You are on page 1of 182

RADIO FREQUENCY MINIATURE PASSIVE DEVICES: FERRITE ANTENNAS AND

INDUCTORS

by

JAEJIN LEE

YANG-KI HONG, COMMITTEE CHAIR

SHUHUI LI
SEONGSIN M. KIM
SUBHADRA GUPTA
GARY J. MANKEY

A DISSERTATION

Submitted in partial fulfillment of the requirements


for the degree of Doctor of Philosophy
in the Department of Electrical and Computer Engineering
in the Graduate School of
The University of Alabama

TUSCALOOSA, ALABAMA

2015
Copyright Jaejin Lee 2015
ALL RIGHTS RESERVED
ABSTRACT

Recent advances in mobile electronic devices demand the integration of more functions

and a further decrease in the overall device size. Device miniaturization and compact design,

therefore, have become important factors. Radio frequency (RF) passive devices, including

antennas and inductors, are key components in mobile telecommunication systems. The

interaction of magnetization dynamics with electromagnetic waves in the RF devices allows

electromagnetic radiation, wave propagation, and RF signal processing. Ferrites possess both

frequency-dependent permeability μr and permittivity εr. Thus, miniaturization and

electromagnetic scaling of antennas and inductors can be realized with the magnetic and

dielectric properties of ferrites. In addition, the unique features of ferrites, such as high electrical

resistivity, excellent chemical stability, and mechanical hardness, provide a low eddy current loss

at high frequencies and a variety of applications in power and microwave systems. In designing

ferrite passive devices, the frequency-dependent permeability and magnetic loss of ferrites play

important roles in device form factor and efficiency.

This dissertation focuses on the magnetic loss mechanism of ferrites, miniaturization, and

performance improvement of RF antennas and inductors. In this dissertation, the effectiveness of

small relative permeability in improvement of gigahertz (GHz) antenna performance is

investigated. In addition, the magnetic loss reduction of ferrites is described using a different

approach from material processing aspects. Both the relative permeability higher than unity and

the low-magnetic loss of ferrites are presented with a small direct current (DC) magnetic field,

ii
therefore, antenna miniaturization and gain improvement are achieved. A high-isolation multi-

input multi-output (MIMO) antenna array design is demonstrated with miniature ferrite antenna

elements for long-term evolution (LTE) mobile technology. Furthermore, micron-thick ferrite

film deposition process is developed using DC magnetron sputtering for high inductance and

quality-factor ferrite inductors. The measured and simulated results suggest that high resistivity

and large magnetic-anisotropy ferrites in combination with high-efficiency RF device design can

lead to the development of compact high-performance antennas and inductors.

iii
DEDICATION

This dissertation is dedicated to everyone who helped me and guided me through the

trials and tribulations of creating this manuscript. In particular, my advisor, Professor Yang-Ki

Hong and MMDL (magnetic materials and device laboratory) members who stood by me

throughout the time taken to complete this masterpiece.

iv
LIST OF ABBREVIATIONS AND SYMBOLS

RF Radio frequency

fFMR Ferromagnetic resonance frequency

μr Relative permeability

GHz Gigahertz

MIMO Multi-input multi-output

m Magnetic moment

S Spin angular momentum

γ Gyromagnetic ratio (   m / s )

T Torque

M Total magnetization

N Number of magnetic dipoles per unit volume

AC Alternating current

DC Direct current

H0 DC magnetic bias field

H AC magnetic field

Ht Total magnetic field

Mt Total magnetization

Ms DC saturation magnetization

0 Precession frequency ( 0  0H0 )

v
m Magnetization frequency ( m  0Ms )

ejωt Time-harmonic dependent term

  Tensor susceptibility

  Tensor permeability

B Magnetic flux

µ Diagonal term of tensor permeability

κ Off-diagonal term of tensor permeability

α Damping factor of spin rotation

ΔH Ferromagnetic resonance linewidth

χd′ and χd′′ Real and imaginary part of susceptibility of the domain wall motion

χd0 Static magnetic susceptibility of the domain wall motion

ωd (= 2πfd) Domian wall resonance frequency

β Damping factor of domain wall motion

χs′ and χs′′ Real and imaginary part of susceptibility of the spin rotation

χs0 Static magnetic susceptibility of the spin rotation

ωs (= 2πfs) Spin resonance frequency

Ms Saturation magnetization

D Grain size

γ′ Wall energy

Ha Magnetic anisotropy field

Hd Demagnetizing field

μ′ Real part of the permeability

μ″ Imaginary part of the permeability

vi
ωs_α High α spin resonance frequency

tan δμ Magnetic loss tangent (tan δμ = μ″/μ′)

tan δh Hysteresis loss

tan δe Eddy current loss

tan δr Residual loss

tan δd Magnetic loss tangent of the domain wall motion

tan δs Magnetic loss tangent of the spin rotation

RHCP Right-hand circularly polarization

 Effective permeability for the RHCP field

LHCP Left-hand circularly polarization

 Effective permeability for the LHCP field

E Electric field

ε Dielectric constant

β Propagation constant

Y+ Wave admittance for the plane wave propagation in a magnetized ferrite with β+

Y− Wave admittance for the plane wave propagation in a magnetized ferrite with β−

λ0 Wavelength of an electromagnetic wave in vacuum

c Speed of light in vacuum (3 × 108 m/s)

f Frequency of the wave in Hertz (Hz)

λeff Effective wavelength in a medium

n Refractive index of the medium ( n  r r )

BW Bandwidth

Q Quality factor

vii
ω Angular frequency

WT Total stored energy

Prad Radiated power

Ploss Dissipated power loss

kg Wave number in a medium (kg = k0(μrεr)0.5)

k0 Wave number in a free-space

PD Dielectric power loss

ε0 Permittivity of free-space (= 8.85 × 10-12 F/m)

tan δε Dielectric loss tangent (tan δε = ε″/ε′)

PM Magnetic power loss

Gr Radiation conductance of the patch antenna

Zg Internal impedance of a generator

ZA Impedance of a transmitting antenna

Rg Resistance of generator impedance

Xg Reactance of generator impedance

RA Antenna resistance (RA = Rr + RL)

Rr Radiation resistance of the antenna

RL Loss resistance of the antenna

Pr Radiated power for the antenna

Vg Peak generator voltage

Ig Generated current in the equivalent circuit

PL Power loss of the antenna

η Radiation efficiency of the antenna

QT Total quality factor of the antenna

viii
Qr Radiated quality factor of the antenna

Qμ Quality factor due to magnetic loss (= 1/tan δμ)

Qε Quality factor due to dielectric loss (= 1/tan δε)

Ei Incident electric field

Hi Incident magnetic field

η0 Wave impedance of free-space

Er Reflected electric field

Hr Reflected magnetic field

Γ Reflection coefficient of the electric field

T Transmission coefficient of the electric field

Et Transmitted electric field

Ht Transmitted magnetic field

ZL Load impedance of a lossy transmission line

V 0 Amplitude of the incident voltage

V 0 Amplitude of the reflected voltage

Z0 Characteristic impedance of the transmission line

Zin Input impedance of the transmission line at a distance from the load

L Inductance

Φ Magnetic flux

LAC Air-core inductance for an inductor

LMC Magnetic-core inductance for an inductor

μeff Effective permeability depending on magnetic film geometry

EMagnetic Magnetic energy stored in an inductor

ix
EElectric Electric energy stored in an inductor

ELoss Energy loss per cycle dissipated by a resistor

QMC Quality factor for a magnetic-core inductor

Ctotal Total capacitance of the parallel capacitors

Cp Parasitic capacitance in the substrate

Cs Spiral capacitance between the coils and underpass of the center pad

Rp Parasitic resistance in the substrate

Rs Series resistance of the conductor coils

Hθ Out-of-plane anisotropy field

Hϕ In-plane anisotropy field

Co2Z Ba3Co2Fe24O41 hexaferrite

Co2Y Ba2Co2Fe12O22 hexaferrite

BaM BaFe12O19 hexaferrite

HFSS High frequency structure simulator

VSWR Voltage standing wave ratio

FBW Fractional bandwidth

RE Radiation efficiency

EM Electromagnetic

CCL Copper clad laminate

3D 3-dimensional

Co/Ti-BaM BaFe9.6Co1.2Ti1.2O19 hexaferrite

tan δDW Magnetic loss tangent of domain wall motion

tan δsp Magnetic loss tangent of spin rotation

FMR Ferromagnetic resonance

x
Ni-Mn-Co Ni0.5Mn0.2Co0.07Fe2.23O4 ferrite

S-parameters Scattering parameters

RL Return loss

LTE Long term evolution

ρe Correlation coefficient of the antenna

PLD Pulsed laser deposition

Cu Copper

Si Silicon

Ni-Zn-Cu Ni0.5Zn0.3Cu0.2 Fe2O4 ferrite

SEM Scanning electron microscopy

Ms Saturation magnetization of a magnetic film (emu/cm3)

Hc Coercivity of a magnetic film

RFICs Radio frequency integrated circuits

MEMS Microelectromechanical systems

Zn-Co-Ni Zn0.13Co0.04Ni0.63Fe2.2O4 ferrite

σ Conductivity (S/m)

PR Photoresist

IBE Ion beam etching

SOLT Short-open-load-through

σs Saturation magnetization of a magnetic powder (emu/g)

Hk Magnetic anisotropy field of a magnetic film

BaZn2 W BaZn2 Fe10O27

Co2X Ba2Co2Fe28O46

xi
ACKNOWLEDGMENTS

I would like to express my sincere gratitude and appreciation to my advisor, Professor

Yang-Ki Hong. He has been giving me all of his professional knowledge, which is not only

academic advice, but also advice for life, career paths, and planning for the future. I respect him

and his principle so much and will keep it along with my future career.

I would also like to thank all of my committee members, Dr. Shuhui Li, Dr. Seongsin M.

Kim, Dr. Subhadra Gupta, and Dr. Gary Mankey for their invaluable input and inspiring

questions of both the dissertation and my academic progress.

This dissertation would have been possible with the support of my fellow graduate

students, Gavin Abo, Jeevan Jalli, Jihoon Park, and Woncheol Lee and of course of my family

who never stopped encouraging me to persist.

xii
CONTENTS

ABSTRACT ............................................................................................ ii

DEDICATION ........................................................................................iv

LIST OF ABBREVIATIONS AND SYMBOLS....................................... v

ACKNOWLEDGMENTS ...................................................................... xii

LIST OF TABLES ...............................................................................xvii

LIST OF FIGURES ............................................................................ xviii

1. INTRODUCTION ................................................................................ 1

1.1. Motivation .................................................................................... 1

1.2. Ferromagnetic resonance frequency and magnetic permeability..... 2

1.3. Outline of dissertation ................................................................... 4

References ........................................................................................... 6

2. FUNDMENTAL PROPERTIES OF FERRITES .................................. 8

2.1. Permeability tensor........................................................................ 8

2.2. Magnetization processes and permeability dispersion .................. 17

2.3. Magnetic loss of polycrystalline ferrites ...................................... 22

2.4. Interaction of AC field with permeability tensor .......................... 25

2.5. Plane wave propagation in magnetized ferrites ............................ 30

References ......................................................................................... 36

3. ROLES OF FERRITES IN RF ANTENNAS ...................................... 37

3.1. Introduction to ferrite antennas .................................................... 37

3.2. Effective wavelength and antenna miniaturization ....................... 38

3.3. Bandwidth of ferrite antennas ...................................................... 41

xiii
3.4. Radiation efficiency of ferrite antennas ....................................... 47

3.5. Reflection loss and impedance matching to free-space ................ 50

References ......................................................................................... 58

4. FUNDAMENTALS OF FERRITE INDUCTORS .............................. 59

4.1. Introduction to ferrite inductors ................................................... 59

4.2. Inductance ................................................................................... 60

4.3. Quality factor .............................................................................. 62

4.4. Design of integrated ferrite inductors........................................... 66

References ......................................................................................... 69

5. SMALL PERMEABILITY IN ANTENNA PERFORMANCE ........... 71

5.1. Introduction................................................................................. 71

5.2. Hexaferrite substrate ................................................................... 72

5.3. Simulation of GHz hexaferrite chip antenna ................................ 73

5.4. Fabrication and measurement results ........................................... 78

5.5. Conclusion .................................................................................. 80

References ......................................................................................... 82

6. CONTROL OF MAGNETIC LOSS OF FERRITES ........................... 83

6.1. Introduction................................................................................. 83

6.2. Experiment .................................................................................. 84

6.3. Experimental results .................................................................... 85

6.4. Discussion ................................................................................... 89

6.5. Antenna simulation ..................................................................... 93

6.6. Conclusion .................................................................................. 96

xiv
References ......................................................................................... 98

7. FERRITE ANTENNA GAIN WITH DC MAGNETIC FIELD ......... 100

7.1. Introduction............................................................................... 100

7.2. Antenna design and fabrication ................................................. 101

7.3. Results and discussion ............................................................... 103

7.3.1. Magnetic-field dependence of magnetic loss of Ni-Mn-Co


ferrite ............................................................................ 103

7.3.2. Magnetic-field dependence of Ni-Mn-Co ferrite antenna


performance .................................................................. 104

7.4. Conclusion ................................................................................ 108

References ....................................................................................... 109

8. MINIATURE LTE MIMO FERRITE ANTENNA............................ 111

8.1. Introduction............................................................................... 111

8.2. Design of LTE MIMO ferrite antenna ....................................... 111

8.3. Fabrication and measurement results ......................................... 114

8.4. Conclusion ................................................................................ 120

References ....................................................................................... 121

9. INTEGRATED FERRITE INDUCTOR FOR POWER SYSTEM-ON-


CHIP ................................................................................................ 122

9.1. Introduction............................................................................... 122

9.2. Experiment ................................................................................ 122

9.2.1. Design and fabrication ..................................................... 122

9.2.2. Measurement ................................................................... 124

9.3. Results and discussion ............................................................... 125

9.4. Conclusion ................................................................................ 131

xv
References ....................................................................................... 132

10. GIGAHERTZ (GHZ) FERRITE INDUCTOR FOR RFIC .............. 133

10.1. Introduction............................................................................. 133

10.2. Experiment .............................................................................. 134

10.2.1 Ferrite inductor design and fabrication ............................ 134

10.2.2 Preparation of high magnetic anisotropy Zn-Co-Ni ferrite


film and Co2Z ferrite particles........................................ 137

10.3. Results and discussion ............................................................. 138

10.3.1. Characteristics of the Zn-Co-Ni ferrite film and Co2Z


ferrite particles ............................................................. 138

10.3.2. Performance of high anisotropy ferrite film inductor...... 143

10.3.3. Dependence of dynamic magnetic properties of ferrites on


L and Q factor ............................................................... 145

10.3.4. Comparison of high anisotropy ferrite inductor to previous


integrated inductors ....................................................... 146

10.4. Conclusion .............................................................................. 148

References ....................................................................................... 150

11. CONCLUSION .............................................................................. 152

12. FUTURE WORK ........................................................................... 156

APPENDIX .......................................................................................... 158

xvi
LIST OF TABLES

3.1 Frequency bands for various wireless communications ..................... 38

5.1 Simulated resonance frequency, VSWR, and fractional bandwidth for


hexaferrite (tan δμ = 0.05 and tan δε = 0.008) and dielectric antennas . 75

6.1 Measured µ′ and tan δµ for Co/Ti-substituted BaM with various


applied magnetic fields ..................................................................... 89

6.2 Curve-fitted parameters for measured complex permeability of Co/Ti-


substituted BaM with various magnetic fields ................................... 91

6.3 Simulated hexaferrite antenna performance at various applied


magnetic fields ................................................................................ 95

7.1 Measured permeability and magnetic tangent loss of Ni-Mn-Co ferrite


at various DC magnetic fields ........................................................ 104

7.2 Measured resonant frequency, return loss and 2D gain of Ni-Mn-Co


ferrite antenna ................................................................................ 108

8.1 Summary of LTE ferrite antenna characteristics .............................. 118

8.2 Comparison of performance for LTE MIMO antennas .................... 119

9.1 Design parameters of ferrite inductor .............................................. 123

9.2 Summary of the inductance, Q-factor, and superimposed DC current


for the fabricated inductors ............................................................. 129

10.1 Measured characteristics of the fabricated air-core and ferrite


inductors at 2 GHz ................................................................. 148

10.2 Performance comparison of various integrated inductors............... 148

xvii
LIST OF FIGURES

1.1 Ferromagnetic resonance (FMR) frequency with various anisotropy


fields (Ha) ........................................................................................... 2

1.2 Initial permeability as a function of sintered density ............................ 4

2.1 Schematic illustration of motion of a magnetic dipole moment in a


magnetic field ..................................................................................... 8

2.2 Total magnetization and precession of magnetic dipole moments


within a ferrite by bias magnetic field along the z-axis ........................ 9

2.3 Magnetization processes of a polycrystalline ferrite consist of domain


wall motion and spin rotation ............................................................ 17

2.4 Frequency dispersion of complex susceptibility of (a) domain wall


motion (resonance-type) and (b) spin rotation (relaxation-type) ........ 22

2.5 Three main components of magnetic loss of a ferrite: hysteresis loss,


eddy current loss, and residual loss ................................................... 23

2.6 Illustration of induced magnetic flux of ferrite with interaction of a


time varying magnetic field............................................................... 23

2.7 Complex effective permeability for right hand circularly polarized


(RHCP) and left hand circularly polarized (LHCP) fields as a function
of the gyromagnetic frequency ω0. .................................................... 29

3.1 Schematic illustration of the effective wavelength λeff in a ferrite


medium in comparison to the wavelength λ0 in vacuum .................... 39

3.2 Miniaturization of monopole antenna. (a) Air-core and (b) ferrite


monopole antenna ............................................................................. 40

3.3 Geometry of rectangular microstrip patch antenna with a ferrite


substrate possessing both relative permeability and permittivity higher
than unity.......................................................................................... 42

3.4 Transmitting antenna and its equivalent circuit. The radiated waves are
in the form of spherical wave in near-field region and become a plane
wave in far-field region ..................................................................... 48

xviii
3.5 Illustration of incident, reflected, and transmitted electric fields at the
interface of two different characteristic impedances .......................... 51

3.6 Metal-backed ferrite structure and its transmission line model


terminated with a short circuit ........................................................... 54

3.7 A lossy transmission line terminated in a load impedance ZL............. 54

4.1 Various ferrites for RF applications at wide operating frequency ...... 60

4.2 Series RLC circuit for a simple inductor model ................................. 62

4.3 Equivalent circuit of an integrated spiral inductor ............................. 64

4.4 Various ferrite-core constructions for integrated RF inductors: (a) air-


core, (b) one-side ferrite (bottom), (c) two-side ferrite (top and
bottom), and (d) closed ferrite structure............................................. 66

4.5 Simulated magnetic-field distributions of ferrite inductor with various


ferrite-core constructions: (a) air-core, (b) one-side ferrite (bottom), (c)
two-side ferrite (top and bottom), and (d) closed ferrite structure in
accordance with Figure 4.4 ............................................................... 67

4.6 Simulated inductance of the ferrite inductors with various ferrite-core


constructions in accordance with Figure 4.4 ...................................... 67

5.1 Measured complex permeability and permittivity of Co 2Z hexaferrite-


glass composite ................................................................................. 73

5.2 Designed antenna geometry and structure of hexaferrite-glass


composite chip antenna ..................................................................... 74

5.3 Simulated VSWR of the designed hexaferrite and dielectric antennas75

5.4 Simulated radiation efficiency of the hexaferrite (closed symbol) and


dielectric (open symbol) antennas in terms of magnetic or dielectric
losses ................................................................................................ 77

5.5 (a) Fabricated Co2Z hexaferrite-glass composite chip antenna and (b)
measured and simulated return losses. ............................................... 79

5.6 Measured (closed symbol) and simulated (open symbol) 3-dimensional


peak- and average-gains and radiation efficiency of the fabricated
hexaferrite composite antenna ........................................................... 80

6.1 (a) Measured magnetization virgin curve and hysteresis loop (inset) of
BaFe9.6Co1.2Ti1.2O19 hexaferrite and (b) illustration of magnetization
processes .......................................................................................... 86

xix
6.2 Measured and curve-fitted complex permeability spectra of
BaFe9.6Co1.2Ti1.2O19 hexaferrite at various applied magnetic fields: (a)
real and (b) imaginary part of permeability ....................................... 88

6.3 Calculated complex susceptibility of (a) domain wall motion and (b)
spin rotation for various DC magnetic fields ..................................... 91

6.4 Magnetic-field dependence of (a) real part of susceptibility and (b)


magnetic loss tangent of domain wall motion and spin rotation ......... 92

6.5 Magnetic-field dependence of resonance frequency of domain wall


motion and spin rotation ................................................................... 93

6.6 Designed ferrite helical antenna for performance simulation ............. 94

6.7 Simulated radiation efficiency and 3-dimensional peak gain of the


designed ferrite helical antenna as a function of applied magnetic field
(i.e. magnetic loss tangent) ................................................................ 95

7.1 Designed geometry and dimension of Ni-Mn-Co ferrite antenna ..... 101

7.2 Fabricated Ni-Mn-Co ferrite antenna with Nd-Fe-B permanent


magnets .......................................................................................... 102

7.3 Measurement system of permeability spectra at various DC magnetic


fields ............................................................................................... 103

7.4 Measured magnetic-field dependent magnetic loss tangent (tan δμ) and
real part of permeability (µr) of Ni-Mn-Co ferrite at 890 MHz ........ 104

7.5 Measured and simulated scattering parameters of Ni-Mn-Co ferrite


antenna with and without applied DC magnetic field....................... 105

7.6 Measured magnetic-field dependent 2D gain of Ni-Mn-Co ferrite


antenna ........................................................................................... 106

7.7 Normalized measured and simulated radiation patterns (in yz plane) at


900 MHz for Ni-Mn-Co ferrite antenna without and with magnetic
field. (a) co-polarization and (b) cross-polarization ......................... 107

8.1 Geometry and dimensions of simulated LTE MIMO ferrite antenna 112

8.2 (a) Simulated return loss (S11) and (b) isolation (S21) of LTE MIMO
ferrite antenna for various separations (d). Material parameters used in
the simulation: μ = 3.7; tan δμ = 0.08; ε = 7.8; tan δε = 0.01 ............. 113

8.3 (a) MIMO antenna structure and (b) photograph of the fabricated LTE
MIMO ferrite antenna ..................................................................... 115

xx
8.4 (a) Measured S-parameters and (b) calculated correlation coefficient of
the fabricated LTE MIMO ferrite antenna ....................................... 116

8.5 Measured 3-dimensional average and peak gain of the fabricated LTE
MIMO ferrite antenna: (a) antenna 1 and (b) antenna 2. Experimental
property of Ni0.5Mn0.2Co0.07Fe2.23O4 ferrite: μ = 3.7; tan δμ = 0.2; ε =
7.8; tan δε = 0.05. Separation distance = 41 mm. ............................. 117

8.6 Measured 3D radiation patterns of the fabricated LTE MIMO ferrite


antenna at 746 MHz: (a) antenna 1 and (b) antenna 2 ...................... 118

8.7 Simulated peak gain (antenna 1) of the LTE MIMO ferrite antenna at
746 MHz as a function of magnetic loss .......................................... 119

9.1 Sputter deposition process of Ni0.5Zn0.3Cu0.2 Fe2O4 ferrite film.


Multilayered oxide films are deposited by DC magnetron sputtering
using Ni, Fe, Zn, and Cu metal targets in a mixture of Ar and O2 gases
....................................................................................................... 124

9.2 Fabrication process for ferrite inductor array on 4 inch Si wafer...... 125

9.3 (a) Magnetic hysteresis loop and (b) cross-section of the 1 μm and 2.5
μm Ni0.38Zn0.31Cu0.27Fe2O3.81 film ................................................... 126

9.4 The fabricated ferrite inductors with 2.5, 3.5, and 4.5 turns of Cu coil
and cross-sectional image of the ferrite inductor with 65 μm thick Cu
coil. ................................................................................................ 127

9.5 (a) Inductance and (b) quality factor of 1 μm, 2.5 μm thick ferrite and
air-core inductors ............................................................................ 128

9.6 Superimposed DC current characteristic of 1 μm and 2.5 μm thick


ferrite inductors at 10 MHz. L0 represents the inductance at zero
superimposed DC current. ............................................................... 129

9.7 Circuit diagram of the buck DC-DC converter for the calculation of
power efficiency ............................................................................. 130

9.8 Calculated power efficiency of the DC-DC converter based on the


studied ferrite inductors as a function of load current ...................... 130

10.1 (a) Designed one-turn spiral inductor and (b) optical microscope
image of the fabricated air-core inductor ....................................... 135

10.2 Photo-images of (a) double-side ferrite inductor (bottom Zn-Co-Ni


ferrite film + top Co2Z ferrite particles), (b) bottom-type Zn-Co-Ni
ferrite film inductor, (c) short de-embedding, and (d) open de-
embedding structures ..................................................................... 136

xxi
10.3 X-ray diffraction pattern of the fabricated Zn-Co-Ni ferrite film on
silicon substrate ............................................................................. 138

10.4 Scanning electron microscope (SEM) cross-sectional image of the


Zn-Co-Ni ferrite film. .................................................................... 139

10.5 Measured magnetic hysteresis loop of the Zn-Co-Ni ferrite film ... 139

10.6 Measured magnetic hysteresis loops of the non-milled and 10-hour


milled Co2Z ferrite particles .......................................................... 140

10.7 Calculated complex permeability of Zn-Co-Ni ferrite film using the


experimental parameters Ms = 334 emu/cm3, Hk = 420 Oe, α = 0.15,
and designed parameters Ms = 430 emu/cm3, Hk = 520 Oe, α = 0.05141

10.8 Measured complex permeability of the non-milled and 10-hour


milled Co2Z ferrites ....................................................................... 142

10.9 Measured and simulated (a) L and (b) Q factor of the air-core and
bottom-type ferrite film inductors .................................................. 144

10.10 Measured (a) L and (b) Q factor of the air-core, Zn-Co-Ni ferrite
(bottom), Zn-Co-Ni ferrite (bottom) + non-milled Co2Z ferrite (top),
and Zn-Co-Ni ferrite (bottom) + 10-hour milled Co2Z ferrite (top)
inductors ....................................................................................... 147

12.1 Permeability of samples of Co2Z and CoZnZ with different degrees


of crystal orientation...................................................................... 156

xxii
CHAPTER 1

INTRODUCTION

1.1. Motivation

The next generation of mobile network technology requires small form factor of

telecommunication systems. Accordingly, miniature radio frequency (RF) passive devices, such

as antennas, isolators, circulators, and inductors, are needed, while maintaining high performance.

