, no. 5, pp.

25–39, Oct. 2008.

[22] D. R. Jackson, C. Caloz, and T. Itoh, “Leaky-wave antennas,” Proc.

IEEE, vol. 100, no. 7, pp. 2194–2206, Jul. 2012.

[23] C. A. Balanis, Antenna Theory: Analysis and Design. New York: Wiley,

2005.

[24] K. M. Luk and B. Wu, “The magnetoelectric dipole—A wideband

antenna for base stations in mobile communications,” Proc. IEEE, vol.

100, no. 7, pp. 2297–2307, Jul. 2012.

[25] H. Gutton and G. Baissinot, “Flat aerial for ultra high frequencies,”

French Patent No. 703113, 1955.

[26] K. Carver and J. Mink, “Microstrip antenna technology,” IEEE Trans.

Antennas Propag., vol. 29, no. 1, pp. 2–24, Jan. 1981.

136
[27] J. Bahl and P. Bhartia, Microstrip Antennas, Dedham, MA, USA: Artech

House, 1982.

[28] L. J. Chu, “Physical limitations of omni-directional antennas”. J.

Applied Phys., vol. 19, 1163–1175, Dec. 1948.

[29] R. F. Harrington, “Effects of antenna size on gain, bandwidth, and

efficiency,” J. Research Nat Bur. Stand. D. Radio Propag., vol. 64D, no.

1, pp. 1–12. Jan. 1960.

[30] G. V. Eleftheriades and K. G. Balmain, Negative-Refraction

Metamaterials: Fundamental Principles and Applications. Hoboken, NJ,

USA: Wiley-IEEE Press, 2005.

[31] C. Caloz and T. Itoh, Electromagnetic Metamaterials Transmission Line

Theory and Microwave Applications. New York: Wiley, 2006.

[32] T. J. Cui, R. Liu, and D. R. Smith, Metamaterials: Theory, Design, and

Applications. New York: Springer Science, 2010.

[33] N. Engheta and R. W. Ziolkowski, Metamaterials Physics and

Engineering Explorations. Wiley-IEEE Press, 2006.

[34] D. H. Werner and D. H. Kwon, Transformation Electromagnetics and

Metamaterials. Springer, 2015.

[35] D. Lu and Z. Liu, “Hyperlenses and metalenses for far-field super-

resolution imaging,” Nat. Comm., vol. 3, no. 1205, pp. 1–9, Nov. 2012.

[36] R. Liu et al., “Broadband ground-plane cloak,” Science, vol. 323, no.

5912, pp. 366–369, Jan. 2009.

[37] M. Silveirinha and N. Engheta, “Tunneling of electromagnetic energy

through subwavelength channels and bends using ε-near-zero

materials,” Phys. Rev. Lett., vol. 97, no. 15403, pp. 1–4, Oct. 2006.

137
[38] K. Yao and X. Jiang, “Designing feasible optical devices via conformal

mapping,” JOSA B, vol. 28, no. 5, pp. 1037–1042, May 2011.

[39] J. B. Pendry, “Negative refraction makes a perfect lens”, Phys. Rev. Lett.,

vol. 85, no. 18, pp. 3966–3969, 2000.

[40] J. B. Pendry, D. Schurig, and D. R. Smith, “Controlling electromagnetic

fields,” Science, vol. 312, no. 5781, pp. 1780-1782, Jun. 2006.

[41] H. F. Ma and T. J. Cui, “Three-dimensional broadband and broad-angle

transformation-optics lens,” Nat. Comm., vol. 1, no. 124, pp. 1–7, Nov.

2010.

[42] H. Chen, C. T. Chan, and P. Sheng “Transformation optics and

metamaterials,” Nat. Mat., vol. 9, no. 5, pp. 387–395, Apr. 2010.

[43] D. Sievenpiper, L. Zhang, R. F. J. Broas, N. G. Alexopolous, and E.

Yablonovitch, “High-impedance electromagnetic surfaces with a

forbidden frequency band,” IEEE Trans. Microw. Theory Tech., vol. 47,

no. 11, pp. 2059–2074, Jan. 1999.