The passive devices in RF front-end tranceiver enable fundamental functions of electromagnetic

radiation, directional wave propagation, and RF signal processing. The functions arise from

electromagnetic wave interaction with magnetization dynamics. Thus, theoretical understanding

of the role of magnetism and magnetic materials in RF passive devices is of great importance for

advancement of device design and performance. High-frequency ferrites have been widely

investigated and applied in antennas [1−8], circulators [9−15], isolators [16−23], and inductors

[24−30]. Ferrites are electrically insulating magnetic materials and possess high magnetic

permeability, high electrical resistivity, and moderate dielectric permittivity [31]. In addition,

ferrites show excellent chemical stability, good mechanical hardness, and low eddy current loss

at high frequencies [32]. These magnetic and electrical properties of ferrites allow a variety of

applications in power electronics and microwave communication systems.

Ferrites are classified into cubic and hexagonal ferrites according to the crystalline

structure. The cubic ferrites include garnets and spinels and exhibit highly symmetric structures,

resulting in a low magnetic anisotropy field. On the other hand, the hexagonal ferrites hold a

magnetoplumbite structure with various metal and oxygen stoichiometries and possess a high

1
magnetic anisotropy field [33, 34]. Different intrinsic magnetic properties of ferrites, therefore,

contribute to various RF device applications from very low to microwave frequencies. Dynamic

magnetic properties of ferrites are governed by Snoek’s limit [35]. This limit describes that a

ferrite permeability linearly decreases with increasing ferromagnetic resonance frequency (fFMR).

The fFMR is dependent on the anisotropy field, demagnetization field, magnetization, and bias

magnetic field [36]. Thus, physical geometries, chemical compositions, and microstructures of

ferrites make a significant impact on relative permeability, magnetic loss tangent, ferromagnetic

resonance frequency, and, consequently, RF device performances.

1.2. Ferromagnetic resonance frequency and magnetic permeability

Ferromagnetic resonance frequency of ferrites can be expressed by Eq. (1.1).

fFMR = (2.8MHz/Oe) × (Ho + Ha), (1.1)

where H0 is the applied DC magnetic field and Ha is the magnetic anisotropy field of a ferrite.

Accordingly, the fFMR is proportional to Ha and H0. Fig. 1.1 shows Ha dependence of fFMR at

FIG. 1.1. Ferromagnetic resonance (FMR) frequency with various anisotropy fields (Ha).

2
various H0. M-type hexaferrite (BaM: BaFe12O19) has about 17 kOe of Ha, therefore, 45 GHz of

self-biased fFMR as seen in Fig. 1.1. Also, W-type (BaZn2 W: BaZn2Fe10O27) and X-type (Co2X:

Ba2Co2Fe28O46) hexaferrites have high Ha of 10 kOe and 9.5 kOe, respectively. Their

corresponding ferromagnetic resonance frequencies are 28 GHz and 26.6 GHz. On the other

hand, Z-type (Co2Z: Ba3Co2Fe24O41) hexaferrite has in-plane anisotropy with Hϕ of 112 Oe and

Hθ of 13 kOe [37] and a fFMR of about 3.4 GHz according to fFMR = (2.8 MHz/Oe) × (Hϕ × Hθ)0.5.

Thus, the hexagonal ferrites are suitable for GHz device applications due to their high magnetic

anisotropy fields.

The initial permeability (μi) for polycrystalline ferrites is given by Eq. (1.2).

(μi – 1)/4π = (1/3) (Ms/Hθ +Ms/Hϕ), (1.2)

where Ms is the saturation magnetization, Hθ is the out-of-plane anisotropy, and Hϕ is the in-

plane anisotropy. The initial permeability is proportional to Ms and inversely proportional to Ha.

For the hexagonal ferrites having the c-axis as the prefered direction of magnetization, the initial

permeability can be written as Eq. (1.3) due to Hθ = Hϕ = Ha.

(μi – 1)/4π = (2/3) (Ms/Ha). (1.3)

For the hexagonal ferrites with a preferred plane of magnetization, however, the in-plane

anisotropy field (H ) is often a factor of 1000 smaller than out-of-plane anisotropy field (Hθ),

which is Hθ ≪ H . Accordingly, the initial permeability can be written by Eq. (1.4).

(μi – 1)/4π = (1/3) (Ms/Hϕ). (1.4)

Based on the Eqs. (1.3) and (1.4), the μi of BaM, BaZn2 W, Co2X, and Co2Z hexaferrites are

calculated. The calculated μi is 1.18 for BaM (Ms = 380 emu/cm3, Ha = 17 kOe), 1.33 for

BaZn2 W (Ms = 392 emu/cm3, Ha = 10 kOe), 1.32 for Co2X (Ms = 365 emu/cm3, Ha = 9.5 kOe),

3
and 11 for Co2Z (Ms = 265 emu/cm3, H = 112 Oe). The low values of μi for BaM, BaZn2W, and

Co2X hexaferrites are attributed to their large Ha.

Figure 1.2 shows the calculated initial permeability as a function of the density of

sintered ferrite since the Ms is given in a unit of volume (in emu/cm3). The saturation

magnetization is increased with the sintered density through the Eq. (1.2). Therefore,

densification of a ferrite plays a role in achieving desired magnetization. As a result of the

fundamental relation of the magnetic properties, moderately high Ha and Ms of ferrites are

required to provide both high fFMR and permeability for miniature and efficient RF devices.

10
9
8
7
6
(i - 1)

5
4
3 s = 51 emu/g (Co2Z)

2 Ha = 120 Oe
3
1 Density = 5.3 g/cm
0
0.0 0.2 0.4 0.6 0.8 1.0
Sintered density
FIG. 1.2. Initial permeability as a function of sintered density.

1.3. Outline of dissertation

In this dissertation research, miniaturization and performance improvement of RF

antennas and inductors are emphasized by incorporating magnetic materials into RF device

designs. The dissertation ranges from fundamental magnetism, electromagnetics, and applied

physics of ferrites to design, performance simulation, and testing of antennas and inductors. This

4
dissertation is written in an article style format and organized as follows. In chapter 2, the

dynamic magnetic properties and non-reciprocal propagation characteristics of ferrites are

introduced. In chapter 3, the roles of ferrites in RF antennas are discussed in various aspects such

as antenna miniaturization, bandwidth, radiation efficiency, and impedance matching to free-

space. Chapter 4 discusses the key inductor parameters and effect of the ferrites on the quality

factor of the inductors. In addition, the design of integrated ferrite inductors is presented. In

chapter 5, the effectiveness of small relative permeability (μr < 2) in gigahertz (GHz) antenna

performance is discussed. In chapter 6, magnetic loss mechanism of ferrites is described, and a

large reduction of ferrite magnetic loss is experimentally demonstrated. Chapter 7 discusses

electrically small ferrite antenna gain with DC magnetic fields. A significant increase in ferrite

antenna gain is presented with a small DC magnetic field of compact Nd-Fe-B permanent

magnets. In chapter 8, a miniature and high-isolation multi-input multi-output (MIMO) ferrite

antenna array is designed and characterized for scattering parameters, correlation coefficient, and

gains. Chapter 9 discusses integrated ferrite inductors for power system-on-chip applications and

describes ferrite thickness-dependent inductance, quality factor, and DC current characteristics.

In chapter 10, high inductance-density ferrite inductors are demonstrated at GHz range.

Furthermore, the roles of dynamic magnetic properties in inductance and quality factor are

discussed. Lastly, chapter 11 gives the conclusion of the dissertation.

5
References

1. S. Bae, Y. K. Hong, J. Lee, W. M. Seong, J. S. Kum, W. K. Ahn, S. H. Park, G. S. Abo, J.


Jalli, and J. H. Park, IEEE Trans. Magn. 46, 2361 (2010).

2. A. D. Brown, J. L. Volakis, L. C. Kempel, and Y. Y. Botros, IEEE Trans. Antennas


Propag.47, 26 (1999).

3. F. Farzami, K. Forooraghi, and M. Norooziarab, IEEE Antennas Wirel. Propag. Lett. 10,
1540 (2011).

4. P. M. T. Ikonen, K. N. Rozanov, A. V. Osipov, P. Alitalo, and S. A. Tretyakov, IEEE Trans.


Antennas Propag. 54, 3391 (2006).

5. R. C. Hansen and M. Burke, Microw. Opt. Technol. Lett. 26, 75 (2000).

6. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag. 52, 1558 (2004).

7. J. Lee, Y. K. Hong, S. Bae, G. S. Abo, J. Jalli, J. Park, W. M. Seong, S. H. Park, W. G. An,


and G. H. Kim, Microwave Opt. Technol. Lett. 53, 1222 (2011).

8. G. M. Yang, X. Xing, A. Daigle, O. Obi, M. Liu, J. Lou, S. Stoute, K. Naishadham, and N. X.


Sun, IEEE Trans. Antennas and Prop. 58, 648 (2010).

9. E. Schloemann and R. E. Blight, IEEE Trans. Microwave Theory Tech. MTT-34, 1394
(1986).

10. I. M. Alexander and F. M. Aitken, IEEE Trans. Magn. Mag-11, 1267 (1975).

11. P. Gelin and P. Quéffélec, IEEE Trans. Magn. 44, 24 (2008).

12. S. D. Yoon, J. Wang, N. Sun, C. Vittoria, and V. G. Harris, IEEE Trans. Magn. 43, 2639
(2007).

13. T. Miura and Y. Konishi, IEEE Trans. Broadcast. 41, 101 (1995).

14. J. Wang, A. Yang, Y. Chen, Z. Chen, A. Geiler, S. M. Gillette, V. G. Harris, and C. Vittoria,
IEEE Microw.Wireless Compon. Lett. 21, 292 (2011).

15. B. K. O’Neiland J. L. Young, IEEE Trans. Microwave Theory Tech. 57, 1669 (2009).

16. C. K. Seewald and J. R. Bray, IEEE Trans. Microwave Theory Tech. 58, 1493 (2010).

17. J. Wu, M. Li, X. Yang, S. Beguhn, and N. X. Sun, IEEE Trans. Magn. 48, 4371 (2012).

18. S. Capraro, T. Rouiller, M. L. Berre, J. P. Chatelon, B. Bayard, D. Barbier, and J. J.


Rousseau, IEEE Trans. Compon. Packag. Technol. 30, 411 (2007).

6
19. S. Takeda, H. Mikami, and Y. Sugiyama, IEEE Trans. Microwave Theory Tech. 52, 2697
(2004).

20. B. N. Enander, Proceedings of the IRE 44, 1421 (1956).

21. P. H. Vartanian, J. L. Melchor, and W. P. Ayres, Microwave Theory and Techniques, IRE
Transactions on, 4, 8 (1956).

22. A. N. Friedman, J. Appl. Phys. 38, 1419 (1967).

23. A. Beyer and K. Solbach, IEEE Trans. Microwave Theory Tech. MTT-29, 1344 (1981).

24. J. Lee, Y. K. Hong, S. Bae, J. Jalli, J. Park, G. S. Abo, G. W. Donohoe, and B. C. Choi, IEEE
Trans. Magn. 47, 304 (2011).

25. C. Yang, F. Liu, X. Wang, J. Zhan, A. Wang, T. L. Ren, L. T. Liu, H. Long, Z. Wu, and X.
Li, IEEE Trans. Electron Devices 56, 3133 (2009).

26. K. Naishadham, IEEE Trans. Electromagn. Compat. 53, 923 (2011).

27. I. Kowase, T. Sato, K. Yamasawa, and Y. Miura, IEEE Trans. Magn. 41, 3991 (2005).

28. Z. Hayashi, Y. Katayama, M. Edo, and H. Nishio, IEEE Trans. Magn. 39, 3068 (2003).

29. H. M. Sung, C. J. Chen, W. S. Ko, and H. C. Lin, IEEE Trans. Magn. 30, 4906 (1994).

30. Y. Fukuda, T. Inoue, T. Mizoguchi, S. Yatabe, and Y. Tachi, IEEE Trans. Magn. 39, 2057
(2003).

31. J. Smit and H. P. J. Wijn, Ferrites, Wiley, New York (1959).

32. A. Goldman, Modern Ferrite Technology, Springer, New York (2006).

33. G. H. Jonker, H. P. J. Wijn, and P. B. Braun, Philips Technol. Rev. 18, 145 (1956-1957).

34. D. Lisjakand M. Drofenik, J. Am. Ceram. Soc. 90, 3517 (2007).

35. J. L. Snoek, Physical 14, 207 (1948).

36. C. Kittel, Phys. Rev. 73, 155 (1948).

37. J. Smit and H. P. J. Wijn Ferrites, John Wiley & Sons, New York, 279 (1959)

7
CHAPTER 2

FUNDAMENTAL PROPERTIES OF FERRITES1

2.1. Permeability tensor

The magnetic dipole moment of an electron can be written in terms of the vector relation

between the magnetic moment m and the spin angular momentum S as [1]

m   S , (2.1)

where γ is the gyromagnetic ratio (   m / s ). Note that the negative sign in Eq. (2.1) implies that

the magnetic moment is oppositely directed to the spin angular momentum as illustrated in Fig.

2.1. In the presence of the magnetic bias field H 0  zˆH 0 applied along the z-direction, the

magnetic dipole moment experiences a torque that is equal to the time rate of change of angular

momentum. Accordingly, the torque can be expressed with respect to the magnetic dipole

moment of an electron as

FIG. 2.1 Schematic illustration of motion of a magnetic dipole moment in a magnetic field.

1
This work was published as “Chapter 8 – Ferrites for RF passive devices,” in Solid State Physics 64, 237
(2013) by Yang-Ki Hong and Jaejin Lee.

8
ds  1 dm
T  m  B0   0 m  H 0   . (2.2)
dt  dt

As a result of the interaction of a magnetic dipole moment and the bias magnetic field, the

equation of motion for the magnetic dipole moment is obtained from Eq. (2.2) as

dm
   0  m  H 0  . (2.3)
dt

Equation (2.3) yields that the magnetic dipole moment m precesses around the H0-field vector

as illustrated in Fig. 2.1.

Now, the total magnetization M of a ferrite is considered and applied to the equation of

motion of Eq. (2.3). Assume that there is N number of magnetic dipoles per unit volume and the

bias magnetic field is strong enough to align all magnetic dipole moments. Then, the total

magnetization M of a saturated ferrite is given by

M  Nm. (2.4)

Applying Eq. (2.4) into (2.3) gives the equation of motion for the magnetic dipole moment in

terms of the total magnetization as

dM
   0  M  H  , (2.5)
dt

where H is the net internal magnetic field of a ferrite. Equation (2.5) implies that the magnetic

dipole moments are aligned and precess around the H field vector as illustrated in Fig. 2.2.

FIG. 2.2. Total magnetization and precession of magnetic dipole moments within a ferrite by
bias magnetic field along the z-axis.

9
In order to understand the fundamental operation mechanisms of RF ferrite devices, the

interaction of the magnetic dipoles with an AC (alternating current) magnetic field is considered.

Assume that the AC magnetic field is much smaller than the DC (direct current) magnetic field

(i.e. H  H 0 ) and that the DC magnetic field H0 is biased in the ẑ direction and strong enough

to saturate the ferrite. The total magnetic field Ht is then expressed as the sum of each of the

components of the DC and AC magnetic fields

H t  H 0 zˆ  H . (2.6)

Likewise, the resulting total magnetization M t of the ferrite caused by the total magnetic field is

given in terms of DC and AC components

M t  M s zˆ  M , (2.7)

where Ms is the DC saturation magnetization produced by the DC magnetic field ( H0 zˆ ) and M is

the AC magnetization created by the AC magnetic field ( H ). Applying the total magnetic field

of Eq. (2.6) and the total magnetization of Eq. (2.7) into the equation of motion of Eq. (2.5), the

equation of motion for magnetic dipoles, interacting with the AC magnetic field, can be written

as

dM
   0  M  H     0  M s zˆ  M  H 0 zˆ  H  . (2.8)
dt

Equation (2.8) can be rewritten in terms of three vector components by

xˆ yˆ zˆ
dM dM x dM y dM z
     0 M x My Ms  Mz
dt dt dt dt
Hx Hy H0  H z
 
 M y H z  H 0   M z  M s H y xˆ  M x H z  H 0   M z  M s H x yˆ 
  0    (2.9)
 
 M x H y  M y H x zˆ 
.

10
It is shown that when the DC bias magnetic field is in the ẑ -direction, the z-component equation

of motion must consist purely of the AC magnetic field and the AC magnetization due to the

dMs / dt  0. Similarly, when the ferrite is biased in the x̂ direction, the x-component equation of

motion will have only AC magnetic components.

Therefore, the three component equations of motion of magnetic dipole moments for the

z-axis biased ferrite from Eq. (2.9) can be written as follows

dM x
   0 M y H z  H 0    0  M z  M s H y , (2.10)
dt

dM y
  0M x  H z  H 0    0 M z  M s H x , (2.11)
dt

dM z
   0 M x H y   0 M y H x . (2.12)
dt

In Eqs. (2.10) − (2.12), the products of all the AC components can be ignored, since the AC

magnetic field is assumed to be much smaller than the DC magnetic field (i.e. H  H 0 ). Then,

the following three equations can be obtained

dM x
   0  M y H z   0  M y H 0   0  M z H y   0  M s H y   0 M y   m H y , (2.13)
dt

dM y
  0M x H z   0M x H 0   0 M z H x   0M s H x  0 M x   m H x , (2.14)
dt

dM z
   0M x H y   0M y H x  0 , (2.15)
dt

where the precession frequency is 0  0H0 , and the magnetization frequency is m  0Ms .

After solving for Mx and My in Eqs. (2.13) and (2.14), by taking the second derivative, the

following equations are obtained

11
d 2M x dM y dH y dH y
 0  m  0 0 M x  m H x   m ,
dt 2 dt dt dt

d 2M x dH y
 02 M x  0m H x  m . (2.16)
2
dt dt

d 2M y dM x dH x Hx
 0  m  0  0 M y  m H y   m ,
2
dt dt dt dt

d 2M y dH x
 02 M y  0m H y  m . (2.17)
dt 2 dt

Since the AC components of magnetic field and magnetization are time-harmonic dependent ejωt,

the AC vector components can be expressed by

M x  M x e j t , (2.18a)

M y  M y e jt , (2.18b)

Mz  Mz e jt , (2.18c)

H x  H x e jt , (2.19a)

H y  H y e jt , (2.19b)

Hz  Hz e jt . (2.19c)

Applying Eqs. (2.18a, b, c) and (2.19a, b, c) to (2.16) and (2.17) gives the following phasor

equations

d2 d
2
   
M x e jt   02 M x e jt   0 m H x e jt   m
dt

H y e jt ,   
dt

  
j 2 2 M x e jt  02 M x e jt  0m H x e jt  jm H y e jt , 
 2
0 
  2 M x  0m H x  jm H y , (2.20)

12
d2 d
and 2
   
M y e jt   02 M y e jt   0 m H y e jt   m
dt

H x e jt ,   
dt

  
j 2 2 M ye jt  02 M ye jt  0m H ye jt  jm H ye jt , 
 2
0 
  2 M y   jm H x  0m H y . (2.21)

After rearranging Eqs. (2.20) and (2.21), the x and y components of AC magnetization (Mz = 0

from Eq. (2.15)) can be obtained by

0m jm
Mx  Hx  Hy , (2.22)

2
0  2   2
0  2 
 jm 0m
My  Hx  Hy , (2.23)
 2
0  2   2
0  2 
Mz  0 . (2.24)

Therefore, Eqs. (2.22) − (2.24) can be simply rewritten in matrix form as

 0m j m 
 2 0

 M x   0  
2
  2
0  2   H x 
    j m 0m  
M y    2 0  H y  , (2.25)

 M   0  
2
  2
0  2   
H
 z  0 0 0  z 
 
 

which shows the linear relation between the magnetization M and the magnetic field H . The

magnetization M is then expressed by the tensor susceptibility   and magnetic field H as

follows

  xx  xy 0  H x 
  
M   H    yx  yy 0  H y  , (2.26)
 0 0 0   H z 

where the elements of the tensor susceptibility   are

13
0m
 xx   yy  , (2.27)

2
0  2 
jm
 xy   yx  . (2.28)
 2
0  2 
From the linear relationship between B and H , the tensor permeability   can be found

B  M  H   H  H     1H  H , (2.29)

where   is the tensor permeability. The tensor permeability is given in the following form [2]

  j 0
   1      j 

0 , (2.30)
 0 0 1 

where

0m
  1   xx  1   yy  1  , (2.31)
 2
0  2 
m
and    j xy  j yx  . (2.32)
2
0  2 
µ is the diagonal term and κ is the off-diagonal term. The above tensor permeability   is

derived for the magnetized ferrite with a static magnetic field along the ẑ direction.

In order to understand the physical meaning of the permeability tensor, the magnetic flux

B induced by the permeability tensor is given by

  j 0  H x 
  
B   H   j  0  H y  , (2.33)
 0 0 1   H z 

Bx  H x  jH y ,

14
B y   jH x  H y ,

Bz  H z .

It is seen that the asymmetrical permeability tensor gives rise to a 90° phase delay between x̂

and ŷ components of B produced from the x̂ (or ŷ ) component of H . This characteristic of

the permeability tensor causes non-reciprocal wave propagation in the ferrite medium. Thus,

non-reciprocal RF devices having unidirectional wave transmission can be realized. When it

comes to the symmetrical permeability tensor arising from the non-magnetized ferrite, the

frequencies ω0 (  0H0 ) and ωm (  0Ms ) will be zero because there are no bias field and

induced magnetization. Accordingly, the diagonal term µ becomes one and the off-diagonal term

κ goes to zero. Therefore, the symmetrical tensor permeability is expected. This permeability

tensor will generate both the x̂ and ŷ components of B in phase in contrast to the

asymmetrical tensor permeability.

The elements of permeability tensor appeared in Eqs. (2.31) and (2.32) are derived

without consideration of magnetic loss of ferrimagnetic materials. However, various losses,

including the hysteresis loss, eddy current loss, and ferromagnetic resonance loss, are existing in

the ferrites. Such magnetic losses reduce the spin precessional motion. Accordingly, a damping

factor α is introduced to Eqs. (2.31) and (2.32) by simply substituting ω0 with a complex

resonant frequency ω0 + jαω [1]. Thus, the following equations can be obtained for the complex

elements of the permeability tensor which takes into account the magnetic loss

15
0  j m 0  j m
     j   1   1
0  j 2   2 02  j 20   2 2   2
0  j m 02   2 1   2  j 20 
 1
 2
0    
  2 1   2  j 20 02   2 1   2  j 20
   j        j 2  
0 m m
2
0
2 2 2
0
1
   1     4   2
0
2 2 2 2 2
0
2

          2   j 2    j       


0 m
2
0
2 2 2
m 0 0 0 m m
2
0
2 2 2

1 
   1     4   2
0    1     4  
2 2 2 2 2
0
2 2
0
2 2 2 2 2
0
2

         
0 m     1   
2
0
2 2 2
m
2
0
2 2

1 j
   1     4      1     4  
2
0
2 2 2 2 2
0
2 2
0
2 2 2 2 2
0
2

(2.34)

 m m
     j   
0  j 2   2 02  j 20   2 2   2


  
    1    j 2  
m 02   2 1   2  j 20 m
2
0
2 2
0

   1    j 2     1    j 2      1     4  


2
0
2 2
0
2
0
2 2
0
2
0
2 2 2 2 2
0
2

        j 2 
m
2
0
2 2 2
0 m

        4  
2
0
2 2 2 2 2
0
2

    1   
m
2
0
2
2   2
0 m
2
 j
        4      1     4  
2
0
2 2 2 2 2
0
2 2
0
2 2 2 2 2
0
2

(2.35)

The damping factor, α, is related to the ferromagnetic resonance linewidth ΔH and can be

determined using the expression

H
 . (2.36)
2

The definition of the linewidth ΔH is the width of the curve of μ″ versus H0 (= ω0/ω) where μ″

decreases to half of its maximum value. A wide ΔH implies a large damping factor α. As a result,

16
the large damping factor leads to relaxation-type permeability dispersion. High frequency

dispersion of magnetic permeability of a polycrystalline ferrite is discussed in the next section.

2.2. Magnetization processes and permeability dispersion of ferrites

The magnetization process of a polycrystalline ferrite generally involves domain wall

motion, followed by spin rotation [3−5], as illustrated in Fig. 2.3. The domain wall motion is a

major process up to the knee of the magnetization curve of a magnetic hysteresis loop, which

shows a large increase in magnetization with small applied magnetic field [6]. By further

increasing the applied DC magnetic field beyond the knee, the domain wall motion can be

completed, and the spin rotation becomes predominant. Correspondingly, the permeability

dispersion of a polycrystalline ferrite is associated with the domain wall motion and spin rotation

and can be given by [7, 8]

FIG. 2.3. Magnetization processes of a polycrystalline ferrite consist of domain wall motion
and spin rotation.

17
 r     j   1   d   s  1   d  j d    s  j s
d2  d 0 s  j s  s 0 , (2.37)
1 
2 2
    j
d s  j 2   2
where χd′ and χd′′ are the real and imaginary part of susceptibility of the domain wall motion, χd0

is its static magnetic susceptibility, ωd (= 2πfd) is the domian wall resonance frequency, and β is

the damping factor of domain wall motion. χs′ and χs′′ are the real and imaginary part of

susceptibility of the spin rotation, χs0 is its static magnetic susceptibility, ωs (= 2πfs) is the spin

resonance frequency, and α is the damping factor of spin rotation.

The static magnetic susceptibility of domain wall motion (χd0) and that of spin rotation

(χs0) in Eq. (2.37) can be given by [7]

3M s2 D
d 0  , (2.38)
4 

4M s
s0  , (2.39)
H a  H d 
where Ms is the saturation magnetization, D is the grain size, γ′ is the wall energy, Ha is the

magnetic anisotropy field, and Hd is the demagnetizing field. According to Eqs. (2.37) − (2.39),

it is seen that the frequency dispersion of domain wall motion is mainly associated with the

microstructure and grain size. On the other hand, the spin rotational permeability dispersion of

ferrite is related to the magnetic anisotropy field and demagnetization field. It is well-known that

domain wall resonance contributes to low-frequency permeability dispersion, while spin rotation

resonance is dominant in the high frequency permeability dispersion [5]. Solving Eq. (2.37), the

complex suceptibility of domain wall motion and spin rotation are obtained as follows

18
d2  d 0
 d   d  j d 
d2   2  j



d2  d 0 d2   2   j  
d2  d 0 d2   2   jd2  d 0
(2.40)

2
d      j 
  2  j  2
d
2 2
d 2
 2   2 2 
     
2
d d0
2
 
d
2 2
d d0
 j
               
2
d
2 2 2 2 2
d
2 2 2 2

s  j s  s 0


 s   s  j s 
s  j 2   2
s  j s  s 0 s2   2 1   2   2 js 

 2
s     
  2 1   2  2 js s2   2 1   2  2 js
  j      1     2 j  
s s s0
2
s
2 2
s

   1     4  
2
s
2 2 2 2
s
2 2

          2  
2
s s0
2
s  2 j    j    
2 2 2
s s s0 s
2
s s0 s s0
2
s
2
  2 2 
 
   1     4  
2
s
2
   1     4  
2 2 2
s
2 2 2
s
2 2 2 2
s
2 2

         
2
s s0      1   
2
s
2 2 2
s s0
2
s
2 2

 j
   1     4      1     4  
2
s
2 2 2 2
s
2 2 2
s
2 2 2 2
s
2 2

(2.41)

Applying Eqs. (2.40) and (2.41) to Eq. (2.37) leads to the real (μ′) and imaginary (μ″) part of the

permeability of a polycrystalline ferrite

r    j  1 d  jd   s  js  1 d  s   jd  s , (2.42)

  1 
d2 d 0 d2   2 

 
 s 0s2 s2   2    2 2
, (2.43)
 2
d  2 2
   2 2
   1     4  
2
s
2 2 2 2 2
s
2

  
d 0d2


 s 0s s2   2 1   2   . (2.44)
 2
d      
2 2 2 2 2
s  2 1    2 2
  4   2 2
s
2

The damping factors, β and α, imply the resonance state of magnetic domain walls and

gyromagnetic spin, respectively. In the presence of a high damping factor, the frequency

19
dispersion of permeability spectra is the relaxation-type dispersion as mentioned previously.

Since a ferrite has a large damping factor of spin rotation, the spin resonance can be

approximated to the relaxation-type frequency dispersion [5, 9]. Therefore, taking a derivative of

Eq. (2.41) leads to the following relation between low α spin resonance (ωs) and high α spin

resonance frequency (ωs_α)

d s
0
d   s _ 

 j s  s 0 s  j 2   2  s  j s  s 0 2 j s  j   2 



s  j 2   2 2   s_
(2.45a)

  j s  s 0  s  js _    
2 2
s _    j
s s _  s  s 0 2 j s  js _    2s _  

 js  s 0 s2  2 jss _   s2_  2  s2_  

 s  s 0 2 j s  js _  s  js _    2s _  s  js _   
 s2   2s2_   s2_ 

s
s _   . (2.45b)
 2 1

Applying Eq. (2.45b) to Eq. (2.41) gives Eq. (2.45c)

s  j  s  s 0
s 
s  j 2   2
 s0  s0 ,
 
 s  j   2
  2 s  j 
 1 j 
s s  j  s s  s  j  s  j  s
s0 s0
s  2 3
 ,
 s  j    2    3
 
1 j   1    j   
s s    s  s      s s    s 
2 2 2  2 2 2   2 2 2

s0
s  ,
  2    2
 1   2 
1   j  1  
  2   2 2  2 
s   1  s     2 2 2 
 s 

20
s0
s  ,
 2   2 1  2 
1   j  1  
  2   2 2  s 1  1 /  2     2   2 2 
 s   s 

 s0
lim  s  lim ,
     2   2 
1   j  1 1  
  2   2 2  s _  1  1 /  2     2   2 2 
 s   s 
 s0
s  . (2.45c)

1 j
s _ 

Therefore, when the spin resonance takes the form of relaxation-type frequency dispersion and

the domain wall resonance takes the resonance-type dispersion, the complex permeability of a

polycrystalline ferrite of Eqs. (2.42) − (2.44) becomes

 r     j   1   d  j d    s  j s
 d2  d 0  s0
1  ,
 d2    j 
1 j
s _ 

d2  d 0 d2   2   s0s2_    d 0d2  s 0s _  


r  1    j  . (2.46)
 2 2 
2
d  2
  2   2 2 s2_   2 
 d  
 2 2
  
2 2
 2
s _   

Figure 2.4 shows the resonance-type and relaxation-type dispersion of the complex

susceptibility for domain wall motion and spin rotation. For the resonance-type dispersion, the

real part (χ′) of susceptibility passes through a maximum and becomes larger than initial

susceptibility. The imaginary part (χ″) of susceptibility is sharp and narrow. On the other hand,

the relaxation-type dispersion shows broad spectra of the imaginary part (χ″) of susceptibility.

Thus, high magnetic loss appears over a broad frequency range in the relaxation-type dispersion

[9]. It is noted that magnetic loss needs to be minimized to achieve high performance RF devices.

21
In the next section, relation of magnetic loss to resonance frequencies of domain wall motion and

spin rotation is studied.

d' s'
Resonance type Relaxation type
d'' s''

Frequency Frequency
(a) (b)
FIG. 2.4. Frequency dispersion of complex susceptibility of (a) domain wall motion
(resonance-type) and (b) spin rotation (relaxation-type).

2.3. Magnetic loss of polycrystalline ferrites

The complex permeability of a ferrite is frequency dependent and expressed in terms of

the real (μ′) and imaginary (μ″) parts of permeability. μ′ represents the energy stored, while μ″ is

the energy loss dissipated within the ferrite. The ratio of μ″ over μ′ is defined as magnetic loss

tangent (tan δμ = μ″/μ′), which describes the power lost versus power stored in a ferrite. The

magnetic loss tangent consists of three main components [10]: the hysteresis loss (tan δh), eddy

current loss (tan δe), and residual loss (tan δr) tangents, as shown in Fig. 2.5.

tan    tan  h  tan  e  tan  r . (2.47)

The hysteresis loss is the energy dissipation of a static hysteresis loop, and the eddy current loss

is determined by the electrical resistivity of the ferrite. Due to high electrical resistivity, ferrites

have almost negligible eddy current loss at high frequencies. The residual loss is associated with

magnetic domain wall and spin rotational resonances [3, 4]. When a low-frequency AC magnetic

22
FIG. 2.5. Three main components of magnetic loss of a ferrite: hysteresis loss, eddy current
loss, and residual loss.

FIG. 2.6. Illustration of induced magnetic flux of ferrite with interaction of a time varying
magnetic field.

field H is applied to a ferrite, a magnetic flux B is induced in phase in response to the applied

magnetic field. However, as the frequency of the AC magnetic field increases, the induced

magnetic flux B is not able to follow the AC field instantaneously due to domain wall resonance.

This results in a lag between the magnetic flux and the AC magnetic field, causing a phase delay

δ. The spin rotation resonance is attributed to the fact that the driving frequency of the AC

magnetic field is at resonance with the natural precession frequency of the ferrite, resulting in

increased power absorption. At the resonance, energy loss occurs in the form of heat. Figure 2.6

illustrates that the magnetic loss tangent takes place due to a time varying magnetic field. The

complex relative permeability μr is expressed in terms of the real and imaginary parts of

permeability (μr = μ′ − jµ″).

23
From the linear relationship of B and H (see Eq. (2.29)), the complex relative

permeability μr for the ferrite interacting with the AC magnetic field can be written as phasor,

B B0e j t   B0 B0
r    j     e j  cos  j sin  
H H 0e jt H0 H0
, (2.48)
B0 B0
 cos  j sin 
H0 H0

where the real and imaginary parts of the permeability are

B0
  cos , (2.49)
H0

B0
   sin  . (2.50)
H0

Thus, one has

  B0 / H 0 sin 
tan    . (2.51)
 B0 / H 0 cos
The real (μ′) and imaginary (μ″) parts of the permeability imply that the magnetic induction B is

in-phase and out-of-phase with the AC magnetic field, respectively.

As discussed in the previous section, the frequency dispersion of complex permeability

(μ′ − jμ″) of a polycrystalline ferrite is related to domain wall motion and spin rotation.

Therefore, using Eq. (2.46), the individual magnetic loss tangent of the domain wall motion (tan

δd) and spin rotation (tan δs) can be given by

 r     j    1   d  j d    s  j s
 d2  d 0  d2   2  
  s 0 s2
 d 0 d2  s 0 s  , (2.52)
1   j 
 2

 d2     2  2  s2   2   2   2   2 2  2   2 
 
 d s 

24
tan d 
 d


 d 0d2  / d d 0 d 
  2 2  2  
   , (2.53)
 d  d2     2  2   d2     2  2  d2   2
   
2 2

 s   s 0s    s 0s2  


tan s   /  . (2.54)
 s  s2   2   s2   2  s

Equations (2.53) and (2.54) suggest that both tan δd and tan δs can be reduced by increasing the

resonance frequencies ωd and ωs. Thus, one of the approaches to increase the resonance

frequencies is to apply a DC magnetic field to a ferrite. It is experimentally observed that

ferrimagnetic resonance frequency and magnetic loss tangent of a ferrite increases and decreases,

respectively, with applied magnetic fields [11].

Magnetic loss tangent of ferrite is an important parameter for RF device applications.

Since the magnetic loss tangent causes energy dissipation within ferrites, performance of RF

devices is thus deteriorated. For example, antenna radiation efficiency, defined as the ratio of

radiated power to input power, decreases due to the magnetic loss tangent of the ferrite substrate.

Therefore, magnetic loss tangent needs to be reduced in order to realize high-efficiency RF

devices. To address this issue, the control of grain size and microstructure, and the optimization

of sintering process are investigated [12−14].

2.4. Interaction of AC field with permeability tensor

Recall that the saturated ferrite with a bias magnetic field gives rise to the permeability

tensor of Eq. (2.30) and sets up a preferential direction of magnetic dipole moments interacting

with electromagnetic waves. This phenomenon enables unidirectional transmission characteristic

of RF signals [15−17]. In this section, the interaction of a circularly polarized AC field with a

saturated ferrite is discussed.

25
A circularly polarized wave is formed when the x and y components with an equal

amplitude of a linearly polarized wave are separated in phase by 90°. Accordingly, a right-hand

circularly polarized (RHCP) field can be written in phase form as [1]

H   H   xˆ  jyˆ  , (2.56)

where the H+ is the amplitude of RHCP magnetic field. From Eq. (2.56), each vector component

of the RHCP field is obtained as

H   H x  H y  H z  H  xˆ  jH  yˆ , (2.57)

where

H x  H  , (2.58a)

H y   jH  , (2.58b)

Hz  0 . (2.58c)

By applying the components of the RHCP field of Eqs. (2.58a, b, c) to the magnetization

components of Eqs. (2.20) and (2.21), the following relation between effective permeability and

frequency can be obtained

 2
0 
  2 M x  0m H x  jm H y  0m H   jm  jH  
, (2.59)
 m 0   H 

2
0 
  2 M y   jm H x  0m H y   jm H   0m  jH  
. (2.60)
  jm 0   H 

The magnetization components interacting with the RHCP magnetic field are then given by

 m 0   H  m
M x  H, (2.61)
0   0    0   

26
 jm 0   H   jm
M y   H. (2.62)
0   0    0   
Therefore, the total magnetization interacting with the RHCP magnetic field H  is written as

M   M x xˆ  M y yˆ
m  m
 H x  j H  yˆ (2.63)
0    0   
m
 H   xˆ  jyˆ 
0   
Note that the resulting magnetization ( M  ) from the RHCP magnetic field is also RHCP.

Therefore, the magnetization rotates with an angular frequency ω of the driving field ( H  ).

Using the linear relationship between B and H of Eq. (2.29) and the total magnetization of


(2.63), the effective permeability (  ) for the RHCP magnetic field is obtained as


M m
  1  1 . (2.64)
H 0   
Likewise, a left-hand circularly polarized (LHCP) field has the following phase form

H   H  xˆ  jyˆ  , (2.65)

where the H− is the amplitude of LHCP magnetic field. Similarly, each vector component of the

LHCP magnetic field can be obtained from Eq. (2.65)

H x  H  , (2.66a)

H y  jH  , (2.66b)

Hz  0 . (2.66c)

Applying the vector components of the LHCP magnetic field of Eq. (2.66a, b, c) to the

magnetization components of (2.20) and (2.21) yields the following expressions

27
2
0 
  2 M x  0m H x  jm H y  0m H   jm jH   
, (2.67)
 m 0   H 

2
0 
  2 M y   jm H x  0m H y   jm H   0m jH   
. (2.68)
 jm 0   H 

After rearranging Eqs. (2.67) and (2.68), the magnetization components interacting with the

LHCP magnetic field can be given by

 m 0   H  m
M x  H, (2.69)
0   0    0   

 jm 0   H  jm


M y  H. (2.70)
0   0    0   
Therefore, the total vector magnetization for the LHCP magnetic field is represented by the sum

of the x and y components of the magnetization of Eqs. (2.69) and (2.70)

M   M x xˆ  M y yˆ
m m
 xˆ H   jyˆ H (2.71)
0    0   
m
 H   xˆ  jyˆ 
0   
Note that the magnetization ( M  ) produced by the interaction of the LHCP field is also in the

form of the LHCP field. Accordingly, this indicates that the resulting magnetization is strongly

coupled with the driving field ( H  ) at the same angular frequency ω. The effective permeability


(  ) for the LHCP field is then given by Eq. (2.72)

M m
 1  1 . (2.72)
H 
0   

28
The expressions of the effective permeability for RHCP and LHCP magnetic fields were

derived without considering the effect of loss. The ω0 is, therefore, substituted with ω0 + jαω in

Eqs. (2.64) and (2.72) in order to calculate effective permeability that includes the damping

factor. Then, Eqs. (2.64) and (2.72) can be written for the complex effective permeability for

RHCP and LHCP magnetic fields

m
  1
0  j   
  m   0   m
 1 j (2.73)
2
  0    
2 2
   0 2   2 2
    j 

m
   1
 0  j   
 m   0   m
 1 j (2.74)
2
   0    
2 2
  0 2   2 2
    j 

The complex effective permeability for RHCP and LHCP fields are plotted as a function

of ω0 in Fig. 2.7 using Eqs. (2.73) and (2.74). It can be seen that the   has a frequency

'+
 "+ "+

'-
(arbitrary units)

 "-

'+
'-

"-


0 (arbitrary units)
FIG. 2.7. Complex effective permeability for right hand circularly polarized (RHCP) and left
hand circularly polarized (LHCP) fields as a function of the gyromagnetic frequency ω0.

29

dispersive property, while the  is insensitive to variations in frequency [18]. Furthermore, it is

seen that the imaginary part of   shows the peak at the frequency ω. Since the μ′′ represents the

dissipated energy loss, a large attenuation occurs for the RHCP (   ), but it is negligible for the


LHCP (  ) as shown in Fig. 2.7. In section 2.1, it is demonstrated that preferential precession

direction of magnetic dipole moment is determined by the direction of DC bias magnetic field.

Thus, circularly-polarized field propagation in the magnetized ferrite strongly interacts with

precessing magnetic dipole moments in the same direction as the precession, but weak

interaction in the opposite direction of precession. These phenomena lead to non-reciprocal

propagation characteristics.

2.5. Plane wave propagation in magnetized ferrites

In this section, electromagnetic wave propagation in a magnetized ferrite is discussed.

Assume that a DC bias field H 0  zˆH 0 is applied to an infinite ferrite medium. There are no free

charges or conduction currents in the medium. Therefore, Maxwell’s equations are given by

  E   jB   j  H , (2.75a)

  H  j  E , (2.75b)

D  0, (2.75c)

B 0. (2.75d)

Then, if the plane wave propagation is considered in the z direction, the electric and magnetic

fields can be expressed in the following form

E  xˆEx  yˆE y e  jz , (2.76a)

H  xˆH x  yˆH y e  jz . (2.76b)

30
In order to express Eq. (2.75a) in terms of tensor permeability, Eqs. (2.76a, b) are substituted into

(2.75a). Then, Eq. (2.75a) becomes Eq. (2.77).

  E   j  H

xˆ yˆ zˆ   j 0  H xe  jz 
     . (2.77)
   j   j  0  H y e  jz 
x y z   
E x e  jz E y e  jz 0  0 0 1  0 

Solving the Eq. (2.77) leads to the expressions below

 
 xˆ Ey e jz  yˆ Exe jz   je jz H x  jH y xˆ  je jz  jH x  H y yˆ ,
z z

jEy e jz xˆ  jExe jz yˆ   je jz H x  jH y xˆ  je  jz  jH x  H y yˆ ,

j E y xˆ  j E x yˆ   j H x  jH y xˆ  j  jH x  H y yˆ . (2.78)

The x̂ and ŷ components of Eq. (2.78) are

j E y   j H x  jH y  , (2.79a)

j E x  j  jH x  H y  . (2.79b)

Now, applying the electric and magnetic fields of Eq. (2.76a, b) to the Maxwell equation

(2.75b) gives

  H  j E

xˆ yˆ zˆ
   (2.80)
  j xˆE x  yˆE y e  jz
x y z
H x e  jz H y e  jz 0

Similarly, solving the cross product of Eq. (2.80) gives the following expressions

 
 xˆ H ye jz  yˆ H xe jz  jExe jz xˆ  jEye jz yˆ ,
z z

31
jH ye  jz xˆ  jH xe jz yˆ  jExe  jz xˆ  jEy e jz yˆ ,

j H y xˆ  j H x yˆ  j E x xˆ  j E y yˆ . (2.81)

The relation between the transverse magnetic and electric fields is obtained from Eq. (2.81) as

j H y  j E x , (2.82a)

 j H x  j E y . (2.82b)

Finally, the rearrangement of Eq. (2.82a, b) gives the relation between the magnetic and electric

field components as


Hx   Ey , (2.83a)


Hy  Ex . (2.83b)

Similarly, after substituting Eq. (2.83a, b) into (2.79a, b), the following equations are obtained

   
jE y   j   Ey  j Ex ,
   
 

 2Ex  j 2   2 Ey  0 , (2.84)

   
 jEx   j j Ey   Ex ,
   
 

    E
2 2
x  j 2E y  0 . (2.85)

Equations of (2.84) and (2.85) can be written in matrix form as

  2  
j  2    2   Ex  0
      . (2.86)

  2    2  j 2   E y  0

32
For the non-trivial solution above, the determinant of this set of Eq. (2.86) must be zero [1].

Accordingly, the following equations, based on the determinant of 2 × 2 matrices in Eq. (2.86),

are given by

  2  
j  2    2 
det   0,

  2    2  j 2 

             0 .


2 2 2 2 2 2
(2.87)

Rearrangement of Eq. (2.87) reduces to Eq. (2.88)

 4 2 2   2   2   0 ,
2

 4  2 2  2   4  2 2   4 2 2  0 . (2.88)

It is noted that the expression of Eq. (2.88) is in the quadratic form. By solving Eq. (2.88) for β2,

Eq. (2.89) is obtained

2 

 2   4 4  2 2  4  4  2 2   4 2 2   2  
2
4 4 2 2
2 2 . (2.89)
2 2
2   2 

2

It is seen that there are two possible propagation constants, β+ and β−, when a plane wave is

propagating through a magnetized ferrite having the asymmetrical permeability tensor. The two

constants are

    2    2        . (2.90)

Applying the constant β+ of Eq. (2.90) to (2.84), the relation between Ey and Ex is obtained

below.

 2
 2 E x  j   2          E y  0 ,
 

33
 2Ex  j 2Ey  j 2    Ey  0 ,

 2Ex  j 2E y  0 ,

E y   jE x . (2.91)

Then, substituting Eq. (2.91) to (2.76a) gives Eq. (2.92), in which electric field is expressed in

terms of propagation constant β+

Ex  xˆEx  yˆEy e j  z   xˆEx  jyˆEx e  j  z  Ex  xˆ  jyˆ e j  z , (2.92)

where Ex is the arbitrary amplitude of the electric field. It can be seen that the electric field

associated with β+ is in the form of RHCP plane wave. Also, applying Eq. (2.91) to (2.83a) for

the magnetic field associated with β+ leads to the following equations

 
H x   E y    jE   j  E   j  E xˆ  jyˆ e
x
 
x x
 j  z

   
, (2.93)

 E x  jxˆ  yˆ e  j  z  Y E x  jxˆ  yˆ e  j  z


where Y+ is the wave admittance for the plane wave propagation in a magnetized ferrite medium

with β+. The wave admittance Y+ is defined by

 
Y   . (2.94)
  

Similarly, substitution of Eq. (2.90) for β− into (2.84) gives the following equations

 2

 2 E x  j   2          E y  0 ,
 

 2Ex  j 2 E y  j 2    Ey  0 ,

 2Ex  j 2E y  0 ,

E y  jE x . (2.95)

34
Then, applying Eq. (2.95) to (2.76a) leads to the equation for the electric field associated with β−

Ex  xˆEx  yˆE y e  j  z   xˆEx  jyˆEx e j  z  Ex xˆ  jyˆ e  j  z . (2.96)

It is seen that the electric field associated with β− is in the form of LHCP plane wave.

Combination of Eqs. (2.83a) and (2.95) gives the corresponding magnetic field

 
H x   E y    jE    j  E    j  E xˆ  jyˆ e

x x

x
 j  z

   
. (2.97)

 E x  jxˆ  yˆ e  j  z  Y E x  jxˆ  yˆ e  j  z


The Y− is the wave admittance for the plane wave propagation in the magnetized ferrite with β−.

 
Y   . (2.98)
  

Based on the above derivations, it is found that there are two propagation constants when a

linearly polarized wave propagates in the magnetized ferrite. The electric field associated with

the two propagation constants are in the form of either right-hand circularly polarized (RHCP) or

left-hand circularly polarized (LHCP) plane waves. As discussed earlier, the circularly polarized

waves interact differently with magnetic dipole moments, depending on the direction of

magnetization of the magnetized ferrite. This is because a bias magnetic field creates a

preferential direction of precession, synchronizing with the sense of the polarization. As shown

in Fig. 2.7, large attenuation occurs when a circularly polarized field rotates in the same direction

as the magnetic dipole precession, while small attenuation occurs when the field rotates in the

opposite direction. In practice, this non-reciprocal propagation characteristic gives the

fundamental operation concepts of RF devices such as circulators, isolators, and phase shifters.

Equations (2.73) and (2.74) give key design parameters of the non-reciprocal devices. The

parameters include saturation magnetization, internal bias field, and FMR linewidth.

35
References

1. D. M. Pozar, Microwave Engineering, Wiley, New Jersey (2004).

2. D. Polder, Philos. Mag. 40, 99 (1949).

3. S. Chikazumi, Physics of Ferromagnetism (John Wiley & Sons, 1964).

4. D. Stoppels, J. Magn. Magn. Mater. 160, 323 (1996).

5. G. T. Rado, R. W. Wright, and W. H. Emerson, Phys. Rev. 80, 273 (1950).

6. B. D. Cullity and C. D. Graham, Introduction to Magnetic Materials, Wiley, New Jersey


(2009).

7. T. Tsutaoka, M. Ueshima, T. Tokunaga, T. Nakamura, and K. Hatakeyama, J. Appl. Phys.78,


3983 (1995).

8. T. Tsutaoka, J. Appl. Phys. 93, 2789 (2003).

9. G. T. Rado, V. J. Folen, and W. H. Emerson, Proceedings of the IEE - Part B: Radio and
Electronic Engineering 104, 198 (1957).

10. A. J. Moulson and J. M. Herbert, Electroceramics: Materials, Properties, Applications,


Wiley, New Jersey (2003).

11. J. Lee, Y. K. Hong, W. Lee, G. S. Abo, J. Park, W. M. Seong, and W. K. Ahn, J. Appl. Phys.
113, 073909 (2013).

12. J. Lee, Y. K. Hong, S. Bae, J. Jalli, G. S. Abo, J. Park, W. M. Seong, S. H. Park, and W. K.
Ahn, J. Appl. Phys. 109, 07E530 (2011).