[44] C. L. Holloway, “An overview of the theory and applications of

metasurfaces: the two-dimensional equivalents of metamaterials,” IEEE

Antennas Propag. Mag., vol. 54, no. 2, pp. 10–35, Apr. 2012.

[45] M. Bosiljevac, M. Casaletti, F. Caminita, Z. Sipus, and S. Maci, “Non-

uniform metasurface Luneburg lens antenna design,” IEEE Trans.

Antennas Propag., vol. 60, no. 9, pp. 4065–4073, Sep. 2012.

[46] A. M. Patel and A. Grbic, “A printed leaky-wave antenna based on a

sinusoidally-modulated reactance surface,” IEEE Trans. Antennas

Propag., vol. 59, no. 6, pp. 2087–2096, Jun. 2011.

138
[47] T. Cai, G. M. Wang, X. L. Fu, J. G. Liang, and Y. Q. Zhuang, “High-

efficiency metasurface with polarization-dependent transmission and

reflection properties for both reflectarray and transmitarray,” IEEE

Trans. Antennas Propag., vol. 66, no. 6, pp. 3219–3224, Jun. 2018.

[48] F. Yang, R. Deng, S. Xu, and M. Li, “Design and experiment of a near-

zero-thickness high-gain transmit-reflect-array antenna using anisotropic

metasurface,” IEEE Trans. Antennas Propag., vol. 66, no. 6, pp. 2853–

2861, Jun. 2018.

[49] D. F. Sievenpiper, J. H. Schaffner, H. J. Song, R. Y. Loo, and G.

Tangonan, “Two-dimensional beam steering using an electrically

tunable impedance surface,” IEEE Trans. Antennas Propag, vol. 51, no.

10, pp. 2713–2722, Oct. 2003.

[50] D. González-Ovejero, G. Minatti, G. Chattopadhyay, and S. Maci,

“Multibeam by metasurface antennas,” IEEE Trans. Antennas Propag.,

vol. 65, no. 6, pp. 2923–2930, Jun. 2017.

[51] C. Pfeiffer and A. Grbic, “Millimeter-wave transmitarrays for wavefront

and polarization control,” IEEE Trans. Microw. Theory Tech., vol. 61,

no. 12, pp. 4407–4417, Dec. 2013.

[52] Y. Liu, K. Li, Y. Jia, Y. Hao, S. Gong, and Y. J. Guo, “Wideband RCS

reduction of a slot array antenna using polarization conversion

metasurfaces,” IEEE Trans. Antennas Propag., vol. 64, no. 1, pp. 326–

331, Jan. 2016.

[53] H. M. Bernety and A. B. Yakovlev, “Reduction of mutual coupling

between neighboring strip dipole antennas using confocal elliptical

139
metasurface cloaks,” IEEE Trans. Antennas Propag., vol. 63, no. 4, pp.

1554–1563, Apr. 2015.

[54] H. Mosallaei and K. Sarabandi, “Antenna miniaturization and bandwidth

enhancement using a reactive impedance substrate,” IEEE Trans.

Antennas Propag., vol. 52, no. 9, pp. 2403–2414, Sep. 2004.

[55] T. Yue, Z. H. Jiang, and D. H. Werner, “Compact, wideband antennas

enabled by interdigitated capacitor-loaded metasurfaces,” IEEE Trans.

Antennas Propag., vol. 64, no. 5, pp. 1595–1606, May 2016.

[56] R. Quarfoth and D. Sievenpiper, “Artificial tensor impedance surface

waveguides,” IEEE Trans. Antennas Propag., vol. 61, no. 7, pp. 3597–

3606, Jul. 2013.

[57] R. E. Collin, Field Theory of Guided Waves, Wiley-IEEE Press, 1990.

[58] J. C. Soric, A. Monti, A. Toscano, F. Bilotti, and A. Alù, “Dual-

polarized reduction of dipole antenna blockage using mantle cloaks,”

IEEE Trans. Antennas Propag., vol. 63, no. 11, pp. 4827–4834, Nov.

2015.

[59] F. Costa et al., “TE surface wave resonances on high-impedance surface

based antennas: analysis and modeling,” IEEE Trans. Antennas Propag.,

vol. 59, no. 10, pp. 3588–3596, Oct. 2011.