13. J. Lee, Y. K. Hong, W. Lee, G. S. Abo, J. Park, N. Neveu, W. M. Seong, S. H. Park, and W.
K. Ahn, J. Appl. Phys. 111, 07A520 (2012).

14. H. Su, X. Tang, H. Zhang, Z. Zhong, and J. Shen, J. Appl. Phys. 109, 07A501 (2011).

15. C. E. Fay and R. L. Comstock, IEEE Trans. Microwave Theory Tech. 13, 15 (1965).

16. H. Bosma, Proceedings of the IEE - Part B: Electronic and Communication Engineering 109,
137 (1962).

17. J. D. Adam, L. E. Davis, G. F. Dionne, E. F. Schloemann, and S. N. Stizer, IEEE Trans.


Microwave Theory Tech. 50, 721 (2002).

18. G. F. Dionne, G. A. Allen, P. R. Haddad, C. A. Ross, and B. Lax, Lincoln Laboratory


Journal 15, 323 (2005).

36
CHAPTER 3

ROLES OF FERRITES IN RF ANTENNAS2

3.1. Introduction to ferrite antennas

An antenna is an indispensible device designed to transform an RF electric signal into

electromagnetic waves in free-space and widely used for transmitting and receiving the RF

signals in a wireless communication system. Modern electronic devices and mobile

communication systems increasingly demand a variety of wireless services, including AM/FM

radio, satellite radio, cellular phone communications, intelligent transportation systems, etc. as

given in Table 3.1. This is because mobile multimedia applications are a rapidly growing market.

Accordingly, a large number of RF antennas are demanded, and all of which need to be

embedded in the mobile devices. Thus, miniaturization, compact design, and easy integration

become important factors for RF antenna applications. At the same time, broad bandwidth and

high gain of the antenna are indispensable characteristics, because they ensure robust operation

and reliable communications.

Ferrites possess both relative permeability and permittivity greater than unity. Thus, those

attempts to achieve ferrite antenna miniaturization and the widening of bandwidth are receiving a

great deal of attention. In this chapter, fundamental characteristics of ferrite antennas are

described in terms of antenna miniaturization, radiation efficiency, bandwidth, and impedance

matching to free-space.

2
This work was published as “Chapter 8 – Ferrites for RF passive devices,” in Solid State Physics 64, 237
(2013) by Yang-Ki Hong and Jaejin Lee.

37
Table 3.1. Frequency bands for various wireless communications
Frequency bands Applications Frequency range (MHz)
VHF FM broadcasting radio 87.5-108
VHF & UHF Terrestrial digital video broadcasting 110-270
350-870
UHF GSM 850 824-894
L PCS 1900 1850-1990
UHF, L & S 4G LTE 746-787, 704-746, 1710-
1880, 1710-2155, 1920-
2170
L GPS 1574-1576
S XM satellite radio 2332.5-2345
S&C Wireless LAN (WLAN) 2400-2483
5150-5250
C Dedicated short-range communication 5850-5925
(DSRC)

3.2. Effective wavelength and antenna miniaturization

The wavelength (λ0) of an electromagnetic wave in vacuum is given as the spatial period

of the wave and is related to the frequency by the relation

c
0  , (3.1)
f

where c is the speed of light in vacuum (3 × 108 m/s) and f is the frequency of the wave in Hertz

(Hz). According to Eq. (3.1), the wavelength in vacuum is only determined by the frequency of

the electromagnetic wave and is inversely proportional to the frequency. As a result, the low-

frequency wave generates a longer wavelength, and the high-frequency wave has a shorter

wavelength.

When the electromagnetic wave travels in a uniform medium, the wavelength of Eq. (3.1)

becomes

38
0
eff  , (3.2)
n

where λeff is the effective wavelength in a medium and n is the refractive index of the medium.

Therefore, Eq. (3.2) suggests that the wavelength in a medium can be reduced by the refractive

index n of a medium as compared to the wavelength in the vacuum. The refractive index n

describes how the electromagnetic wave propagates through the medium and is given by

n  r r , (3.3)

where μr is the relative permeability and εr is the relative permittivity of the medium. Ferrites are

magneto-dielectric materials that possess both μr and εr higher than unity. Accordingly, the

wavelength associated with a ferrite is obtained by combining Eqs. (3.2) and (3.3)

0 c
eff   . (3.4)
r r f  r r

Note that the wavelength in a ferrite shortened by r r at a constant frequency in comparison

to the wavelength in the vacuum (μr = 1, εr =1). A decrease in the wavelength of the wave

propagating in a ferrite is illustrated in Fig. 3.1.

FIG. 3.1. Schematic illustration of the effective wavelength λeff in a ferrite medium in
comparison to the wavelength λ0 in vacuum.

39
Antenna size (or length) is determined by the wavelength of the electromagnetic wave at

a constant operating frequency of the antenna [1]. For example, a patch antenna, formed with

two conductive radiating and ground planes, has a resonant length of approximately one-half

wavelength of the wave. A monopole antenna consists of a straight conductor mounted vertically

over a ground plane, whose length is approximately a quarter of a wavelength. Therefore,

integration of a ferrite of higher μr and εr than unity into the antenna structure results in a

decreased wavelength, and consequently, an antenna size reduction. According to Eqs. (3.3) and

(3.4), the size of a monopole antenna surrounded by a ferrite in Fig. 3.2 can be given by

eff 0 c
l   , (3.5)
4 4 r  r 4 f r  r

where the length l of a quarter-wavelength monopole antenna is decreased by the refractive index

n of ferrite. The refractive index n is, therefore, referred to as the miniaturization factor. It is

demonstrated that the size of antennas can be effectively reduced by the miniaturization factor of

ferrite. Accordingly, high permeability ferrites are needed for the miniaturization of RF antennas.

In the next section, the effectiveness of ferrite permeability in the antenna bandwidth is

discussed.

(a) (b)
FIG. 3.2. Miniaturization of monopole antenna. (a) Air-core and (b) ferrite monopole antenna.

40
3.3. Bandwidth of ferrite antennas

Antenna bandwidth is a fundamental parameter and describes the range of frequencies

over which the antenna can properly transmit and receive the RF signals. In general, broadband

antennas are desired because a wide range of frequencies can be covered with a single antenna.

Therefore, there are many efforts to improve the antenna bandwidth. One approach is to add

impedance matching circuits to the antenna system in order to maximize the power transfer from

the source to the load. Also, the acceptable frequency range of an antenna can be increased by

optimizing of the design of antenna geometries. These techniques, however, are rather

complicated. In the last section, the capability of high permeability ferrite used for antenna

miniaturization was introduced. Furthermore, the ferrite can improve impedance matching, and

increase electromagnetic energy radiation due to its magneto-dielectric properties [2], resulting

in an increase in antenna bandwidth. In this section, the bandwidth equation of ferrite antenna in

association with magnetic and dielectric properties is derived to discuss the role of the ferrite in

antenna bandwidth.

A rectangular ferrite patch antenna in Fig. 3.3 is considered for the calculation of the

bandwidth. A ferrite substrate has the width of W, length of L, and thickness of d. The ground

plane is assumed to be infinite. The bandwidth (BW) of the patch antenna is related to the total

quality factor Q and can be given by [3]

1
BW  . (3.6)
2Q

The quality factor (Q) is defined by the ratio of the energy stored to the energy supplied by a

generator as given by

 WT
Q , (3.7)
Prad  Ploss

41
FIG. 3.3. Geometry of rectangular microstrip patch antenna with a ferrite substrate possessing
both relative permeability and permittivity higher than unity.

where ω is the angular frequency, WT is the total stored energy, Prad is the radiated power, and

Ploss is the dissipated power loss. In the case of a ferrite antenna, the power losses are mainly

associated with the magnetic and dielectric losses of the ferrite substrate. Therefore, the magnetic

and dielectric power losses of the ferrite patch antenna are first derived to calculate the antenna

bandwidth. From the cavity model [3], the electric and magnetic field components can be

expressed as [1]

E z  E 0 cos k g x  , (3.8a)

H y  H 0 sin k g x  , (3.8b)

Ex  Ey  H x  H z  0 , (3.8c)

n
where the wave number kg   k0 r r , and k0 is the free-space propagation constant (=
L

2 /  ).

The dielectric power loss of the ferrite antenna is obtained by the integral of the electric field

over the volume of the microstrip cavity. Then, the dielectric power loss PD is given by [3]

42
 " 2
PD   E dv , (3.9)
2

where ε″ is the imaginary part of permittivity. Applying Eq. (3.8) to (3.9), the dielectric power

loss can be obtained as follows

LW d
 "  0 r tan  
cos2 k g x dxdydz ,
2 2
PD   E dv   E 0 (3.10)
2 v 2 0 0 0

where the permittivity is     j  0r 1 j tan  . Applying the boundary conditions of

Ez 0, z   Ez L, z   Ez , Eq. (3.10) becomes

L
 0 r tan  
 cos k x dx .
2 2
PD   W  d  E0 g (3.11)
2 0

L
Solving  cos 2 k g x dx by letting k g x  t in Eq. (3.11) leads to the following equations
0

L kg L kg L kg L
1 1 L 1
 cos k g x dx 
2
 cos 2 t  dt   cos 2t  1dt    cos 2tdt . (3.12a)
0 0 kg 2k g 0 2 2k g 0

kg L
Again, solving for  cos 2t dt with 2t = u, Eq. (3.12a) can be reduced to
0

 n 
sin  2    L 
L sin 2k g L  L
kg L 2k g L
L 1 L 1 1
  cos 2tdt    cos udu      L 
2 2k g 0 2 2k g 0 2 2 4k g 2 n
4  , (3.12b)
L
n 0
 
2k 0  r  r 4  r  r

where sin (2nπ) is zero (for n =0,1,2,3.....).

Therefore,

L
2 n 2 0 2
E
0
dx 
2k 0  r  r
E0 
4 r r
E0 . (3.13)

43
Applying Eq. (3.13) to (3.11), the dielectric power loss is obtained as follows

 0 r tan   2 0 1 2 Wd tan   r
PD   W  d  E0  E0  2   0  c  ,
2 4 r r 2 4 r

1 2 Wd tan   r
PD  E0 , (3.14)
2 240 r

where ε0 is the permittivity of free-space (= 8.85 × 10-12 F/m). According to Eq. (3.14), the

dielectric power loss is proportional to the dielectric loss tangent (tan δε = ε″/ε′) of the ferrite

substrate.

Similarly, the magnetic power loss PM can be obtained by the integral of the magnetic

field over the volume of the microstrip cavity

" 2
PM   H dv, (3.15)
2 v

where μ″ is the imaginary part of permeability. Applying Eq. (3.8) to (3.15), the magnetic power

loss of the ferrite antenna is given by

LW d
 "  0  r tan  
sin 2 k g x dxdydz ,
2 2
PM   H dv   H 0 (3.16)
2 v 2 0 0 0

where the complex permeability is      j    0  r 1  j tan    . Using the boundary

conditions of H y 0, y   H y L, y   0 , Eq. (3.16) is written as

L
 0  r tan  
 sin k x dx .
2 2
PM  W  d  H 0 g (3.17)
2 0

L
By letting kgx = t, the solution of  sin 2 k g x dx can be obtained as follows
0

L kg L kg L kg L
1 1 L 1
 sin k x dx   sin t  k
2 2

0
g
0
dt 
2k g
 1  cos2t dt  2  2k  cos 2tdt .
0 0
(3.18a)
g g

44
kg L
Solving  cos 2t dt with 2t = u, Eq. (3.18a) can be reduced to
0

 n 
sin  2   L 
L sin 2k g L  L
kg L 2k g L
L 1 L 1 1
  cos 2tdt    cos udu      L 
2 2k g 0 2 4k g 0 2 2 4k g 2 n 
4 , (3.18b)
L
n 0
 
2k 0  r  r 4  r r

where sin (2nπ) is zero (for n =0,1,2,3.....).

Thus,

L
2 n 2 0 2
H
0
dx 
2k0  r  r
H0 
4  r r
H0 . (3.19)

Substitution of Eq. (3.19) into (3.17) gives the magnetic power loss PM by

0 r tan   2 0 1 2 Wd tan   r


PM  W  d  H 0  H0  2   0  c  , (3.20)
2 4 r  r 2 4 r

where H0 is the amplitude of the magnetic field and is expressed in terms of the electric field E0

as

 0 r
H 0  jE0 . (3.21)
0  r

Applying Eq. (3.21) to (3.20), the magnetic power loss PM of the ferrite antenna can be written in

terms of the electric field component as

1  0 r 2 Wd tan   r
PM  E0  2   0  c  ,
2 0  r 4 r

1 2 Wd tan   r
PM  E0 . (3.22)
2 240 r

45
By combining Eqs. (3.14) and (3.22), the total power loss of the ferrite patch antenna is given by

the sum of the magnetic and dielectric power losses

1 Wd r
Ploss  PM  PD  E0
2
tan    tan   . (3.23)
2 240  r

Now, the total stored energy WT of the ferrite patch antenna is given by [4]

LW d
1 L 1  1 L L

    
2 2 2 2 2 2
WT   ' E   ' H dv  Wd    ' E   ' H dx   Wd    ' E dx    ' H dx  ,
40 0 0 4 0  4 0 0 

(3.24)

where ε′ and µ′ are the real parts of permittivity and permeability, respectively. According to

Eqs. (3.13) and (3.19), the integrals of the electric and magnetic fields over the volume are given

by

L
2 n 2 0 2
E
0
dx 
2k 0  r  r
E0 
4 r r
E0 , (3.25a)

L
2 n 2 0 2 0  0 r 2
H
0
dx 
2k 0  r  r
H0 
4 r r
H0 
4 r r 0 r
E0 . (3.25b)

Substitution of Eqs. (3.25a) and (3.25b) into (3.24) gives the following equations for the total

stored energy of the ferrite patch antenna

1  0  0  1 20
WT  Wd   0 r E0
2 2 2
  0  r 0 r E0  Wd 0 r E0
4  4  r r 0  r 4  r  r  4 4  r r
 ,
1 2  r  0  c  2
 Wd E 0 
2 r 4

1 2 Wd r
WT  E0 . (3.26)
2 240 r

The radiated power of the patch antenna defined by the cavity model is given by [1]

46
1 2
Prad  E0 d 2 G r , (3.27)
2

where Gr is the radiation conductance of the patch antenna. Applying Eqs. (3.23), (3.26), and

(3.27) to (3.6), the bandwidth of ferrite patch antenna can be obtained as

1 P  PMD
BW   rad
2Q 2WT
1 
 E0 2 d 2 Gr  1 E0 2 Wd  r tan    tan    ,
1  2 2 240  r 


2 1 2 Wd r
E0
2 240  r

1  dGr r 
BW  240  tan   tan   . (3.28)
2  W r 

Equation (3.28) implies that the bandwidth of the ferrite antenna increases with r /  r .

Accordingly, an increase in the relative permeability μr and a decrease in the relative permittivity

εr of ferrite can enlarge the antenna bandwidth. It is also seen that broadband antenna can be

realized with increasing the magnetic loss (tan δμ) and dielectric loss (tan δε) of the ferrite.

However, a high loss ferrite in antennas also increases the power loss, thereby decreasing the

antenna radiation efficiency. Thus, the low-loss ferrite of high μr and low εr could be utilized to

devise antennas of attractive features such as broad bandwidth, miniature structure, and high

radiation efficiency. In the next section, the antenna radiation efficiency is discussed in terms of

the magnetic loss of ferrite.

3.4. Radiation efficiency of ferrite antennas

Radiation efficiency is another important parameter to describe how efficiently an

antenna transmits and receives RF signals, which is defined as the ratio of the total power

47
radiated by an antenna to the total input power received from the generator. An antenna with

high radiation efficiency efficiently radiates the input power to free-space. In the case of low

radiation efficiency, the input power is mostly dissipated because of the internal losses such as

metal conduction, dielectric and magnetic losses within the antenna. As presented in the last

section, the power loss of the ferrite antenna is closely related to the loss factors of tan δε and tan

δμ of the ferrite. Other contributions to antenna loss include input impedance mismatch and

polarization loss between transmitting and receiving antennas. However, these degradation

factors can be minimized by using impedance matching circuitry and proper positioning of the

antennas.

The basic operation of a transmitting antenna is illustrated in Fig. 3.4. A generator with

internal impedance Zg is connected to the transmitting antenna with impedance ZA

Z g  R g  jX g
, (3.29)

Z A  RA  jX A   Rr  RL   jX A . (3.30)

Rg and Xg are the resistance and reactance of generator impedance, respectively, RA is the antenna

resistance being the sum of radiation resistance Rr and loss resistance RL of the antenna, and XA is

FIG. 3.4. Transmitting antenna and its equivalent circuit. The radiated waves are in the form
of spherical wave in near-field region and become a plane wave in far-field region.

48
the antenna reactance. The generator transmits the input power to the antenna for

electromagnetic radiation. The radiated power Pr for the antenna can be written as

Pr  I g2 Rr , (3.31)

where

Vg
Ig  . (3.32)
Zg  ZA

Vg is the peak generator voltage and Ig is the generated current in the equivalent circuit. Since the

antenna has internal losses, some of the input power delivered to the antenna is dissipated as

heat. The power loss PL is given by

PL  I g2 RL . (3.33)

Therefore, the total input power of the antenna is then obtained as

PT  Pr  PL  I g2 Rr  I g2 RL . (3.34)

The radiation efficiency η is defined by the ratio of the power radiated over the input power

Pr Pr
  . (3.35)
PT Pr  PL

Applying Eqs. (3.31) and (3.33) to (3.35) leads to the following equation of the radiation

efficiency

I g2 Rr Rr 1
   . (3.36)
I g2 Rr  I g2 RL Rr  RL 1  RL
Rr

This demonstrates that high radiation efficiency can be achieved by decreasing the loss

resistance and increasing the radiation resistance.

49
The radiation efficiency η can also be expressed in terms of quality factor Q by Eq. (3.37)

[5]

Rr 1 Qr QT
   , (3.37)
Rr  RL 1 QT Qr

where QT is the total quality factor of antenna, and Qr is the radiated quality factor (= 1/Rr). The

total quality factor QT describes the total antenna losses including radiation loss, conduction loss,

dielectric and magnetic losses and is given by [1, 2]

1 1 1 1 1
    ,
QT Qr Qc Q Q

1 1 1
   tan    tan   , (3.38)
QT Qr Qc

where Qμ is the quality factor due to magnetic loss (= 1/tan δμ), and Qε is the quality factor due to

dielectric loss (= 1/tan δε). Equations (3.37) and (3.38) demonstrate that magnetic and dielectric

losses lower the radiation efficiency. Therefore, a ferrite of low magnetic and dielectric losses is

desirable in order to increase the antenna radiation efficiency. It was shown, in the preceding

sections, that ferrite antenna size and bandwidth are reduced and increased, respectively,

according to Eqs. (3.4) and (3.28). To realize a high efficiency, broadband, and miniature RF

ferrite antenna for wireless communication systems, the high permeability and low-loss ferrites

are imperative, and also antenna structure needs to be optimized.

3.5. Reflection loss and impedance matching to free-space

Recall that a ferrite having both relative permeability and permittivity gives rise to size

reduction and bandwidth broadening for the antenna. In this section, enhancement of the

characteristic impedance matching of the antenna to free-space is discussed. In a ferrite antenna,

50
the ferrite permeability can reduce the capacitive property, which a dielectric antenna holds.

Therefore, better characteristic impedance matching to the free-space is expected for a ferrite

antenna. The improved impedance matching of the ferrite antenna enables an efficient radiation

of electromagnetic energy to free-space [2, 6]. In addition, the quality factor of a resonant

antenna can be decreased by reducing of the stored electric energy within the antenna, which

leads to an increase in radiation efficiency and bandwidth. In this section, the reflection loss and

input impedance of a ferrite will be derived to discuss the characteristic impedance matching

between a ferrite and the free-space.

As shown in Fig. 3.5, a plane wave is generated in the free-space region z < 0. The lossy

medium region z > 0 has both relative permeability μr and relative permittivity εr higher than

unity. Due to the characteristic impedance mismatch at the interface, a part of the incident wave

is transmitted to the region 2, namely the ferrite, and another part is reflected back to the region

1, namely the free-space. Figure 3.5 illustrates the incident, reflected, and transmitted fields. The

incident electric and magnetic fields can be expressed as [7]

FIG. 3.5. Illustration of incident, reflected, and transmitted electric fields at the interface of
two different characteristic impedances.

51
E i  xˆ E 0 e j  t  k 0 z  , (3.39a)

1
Hi  yˆ E0e jt k0z  , (3.39b)
0

where k0 is the propagation constant of free-space, η0 is the wave impedance of free-space, and

E0 is the amplitude of electric field. The reflected fields at the interface between free space and

ferrite are given by

E r  xˆE0 e j t  k0 z  , (3.40a)


H r   yˆ E0 e j t k0 z  , (3.40b)
0

where Γ is the reflection coefficient of the electric field. Note that since the incident and reflected

fields travel in the opposite direction, the sign of propagation constant k0 is positive for the

reflected field. The transmitted fields in a lossy ferrite can be given with the transmission

coefficient T by

Et  xˆTE0 e j t  "z  , (3.41a)

T
H t  yˆ E0e j t  "z  , (3.41b)

where γ″ is the propagation constant and η is the intrinsic impedance of the lossy ferrite.

According to the boundary conditions, the tangential components of the electric and magnetic

fields must be continuous at the interface z = 0. Since the interface between the two regions is

located at z = 0, the exponent of e j  t  kz  will reduce to e j t . Therefore, at the boundary z = 0,

the electric and magnetic fields become

Ei  Er  Et , (3.42a)

Hi  Hr  Ht . (3.42b)

52
Substituting Eqs. (3.39a, b), (3.40a, b) and (3.41a, b) into (3.42a, b) gives the following

equations

E 0 e  t   E 0 e  t  TE 0 e t , (3.43a)

1  T
E0et  E0et  E0et . (3.43b)
0 0 

After simple rearrangement of Eqs. (3.43a, b), the following expressions are obtained.

1    T , (3.44a)

1  T
 . (3.44b)
0 

According to Eqs. (3.44a, b), the reflection Γ and transmission T coefficients can be specified as

  0
 , (3.45)
  0

2
T  1   . (3.46)
  0

It is noted that if the characteristic impedances of the two media (regions) are known, the

propagation characteristics of the waves can be determined. According to Eqs. (3.45) and (3.46),

all electromagnetic energy can be transmitted without reflection from region 2 to region 1 when

the characteristic impedances of both regions are equal. Since reflection loss (RL) is the

magnitude of the reflection coefficient (Γ) in dB, the reflection loss is

  0 Z  Z0
RL  20 log10   20 log10  20 log10 , (3.47)
  0 Z  Z0

0  r
where   Z  .
 0 r

53
FIG. 3.6. Metal-backed ferrite structure and its transmission line model terminated with a
short circuit.

FIG. 3.7. A lossy transmission line terminated in a load impedance ZL.

In order to calculate the input impedance between the ferrite and the free-space for the

metal-backed ferrite structure presented in Fig. 3.6, a lossy transmission line terminated in load

impedance ZL [7] is first considered as shown in Fig. 3.7. Assume that an incident wave is

generated from the source in the free-space region z < 0 and propagating into the lossy region.

Since there is an impedance mismatch between the two regions, the total voltage on the line is

given as a sum of incident and reflected waves by

V  z   V0 e   "z  V0 e  " z , (3.48)

54
where V 0 and V 0 are the amplitude of the incident and reflected voltage, respectively. The

propagation constant γ″ is given by

 "    j , (3.49)

where α is the attenuation constant and β is the propagation constant for the lossless case. The

total current on the line can be obtained by dividing the total voltage of (3.48) by the

characteristic impedance Z0

V0 V0
I z   e  " z  e "z . (3.50)
Z0 Z0

Next, the load impedance ZL is defined by the ratio of the total voltage to the current at the load z

= 0. Using Eqs. (3.48) and (3.50), the load impedance ZL can be given by

V 0  V0  V0
ZL   Z0 . (3.51)
I 0  V0  V0

After rearranging Eq. (3.51), the amplitude of the reflected voltage V 0 can be obtained as

Z L  Z0
V0  V0 . (3.52)
Z L  Z0

Since the characteristic impedance, Z0 is not matched to the load impedance ZL, a part of the

waves is reflected at the interface of the load. The reflection coefficient Γ at the load is then

given by the ratio of the reflected to the incident voltage wave amplitudes at z = 0

V0 ZL  Z0
  . (3.53)
V0 ZL  Z0

After substitution of Eq. (3.53) into (3.48) and (3.50), the total voltage and current on the line

can be written as

V  z   V0 e   " z  e  " z , (3.54)

55
V0
I z   e  "z
 e  " z  . (3.55)
Z0

Using Eqs. (3.54) and (3.55), the input impedance Zin at z = −l, a distance l from the load,

z = 0, can be calculated as follows

V0 e "l  e  "l 


V  l  e "l   e  "l
Z in     Z0 . (3.56)
I  l  V  "l  "l
0
e  e  e  e
 "l   "l

Z0

Substitution of the reflection coefficient Γ (= ZL  Z0  /ZL  Z0  ) into Eq. (3.56) gives the input

impedance in the following form

Z L  Z 0 e "l  Z L  Z 0 )e  "l  Z e "l  e  "l   Z 0 e  "l  e  "l 


Z in  Z 0  Z0 L . (3.57)
Z L  Z 0 e "l  Z L  Z 0 )e  "l  Z L e  "l  e  "l   Z 0 e "l  e  "l 

Dividing Eq. (3.57) by 1/ e   l


 e  l  gives

Z e  "l
 e "l   Z 0 e "l  e  "l 
1 
Z L  Z 0
e "l  e  "l 
L

Z in  Z 0
e  "l
 e "l  
 Z0 
e "l  e  "l  . (3.58)
1  e "l  e  "l  
Z e
L
 "l
 e  "l   Z 0 e "l  e  "l  Z L  Z0 
e  "l
 e "l 
 e  e 
 "l  "l


After applying tanhx  e  e  x x


/e  e  to
x x
Eq. (3.58), the input impedance of a lossy

transmission line terminated in a load impedance ZL can be reduced to

Z L  Z 0 tanh " l 
Z in  Z 0 . (3.59)
Z 0  Z L tanh " l 

Applying the above input impedance of Eq. (3.59) to the earlier example of the metal-backed

ferrite structure (see Fig. 3.6), this structure can be considered as a lossy microstrip line

terminated in a short circuit. Thus, the load impedance ZL must be set to zero. Applying ZL = 0 to

Eq. (3.59) gives the input impedance Zin of the lossy transmission line

56
Zin  Z tanh "l  , (3.60)

where the characteristic impedance of the ferrite Z is given by

0  r r
Z  Z0 . (3.61)
 0 r r

After applying the propagation constant    j 0 r  0 r  j 2f / c r r to Eq. (3.60), the

input impedance Zin can be written in the final form as

r  2fl 
Z in  Z tanh " l   Z 0 tanh j r  r  , (3.62)
r  c 

where l is the thickness of the ferrite.