[60] M. Paquay, J. C. Iriarte, I. Ederra, R. Gonzalo, and P. de Maagt, “Thin

AMC structure for radar cross-section reduction,” IEEE Trans. Antennas

Propag., vol. 55, no. 12, pp. 3630–3638, Dec. 2007.

[61] A. Grbic, L. Jiang, and R. Merlin, “Near-field plates: subdiffraction

focusing with patterned surfaces,” Science, vol. 320, no. 5875, pp. 511–

513, Apr. 2008.

140
[62] A. Grbic, R. Merlin, E. M. Thomas, and M. F. Imani, “Near-field plates:

metamaterial surfaces/arrays for subwavelength focusing and

probing,” Proc. IEEE, vol. 99, no. 10, pp. 1806–1815, Oct. 2011.

[63] M. Ettorre and A. Grbic, “Generation of propagating Bessel beams using

leaky-wave modes,” IEEE Trans. Antennas Propag., vol. 60, no. 8, pp.

3605–3613, Aug. 2012.

[64] B. A. Munk, Frequency Selective Surfaces: Theory and Design, New

York: Wiley, 2000.

[65] R. Mittra, C. H. Chan, and T. Cwik, “Techniques for analyzing

frequency selective surfaces-a review,” Proc. IEEE, vol. 76, no. 12, pp.

1593–1615, Dec. 1988.

[66] Chen, S. W. Qu, B. J. Chen, X. Bai, K. B. Ng, and C. H. Chan,

“Terahertz metasurfaces for absorber or reflectarray applications,” IEEE

Trans. Antennas Propag., vol. 65, no. 1, pp. 234–241, Jan. 2017.

[67] M. Yazdi et al., “A bianisotropic metasurface with resonant asymmetric

absorption,” IEEE Trans. Antennas Propag., vol. 63, no. 7, pp. 3004–

3015, Jul. 2015.

[68] D. M. Pozar and S. D. Targonski, “A shared-aperture dual-band dual-

polarized microstrip array,” IEEE Trans. Antennas Propag., vol. 49, no.

2, pp. 150–157, Feb. 2001.

[69] S. A. Tretyakov, Analytical Modeling in Applied Electromagnetics,

Boston, MA: Artech House, 2003.

[70] R. G. Rojas and Z. Al-hekail, “Generalized impedance/resistive

boundary conditions for electromagnetic scattering problems,” Radio

Science, vol. 24, no. 1, pp. 1–12, Jan. 1989.

141
[71] K. W. Whites, “Electromagnetic scattering simulations using equivalent

boundary condition models,” Ph.D. dissertation, Graduate College,

University of Illinois at Urbana-Champaign, IL, USA, 1991.

[72] A. W. Glisson, “Electromagnetic scattering by arbitrarily shaped

surfaces with impedance boundary conditions,” Radio Science, vol. 27,

no. 6, pp. 935–943, Jan. 1992.

[73] B. Stufpfel and Y. Pion, “Impedance boundary conditions for finite

planar and curved frequency selective surfaces,” IEEE Trans. Antennas

Propag., vol. 53, no. 4, pp. 1415–1425, Apr. 2005.

[74] Z. G. Qian, W. C. Chew, and R. Suaya, “Generalized impedance

boundary condition for conductor modeling in surface integral equation,”

IEEE Trans. Microw. Theory Tech., vol. 55, no. 11, pp. 2354–2364, Nov.

2007.

[75] M. Albooyeh, D. H. Kwon, F. Capolino, and S. A. Tretyakov,

“Equivalent realizations of reciprocal metasurfaces: role of tangential

and normal polarization,” Phys. Rev. B, vol. 95, no. 11, Mar. 2017.

[76] X. C. Wang, A. Diaz-Rubio, and S. A. Tretyakov, “An accurate method

for measuring the sheet impedance of thin conductive films at

microwave and millimeter-wave frequencies,” IEEE Trans. Microw.

Theory Tech., vol. 65, no. 12, pp. 5009–5018, Dec. 2017.

[77] R. J. Garbacz, “Modal expansions for resonance scattering phenomena,”

Proc. IEEE, vol. 53, no. 8, pp. 856–864, Aug. 1965.

[78] R. F. Harrington and J. R. Mautz, “Theory of characteristic modes for

conducting bodies," IEEE Trans. Antennas Propag., vol. 19, no. 5, pp.

622–628, Sep. 1971.

142
[79] R. F. Harrington and J. R. Mautz, “Computation of characteristic modes

for conducting bodies,” IEEE Trans. Antennas Propag., vol. AP-19,

no.5, pp. 629–639, Sep. 1971.

[80] R. F. Harrington, Field Computation by Moment Methods, Wiley-IEEE

Press, 1993.

[81] R. F. Harrington, J. R. Mautz, and Y. Chang, “Characteristic modes for

dielectric and magnetic bodies,” IEEE Trans. Antennas Propag., vol. 20,

no. 2, pp. 194–198, Mar. 1972.

[82] Y. Chang and R. F. Harrington, “A surface formulation for characteristic

modes of material bodies,” IEEE Trans. Antennas Propag., vol. 25, no.

6, pp. 789–795, Nov. 1977.

[83] R. T. Maximidis, C. L. Zekios, T. N. Kaifas, E. E. Vafiadis, and G. A.

Kyriacou, “Characteristic mode analysis of composite metal-dielectric

structure, based on surface integral equation/moment method,” in Proc.

IEEE Eur. Conf. Antennas Propag. (EuCAP), Hague, Netherlands, Apr.

2014, pp. 2822–2826.

[84] Y. Chen, L. Guo, and S. Yang, “Mixed-potential integral equation based

characteristic mode analysis of microstrip antennas,” Int. J. Antennas

Propag., vol. 2016, pp. 1–8, Nov. 2016.

[85] L. Guo, Y. Chen, and S. Yang, “Generalized characteristic-mode

formulation for composite structures with arbitrarily metallic-dielectric

combinations”, IEEE Trans. Antennas Propag., vol. 66, no. 7, pp. 3556–

3566, Jul. 2018.

143
[86] R. F. Harrington and J. R. Mautz, “Control of radar scattering by

reactive loading,” IEEE Trans. Antennas Propag., vol. AP-20, no. 4, pp.

446–454, Jul. 1972.

[87] J. Mautz and R. Harrington, “Modal analysis of loaded N-port scatters,”

IEEE Trans. Antennas Propag., vol. AP-21, no. 2, pp. 188–199, Mar.

1973.

[88] R. F. Harrington and J. R. Mautz, “Pattern synthesis for loaded N-port

scatterers,” IEEE Trans. Antennas Propag., vol. AP-22, no. 2, pp. 184–

190, Mar. 1974.

[89] R. F. Harrington and J. Mautz, “Optimization of radar cross section of

N-port loaded scatterers,” IEEE Trans. Antennas Propag., vol. AP-22,

no. 5, pp. 697–701, Sep. 1974.

[90] R. J. Garbacz and D. M. Pozar, “Antenna shape synthesis using

characteristic modes,” IEEE Trans. Antennas Propag., vol. AP-30, no. 3,

pp. 340–350, May 1982.

[91] Y. Chen and C. F. Wang, Characteristics Modes Theory and

Applications in Antenna Engineering, Hoboken, New Jersey Wiley,

2015.

[92] Y. Chen and C. F. Wang, “Synthesis of reactively controlled antenna

arrays using characteristic modes and DE algorithm,” IEEE Antennas

Wireless Propag. Lett., vol. 11, pp. 385–388, Mar. 2012.

[93] C. Zhao and C. F. Wang, “Characteristic mode design of wide band

circularly polarized patch antenna consisting of H-shaped unit cells,”

IEEE Access, vol. 6, pp. 25292–25299, Apr. 2018.

144
[94] M. Cabedo-Fabres, E. Antonino-Daviu, A. Valero-Nogueira, and M. F.

Bataller, “The theory of characteristic modes revisited: a contribution to

the design of antennas for modern applications,” IEEE Antennas Propag.

Mag., vol. 49, no. 5, pp. 52–68, Oct. 2007.