Recall (for Eq. (3.47)) that all energy can be transmitted without reflection from region 1

(free-space) to region 2 (ferrite) when the characteristic impedances of both media are equal.

The impedance matching between the ferrite and the free-space can be improved according to

Eq. (3.62), because the ferrite possesses both permeability and permittivity, in contrast to the

impedance matching between the free-space and dielectric medium. Therefore, the ferrite

medium allows good impedance matching and electromagnetic energy radiation over wide

frequency range.

It was demonstrated that ferrite provides high miniaturization factor ( r r ), wide

bandwidth, and good impedance matching to free-space for RF antenna applications. Antenna

radiation efficiency is inversely proportional to magnetic and dielectric losses of ferrite.

Therefore, high permeability and low loss ferrite are desired for constructing miniature and high-

efficiency RF antennas.

57
References

1. C. A. Balanis, Antenna Theory: Analysis and Design, Wiley, New Jersey (2005).

2. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag. 55, 45 (2007).

3. R. Garg, P. Bhartia, I. Bahl, and A. Ittipiboon, Microstrip Antenna Design Handbook, Artech
House, Massachusetts (2001).

4. C. Niamien, S. Collardey, A. Sharaiha, and K. Mohdjoubi, IEEE Antennas Wirel. Propag.


Lett. 10, 63 (2011).

5. K. R. Carver, J. W. Mink, IEEE Trans. Antennas Propag. AP-29, 2 (1981).

6. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag. 52, 1558 (2004).

7. D. M. Pozar, Microwave Engineering, Wiley, New Jersey (2004).

58
CHAPTER 4

FUNDAMENTALS OF FERRITE INDUCTORS3

4.1. Introduction to ferrite inductors

An inductor is an important passive component used in many RF devices, such as DC-

DC converters, power delivery systems, voltage regulation, microwave and RF circuitry. There

have been a lot of efforts to increase the inductance density and quality-factor of the inductors

[1−9]. For magnetic inductors, metallic magnetic materials [10−12] are widely used as magnetic-

core to increase the inductance and are easily fabricated by physical vapor deposition techniques.

The magnetic metal-core inductor is usually fabricated on a silicon wafer and shows a significant

increase in inductance due to high magnetic permeability of the magnetic-core as compared to

air-core inductor. However, the quality-factor of this magnetic metal-core inductor is limited to

less than 10 at high frequencies [10, 11]. This is because there is large eddy current loss due to

the nature of high electrical conductivity of metal and reasonably high conductivity of silicon

substrate. To reduce the eddy current loss, MEMS (microelectromechanical systems) is

employed to eliminate the inductor’s underlying silicon substrate and build a suspended inductor

structure. In this way, reduction of eddy current loss and minimal parasitic capacitance in the

silicon substrate are possible, thereby resulting in a quality-factor higher than 25 at high

frequencies [13]. However, the absence of magnetic core in MEMS inductors results in low

inductance because there is no enhanced magnetic flux density. To overcome these low

inductance and quality factor issues, many researchers have considered ferrite film inductors

because of its high magnetic permeability and electrical resistivity, and good chemical stability.

Therefore, ferrites can facilitate high inductance and quality-factor in inductors [14−17]. Various

3
This work was published as “Chapter 8 – Ferrites for RF passive devices,” in Solid State Physics 64, 237
(2013) by Yang-Ki Hong and Jaejin Lee.

59
FIG. 4.1. Various ferrites for RF applications at wide operating frequency.

ferrites are available for RF applications at wide operating frequencies as presented in Fig. 4.1.

This chapter focuses on the role of ferrites in integrated inductor performance and the design of

the integrated ferrite inductors.

4.2. Inductance

Inductance is defined as the ratio of magnetic flux to current and describes how much

magnetic energy can be stored in an inductor. The inductance L of a conductor carrying the

current I can be expressed as


L , (4.1)
I

where the magnetic flux Φ is given by

   0  r  H  ds
. (4.2)

60
It is seen that the inductance is proportional to the magnetic flux, which is proportional to the

relative permeability μr of a magnetic core. A wire inductor with a rectangular cross-section area

and magnetic film is taken for expression of magnetic inductance. The inductance of the

magnetic core inductor can be expressed by the sum of the air-core inductance LAC and the

additional contribution ΔL of the magnetic film to inductance [18]

LMC  LAC  L . (4.3)

The air-core inductance LAC for a planar inductor with rectangular cross section depends

primarily on the geometry of a coil structure, which is given by [19]

 2l wt 
L AC  2l  ln  0.5  , (4.4)
 wt 3l 

where l is the length, w is the width, and t is the thickness of the coil. The additional inductance

ΔL from the magnetic film can be expressed by the relation [20−22]

tm  wm 
L   0  eff   , (4.5)
2  lm 

where tm, wm and lm are the thickness, width, and length of the magnetic film, respectively. μeff is

the effective permeability, which is dependent on magnetic film geometry

r
 eff  . (4.6)
1  N d  r  1

Equations (4.3), (4.4) and (4.5) imply that the inductance of a magnetic inductor can be

significantly enhanced by the effective permeability of the magnetic film as compared to the air-

core inductor. Consequently, high inductance density can be achieved, leading to small area

consumption in integrated power modules and RF circuits. In the next section, Q-factor

enhancement with a ferrite based on lumped model of planar spiral inductor is discussed.

61
4.3. Quality factor

One of the important performance parameters of an inductor is the quality (Q) factor. The

Q is defined as the ratio of the magnetic energy stored in an inductor to the energy loss per cycle

Energy stored Magnetic energy  Electric energy


Q  2   2  . (4.7)
Energy loss per cycle Energy loss per cycle

Equation (4.7) implies that with higher Q-factor, there is higher stored magnetic energy in an

inductor. The high Q means low energy loss. Thus, Q-factor of inductor needs to be increased for

RF integrated circuits. The series RLC circuit for a simple inductor model is shown in Fig. 4.2. In

the series RLC circuit, the resistor R and capacitor C are the loss factor and parasitic capacitance

of the inductor, respectively. The magnetic energy stored in an inductor can be given by the

integral of the instantaneous power

I
t t t di I 1 L 2
E Magnetic   pdt   ivt dt   iL dt  L  idi  L i 2
0 0 0 dt 0 2

2
I 0
0 , (4.8)
1
 LI 2
2

where L is the inductance and I is the current. Similarly, the electric energy stored in a capacitor

can be written as

2
1 1 1 2 1 1
CVc 2  1 CVc2  C  I  1 
Q Q
E Electric 
C 
0
qdq 
C2
q
0

2C

Q2 0 2C 2 2  C 
. (4.9)
1 I2

2  2C

FIG. 4.2. Series RLC circuit for a simple inductor model.

62
Then, the energy loss per cycle dissipated by a resistor is given by

1  cos2t dt   T  sin T  


T T V 02 V02 T 1 V 02  1
E Loss   pdt   sin t dt 
2

0 0 R R 0 2 R 2 4 
. (4.10)
2
2 V0 2 1 2
    I R
 2R  2

Applying Eqs. (4.8), (4.9), and (4.10) to (4.7) gives the following equation of the Q-factor for an

inductor

1 2 1 I2
LI 
Magnetic energy  Electric energy 2 2  2C
Q  2   2 
Energy loss per cycle 2 1 2
 I R
 2
LI 2  2 C  I 2
 2  2 2 C  2 

 LI 2  2 C  I 2

 
LI 2  2 C  I 2

 
L 2 C  1
,
 (4.11)
I 2 R I 2 R 2 2 C I 2 RC RC

2
L    0  
  1    
R     

where ω0 is the LC resonance frequency (= 1 LC ). From Eq. (4.11), it is noted that the

inductance L needs to be larger than the resistance R in order to achieve a high Q-factor. Also,

when the self-resonance frequency ω of the inductor is equal to the LC resonance frequency ω0,

the Q-factor tends to decrease to zero.

Equation (4.11) is employed to discuss the Q-factor of a magnetic-core inductor (QMC).

The Q-factor of the magnetic inductor is given in terms of air-core component and magnetic core

contributions [18]

L MC   L AC  L 
Q MC   . (4.12)
R MC R AC  R 
Specifically, the inductance of the magnetic-core inductor (LMC) in Eq. (4.12) is the sum of air-

core inductance and the additional inductance ΔL from the magnetic core of Eq. (4.3). Similarly,

63
the resistance of the magnetic-core inductor results from both conductor power and magnetic-

core losses. The loss of the air-core inductor is mainly attributed to eddy current of conductor

coils at high frequencies. On the other hand, the magnetic-core loss is related to magnetic loss

tangent (tan δμ) (recall Eq. (3.22)). Since the magnetic loss tangent is a frequency dispersive

characteristic, the Q-factor of the magnetic-core inductor tends to decrease significantly at high

frequencies as compared to the air-core inductor. Approaches to enhance the Q-factor of ferrite

inductor are discussed based on the equivalent circuit of an integrated spiral inductor.

To derive the Q-factor of integrated spiral inductor, the definition of Q-factor in Eq. (4.7)

and the equivalent circuit [23] in Fig. 4.3 are used. First, the stored magnetic energy of an

inductor in the equivalent circuit is given by

1  
2
1 2 1  V0  V0   V 02 Ls
E Magnetic  Ls I s  Ls    Ls , (4.13)
2 2  Z L  2 


Ls 2  R s2  2 Ls   R s
2 2

where Ls is the spiral inductance, Rs is the series resistance of the conductor coils, and V0 is the

peak voltage across the inductor. The electric energy stored in the parallel capacitors is expressed

in terms of the total capacitance as

1 2 1
E Electric  V 0 C  V 02 C p  C s  , (4.14)
2 2

where Ctotal is the total capacitance of the parallel capacitors, Cp is the parasitic capacitance in the

FIG. 4.3. Equivalent circuit of an integrated spiral inductor.

64
substrate, and Cs is the spiral capacitance between the coils and underpass of the center pad. The

energy loss per cycle dissipated by resistors is given by [23]

2
2 V0  1 Rs 
E Loss     , (4.15)
 2  R p Ls 2  Rs2 

where Rp is the parasitic resistance in the substrate. Applying Eqs. (4.13), (4.14), and (4.15) to

(4.7) gives the expression for the Q-factor of integrated spiral inductor

V02 L s V 02 C p  C s 

Q  2 
Magntic energy  Electric energy
 2 
2 L s   2
 R s2  2
Energy loss per cycle 2 V 0
2
 1 Rs 
   2 
 R p Ls   R s 
2
 2
Ls Ls
 C p  C s    C p  C s 


L s   R
2 2
s  
 L2s  R s2  2

1 1 Rs   2 L2s  R s2  R p R s
  
  R p L s  2  R s2  R p  2 L2s  R s2 
Ls R p R p C p  C s  2 L2s  R s2 
 
 2 L2s  R s2  R p R s  2 L2s  R s2  R p R s
Ls Rp 
 C p  C s   2 L2s  R s2  
   1  
Rs  2 L2s 
 Ls 
Rp   Rs
Rs
Ls Rp  R s2 C p  C s    2 L2s C p  C s 
   1  
Rs  2 L2s  L 
Rp   Rs  s

Rs

L s Rp  R s2 C p  C s   2 L2s C p  C s 
Q  2
 1   . (4.16)
Rs  L    Ls Ls 
R p   s   1 R s
 R s  

Equation (4.16) suggests that Rs and Cp need to be minimized, but large Rp is required for high

Q-factor inductors. The decrease in Rs and Cp can be realized using high conductive spiral coils

and thick oxide layers on the substrate, respectively. Regarding the parasitic resistance Rp in the

65
substrate, the use of high resistivity substrate leads to an increase in Rp, resulting in a low eddy

current loss. It is noted that the magnetic insulation layer that possesses high permeability, and

electrical resistivity, and good insulation property can increase both inductance and Q-factor of

the integrated inductor (refer to Eqs. (4.3) and (4.16)). Ferrites are insulating magnetic materials

with high permeability and electrical resistivity. Therefore, high inductance density and Q-factor

at high frequencies are expected.

4.4. Design of integrated ferrite inductors

RF designers and engineers have strived to enhance performances of integrated ferrite

inductors. The figure of merit for integrated ferrite inductors includes inductance, quality factor,

inductance density, and maximum rated DC current. To meet the desired performance of

inductor for RF applications, the design parameters such as coil thickness, number of coil turns,

coil width, coil area, and space between neighboring coils needs to be carefully optimized. Also,

the construction of the ferrite core for the integrated inductor is important to increase the

magnetic flux and confine it closely to the inductor, thereby effectively increasing the inductance

per a given area. Figure 4.4 shows different magnetic-core constructions: (a) air-core, (b) one-

side ferrite (bottom), (c) two-side ferrite (top and bottom), and (d) closed ferrite structure. The

air-core inductor does not use a magnetic core and consists of air as an inductor-core. For the

(a) (b) (c) (d)

FIG. 4.4. Various ferrite-core constructions for integrated RF inductors: (a) air-core, (b) one-
side ferrite (bottom), (c) two-side ferrite (top and bottom), and (d) closed ferrite structure.

66
(a) (b) (c) (d)

FIG. 4.5. Simulated magnetic-field distributions of ferrite inductor with various ferrite-core
constructions: (a) air-core, (b) one-side ferrite (bottom), (c) two-side ferrite (top and bottom),
and (d) closed ferrite structure in accordance with Figure 4.4.

closed ferrite structure, the conductive coils are surrounded by the ferrite core, resulting in the

formation of a closed magnetic circuit. In this configuration, the magnetic flux passes primarily

through high permeability ferrite core and is therefore confined within the ferrite without

escaping outside the ferrite core. As a result, a higher magnetic flux and inductance can be

achieved, since inductance is proportional to the magnetic flux according to Eq. (4.1). Figure 4.5

shows the simulated magnetic-field distribution of the magnetic-core structures presented in Fig.

10000
Air-core
1Side
2Side
Closed
Inductance (nH)

1000

100

10 100
Frequency (MHz)
FIG. 4.6. Simulated inductance of the ferrite inductors with various ferrite-core constructions
in accordance with Figure 4.4.

67
4.4. It is seen that uniform magnetic field distribution and small leakage of the magnetic field are

observed from the two-side and the closed ferrite inductors compared to the air-core and the one-

side ferrite inductors. The ferrite inductors, therefore, show larger inductance enhancement than

air-core inductor as shown in Fig. 4.6.

68
References

1. S. Bae, Y. K. Hong, J. Lee, G. Abo, J. Jalli, A. Lyle, H. Han, and G. W. Donohoe, J. Magn.
13, 37 (2008).

2. R. F. Jiang, N. N. Shams, M. T. Rahman, and C. H. Lai, IEEE Trans. Magn. 43, 3930 (2007).

3. M. Yamaguchi, M. Baba, and K. I. Arai, IEEE Trans. Microwave Theory Tech. 49, 2331
(2001).

4. D. S. Gardner, G. Schrom, F. Paillet, B. Jamieson, T. Karnik, and S. Borkar, IEEE Trans.


Magn. 45, 4760 (2009).

5. Y. J. Kim and M. G. Allen, IEEE Trans. Compon. Packag. Manuf. Technol. Part C 21, 26
(1998).

6. W. Wu, F. Huang, Y. Li, S. Zhang, X. Han, Z. Li, Y. Hao, and Y. Wang, IEEE Microwave
Wireless Compon. Lett. 15, 853 (2005).

7. J. B. Yoon, Y. S. Choi, B. I. Kim, Y. Eo, and E. Yoon, IEEE Electron Device Lett. 23, 591
(2002).

8. C. H. Ahn and M. G. Allen, IEEE Trans. Ind. Electron. 45, 866 (1998).

9. J. Lee, S. Park, H. C. Kim, and K. Chun, J. Micromech. Microeng. 19, 085014 (2009).

10. W. Xu, H. Wu, D. S. Gardner, S. Sinha, T. Dastagir, B. Bakkaloglu, Y. Cao, and H. Yu, J.
Appl. Phys. 109, 07A316 (2011).

11. H. Wu, D. S. Gardner, W. Xu, and H. Yu, IEEE Trans. Magn. 48, 4123 (2012).

12. J. W. Park and M. G. Allen, IEEE Trans. Magn. 39, 3184 (2003).

13. E. C. Park, Y. S. Choi, J. B. Yoon, S. Hong, and E. Yoon, IEEE Trans. Microwave Theory
Tech. 51, 289 (2003).

14. J. Lee, Y. K. Hong, S. Bae, J. Jalli, J. Park, G. S. Abo, G. W. Donohoe, and B. C. Choi, IEEE
Trans. Magn. 47, 304 (2011).

15. I. Kowase, T. Sato, K. Yamasawa, and Y. Miura, IEEE Trans. Magn. 41, 3991 (2005).

16. S. Bae, Y. K. Hong, J. Lee, J. Jalli, G. S. Abo, A. Lyle, B. C. Choi, and G. W. Donohoe,
IEEE Trans. Magn. 45, 4773 (2009).

17. M. Saidani and M. A. M. Gijs, Appl. Phys. Lett. 84, 4496 (2004).

18. D. W. Lee, K. P. Hwang, and S. X. Wang, IEEE Trans. Magn. 44, 4089 (2008).

69
19. H. M. Greenhouse, Parts, Hybrids, and Packaging, IEEE Transactions on 10, 101 (1974).

20. D. W. Lee and S. X. Wang, J. Appl. Phys. 103, 07E907 (2008).

21. R. F. Soohoo, IEEE Trans. Magn. MAG-15, 1803 (1979).

22. L. Li, D. W. Lee, K. P. Hwang, Y. Min, T. Hizume, M. Tanaka, M. Mao, T. Schneider, R.


Bubber, and S. X. Wang, IEEE Trans. Adv. Packag. 32, 780 (2009).

23. C. P. Yue and S. S. Wong, IEEE J. Solid-State Circuits 33, 743 (1998).

70
CHAPTER 5

SMALL PERMEABILITY IN ANTENNA PERFORMANCE 4

5.1. Introduction

Ferrite possesses both frequency-dependent permeability (μr) and permittivity (εr). The

wavelength (λ) of incident wave in a ferrite decreases by λ = λ0/(μrεr)0.5, where λ0 is the free space

wavelength. Therefore, miniaturization of antenna and electromagnetic scaling are realized [1, 2].

Furthermore, the permeability can increase the bandwidth of antenna [3]. However, the ferrite

ferromagnetic resonance frequency (fFMR) needs to be higher than the antenna operation

frequency, and high permeability is desired for GHz antenna application. This is because the

magnetic loss tangent (μ″/μ′) of ferrite significantly increases near the fFMR, and bandwidth and

miniaturization factor increase with permeability. In addition, high permeability is desired to

achieve good impedance matching. On the contrary, permeability decreases with increasing fFMR,

according to the Snoek limit in Eq. (5.1) [4].

4
f FMR  r  1  M s , (5.1)
3

where γ is the gyromagnetic constant and Ms is saturation magnetization. Therefore, high

permeability of ferrite is difficult to obtain in the GHz frequency range.

One way to achieve high fFMR is to use a high magneto-crystalline anisotropy hexaferrite.

This can be understood by the relation fFMR = (γ/2π)∙(HθHϕ)0.5, where Hθ is the out-of-plane

anisotropy field and Hϕ is the in-plane anisotropy field. Y-type (Co2Ba2Fe12O22) and Z-type

(Ba3Co2Fe24O41) hexaferrites show 28,000 Oe of Hθ and 155 Oe of Hϕ, and 13,000 Oe of Hθ and

112 Oe of Hϕ, respectively. These anisotropies correspond to fFMR of 5.8 GHz and 3.4 GHz,

4
This work was published as “Role of small permeability in gigahertz ferrite antenna performance,” in
IEEE Magnetics Letters 4, 5000104 (2013) by Jaejin Lee, Yang-Ki Hong, et al.

71
respectively. Recently, gigahertz (GHz) Y-type hexaferrite has been investigated to achieve

small form factor and broadband of antenna for mobile communications [5, 6].

It was, however, reported that the hexagonal ferrites have permeability smaller than 2.5

and permittivity smaller than 13 in the low end of the GHz range [5−7]. Therefore, a specific

question arises as to whether such small permeability of the GHz hexaferrites is effective enough

for improving antenna miniaturization, impedance matching, and bandwidth as compared to high

permittivity dielectrics.

In this chapter, we report experimental and simulation results of a 1.57 GHz Z-type

hexaferrite chip antenna (permeability < 2) and simulated results of high permittivity (9.4 or 11)

dielectric chip antenna to assess the effectiveness of small permeability.

5.2. Hexaferrite substrate

Z-type hexaferrite (Co2Z: Ba3Co2Fe24O41) was synthesized by firing a mixture of Y-type

(Co2Y: Ba2Co2Fe12O22) and M-type (BaM: BaFe12O19) hexaferrites. Y-type hexaferrite was

prepared by conventional ceramic process. BaCO3, CoO, and α-Fe2O3 were used as starting

materials. The mixture of starting materials was wet shake-milled for 2 h with motor speed of

1725 rpm (Spex SamplePrep 8000D). For conversion of starting materials to Y-type hexaferrite,

heat treatment was performed at 1100 oC for 5 h in O2. A mixture of M- (Sigma-Aldrich, -325

mesh, 98 %) and Y-type hexaferrite precursors in equal molar ratio was shake-milled for 4 h and

heated at 1300 oC for 4 h in O2.

Gigahertz Co2Z hexaferrite-glass composite was prepared by compacting a mixture of

Co2Z hexaferrite powder and 2 wt% borosilicate glass into a rectangular or toroidal ring mold for

use in antenna substrate or dynamic magnetic and dielectric measurement. The composite body

72
8
0.14
7

Permeability and permittivity


0.12
6 ' '
" "
0.10

tan  or tan 
5 tan  tan 
0.08
4

3 0.06

2 0.04

1 0.02

0 0.00
1.0 1.2 1.4 1.6 1.8 2.0 2.2 2.4 2.6 2.8 3.0
Frequency (GHz)

FIG. 5.1. Measured complex permeability and permittivity of Co2Z hexaferrite-glass


composite.

was sintered at 950 oC for 1 hr, machined, and grinded. The sintered hexaferrite composite

toroidal ring was precisely polished with the dremel tool and abrasive sand papers to fit into the

airline fixture having 7 mm outer diameter and 3 mm inner diameter. A network analyzer

(Agilent N5230A) with a coaxial airline fixture (Agilent 85051-60007) was used to measure

permeability and permittivity dispersion. Fig. 5.1 shows the measured complex permeability (μ′

− j μ″) and permittivity (ε′ − jε″) of the hexaferrite composite. The real part of permeability and

permittivity was 1.97 and 7.36 at 1.61 GHz, respectively. Magnetic loss tangent (tan δμ) of 0.054

and dielectric loss tangent (tan δε) of 0.008 were obtained at 1.61 GHz. It was found that tan δμ

increased with frequency and was less than 0.1 up to 2.7 GHz. The measured magnetic and

dielectric properties were used to design a GHz mobile hexaferrite chip antenna.

5.3. Simulation of GHz hexaferrite chip antenna

73
A hexaferrite chip antenna was designed and simulated to investigate the effectiveness of

small permeability in GHz antenna performance. Fig. 5.2 shows the designed hexaferrite antenna.

A conductive spiral radiator was disposed on a ferrite substrate (8 × 5 × 1.5 mm3). FR4 board

(100 × 50 mm2) with ground size of 80 × 50 mm2 was employed as a system board. This design

was applied to both hexaferrite and dielectric antennas for performance comparison. The

ANSYS high frequency structure simulator (HFSS v. 11) was used to simulate antenna

performance. In the simulation, various small permeability values (μr = 1.4, 1.6, 1.97, 2.2, and

2.5) of the hexaferrite were used. These small values were chosen because the hexaferrite

possesses small permeability value in the GHz range due to the Snoek limit as previously

described. Also, the following permittivity and loss values were used in the antenna performance

simulation: experimental εr of 7.36 for hexaferrite; 9.4 for alumina; experimental tan δμ of 0.05

and tan δε of 0.008 for hexaferrite; tan δε of 0.01 for alumina [8]. In addition, an artificial alumina

with high tan δε of 0.05 was simulated for the effect of large dielectric loss tangent on voltage

FIG. 5.2. Designed antenna geometry and structure of hexaferrite-glass composite chip
antenna.

74
5.0

4.5

4.0

3.5
VSWR

3.0
r = 1.4; r = 7.36
2.5 r = 1.6; r = 7.36
r = 1.97; r = 7.36
2.0
r = 1.97; r = 9.4
1.5 r = 2.2; r = 9.4
r = 1; r = 9.4
1.0
1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8
Frequency (GHz)
FIG. 5.3. Simulated VSWR of the designed hexaferrite and dielectric antennas.

Table 5.1. Simulated resonance frequency, VSWR, and fractional bandwidth for hexaferrite (tan
δμ = 0.05 and tan δε = 0.008) and dielectric antennas.
FBW (%) at
Substrate Material properties fr (GHz) VSWR at fr VSWR =
2.5:1
Hexaferrite μr = 1.4; εr = 7.36 1.66 1.38 6.3
μr = 1.6; εr = 7.36 1.62 1.32 6.2
μr = 1.97; εr = 7.36 (exp.) 1.58 1.28 6.3
μr = 1.97; εr = 9.4 1.46 1.05 5.9
μr = 2.2; εr = 9.4 1.43 1.06 5.6
μr = 2.5; εr = 9.4 1.41 1.06 5.6
Alumina εr = 9.4; tan δε = 0.01 1.65 1.79 4.1
εr = 9.4; tan δε = 0.05 1.65 1.16 6.0

standing wave ratio (VSWR) and bandwidth. Fig. 5.3 shows the simulated VSWR of the

designed hexaferrite antenna in comparison with the high permittivity antenna. Table 5.1

summarizes the simulated resonance frequency (fr), VSWR, and fractional bandwidth (FBW) of

the hexaferrite and dielectric antennas. It was found that even small permeability (μr > 1.6) of the

75
GHz hexaferrite substrate had greater contribution to lowering of the antenna fr than the high

permittivity antenna. At experimental μr of 1.97 and εr of 7.36, the hexaferrite antenna showed

the fr of 1.58 GHz, while the alumina antenna resonated at 1.65 GHz. Further simulation showed

that the 8 × 5 × 5 mm3 alumina antenna resonates at the same 1.58 GHz as the 8 × 5 × 1.5 mm3

ferrite antenna does. This demonstrates a 70 % reduction of GHz antenna volume when the

hexaferrite substrate is used. Furthermore, VSWR and FBW of the hexaferrite antenna were

found to be lower and wider than the dielectric antenna, respectively. This suggests that VSWR

and bandwidth can be enhanced with even small permeability (< 2) and relatively high loss

tangent of hexaferrite substrate. It is noted that, as given in Table 1, VSWR decreased and FBW

of artificial alumina antenna also increased as tan δε increased to 0.05. However, the FBW of the

hexaferrite antenna decreased from 6.3 % to 5.9 % as its dielectric constant increased from 7.36

to 9.4 at constant μr of 1.97 and tan δμ of 0.05. This is attributed to an increase in capacitive

property due to high permittivity of ferrite [9]. Consequently, the quality factor (Q = 2ωW/Prad,

where W is stored energy and Prad is radiated power) increased, thereby decreasing bandwidth.

Radiation efficiency is another antenna performance characteristic. It is known that

relatively high magnetic loss of ferrite decreases antenna radiation efficiency (RE). Therefore,

RE in terms of the magnetic loss tangent was simulated to gain insight into radiation

performance improvement with smaller permeability than 2 at GHz. The experimental μr of 1.97

(tan δμ = 0.05, 0.03 and 0.01) and εr of 7.36 with experimental tan δε of 0.008 were used in the

simulation. On the other hand, the dielectric constant (εr) for the high permittivity antenna was

chosen to be 11 by parametric simulation to resonate at the same frequency of 1.58 GHz of the

hexaferrite antenna. Various dielectric tan δε values of 0.002, 0.008 and 0.05 were used in RE

simulation. Antenna RE can be expressed by Eq. (5.2) [10]

76
Prad
RE  , (5.2)
Prad  Ploss

where Prad is radiated power and Ploss is power loss associated mainly with loss tan δµ and tan δε

of the antenna substrate. Fig. 5.4 shows the simulated RE of the hexaferrite antenna in the GHz

range in comparison with high permittivity antenna. The simulated maximum RE was 80 % for

the hexaferrite antenna having tan δμ of 0.01 and 87.8 % for the dielectric antenna with tan δε of

0.002. The hexaferrite antenna showed a significant increase in RE from 66 % to 80 % as tan δµ

decreased from 0.05 (experimental) to 0.01. It is noted that much higher RE of 66 % was

obtained for the hexaferrite antenna having tan δμ of 0.05 and tan δε of 0.008 than the RE of 55.4 %

for the lossy dielectric antenna (εr = 11; tan δμ = 0; tan δε = 0.05) despite both magnetic and

dielectric losses of the ferrite. This is because the hexaferrite antenna has low electromagnetic

(EM) energy concentration around the antenna substrate, leading to radiation of EM energy [3].

90
80
Radiation efficiency (%)

70
60
50
40
30
20 tan  = 0.05 tan  = 0.05
10 tan  = 0.03 tan  = 0.008
tan  = 0.01 tan  = 0.002
0
1.50 1.53 1.56 1.59 1.62 1.65 1.68
Frequency (GHz)
FIG. 5.4. Simulated radiation efficiency of the hexaferrite (closed symbol) and dielectric (open
symbol) antennas in terms of magnetic or dielectric losses.

77
It is also noted that broader frequency range of high RE was observed for the hexaferrite antenna

than the dielectric antenna. Simulation results demonstrate that small permeability (< 2) of GHz

hexaferrite significantly contributes to antenna RE and bandwidth.

5.4. Fabrication and measurement results

A GHz hexaferrite chip antenna was fabricated based on the designed structure in Fig. 5.2

to confirm the simulated antenna performance. Performance of the fabricated antenna was

characterized in terms of return loss, RE, and 3-dimensional gains. Fig. 5.5(a) presents a photo-

image of the fabricated hexaferrite antenna. Conductive copper tape (3M copper foil tape 1181)

was disposed on the Co2Z hexaferrite-glass composite substrate having a volume of 8 × 5 × 1.5

mm3. The double-sided copper clad laminate (CCL) FR4 board was used as a testing board and

fabricated by chemical etching process. The hexaferrite composite chip antenna was mounted on

the 100 × 50 mm2 CCL FR4 board and fed by a semi-rigid 50 ohm coaxial cable. Fig. 5.5(b)

shows the measured and simulated return losses of the hexaferrite antenna. The return loss of 13

dB at 1.57 GHz and bandwidth of 110 MHz (1.51 GHz ~ 1.62 GHz) at VSWR of 2.5:1 were

obtained. The measured return loss was in good agreement with the simulated result. The

hexaferrite antenna was characterized in an anechoic chamber system (volume: 10 × 7 × 7 m3 ;

shielding: -105 dB; ripple < ± 1 dB in 0.17 - 3 GHz) with a network analyzer (Agilent

ENA5070B) for the radiation performance. Fig. 5.6 shows the experimental and simulated RE,

and 3-dimensional (3D) peak- and average-gains in the range of 1.51 to 1.63 GHz. The

experimental radiation efficiency values were in good agreement with the simulation results. The

maximum RE was measured to be 66 % at 1.57 GHz. This low RE is attributed to relatively high

magnetic loss tangent (tan δμ = 0.05) of the hexaferrite composite substrate. However, the

78
(a)
0
-2
-4
-6
-8
S11 (dB)

-10
-12
-14
-16 Measured
-18 Simulated
-20
1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0
Frequency (GHz)
(b)

FIG. 5.5. (a) Fabricated Co2Z hexaferrite-glass composite chip antenna and (b) measured and
simulated return losses.

simulated radiation efficiency suggests that the RE can be increased to 80 % with a low tan δμ of

0.01. The fabricated hexaferrite antenna showed an insignificant change in RE in the range of

1.53 to 1.6 GHz. Regarding 3D gains, the hexaferrite antenna had the maximum peak gain of

2.56 dBi at 1.53 GHz and average gain of -1.81 dBi at 1.57 GHz.

79
5 100
4 90
3
80
2

Radiation efficiency (%)


1 70
0
Gain (dBi)

60
-1
50
-2
-3 40
-4 30
-5 Measured 3D peak gain
Simulated 3D peak gain 20
-6 Measured RE
Measured 3D Avg. gain 10
-7 Simulated 3D Avg. gain Simulated RE
-8 0
1.50 1.52 1.54 1.56 1.58 1.60 1.62 1.64
Frequency (GHz)

FIG. 5.6. Measured (closed symbol) and simulated (open symbol) 3-dimensional peak- and
average-gains and radiation efficiency of the fabricated hexaferrite composite antenna.

Our simulation and experimental results demonstrate that even small permeability (< 2)

of GHz hexaferrite substrate significantly contributes to antenna miniaturization, bandwidth, and

impedance matching. Since both domain wall motion and spin rotation are involved in the

permeability dispersion [11], magnetic loss can be reduced by controlling grain size or applying

a small static magnetic field. Therefore, low-loss hexaferrite is under development.

5.5. Conclusion

Gigahertz hexaferrite antenna was designed and fabricated to assess the effectiveness of

small permeability in improvement of GHz antenna performance. Both antenna performance

simulation and experimental results demonstrate that small permeability (< 2) of hexaferrite is

sufficiently effective in antenna miniaturization, impedance matching, and bandwidth. It was

80
found that the radiation efficiency can be increased to 80 % at tan δμ of 0.01 from experimental

66 % at tan δμ of 0.05. Therefore, small permeability and low-loss hexaferrite will impact future

advanced miniature and high-efficiency GHz mobile antennas.

81
References

1. J. Lee, Y. K. Hong, S. Bae, G. S. Abo, W. M. Seong, and G. H. Kim, IEEE Antennas


Wireless Propag. Lett., 10, 603 (2011).

2. H. Moon, G. Y. Lee, C. C. Chen, and J. L. Volakis, IEEE Antennas Wireless Propag. Lett.,
11, 322 (2012).

3. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag., 55, 45 (2007).

4. J. Smit and H. P. J. Wijn, Ferrites, John Wiley & Sons, New York, 271 (1959).

5. Y. Cheon, J. Lee, and J. Lee, IEEE Antennas Wireless Propag. Lett., 11, 137 (2012).

6. J. Lee, J. Heo, J. Lee, and Y. Han, IEEE Trans. Antennas Propag., 60, 2080 (2012).

7. J. Lee, Y. K. Hong, S. Bae, J. Jalli, G. S. Abo, J. H. Park, W. M. Seong, S. H. Park, and W. K.


Ahn, J. Appl. Phys., 109, 07E530 (2011).

8. Y. T. J. Charles, V. Ungvichian, and J. A. Barbosa, J. Comput., 4, 610 (2009).

9. P. M. T. Ikonen, K. N. Rozanov, A. V. Osipov, P. Alitalo, and S. A. Tretyakov, IEEE Trans.


Antennas Propag., 54, 3391 (2006).

10. C. A. Balanis, Antenna Theory, John Wiley & Sons, New Jersey, 85 (2005).

11. G. T. Rado, R. W. Wright, and W. H. Emerson, Phys. Rev., 80, 273 (1950).

82
CHAPTER 6

CONTROL OF MAGNETIC LOSS OF FERRITES5

6.1. Introduction

Modern electronic devices demand compact-size and high-performance antennas for

reliable wireless communications. Accordingly, ferrite antennas have been of great interest [1−4].

This is because miniaturization factor (μrεr)0.5 and bandwidth of ferrite antenna increase with

permeability (μr) [5, 6]. However, relatively high magnetic loss of ferrite is an important issue to

address due to low antenna radiation efficiency. Therefore, the magnetic loss limits wide spread

use of ferrite for RF antenna applications.

For a polycrystalline ferrite, magnetic loss tangent (tan δμ) can be described by three main

contributions in Eq. (6.1) [7].

tan    tan  h  tan  e  tan  r , (6.1)

where tan δh is the hysteresis loss, tan δe is the eddy current loss, and tan δr is the residual loss.

The hysteresis loss (tan δh) is the energy dissipation of static hysteresis loop. The eddy current

loss (tan δe) is dependent on the electrical resistivity of the ferrite. The residual loss tangent (tan

δr) is associated with domain wall and spin rotational resonances [8, 9]. It is known that domain

wall resonance contributes to low-frequency permeability dispersion, while spin rotational

resonance is dominant in high-frequency permeability dispersion [10]. In response to this, the

control of grain size and microstructure, and optimization of sintering process were investigated

to achieve low-loss ferrite [11-13]. It was reported that domain wall resonance was suppressed

and magnetic loss tangent (tan δμ) decreased. Nevertheless, magnetic loss of ferrite was still high

compared to low-loss dielectric materials for RF antenna applications.

5
This work was published as “Control of magnetic loss tangent of hexaferrite for advanced RF antenna
applications,” Journal of Applied Physics 113, 073909 (2013) by Jaejin Lee, Yang-Ki Hong, et al.

83
DC magnetic field effect on complex permeability of hexagonal and spinel ferrites, and

garnet was studied [14-17]. Ferromagnetic resonance frequency (fFMR), according to Snoek’s

product, was found to be controllable via applied DC magnetic field. Accordingly, we have

applied these results to reduce and understand the magnetic loss tangent of soft

BaFe9.6Co1.2Ti1.2O19 hexaferrite with a small DC magnetic field for magneto-dielectric antenna

applications. The significance of our work is to realize and overcome high magnetic loss of

ferrite for antenna applications using a different approach from material processing aspects. Both

permeability higher than unity and low-magnetic loss of the ferrite are capable of antenna size

reduction, widening of bandwidth, and an increase in antenna radiation efficiency.

In this chapter, we report control of the magnetic loss tangent of soft BaFe9.6Co1.2Ti1.2O19

hexaferrite with a DC magnetic field and ferrite antenna performance. Ferrite antenna simulation

based on experimental magnetic properties confirmed that RF ferrite antenna performance can be

improved with a small DC magnetic field.

6.2. Experiment

Conventional ceramic process was used to prepare BaFe9.6Co1.2Ti1.2O19 (Co/Ti-

substituted BaM) powder. Raw materials of BaCO3, Fe2O3, CoO, and TiO2 were mixed by

shake-milling and followed by calcination of the mixture at 1100 °C for 5 hr. The resulting

Co/Ti-substituted BaM powder was shake-milled for 8 hr prior to sintering. Ferrite green bodies

were fabricated by compacting the Co/Ti-substituted BaM powder. The green bodies were then

heated to 1100 °C (first-step temperature) and rapidly quenched to 930 °C (second-step

temperature). The quenched sample was held at 930 °C for 8 hr before cooling to room

temperature.

84
Static magnetic properties of the sintered Co/Ti-substituted BaM were measured with a

vibrating sample magnetometer (MicroSense EV9). An impedance/material analyzer (Agilent

E4991A) with a magnetic material test fixture (Agilent 16454A) was used to characterize

dynamic magnetic properties in the range of 10 MHz to 1 GHz. Various DC magnetic fields

were produced with an electromagnet and applied to a ferrite toroidal ring sample to measure

magnetic-field dependence of complex permeability (µ′ - jµ′′). The ferrite ring had 7 mm of outer

diameter and 3 mm of inner diameter. An impedance/material analyzer with a dielectric material

test fixture (Agilent 16453A) was also used for complex permittivity measurement.

6.3. Experimental results

Figure 6.1(a) shows the measured magnetization virgin curve and hysteresis loop (inset)

of the sintered Co/Ti-substituted BaM. The measured virgin curve shows a sharp increase in

magnetization in the range of 0 to 400 Oe. This suggests that the magnetization is due to a

dominant domain wall motion. As illustrated in Fig. 6.1(b), domain wall motion is the main

process up to about the knee of the magnetization curve [18]. By further applying DC magnetic

field, domain wall motion can be complete and spin rotation becomes predominant. Saturation

magnetization and coercivity of Co/Ti-substituted BaM were measured to be 44.5 emu/g and 29

Oe, respectively. The coercivity dramatically decreased to 29 Oe from 5 kOe of pure BaFe12O19

(BaM) [19]. This is because cobalt (Co2+) and titanium (Ti4+) cations preferentially occupy the

2b (magnetic spin-up), 4f2 (magnetic spin-down), and 12k (magnetic spin-up) crystallographic

sites of BaM [20]. It has been known that the 2b magnetic site in BaM hexaferrite has the largest

magneto-crystalline anisotropy.

85
40
35

Magnetization (emu/g)
50
30 40

Magnetization (emu/g)
30
25 20
10
20 0
-10
15 -20
-30
10 -40
-50
5 -10 -8 -6 -4 -2 0 2 4 6 8 10
Applied field (kOe)
0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Applied field (kOe)
(a)

(b)

FIG. 6.1. (a) Measured magnetization virgin curve and hysteresis loop (inset) of
BaFe9.6Co1.2Ti1.2O19 hexaferrite and (b) illustration of magnetization processes.

86
Magnetic-field dependence of complex permeability was measured to investigate the DC

magnetic field effect on magnetic loss tangent. Fig. 6.2(a) and (b) show measured real (µ′) and

imaginary (µ″) part of permeability of Co/Ti-substituted BaM at various DC magnetic fields.

Real part of permeability decreased from 10.5 to 3.4 at 200 MHz with increasing the magnetic

field from 0 to 400 Oe. The FMR frequency, i.e. at the maximum µ″, shifted from 917 MHz to

higher frequency (not shown in Fig. 6.2(b)). According to Snoek’s limit in Eq. (6.2), the product

of the ferromagnetic resonance frequency (ωr) and permeability (μr) becomes constant. As the

applied magnetic field increases, the resonance frequency increases. Consequently, the

corresponding permeability decreases.

2M s
r  μr  1  , (6.2)
3

2M s
where r   Ha  H0  and  μr  1  .
3H a  H 0 

The experimental results are in good agreement with the Snoek’s limit. Furthermore, it was

observed that magnetic loss tangent (tan δµ = µ″/µ′) at 200 MHz decreased significantly to less

than 1 % at 400 Oe from 11.8 % without a DC magnetic field. Measured µ′ and tan δµ at 200,

500, and 800 MHz are summarized in terms of various DC magnetic fields in Table 6.1.

The experimental results suggest that permeability higher than unity and a decrease in

magnetic loss tangent are achievable with small DC magnetic field, leading to antenna

miniaturization and high radiation efficiency.

87
Exp. H = 0 Oe Curve fitted H = 0 Oe
14
Exp. H = 100 Oe Curve fitted H = 100 Oe
Exp. H = 200 Oe Curve fitted H = 200 Oe

Real part of permeability


12 Exp. H = 400 Oe Curve fitted H = 400 Oe

10

0
10 100 1000
Frequency (MHz)
(a)
Exp. H = 0 Oe Curve fitted H = 0 Oe
8 Exp. H = 100 Oe
Imaginary part of permeability

Curve fitted H = 100 Oe


Exp. H = 200 Oe Curve fitted H = 200 Oe
Exp. H = 400 Oe Curve fitted H = 400 Oe

0
10 100 1000
Frequency (MHz)
(b)

FIG. 6.2. Measured and curve-fitted complex permeability spectra of BaFe9.6Co1.2Ti1.2O19


hexaferrite at various applied magnetic fields: (a) real and (b) imaginary part of permeability.

88
Table 6.1. Measured µ′ and tan δµ for Co/Ti-substituted BaM with various applied magnetic
fields.
Frequency
0 Oe 100 Oe 200 Oe 400 Oe
(MHz)
200 µ′ 10.5 8.3 5.5 3.4
tan δµ 0.118 0.086 0.038 Negligible
500 µ′ 10.3 8.53 5.82 3.62
tan δµ 0.462 0.349 0.197 0.082
800 µ′ 7.16 6.81 5.5 3.71
tan δµ 1.016 0.763 0.445 0.209

6.4. Discussion

The experimental permeability spectra in Fig. 6.2 were curve-fitted with Eqs. (6.3), (6.4),

and (6.5) [21] in order to separate the domain wall contribution to permeability from spin

rotation.


μ r  μ   jμ  1   DW   sp  1   ' DW  j ' ' DW   ' sp  j ' ' sp   
  
 1   ' DW   ' sp  j  ' ' DW   ' ' sp  , (6.3)

1
2
 K DW
DW

 sp  i  sp K sp
2
 DW   2  i  sp  i 2   2

μ'  1 
K spsp2

2
K DW DW 2
DW 
2 
, (6.4)
2
 
sp
2
 2
DW  2 2
  β 2 2

2
K spsp K DW  DW  
μ"   , (6.5)
 sp2   2  2
DW 2  2
  2 2

where DW′ and DW′′ are real and imaginary susceptibility, KDW is static magnetic susceptibility,

ωDW (=2πfDW) is angular resonance frequency, and β is damping factor of domain wall motion.

sp′ and sp′′ are real and imaginary susceptibility, Ksp is static magnetic susceptibility, and ωsp

89
(=2πfsp) is angular resonance frequency of spin rotation. The curve-fitted real and imaginary

parts of permeability in Fig. 6.2 are in good agreement with the measured permeability spectra.

The fitted parameters are given in Table 6.2. Based on the fitted parameters in Table 6.2,

complex susceptibility of domain wall motion and spin rotation was calculated in Fig. 6.3. It is

found that low-frequency permeability is mainly attributed to the domain wall motion, while spin

rotation is dominant at high frequency. As the DC magnetic field increases from 0 to 400 Oe, the

contribution of domain wall motion decreases and spin rotation becomes dominant.

The calculated real part of susceptibility and magnetic loss tangent of domain wall

motion and spin rotation at 200, 500, and 800 MHz are plotted as a function of DC magnetic

field in Fig. 6.4. The DW′ at 200 and 500 MHz decreased more rapidly than the sp′ with

increasing the magnetic field from 0 to 200 Oe and was followed by a slow decrease up to 400

Oe. This indicates that the domain wall motion in the magnetization process is partially

completed by the magnetic field of 200 Oe. At the magnetic field of 400 Oe, the DW′ became

lower than sp′, which implies that spin rotation is predominant. This is in good agreement with

the experimental magnetization virgin curve in Fig. 6.1. As a result, the contribution of domain

wall motion to permeability dispersion decreased [10, 17]. It was also found that the sp′ at 800

MHz was higher than the DW′. This is because high-frequency permeability is mainly attributed

to spin rotation.

Regarding magnetic loss tangent, the loss tan δDW (= DW′′/ DW′) was observed to be larger

than the loss tan δsp (= sp′′/ sp′) at 0 Oe. As the DC magnetic field increased to 400 Oe, both loss

tan δDW and tan δsp decreased. It was noted that the loss tan δDW at 800 MHz was significantly

reduced with DC magnetic field. This is because of an increase in the resonance frequency of fDW

(at the maximum DW″). Fig. 6.5 shows the magnetic-field dependence of resonance frequency of

90
Table 6.2. Curve-fitted parameters for measured complex permeability of Co/Ti-substituted
BaM with various magnetic fields
H (Oe) KDW fDW (GHz) β (×109) Ksp fsp (GHz)
0 5.3 0.89 4.61 4.4 1.51
100 3.5 0.98 4.48 3.8 2.10
200 1.31 1.12 3.40 3.02 2.39
400 0.6 1.38 2.91 1.82 2.99
Susceptibility of domain wall motion

10
'DW at 0 Oe 'DW at 200 Oe
9
''DW at 0 Oe ''DW at 200 Oe
8 'DW at 100 Oe 'DW at 400 Oe
7 ''DW at 100 Oe ''DW at 400 Oe
6
5
4
3
2
1
0
10 100 1000
Frequency (MHz)
(a)

'sp at 0 Oe 'sp at 200 Oe


Susceptibility of spin rotation

7
''sp at 0 Oe ''sp at 200 Oe
6 'sp at 100 Oe 'sp at 400 Oe
5 ''sp at 100 Oe ''sp at 400 Oe

4
3
2
1
0
10 100 1000
Frequency (MHz)
(b)
FIG. 6.3. Calculated complex susceptibility of (a) domain wall motion and (b) spin rotation
for various DC magnetic fields.

91
6

Real part of susceptibility (')


'DW 'sp at 200 MHz
5 'DW 'sp at 500 MHz
'DW 'sp at 800 MHz
4

0
0 100 200 300 400
Applied magnetic field (Oe)
(a)
4.0
tan DW tan sp at 200 MHz
3.9
tan DW tan sp at 500 MHz
3.8
Magnetic loss tangent

tan DW tan sp at 800 MHz


3.7
3.6

1.5
1.0
0.5
0.0
0 100 200 300 400
Applied magnetic field (Oe)
(b)

FIG. 6.4. Magnetic-field dependence of (a) real part of susceptibility and (b) magnetic loss
tangent of domain wall motion and spin rotation.

domain wall motion and spin rotation. Both the fsp (at the maximum sp″) and fDW increase with

the magnetic field. This suggests that magnetic loss can be reduced, because ferrite becomes very

lossy near the resonance. Accordingly, magnetic loss tangent (tan δµ) was measured to decrease

from 11.8 % at 200 MHz and 101 % at 800 MHz without the magnetic field to less than 1 % and

92
3.5
fDW

Resonance frequency (GHz)


3.0 fsp
Fitted fFMR = Ha + H0)
2.5 (with Ha = 650 Oe)

2.0

1.5

1.0

0.5
0 100 200 300 400
Applied dc magnetic field (Oe)
FIG. 6.5. Magnetic-field dependence of resonance frequency of domain wall motion and spin
rotation.

20 % at 400 Oe, respectively. The fsp can be expressed by the ferromagnetic resonance frequency

(fFMR) [22]. Therefore, the measured fsp was fitted with the equation fFMR = (γ/2π) × (H0 + Ha),

where H0 is an applied DC magnetic field and Ha is the magnetic anisotropy field. Fig. 6.5 shows

that the experimental fsp is in good agreement with the fitted fFMR at Ha of 650 Oe. This magnetic

anisotropy field is almost the same field as reported for polycrystalline BaFe9.6Co1.2Ti1.2O19

hexaferrite with Ha less than 1 kOe [23].