[95] E. Safin and D. Manteuffel, “Manipulation of characteristic wave modes

by impedance loading,” IEEE Trans. Antennas Propag., vol. 63, no. 4,

pp. 1756–1764, Apr. 2015.

[96] E. Safin and D. Manteuffel, “Reconstruction of the characteristic modes

on an antenna based on the radiated far field,” IEEE Trans. Antennas

Propag., vol. 61, no. 6, pp. 2964–2971, Jun. 2013.

[97] D. Manteuffel and R. Martens, “Compact multimode multielement

antenna for indoor UWB massive MIMO,” IEEE Trans. Antennas

Propag., vol. 64, no. 7, pp. 2689–2697, Jul. 2016.

[98] R. Martens and D. Manteuffel, “Systematic design method of a mobile

multiple antenna system using the theory of characteristic modes,” IET

Microw., Antennas & Propag., vol. 8, no. 12, pp. 887–893, Sep. 2014.

[99] Z. Miers, H. Li, and B. K. Lau, “Design of bandwidth-enhanced and

multiband MIMO antennas using characteristic modes,” IEEE Antennas

Wireless Propag. Lett., vol. 12, pp. 1696–1699, Nov. 2013.

[100] H. Li, Z. Miers and B. K. Lau, “Design of orthogonal MIMO handset

antennas based on characteristic mode manipulation at frequency bands

below 1 GHz,” IEEE Trans. Antennas Propag., vol. 62, no. 5, pp. 2756–

2766, May 2014.

145
[101] Q. Wu, W. Su, Z. Li, and D. Su, “Reduction in out-of-band antenna

coupling using characteristic mode analysis,” IEEE Trans. Antennas

Propag., vol. 64, no. 7, pp. 2732–2742, Jul. 2016.

[102] P. Liang and Q. Wu, “Characteristic mode analysis of antenna mutual

coupling in the near-field,” IEEE Trans. Antennas Propag. vol. 66, no. 7,

pp. 3757–3762, Jul. 2018.

[103] F. A. Dicandia, S. Genovesi, and A. Monorchio, “Null-steering antenna

design using phase-shifted characteristic modes,” IEEE Trans. Antennas

Propag., vol. 64, no. 7, pp. 2698–2706, Jul. 2016.

[104] F. A. Dicandia, S. Genovesi, and A. Monorchio, “Advantageous

exploitation of characteristic modes analysis for the design of 3-D null-

scanning antennas,” IEEE Trans. Antennas Propag., vol. 65, no. 8, pp.

3924–3934, Aug. 2017.

[105] F. A. Dicandia, S. Genovesi, and A. Monorchio, “Efficient excitation of

characteristic modes for radiation pattern control by using a novel

balanced inductive coupling element,” IEEE Trans. Antennas Propag.,

vol. 66, no. 3, pp. 1102–1113, Mar. 2018.

[106] E. Antonino-Daviu, M. Cabedo-Fabrés, M. Sonkki, N. Mohamed

Mohamed-Hicho, and M. Ferrando-Bataller, “Design guidelines for the

excitation of characteristic modes in slotted planar structures,” IEEE

Trans. Antennas Propag., vol. 64, no. 12, pp. 5020–5029, Dec. 2016.

[107] N. M. Mohamed-Hicho, E. Antonino-Daviu, M. Cabedo-Fabrés, and M.

Ferrando-Bataller, “Designing slot antennas in finite platforms using

characteristic modes,” IEEE Access, Jun. 2018, in press.

146
[108] T. Y. Shih and N. Behdad, “Bandwidth enhancement of platform-

mounted HF antennas using the characteristic mode theory,” IEEE Trans.

Antennas Propag., vol. 64, no.7, pp. 2648–2659, Jul 2016.

[109] M. Li and N. Behdad, “Dual-band platform-mounted HF/VHF antenna

design using the characteristic mode theory,” IET Microw., Antennas &

Propag., vol. 12, no. 4, pp. 452–458, Mar. 2018.

[110] J. J. Adom, “Characteristic modes for impedance matching and

broadbanding of electrically small antennas,” Ph.D. dissertation,

University of Illinois at Urbana-Champaign, IL, USA, 2011.

[111] B. Yang and J. J. Adams, “Systematic shape optimization of symmetric

MIMO antennas using characteristic modes,” IEEE Trans. Antennas

Propag., vol. 64, no. 7, pp. 2668–2678, Jul. 2016.

[112] M. H. Rabah, D. Seetharamdoo, and M. Berbineau, “Analysis of

miniature metamaterial and magnetodielectric arbitrary-shaped patch

antennas using characteristic modes: evaluation of the Q factor”, IEEE

Trans. Antennas Propag., vol. 64, no. 7, Jul. 2016.

[113] S. Daniel, et al., “Analysis and design of a metamaterial lens antenna

using the theory of characteristic modes,” Int. J. Antennas Propag., vol.

2018, no. 6329531, pp. 1-9, Feb. 2018.

[114] C. Zhao, C.-F. Wang, “Characteristic mode design of wide band

circularly polarized patch antenna consisting of H-shaped unit cells,”

IEEE Access, vol. 6, pp. 25292-25299, Apr. 2018.

[115] X. Yang, Y. Liu, and S. Gong, “Design of a wideband omnidirectional

antenna with characteristic mode analysis,” IEEE Antennas Wireless

Propag. Lett., vol. 17, no. 6, pp. 993-997, Jun. 2018.

147
[116] Z. Liang, J. Ouyang, and F. Yang, “Design and characteristic mode

analysis of a low-profile wideband patch antenna using metasurface,” J.

Electromagnetic Waves and Applications, vol. 32, no. 17, pp. 2304-

2313, Aug. 2018.

[117] T. Li and Z. N. Chen, “Design of dual-band metasurface antenna array

using characteristic mode analysis (CMA) for 5G millimeter-wave

applications,” Proc. IEEE Antennas Propag. Wireless Commun.

(APWC), pp. 721-724, Sep. 2018.

[118] T. Li and Z. N. Chen, “A dual-band metasurface antenna using

characteristic mode analysis,” IEEE Trans. Antennas Propag., vol. 66,

no. 10, pp. 5620-5624, Oct. 2018.

[119] CST Computer Simulation Technology. CST MWS. [Online].

Available: http://www.cst.com/

[120] Altair. (2017) FEKO. Altair. [Online]. Available: www.feko.info

[121] Z. N. Chen, T. S. P. See, and X. Qing, “Cross-band mutual coupling of

monopole antennas on a finite-sized ground plane,” IEEE Trans.

Antennas Propag., vol. 61, no. 8, pp. 4372–4375, Aug. 2013.

[122] R. F. Harrington, “Theory of loaded scatterers,” Proc. Inst. Elec. Eng.,

vol. 111, no. 4, pp. 617–623, Apr. 1964.

[123] A. C. Gately, D. J. R. Stock, and B. R. Cheo, “A network description for

antenna problems,” Proc. IEEE, vol. 56, no. 7, pp. 1181–1193, Jul. 1968.

[124] W. Wasylkiwskyj and W. K. Kahn, “Scattering properties and mutual

coupling of antennas with prescribed radiation pattern,” IEEE Trans.

Antennas Propag., vol. 18, no. 6, pp. 741–752, Nov. 1970.

148
[125] J. F. A. Ormsby, “Antenna, load, and field effects on the bistatic

scattering patterns from a linear dipole array,” IEEE Trans. Antennas

Propag., vol. 27, no. 1, pp. 116–122, Jan. 1979.

[126] R. C. Hansen, “Relationships between antennas as scatterers and as

radiators,” Proc. IEEE, vol. 77, no. 5, pp. 659–662, May 1989.

[127] G. Zhai, Z. N. Chen, and X. Qing, “Mutual coupling reduction of a

closely spaced four-element MIMO antenna system using discrete

mushrooms,” IEEE Trans. Antennas Propag., vol. 64, no. 10, pp. 3060–

3067, Oct. 2016.

[128] W. H. Tu and K. Chang, “Wide-band microstrip-to-coplanar

stripline/slotline transitions,” IEEE Trans. Microw. Theory Tech., vol.

54, no. 3, pp. 1084–1089, Mar. 2006.