6.5. Antenna simulation

Antenna radiation efficiency (RE) is an important characteristic and decreases with

magnetic loss tangent of ferrite antenna substrate. The RE can be expressed by the relation in Eq.

(6.6)

93
Prad Prad
RE   , (6.6)
Ptotal Prad  Ploss

where Ptotal is total input power, Prad is radiated power and Ploss is power loss associated mainly

with loss tan δµ and tan δε of ferrite substrate. In order to demonstrate antenna performance

improvement using a DC magnetic field, a ferrite helical antenna in Fig. 6.6 was designed, and

antenna radiation efficiency and 3-dimensional (3D) peak gain were simulated. A 3D finite

element method simulator (ANSYS HFSS v. 11) was used for the antenna performance

simulation. The ferrite antenna has a ferrite substrate with a volume of 42 mm × 11 mm × 3.8

mm and an 8.5 turns of a helical radiator. As a system board, a FR4 substrate (100 mm × 50 mm)

having conductive ground of 80 mm × 50 mm was used, which is suitable for mobile handheld

devices. In the simulation, experimental magnetic-field dependent permeability of the Co/Ti-

substituted BaM was used with the measured dielectric properties. The measured real part of

permittivity (ε′) and dielectric loss tangent (tan δε) of the Co/Ti-substituted BaM were 11 and 0.6 %

FIG. 6.6. Designed ferrite helical antenna for performance simulation.

94
Table 6.3. Simulated hexaferrite antenna performance at various applied magnetic fields.

H (Oe) fr (MHz) RE (dB) at fr Peak gain (dB) at fr


0 181 -22.9 -21.2
100 194 -20.6 -18.9
200 222 -16.8 -14.9
400 260 -9.2 -7.0

-6 3D peak gain
-8 Radiation efficiency (RE)
RE and 3D peak gain (dB)

-10
-12
-14 tan  = 0.001
(negligible)
-16
-18 tan  = 0.118
-20
tan  = 0.038
-22
-24
0 100 200 300 400
Applied magnetic field (Oe)
FIG. 6.7. Simulated radiation efficiency and 3-dimensional peak gain of the designed ferrite
helical antenna as a function of applied magnetic field (i.e. magnetic loss tangent).

at 200 MHz, respectively. The simulated radiation efficiency and 3D peak gain at antenna

resonant frequency (fr) are summarized in Table 6.3.

Figure 6.7 shows simulated RE and 3D antenna gain as a function of applied magnetic

field. Both antenna RE and 3D gain dramatically increased with increasing magnetic field, which

resulted from the decreasing magnetic loss tangent (tan δµ). The RE increased to -9.2 from -22.9

dB when the field was 400 Oe, and the 3D peak gain increased to -7.0 from -21.2 dB. This is

because of low tan δµ due to the decrease in contribution of domain wall motion and increase in

95
ferromagnetic resonance frequency. It is also noted that the antenna resonant frequency (fr)

shifted from 181 to 260 MHz with increasing DC magnetic field. The resonant frequency shift of

the ferrite antenna can be understood by a reduction in permeability with the high DC magnetic

field according to the relation fr ~ 1/(μrεr)0.5. Furthermore, the ferrite antenna performance was

compared with a FR4 dielectric antenna having εr of 4.4 and tan δε of 0.02. The FR4 dielectric

antenna resonated at 404 MHz with the same antenna structure. This suggests that the hexaferrite

substrate largely contributes to antenna miniaturization factor (n = (μrεr)0.5). To decrease the fr of

the FR4 dielectric antenna to 260 MHz, near the ferrite antenna fr, the number of helical turns

needs to increase from 8.5 to 16. The simulated radiation efficiency of the FR4 dielectric antenna

was -9.5 dB at 260 MHz, which is lower than the ferrite antenna.

The experimental and simulation results demonstrated that magnetic loss tangent and

antenna radiation efficiency of the BaFe9.6Co1.2Ti1.2O19 with c-axis anisotropy can be controlled

by a small applied DC magnetic field, and antenna can be miniaturized. Accordingly, it is

suggested that the application of high-flux permanent magnet can allow the development of

novel magneto-dielectric antenna substrate when in combination with material processing

techniques. The magnetic anisotropy effect on magnetic loss tangent is under investigation for

BaFe12-2xCoxTixO19 (in-plane anisotropy for x > 1.3) [24] antenna substrate.

6.6. Conclusion

M-type BaFe9.6Co1.2Ti1.2O19 was prepared and used for RF antenna design and simulation.

The effects of DC magnetic field on magnetic loss tangent and antenna radiation performance

were investigated. Magnetic loss tangent (tan δµ) decreased dramatically from 11.8 % to less

than 1 % at 200 MHz with increasing DC magnetic field from 0 to 400 Oe. Antenna performance

96
simulation showed that a small DC magnetic field increased antenna radiation efficiency of the

designed ferrite helical antenna from -22.9 to -9.2 dB. Miniature and high-efficiency ferrite

antennas can therefore be realized with a small DC magnetic field.

97
References

1. S. Bae, Y. K. Hong, J. Lee, W. M. Seong, J. S. Kum, W. K. Ahn, S. H. Park, G. Abo, J. Jalli,


and J. H. Park, IEEE Trans. Magn. 46, 2361 (2010).

2. A. D. Brown, J. L. Volakis, L. C. Kempel, and Y. Y. Botros, IEEE Trans. Antennas Propag.


47, 26 (1999).

3. F. Farzami, K. Forooraghi, and M. Norooziarab, IEEE Antennas Wirel. Propag. Lett. 10,
1540 (2011).

4. P. M. T. Ikonen, K. N. Rozanov, A. V. Osipov, P. Alitalo, and S. A. Tretyakov, IEEE Trans.


Antennas Propag. 54, 3391 (2006).

5. R. C. Hansen and M. Burke, Microw. Opt. Technol. Lett. 26, 75 (2000).

6. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag. 52, 1558 (2004).

7. J. Moulson and J. M. Herbert, Elctroceramics: Materials, Properties, Applications, (John


Wiley & Sons, 2003).

8. S. Chikazumi, Physics of Ferromagnetism (John Wiley & Sons, 1964).

9. D. Stoppels, J. Magn. Magn. Mater. 160, 323 (1996).

10. G. T. Rado, R. W. Wright, and W. H. Emerson, Phys. Rev. 80, 273 (1950).

11. J. Lee, Y. K. Hong, S. Bae, J. Jalli, G. S. Abo, J. Park, W. M. Seong, S. H. Park, and W. K.
Ahn, J. Appl. Phys. 109, 07E530 (2011).

12. J. Lee, Y. K. Hong, W. Lee, G. S. Abo, J. Park, N. Neveu, W. M. Seong, S. H. Park, and W.
K. Ahn, J. Appl. Phys. 111, 07A520 (2012).

13. H. Su, X. Tang, H. Zhang, Z. Zhong and J. Shen, J. Appl. Phys. 109, 07A501 (2011).

14. Y. Bai, J. Zhou, Z. Yue, Z. Gui, and L. Li, J. Appl. Phys. 98, 063901 (2005).

15. G. G. Bush, J. Appl. Phys. 64, 5653 (1988).

16. T. Kato, H. Mikami, and S. Noguchi, J. Appl. Phys. 108, 033903 (2010).

17. T. Tsutaoka, T. Kasagi, and K. Hatakeyama, J. Appl. Phys. 110, 053909 (2011).

18. B. D. Cullity and C. D. Graham, Introduction to Magnetic Materials (John Wiley & Sons,
2009).

19. S. Y. An, I. B. Shim and C. S. Kim, J. Appl. Phys. 91, 8465 (2002).

98
20. X. Z. Zhou, A. H. Morrish, Z. W. Li, and Y. K. Hong, IEEE Trans. Magn. 27, 4654 (1991).

21. Mu, Y. Liu, Y. Song, L. Wang, and H. Zhang, J. Appl. Phys. 109, 123925 (2011).

22. T. Tsutaoka, T. Nakamura, and K. Hatakeyama, J. Appl. Phys. 82, 3068 (1997).

23. D. J. Bitetto, J. Appl. Phys. 35, 3482 (1964).

24. A. D. Shchurova, T. M. Perekalina, and V. P. Cheparin, Zh. Eksp. Teor. Fiz. 55, 1197 (1968).

99
CHAPTER 7

FERRITE ANTENNA GAIN WITH DC MAGNETIC FIELD6

7.1. Introduction

The technological trends in mobile communication devices are moving towards a variety

of functions and ensure high data rate and communication reliability. Therefore, size reduction

and control of radiation pattern of high-efficiency radio frequency (RF) antennas have been

issues to address. In response to these trends, both relative permeability (μr) and permittivity (εr)

are considered in miniature antenna design. This is because the antenna miniaturization factor is

(μrεr)0.5. On the other hand, the material losses such as conductive, dielectric, and magnetic

losses play a dominant role in antenna efficiency, i.e., gain.

A ferrite possesses both μr and εr greater than unity. Therefore, integration of the ferrite

into an antenna structure decreases the effective wavelength of an incident wave and low

electromagnetic (EM) energy concentrated around an antenna. Antenna miniaturization and

broadening of bandwidth are realized with ferrite [1-7]. However, relatively high magnetic loss

of ferrite is detrimental to the antenna gain. Therefore, low loss ferrite having large μr and εr is

indispensable [8, 9].

A DC bias magnetic field was applied to a ferrite antenna to lower magnetic tangent loss,

thereby improving antenna gain [10-17]. It was observed that antenna gain was improved with

the DC magnetic field. All studied antenna sizes are in the range of centimeter, and the antenna

radiation patterns are directional in GHz range. Their applications target mobile device

applications, but hand-held mobile devices with omnidirectional radiation patterns.

6
This work was published as “Electrically small ferrite antenna gain with dc magnetic field for mobile
device application,” Microwave and Optical Technology Letters 56, 1531 (2014) by Jaejin Lee, Yang-Ki
Hong, et al.

100
In this chapter, we have used a high loss ferrite for MHz antenna to gain the insight into

the effectiveness of DC magnetic field in antenna gain and demonstrate the electrically small

ferrite antenna with omnidirectional radiation patterns, aiming at hand-held mobile device

application.

7.2. Antenna design and fabrication

We designed the ferrite antenna shown in Fig. 7.1 using a high permeability

Ni0.5Mn0.2Co0.07Fe2.23O4 (Ni-Mn-Co) ferrite substrate [18]. The antenna has a volume of 14.5 mm

× 7.35 mm × 3.25 mm, and the antenna size (i.e. diagonal length = 16.25 mm) is about 0.046λ (λ

= 349 mm at 860 MHz). Therefore, this antenna is electrically small [19]. A folded radiator

increases the electrical length, thereby reducing the antenna size. For a testing board, the 110 mm

× 60 mm double-sided copper clad laminated (CCL) FR4 substrate having a ground of 90 mm ×

60 mm was used. ANSYS High Frequency Structure Simulator (HFSS ver. 11) was used to

obtain scattering parameters (S-parameters) and 2-dimensional (2D) radiation patterns. We have

FIG. 7.1. Designed geometry and dimension of Ni-Mn-Co ferrite antenna.

101
FIG. 7.2. Fabricated Ni-Mn-Co ferrite antenna with Nd-Fe-B permanent magnets.

used the measured values of ferrite permeability and magnetic tangent loss to simulate antenna

performance.

Based on the designed antenna in Fig. 7.1, the antenna was fabricated on a 14.5 mm ×

7.35 mm × 3.25 mm Ni-Mn-Co ferrite substrate and is shown in Fig. 7.2. For a testing board, the

110 mm × 60 mm double-sided copper clad laminated (CCL) FR4 substrate was fabricated by a

precision milling machine (LPKF ProtoMat S62). To provide DC bias magnetic field to ferrite

antenna, compact Nd-Fe-B permanent magnet (9.5 mm × 4.6 mm × 1.5 mm) was attached to the

ferrite antenna, shown in Fig. 7.2. Magnetic field of one pair and two-pairs of the magnets at a

distance of 7.35 mm (equal to a width of ferrite antenna) was measured to be about 700 and

1,600 Gauss with a gaussmeter. The fabricated ferrite antenna was characterized with a network

analyzer (Agilent N5230A) in an anechoic chamber (Raymond RF QuietBox AVS 700) for its

performance. Dynamic material properties of the ferrite were measured with an

impedance/material analyzer (Agilent E4991A) with a test fixture (Agilent 16454A). An

electromagnet was used to measure DC magnetic-field dependence of complex permeability in

102
FIG. 7.3. Measurement system of permeability spectra at various DC magnetic fields.

the range of 10 MHz to 1 GHz. The system shown in Fig. 7.3 was used to measure permeability

spectra at various DC magnetic fields.

7.3. Results and discussion

7.3.1. Magnetic-field dependence of magnetic loss of Ni-Mn-Co ferrite

Figure 7.4 shows magnetic tangent loss (tan δµ = µ″/µ′) and real part of permeability of

Ni-Mn-Co ferrite as a function of applied DC magnetic field. Magnetic tangent loss at 890 MHz

significantly decreased to 0.192 with 750 Gauss from 0.509 without applied DC magnetic field.

The small DC magnetic field of 750 Gauss led to about 62 % reduction in the magnetic tangent

loss, while maintaining µr of 2.4. This is because the FMR frequency, at which frequency the

ferrite becomes lossy, increases with the applied magnetic field [8, 20]. Table 7.1 smmarizes the

103
0.55 6
0.50 tan 
0.45 r
5
0.40
0.35
4
tan 

r
0.30
0.25
3
0.20
0.15
0.10 2
At 890 MHz
0.05
0.00 1
0 100 200 300 400 500 600 700
Applied magnetic field (Oe)
FIG. 7.4. Measured magnetic-field dependent magnetic loss tangent (tan δμ) and real part of
permeability (µr) of Ni-Mn-Co ferrite at 890 MHz.

Table 7.1. Measured permeability and magnetic tangent loss of Ni-Mn-Co ferrite at various
dc magnetic fields.
H (Gauss) µr at 890 MHz tan δμ at 890 MHz
0 3.74 0.509
200 3.20 0.417
400 2.81 0.313
600 2.50 0.225
750 2.39 0.192

real part of permeability and magnetic tangent loss measured at various magnetic fields. The

magnetic tangent loss significantly decreases with increasing the field as previously mentioned.

7.3.2. Magnetic-field dependence of Ni-Mn-Co ferrite antenna performance

Figure 7.5 shows measured and simulated S-parameters of Ni-Mn-Co ferrite antenna with

and without Nd-Fe-B permanent magnets. Measured S-parameters are in close agreement with

simulated results, but there is a small discrepancy between simulated at 750 Gauss and measured

104
0

-5

-10
S-parameters (dB)
-15

-20

-25 Measured
H = 0 Gauss Simulated
H = 700 Gauss H = 0 Gauss
-30 H = 1600 Gauss H = 750 Gauss

600 700 800 900 1000 1100 1200


Frequency (MHz)
FIG. 7.5. Measured and simulated scattering parameters of Ni-Mn-Co ferrite antenna with
and without applied dc magnetic field.

at 700 Gauss. This can be attributed to the inhomogeneity of bias field in the ferrite, ferrite

demagnetization effect, and the material parameters of the antenna simulation, different from the

actual ones at 700 Gauss. The resonant frequency (fr) of the fabricated antenna shifted from 859

MHz without the DC magnetic field to 892 MHz with 1,600 Gauss. This is because the resonant

frequency of ferrite follows fr ~ 1/(μrεr)0.5. The return loss (RL) of the ferrite antenna varies over

the range of 14.8 to 26.2 dB with the DC magnetic field. This can be understood by µr and tan δμ

dependences of antenna input impedance and quality factor [10].

In order to investigate the effect of the applied DC magnetic field on antenna gain, 2D

gain (in yz plane) was measured for Ni-Mn-Co ferrite antenna in an anechoic chamber. Fig. 7.6

shows measured 2D gain as a function of frequency. The antenna shows remarkably improved

2D gains: −10.58 dBi of gain with 0 Gauss, −5.82 dBi with 700 Gauss, and −4.81 dBi with 1,600

Gauss at 900 MHz. The DC magnetic field of 1,600 Gauss contributed to 54 % increase in

105
-3
-4
-5
-6
2D gain (dBi) -7
-8
-9
-10
-11
-12
H = 0 Gauss (fr = 859 MHz)
-13
H = 700 Gauss (fr = 887 MHz)
-14 H = 1600 Gauss (fr = 892 MHz)
-15
870 880 890 900 910 920 930 940 950
Frequency (MHz)

FIG. 7.6. Measured magnetic-field dependent 2D gain of Ni-Mn-Co ferrite antenna.

antenna gain. It is found that higher DC magnetic field is more effective in increasing the

antenna gain than lower field. This is because high magnetic-field shifts the FMR frequency of

ferrite to higher frequency, thereby lowering magnetic tangent loss. Fig. 7.7 shows normalized

measured and simulated radiation patterns (in yz plane) of Ni-Mn-Co ferrite antenna without and

with magnetic field at 900 MHz. The measured and simulated radiation patterns are in good

agreement. The antenna shows omnidirectional radiation patterns that are suitable for hand-held

mobile device applications. It is noted that both co- and cross-polarized radiations increased with

applied magnetic field, but the co-polarized radiation is dominant. The antenna 2D gain, RL and

fr measured at various applied magnetic fields are summarized in Table 7.2. Both experimental

dynamic magnetic properties of Ni-Mn-Co ferrite and ferrite antenna performance results

support that the ferrite antenna gain can be significantly improved with small DC magnetic field.

Our antenna performance simulation (not presented here) indicates that the antenna gain

106
Measured (0 Gauss)
90 Simulated (0 Gauss)
0 120 60 Measured (700 Gauss)
Simulated (750 Gauss)
-10
150 30
Normalized 2D gain
-20

-30

-40 180 0

-30

-20 210 330


-10

0 240 300
270

(a)

Measured (0 Gauss)
90 Simulated (0 Gauss)
0 120 60 Measured (700 Gauss)
Simulated (750 Gauss)
-10
150 30
Normalized 2D gain

-20

-30

-40 180 0

-30

-20 210 330


-10

0 240 300
270

(b)

FIG. 7.7. Normalized measured and simulated radiation patterns (in yz plane) at 900 MHz for
Ni-Mn-Co ferrite antenna without and with magnetic field. (a) co-polarization and (b) cross-
polarization.

107
Table 7.2. Measured resonant frequency, return loss and 2D gain of Ni-Mn-Co ferrite antenna.
H-field 2D gain (dBi)
fr (MHz) RL (dB)
(Gauss) at 900 MHz
0 859 14.8 −10.58
700 887 26.2 −5.82
1,600 892 16.9 −4.81

increases to 1.1 dBi at 900 MHz when 1% loss ferrite is used. The demonstrated results imply

that low profile and high-gain RF antenna can be realized with out-of-plane anisotropy

permanent magnet film rather than bulk magnets.

7.4. Conclusion

The gain of Ni0.5Mn0.2Co0.07Fe2.23O4 (Ni-Mn-Co) ferrite antenna remarkably increased by

about 6 dBi (from −10.58 to −4.81 dBi) at 900 MHz when a DC magnetic field of 1,600 Gauss is

applied to the ferrite antenna. Magnetic loss tangent at 890 MHz was 0.509 without magnetic

field and 0.192 with 750 Gauss. The volume of the ferrite antenna is 0.346 cm3. We demonstrate

that electrically small and high-gain ferrite antenna with permanent magnet has the

omnidirectional radiation pattern, which is suitable for hand-held mobile device application.

108
References

1. H. Mosallaei and K. Sarabandi, IEEE Trans. Antennas Propag., 55, 45 (2007).

2. P. M. T. Ikonen, K. N. Rozanov, A. V. Osipov, P. Alitalo and S. A. Tretyakov, IEEE Trans.


Antennas Propag., 54, 3391 (2006).

3. H. Moon, G. Y. Lee, C. C. Chen and J. L. Volakis, IEEE Antennas Wirel. Propag. Lett., 11,
322 (2012).

4. S. Yoon and R. W. Ziolkowski, IEEE Antennas Propagat. Soc. Int. Symp., Columbus, OH,
Jun. 22–27, 2003, pp. 297–300.

5. S. Yoon, C. R. Birtcher, and C. A. Balanis, IEEE Trans. Antennas Propag., 53, 531 (2005).

6. G. M. Yang, X. Xing, A. Daigle, M. Liu, O. Obi, S. Stoute, K. Naishadham, and N. X. Sun,


IEEE Trans. Antennas Propag., 57, 2190 (2009).

7. G. S. Abo, Y. K. Hong, R. Syslo, J. Lee, W. Lee, H. M. Kwon, and C. K. K. Jayasooriya,


Microwave Opt. Technol. Lett., 55, 551 (2013).

8. J. Lee, Y. K. Hong, W. Lee, G. S. Abo, J. Park, W. M. Seong, and W. K. Ahn, J. Appl. Phys.,
113, 073909 (2013).

9. W. Lee, Y. K. Hong, J. Lee, D. Gillespie, K. G. Ricks, F. Hu, and J. Abu-Qahouq, IEEE


Antennas Wirel. Propag. Lett., 12, 765 (2013).

10. M. A. Amiri, C. A. Balanis, and C. R. Birtcher, IEEE Antennas Wirel. Propag. Lett., 12, 611
(2013).

11. V. G. Kononov, C. A. Balanis, and C. R. Birtcher, IEEE Trans. Antennas Propag., 60, 1717
(2012).

12. S. Yoon, C. R. Birtcher, and C. A. Balanis, IEEE Trans. Antennas Propag., 53, 531 (2005).

13. A. D. Brown, J. L. Volakis, L. C. Kempel, and Y. Y. Botros, IEEE Trans. Antennas Propag.,
47, 26 (1999).

14. H. Y. David Yang, IEEE Trans. Antennas Propag., 44, 1127 (1996).

15. A. Henderson, J. R. James, and A. Fray, Electron. Lett., 24, 45 (1988).

16. H. Haheri, M. Tsutsumi, and N. Kumagai, IEEE Trans. Antennas Propag., 36, 911 (1988).

17. H. How and T. M. Fang, IEEE Trans. Magn., 30, 4551 (1994).

18. J. Lee, S. Bae, Y. K. Hong, J. Jalli, G. S. Abo, W. M. Seong, S. H. Park, C. J. Choi, and J. G.
Lee, J. Appl. Phys., 105, 07A514 (2009).

109
19. G. Breed, J. High Freq. Electronics, 6, 50 (2007).

20. C. Kittel, Phys. Rev., 73, 155 (1948).

110
CHAPTER 8

MINIATURE LTE MIMO FERRITE ANTENNA7

8.1. Introduction

Long term evolution (LTE) is the latest standard in mobile network technology. LTE

provides the ability to deliver high data rate, throughput, and spectrum efficiency. Furthermore,

LTE uses the 700 MHz band, which allows for more efficient RF propagation and greater

structural penetration [1]. To achieve the aforementioned features, the LTE standard requires a

multiple-input multiple-output (MIMO) antenna system. However, limited space of a mobile

handheld device and low correlation coefficient of MIMO configuration are challenging for

development and implementation of the 700 MHz frequency band LTE MIMO antenna.

Therefore, miniaturization and high isolation of MIMO antenna need to be addressed for LTE

application.

In order to achieve high isolation of MIMO antenna, many methods have been developed.

These methods include suspended line [2], band stop filter [3], decoupling technique [4, 5],

magneto-dielectric substrate [6], and matching network [7]. However, these methods require

relatively complicated MIMO antenna systems. In this paper, isolation, gain, and correlation

coefficient of miniature MIMO ferrite antenna are reported and also compared with previously

reported data.

8.2. Design of LTE MIMO ferrite antenna

In MIMO antenna system, high isolation and low mutual coupling are significantly

important. In order to achieve a low mutual coupling, two antennas need to be separated by more

7
This work was published as “Miniature Long-Term Evolution (LTE) MIMO ferrite antenna,” IEEE
Antennas and Wireless Propagation Letters 10, 603 (2011) by Jaejin Lee, Yang-Ki Hong, et al.

111
than a quarter wavelength [8]. Therefore, miniaturization of the MIMO antennas is indispensable

to allow enough separation between the two antennas in the limited space of a mobile device.

According to equation eff  0  r  r (where λ0 is the free space wavelength), effective

wavelength can be reduced by both permeability (μr) and permittivity (εr). Ferrite possesses both

μr and εr. Thus, the ferrite is more effective in the reduction of antenna size than dielectric

material.

Miniature LTE MIMO ferrite antenna was designed as shown in Fig. 8.1 and studied for

the effect of separation (d) on MIMO antenna isolation and performance. The Ansoft HFSS V.

11 (high frequency structure simulator) was used to simulate the antenna characteristics. Material

FIG. 8.1. Geometry and dimensions of simulated LTE MIMO ferrite antenna.

112
0

-5

Return loss (dB)


-10

-15
d = 0.10
d = 0.070
-20 d = 0.040  = 41 cm

0.66 0.68 0.70 0.72 0.74 0.76 0.78 0.80


Frequency (GHz)
(a)
-5
d = 0.10
-10
d = 0.070
-15 d = 0.040
Isolation (dB)

-20

-25

-30

-35  = 41 cm

-40
0.66 0.68 0.70 0.72 0.74 0.76 0.78 0.80
Frequency (GHz)
(b)

FIG. 8.2. (a) Simulated return loss (S11) and (b) isolation (S21) of LTE MIMO ferrite antenna
for various separations (d). Material parameters used in the simulation: μ = 3.7; tan δμ = 0.08;
ε = 7.8; tan δε = 0.01.

parameters used in the antenna performance simulation were μ of 3.7 (tan δμ = 0.08) and ε of 7.8

(tan δε = 0.01) for the ferrite. FR4 substrate (110 × 55 mm2) with ground size of 55× 90 mm2 was

used as a system board. The antenna was designed using the folding technique to further reduce

the antenna size. The folded and meandered radiator increases the electrical length and improves
113
bandwidth [9, 10]. The non-resonating parts of the antenna radiator improve the impedance

matching. The antenna is fed by 50 Ω coaxial cables.