[129] D. M. Pozar, Microwave Engineering, 3rd Ed., New York: John Wiley

& Sons, 2005.

[130] T. B. A. Senior and J. L. Volakis, “Generalized impedance conditions in

scattering,” Proc. IEEE, vol. 79, no. 10, pp. 1413–1420, Oct. 1991.

[131] H. J. Bilow, “Guided waves on a planar tensor impedance surface,”

IEEE Trans. Antennas Propag., vol. 51, no. 10, pp. 2788–2792, Oct.

2003.

[132] O. Luukkonen et al, “Simple and accurate analytical model of planar

grids and high-impedance surfaces comprising metal strips or patches,”

IEEE Trans. Antennas Propag., vol. 56, no. 6, pp. 1624–1632, Jun.

2008.

[133] B. H. Fong, J. S. Colburn, J. J. Ottuch, J. L. Visher, and D. F.

Sievenpiper, “Scalar and tensor holographic artificial impedance

149
surfaces,” IEEE Trans. Antennas Propag., vol. 58, no. 10, pp. 3212–

3221, Jun. 2010.

[134] D. J. Gregoire and A. V. Kabakian, “Surface-wave waveguides,” IEEE

Antennas Wireless Propag. Lett., vol. 10, pp. 1512–1515, Dec. 2011.

[135] A. M. Patel and A. Grbic, “Modeling and analysis of printed-circuit

tensor impedance surfaces,” IEEE Trans. Antennas Propag., vol. 61, no.

1, pp. 211–220, Jan. 2013.

[136] A. M. Patel and A. Grbic, “Effective surface impedance of a printed-

circuit tensor impedance surface (PCTIS),” IEEE Trans. Antennas

Propag., vol. 61, no.7, pp. 1403–1413, Mar. 2013.

[137] F. Elek, B. B. Tierney, and A. Grbic, “Synthesis of tensor impedance

surfaces to control phase and power flow of guided waves,” IEEE Trans.

Antennas Propag., vol. 63, no. 8, pp. 3956–3962, Sep. 2015.

[138] R. G. Quarfoth and D. F. Sievenpiper, “Nonscattering waveguides based

on tensor impedance surfaces,” IEEE Trans. Antennas Propag., vol. 63,

no. 4, pp. 1746–1755, Apr. 2015.

[139] M. Li, S. Xiao, J. Long, and D. F. Sievenpiper, “Surface waveguides

supporting both TM mode and TE mode with the same phase velocity,”

IEEE Trans. Antennas Propag., vol. 64, no. 9, pp. 3811–3819, Aug.

2016.

[140] M. Teniou, H. Roussel, N. Capet, G. P. Piau, and M. Casaletti,

“Implementation of radiating aperture field distribution using tensorial

metasurfaces,” IEEE Trans. Antennas Propag., vol. 65, no. 11, pp.

5895–5907, Aug. 2017.

150
[141] C. S. Lee, S. W. Lee, and S. L. Chuang, “Plot of modal field distribution

in rectangular and circular waveguides,” IEEE Microw. Theory Tech.,

vol. 33, no. 3, pp. 271–274, Mar. 1985.

[142] H. Kang and S. O. Park, “Mushroom meta-material based substrate

integrated waveguide cavity backed slot antenna with broadband and

reduced back radiation,” IET Microw. Antennas Propag., vol. 10, no. 14,

pp. 1598–1603, Nov. 2016.

[143] F. Croq and A. Papiernik, “Stacked slot-coupled printed antenna,” IEEE

Microw. Guided Wave Lett., vol. 1, no. 10, pp. 288–290, Oct. 1991.

[144] D. Sun and L. You, “A broadband impedance matching method for

proximity-coupled microstrip antenna”, IEEE Trans. Antennas Propag.,

vol. 58, no. 4, pp. 1392–1397, Apr. 2010.

[145] M. Caillet, M. Clénet, A. Sharaiha, and Y. M. M. Antar, “A compact

wide-band rat-race hybrid using microstrip lines,” IEEE Microw.

Wireless Comp. Lett., vol. 19, no.4, pp. 191–193, Apr. 2009.

151

You might also like