Figure 8.2(a) shows the simulated return loss (S11) of antenna 1 with various separations

(λ0 = 410 mm). The designed antenna resonates at around 732 MHz and return loss was lower

than -18 dB. It was found that resonance frequency (fr) and impedance matching are almost

unchanged with the antenna separation. Simulated isolation (S21) is presented in Fig. 8.2 (b). The

isolation was improved from -12 dB to -15.5 dB as the separation increased from 0.04λ0 to 0.1λ0.

The isolation of -15 dB is sufficient for antenna diversity [11].

8.3. Fabrication and measurement results

Design and dimension of LTE MIMO ferrite antenna are shown in Fig. 8.3(a), and the

fabricated antenna in Fig. 8.3(b). The 110 × 55 mm2 double-sided copper clad laminate (CCL)

FR4 substrate with permittivity of 4.4 was used as a system board. Copper area of 20 × 15 mm2

was removed at the top corners, and the antennas were mounted on the open areas as shown in

Fig. 8.3(b). The volume of Ni0.5 Mn0.2Co0.07Fe2.23O4 spinel ferrite antenna substrate was 14 × 7 ×

3 mm3. This ferrite has experimental μ of 3.7, tan δμ of 0.2, ε of 7.8, and tan δε of 0.05.

Figure 8.4(a) shows the measured scattering parameters of the fabricated antenna. Return

loss and isolation were measured to be about -26 dB and -16.4 dB at 720 MHz, respectively. This

measured isolation is suitable for the MIMO system [11]. It is noted that the LTE MIMO ferrite

antenna does not include suspended line, band stop filter, or decoupling network, but still shows

good isolation.

Regarding antenna bandwidth, the antenna 1 shows 105 MHz (682 MHz ~ 787 MHz) and

88 MHz (677 ~ 765 MHz) was measured for the antenna 2 at VSWR < 3 as shown in Fig. 8.4(a).

114
(a)

(b)

FIG. 8.3. (a) MIMO antenna structure and (b) photograph of the fabricated LTE MIMO
ferrite antenna.

This bandwidth covers the allocated frequency of LTE band 12 (698 ~ 746 MHz) and part of

LTE band 13 (746 ~ 787 MHz) [12].

In order to verify a low mutual coupling between the antennas, the correlation coefficient

(ρe), in Fig. 8.4(b), was calculated from the experimental complex S-parameters using the

equation (8.1) [13].

2
S11* S12  S 21
*
S 22
e  2 2 2 2 (8.1)
(1  ( S11  S 21 ))(1  ( S 22  S12 ))

115
where * denotes a complex conjugate. Correlation coefficient of the antenna is less than 0.02 in

the LTE band. This correlation coefficient is practically acceptable for MIMO antenna diversity

[14].

-10
S-parameters (dB)

-20

-30 S11
S22
-40 S21
S12

0.5 0.6 0.7 0.8 0.9 1.0


Frequency (GHz)
(a)
0.10
Correlation coefficient (e)

0.08

0.06

0.04

0.02

0.00
0.5 0.6 0.7 0.8 0.9 1.0
Frequency (GHz)
(b)

FIG. 8.4. (a) Measured S-parameters and (b) calculated correlation coefficient of the fabricated
LTE MIMO ferrite antenna.

116
Figure 8.5(a) and (b) show the measured 3-dimensional average and peak gains of the

fabricated LTE MIMO ferrite antenna. Peak and average gain at 746 MHz are -8.83 dBi, and -

-4 3D Average gain Antenna 1


3D Peak gain
-6
Gain (dBi)

-8

-10

-12

-14
740 750 760 770 780 790 800
Frequency (MHz)
(a)

-6 3D Average gain Antenna 2


3D Peak gain

-8
Gain (dBi)

-10

-12

-14
740 750 760 770 780 790 800
Frequency (MHz)
(b)

FIG. 8.5. Measured 3-dimensional average and peak gain of the fabricated LTE MIMO ferrite
antenna: (a) antenna 1 and (b) antenna 2. Experimental property of Ni0.5 Mn0.2Co0.07Fe2.23O4
ferrite: μ = 3.7; tan δμ = 0.2; ε = 7.8; tan δε = 0.05. Separation distance = 41 mm.

117
12.59 dBi for the antenna 1, and -8.32 dBi, -11.67 dBi for antenna 2, respectively. This low

antenna gain is mainly attributed to high magnetic loss (tan δμ = 0.2) of ferrite substrate. Antenna

1 has higher gain at 760 MHz and 794 MHz than antenna 2. This is because antenna 1 shows

better impedance matching over the wider frequency band (BW: 105 MHz) than antenna 2 (BW:

88 MHz) as shown in Fig. 8.4(a). The measured 3D radiation patterns in Fig. 8.6 are nearly

omnidirectional patterns for both antenna 1 and 2.

Characteristics of the fabricated LTE MIMO ferrite antenna are summarized in Table 8.1.

The performance of the LTE MIMO ferrite antenna is compared with previously reported data in

Table 8.2. The fabricated ferrite antenna has considerably small volume and shows competitively

good isolation compared to suspended line [2] and band stop filter [3] LTE MIMO antennas as

(a) (b)

FIG. 8.6. Measured 3D radiation patterns of the fabricated LTE MIMO ferrite antenna at 746
MHz: (a) antenna 1 and (b) antenna 2.

Table 8.1. Summary of LTE ferrite antenna characteristics


Peak gain at 746 MHz (dBi) Bandwidth Isolation
ρe
Measured Simulated (MHz) (dB)
Antenna 1 -8.83 105
-8.65 -16.4 0.009
Antenna 2 -8.32 88

118
Table 8.2. Comparison of performance for LTE MIMO antennas
Dimension Isolation Gain Relative
LTE MIMO 3
(mm ) (dB) (dBi) volume
Suspended line [2] 40 × 20 × 5 -18 -3.42 6.8
Band stop filter [3] 36 × 12 × 8 -15 -7.5 11.7
Ferrite-carbon black composite [6] 18.5×7.5×3.4 -15 -5.1 1.6
Ni0.5Mn0.2Co0.07Fe2.23O4 ferrite
14 × 7 × 3 -16.4 -8.32 1
(this work)

-3 tan  = 0.01
tan  = 0.05
-4
Peak gain (dBi)

-5 Experiment:
tan  = 0.2, tan 
-6

-7

-8

-9
0.00 0.04 0.08 0.12 0.16 0.20
tan 

FIG. 8.7. Simulated peak gain (antenna 1) of the LTE MIMO ferrite antenna at 746 MHz as a
function of magnetic loss.

given in Table 8.2, but the fabricated LTE MIMO ferrite antenna gain is lower. As previously

described, comparatively high gain can be achieved using low loss ferrite. In order to verify this,

peak gain was simulated in terms of magnetic loss at constant dielectric loss. The antenna gains

are presented in Fig. 8.7. The peak gain increases with decreasing magnetic loss of ferrite. Peak

gain of -3.14 dBi at 746 MHz can be achieved with tan δμ of 0.01 and tan δε of 0.01.

Experimental and simulated peak gains at tan δμ of 0.2 and tan δε of 0.05 are -8.83 dBi and -8.65

119
dBi at 746 MHz, which are in close good agreement with each other. Furthermore, it was found

that the simulated isolation was improved from -14.2 dB at tan δμ of 0.03 to -16.5 dB at tan δμ of

0.07 with constant tan δε of 0.01.

8.4. Conclusion

LTE MIMO ferrite antenna was fabricated on Ni0.5Mn0.2Co0.07Fe2.23O4 ferrite substrate

(permeability of 3.7, tan δμ of 0.2, permittivity of 7.8, and tan δε of 0.05) having a volume of 14

× 7 × 3 mm3 and characterized for antenna performance. Return loss and isolation were measured

to be -26 dB and -16.4 dB at 720 MHz, respectively. The bandwidth of the antenna was 105

MHz (682 MHz ~ 787 MHz) for antenna 1 and 88 MHz (677 ~ 765 MHz) for antenna 2 at

VSWR < 3. Correlation coefficient was less than 0.02 in the LTE band. The 3D peak gain of -

8.83 dBi and -8.32 dBi at 746 MHz was measured for antenna 1 and 2, respectively. Low

antenna gain is attributed to high magnetic loss of ferrite substrate. Antenna performance

simulation suggests that higher gain than the measured gain in this paper can be achieved using a

low loss ferrite substrate. Miniature ferrite antenna is most likely to make a significant impact on

improvement of isolation and development of MIMO systems.

120
References

1. F. Hall, “Overlay strategies for 700 MHz LTE deployment,” Radio Frequency Systems, 2009.

2. G. Park, M. Kim, T. Yang, J. Byun, and A. Kim, Antennas and Propagation Society
International Symposium, pp. 1-4 (2009).

3. M. S. Han and J. H. Choi, Antennas and Propagation Society International Symposium, pp.
1-4 (2009).

4. A. C. K. Mak, C. R. Rowell, and R. D. Murch, IEEE Trans. Antennas Propag., 56, 3411
(2008).

5. S. C. Chen, Y. S. Wang, and S. J. Chung, IEEE Trans. Antennas Propag., 56, 3650 (2008).

6. Y. S. Shin and S. O. Park, Microwave Opt. Technol. Lett., 52, 2364 (2010).

7. S. Dossche, S. Blanch and J. Romeu, Electron. Lett., 40, 1164 (2004).

8. P. Tomatta, “Overcoming the LTE handset antenna design problem,” Available:


http://www.eetasia.com/STATIC/PDF/200908/EEOL_2009AUG13_RFD_TA_01.pdf?SOU
RCES=DOWNLOAD.

9. A. R. Razali and M. E. Bialkowski, Microwave Opt. Technol. Lett., 53, 900 (2011).

10. S. M. Ali and H. Kanj, Antennas and Propagation Society International Symposium, 2008.
AP-S 2008. IEEE, vol., no., pp.1-4 (2008).

11. K. J. Kim and K. H. Ahn, Microwave Opt. Technol. Lett., 49, 731 (2007).

12. LTE frequency bands & spectrum allocations. Available: http://www.radio-


electronics.com/info/cellulartelecomms/lte-long-term-evolution/lte-frequency-spectrum.php.

13. S. Blanch, J. Romeu, and I . Corbella, Electron. Lett., 39, 705 (2003).

14. J. Byun, J. H. Jo, and B. Lee, Microwave Opt. Technol. Lett., 50, 2600 (2008).

121
CHAPTER 9

INTEGRATED FERRITE INDUCTOR FOR POWER SYSTEM-ON-CHIP8

9.1. Introduction

High power efficiency of DC-DC converter is significantly important due to the limited

power capacity of battery of smart phone. Also, integration and miniaturization of DC-DC

converter have attracted great attention to power system-on-chip (PowerSoC) [1, 2]. High quality

(Q) factor and magnetic core for power inductor are needed to meet high efficiency and

miniaturization of DC-DC converter. Furthermore, according to the current trend of DC-DC

converter, power inductor should have the capability of large current to respond to low input

voltage of microprocessor [3]. Ferrite film planar inductors have been extensively studied [4−7]

because ferrite has advantages of high resistivity, insulating property, and good chemical

stability that lead to low eddy current, parasitic capacitance, and no passivation layer structure.

However, thick ferrite deposition is still challenging on silicon wafer. RF sputtering [8] and PLD

(pulsed laser deposition) [9] processes have been used to deposit ferrite, but deposition and

deposition area are limited to low rate and small area.

In this chapter, micron thick ferrite film deposition process was performed on 4 inch

silicon wafer using DC magnetron sputtering. Inductance, Q-factor, and current capability of

ferrite inductors are reported in comparison with air-core inductor. Also, inductor performance

was studied in terms of ferrite film thickness.

9.2. Experiment

9.2.1. Design and fabrication


8
This work was published as “Integrated ferrite film inductor for power system-on-chip (PowerSoC)
smart phone applications,” IEEE Transactions on Magnetics 47, 304 (2011) by Jaejin Lee, Yang-Ki
Hong, et al.

122
Table 9.1. Design parameters of ferrite inductor

Type Coil thickness Coil width Coil space Coil area


(turns) ( m) ( m) ( m) (mm2)

Spiral
65 500/400/300 200 5×5
(2.5/3.5/4.5)

Spiral ferrite inductors with 2.5, 3.5, and 4.5 turns of 65 μm thick Cu coil were fabricated

on silicon (Si) wafer. The outer coil area of inductor is 5 × 5 mm2 and the coil width of 2.5, 3.5,

and 4.5 turns is 500, 400, and 300 μm, respectively. The space between two adjacent coils is

fixed at 200 μm. Design parameters of the ferrite inductor are summarized in Table 9.1. For

performance comparison, air-core inductor was fabricated on 3 μm thick SiO2/Si wafer with the

same geometry of ferrite inductor.

One μm thick Ni-Zn-Cu ferrite (Ni0.5Zn0.3Cu0.2Fe2O4) film was prepared by sputtering of

multilayered oxide and consequent heat treatment of the multilayer as shown in Fig. 9.1.

Multilayered oxide films of 130 nm NiO, 700 nm Fe2O3, 100 nm ZnO and 60 nm CuO were

sequentially deposited by DC magnetron sputtering using Ni, Fe, Zn, and Cu metal targets in a

mixture of 75 % of Ar and 25 % of O2 at 5 mTorr of working pressure and 300 W of DC power.

The deposited multilayered oxide films were post-annealed at 800 oC for 1 h to crystallize into

Ni-Zn-Cu ferrite. In order to fabricate the 2.5 μm thick ferrite film, the same multilayered oxide

deposition process of the 1 μm thick ferrite film was performed repeatedly 3 times and post-

annealed at 800 oC for 4 h. It is noted that the 2.5 μm thick ferrite film required longer post-

annealing time than 1 μm thick ferrite film in order to convert from multilayered oxide films to

ferrite. These fabricated ferrite films were used for the fabrication of planar inductor. The

fabrication process for ferrite inductor is depicted in Fig. 9.2 [10]. An Au 100 nm/Ti 20 nm film

is deposited on the ferrite film for seed layer of electroplating. To fabricate the 90 µm thick Cu

123
FIG. 9.1. Sputter deposition process of Ni0.5 Zn0.3Cu0.2Fe2O4 ferrite film. Multilayered oxide
films are deposited by DC magnetron sputtering using Ni, Fe, Zn, and Cu metal targets in a
mixture of Ar and O2 gases [10].

coils, 120 µm thick photo-resist molds is produced on the Au/Ti seed layer by Microchem

KMPR1050, and gold wire bonding follows for electrical connection from center pad to outer

pad. After the wire bonding, 90 µm thick Cu coils is fabricated by electroplating. Gold thin film

is deposited on the electroplated Cu for prevention of Cu oxidation. Finally, Ni0.5Zn0.3Cu0.2Fe2O4

film is deposited on the coil area.

9.2.2. Measurement

X-ray diffractometer was used to characterize crystal phase of the ferrite film. Magnetic

properties were measured by vibrating sample magnetometer. Cross-sectional images of the

ferrite film and inductor were observed by scanning electron microscopy (SEM). Inductance,

quality (Q) factor, and superimposed DC current of the fabricated inductors were evaluated by

124
FIG. 9.2. Fabrication process for ferrite inductor array on 4 inch Si wafer [10].

distributed parameter method with Agilent N5230A network analyzer, Agilent 6631B DC power

supply, microprobe station, and Picobrobe model 7-125 probe.

9.3. Results and discussion

Crystalline phase of the fabricated ferrite film was well indexed to the spinel ferrite, and

the composition of Ni0.38~0.41Zn0.31~0.38Cu0.25~0.27Fe1.98~2.01O3.81~3.98 was confirmed with energy

dispersive X-ray (EDX). Fig. 9.3(a) shows magnetic properties of the ferrite film. Saturation

magnetization and coercivity of the 2.5 µm thick ferrite were measured to be 3.3 kG and 78.06

125
4
2.5 m: 4Ms= 3.3 kG, Hc=78.06 Oe
3 1 m: 4Ms= 2.6 kG, Hc=48.79 Oe

4Ms (kG)
1
0
-1
-2
2.5 m thick ferrite
-3
1 m thick ferrite
-4
-6 -4 -2 0 2 4 6
Applied field (kOe)
(a)

(b)
FIG. 9.3. (a) Magnetic hysteresis loop and (b) cross-section of the 1 μm and 2.5 μm
Ni0.38Zn0.31Cu0.27Fe2O3.81 film.

Oe, while the 1 µm thick ferrite shows 2.6 kG and 48.79 Oe, respectively. The higher saturation

magnetization of 2.5 µm thick ferrite can be attributed to the increase in film density and

crystallinity by longer post-annealing time. The cross-sectional SEM image of the fabricated

ferrite films are shown in Fig. 9.3(b). The left image shows 1 µm thick ferrite thickness and the

right one indicates 2.5 µm thick ferrite thickness on Si wafer. Regarding the permeability (μ) of

the fabricated ferrite film, the real part of μ was measured to be about 28 at 10 MHz by coplanar

waveguide method.

126
FIG. 9.4. The fabricated ferrite inductors with 2.5, 3.5, and 4.5 turns of Cu coil and cross-
sectional image of the ferrite inductor with 65 μm thick Cu coil.

Array of ferrite inductors was fabricated on silicon wafer and is shown in Fig. 9.4. The fabricated

ferrite inductors have 5×5 mm2 of the outer coil area, and 2.5 turns, 3.5 turns, and 4.5 turns of Cu

coil. It is observed that the fabricated inductor has 65 μm thick Cu coil from the cross-sectional

SEM image of the fabricated inductor.

Frequency dependence of inductance (L) and Q-factor for 1 µm, 2.5 µm thick ferrite and

air-core inductors are shown in Fig. 9.5(a) and (b), respectively. The L of the ferrite inductor

increases with the thickness of ferrite film from 45.5 nH for 1 μm thick ferrite to 50 nH for 2.5

μm thick ferrite. This increment of L is attributed to the increase in magnetic flux, which is a

function of 4πMs and ferrite thickness [7]. Compared to L of the air-core inductor, the L of the

ferrite inductors was increased by 11.5 % for 1 μm thick ferrite and 22.5 % for 2.5 μm thick

ferrite. The maximum Q-factor was obtained to be 59 at 2.87 MHz from 2.5 μm thick ferrite

inductor, which is increased by a factor of 1.19 and 2.54 compared to 1 μm thick ferrite and air-

core inductors, respectively. As the ferrite thickness increased, the parasitic capacitance and

substrate loss by eddy current were decreased, and the inductor performance was improved.

127
80
Air-core (4.5 turn)
70
1 m NiZnCu ferrite (4.5 turn)
2.5 m NiZnCu ferrite (4.5 turn)

Inductance (nH)
60
50
40
30
20
10
0
1 10
Frequency (MHz)
(a)
80
Air-core (4.5 turn)
70 1 m NiZnCu ferrite (4.5 turn)
2.5 m NiZnCu ferrite (4.5 turn)
60
50
Q-factor

40
30
20
10
0
1 10
Frequency (MHz)
(b)
FIG. 9.5. (a) Inductance and (b) quality factor of 1 μm, 2.5 μm thick ferrite and air-core
inductors.

Figure 9.6 shows the superimposed DC current characteristic of the ferrite inductors. DC

current ranging from 0 to 2.5 A was superimposed onto the fabricated inductors. Then, L of the

ferrite inductors superimposed on DC current was determined by the measured S-parameters.

Superimposed DC current was estimated corresponding to a 5 % drop in L at 10 MHz. The 2.5

128
100
99
98
97
L/L0 (%)
96
95
94 2.5 m ferrite inductor
1 m ferrite inductor
93
92
0.0 0.5 1.0 1.5 2.0 2.5
Superimposed DC current (A)
FIG. 9.6. Superimposed DC current characteristic of 1 μm and 2.5 μm thick ferrite inductors at
10 MHz. L0 represents the inductance at zero superimposed DC current.

Table 9.2. Summary of the inductance, Q-factor, and superimposed DC current for the
fabricated inductors.
Superimposed DC
Inductor Lmax (nH) Maximum Q Rdc (Ω)
current (A)
Air-core 40.8 23.2 at 1.56 MHz ~ 0.02 -
1μm ferrite 45.5 49.3 at 2.26 MHz ~ 0.02 2.15
2.5μm ferrite 50.0 59 at 2.87 MHz ~ 0.025 ~ 2.5

µm thick ferrite inductor shows 2.5 A of the superimposed DC current, which is higher than 2.15

A of 1 µm thick ferrite inductor. This is because 2.5 μm thick ferrite film has high 4πMs and Hc

(∝ Hk) that result in increase of the magnetic saturation field compared with 1 μm thick ferrite. It

is found that the fabricated ferrite inductors have less than 5 % decrease of L in the DC current

range of 0 A to 2 A as shown in Fig. 9.6. The Lmax, Qmax, and superimposed DC current of the

fabricated inductors are summarized in Table 9.2.

129
Electrical properties of a buck DC-DC converter based on the studied ferrite inductors

were simulated for calculation of power efficiency. Fig. 9.7 shows the circuit diagram of the

buck DC-DC converter that was used for the simulation. The input parameters are 5 MHz

switching frequency, 5 V input voltage, and 20 % duty cycle. For the inductor input parameters,

the measured L and Q at 5 MHz of the fabricated 1 µm and 2.5 µm thick ferrite inductors were

used. The calculated power efficiency (= Pout/Pin) of buck DC-DC converter is shown as a

function of load current in Fig. 9.8. Maximum power efficiency of the 2.5 µm thick ferrite

FIG. 9.7. Circuit diagram of the buck DC-DC converter for the calculation of power efficiency.

100

90
Efficiency (%)

80

70
2.5 m ferrite inductor
1 m ferrite inductor

60
0.5 1.0 1.5 2.0
Load current (A)
FIG. 9.8. Calculated power efficiency of the DC-DC converter based on the studied ferrite
inductors as a function of load current.

130
inductor was calculated to be 91.7 % at load current of 0.647 A. On the other hand, the 1 µm

thick ferrite inductor shows the maximum power efficiency of 90.8 % at load current of 1.18 A.

This performance for the fabricated ferrite inductors is competitive to those previously reported

[2, 4, 5] and the commercial product [11].

9.4. Conclusion

The array of ferrite inductors was fabricated on silicon wafer and characterized in

comparison with air-core inductor. Inductance (L) of 1 μm and 2.5 μm thick ferrite inductors are

45.5 nH and 50 nH, respectively, which are higher than 40.8 nH of air-core inductor. The

maximum Q-factor of the 2.5 μm thick ferrite inductor was obtained to be 59.0 at 2.87 MHz,

which increased by a factor of 1.19 for 1 μm thick ferrite inductor and 2.54 for air-core inductor.

Superimposed DC current of the 2.5 µm thick ferrite inductor was estimated to be 2.5 A

corresponding to a 5 % drop in L at 10 MHz, while the 1 µm thick ferrite inductor showed 2.15

A. It was found that inductor performance was improved by increasing ferrite film thickness.

131
References

1. S. C. O. Mathuna, T. O. Donnell, N. Wang, and K. Rinne, IEEE Trans. Magn., 20, 585
(2005).

2. N. Wang, T. O. Donnell, R. Meere, F. M. F. Rhen, S. Roy, and S. C. O. Mathuna, IEEE


Trans. Magn., 44, 4096 (2008).

3. I. Kowase, T. Sato, K. Yamasawa, and Y. Miura, IEEE Trans. Magn., 41, 3991 (2005).

4. I. Sasada, T. Yamaguchi, K. Harada, and Y. Notohara, IEEE Trans. Magn., 29, 3231 (1993).

5. Y. Fukuda, T. Inoue, T. Mizoguchi, S. Yatabe, and Y. Tachi, IEEE Trans. Magn., 39, 2057
(2003).

6. J. J. Lee, Y. K. Hong, S. Bae, J. H. Park, J. Jalli, G. S. Abo, R. Syslo, B. C. Choi, and G. W.


Donohoe, IEEE Trans. Magn., 46, 2417 (2010).

7. X. L. Tang, H. W. Zhang, H. Su, Y. Shi, and X. D. Jiang, J. Magn. Magn. Mater., 293, 812
(2005).

8. M. R. Koblischka, M. Kirsch, M. Brust, A. K. Veneva, and U. Hartmann, Phys. Status Solidi


A, 205, 1783 (2008).

9. O. F. Caltun, J. Optoelectron. Adv. Mater., 7, 739 (2005).

10. S. Bae, Y. K. Hong, J. J. Lee, J. Jalli, G. S. Abo, A. Lyle, B. C. Choi, and G. W. Donohoe,
IEEE Trans. Magn., 45, 4773 (2009).

11. Data sheet of EN5364QI, Enpirion Inc., New Jersey, 2009 (available at
http://www.enpirion.com/products-en5364qi.htm).

132
CHAPTER 10

GIGAHERTZ (GHZ) FERRITE INDUCTOR FOR RFIC9

10.1. Introduction

Radio frequency (RF) transceivers with fully integrated circuits, including voltage

controlled oscillators, low-noise amplifiers, and frequency filters, are highly demanded for

mobile electronic devices. In response to this, a large number of inductors are integrated into RF

integrated circuits (RFICs). However, the integrated inductors consume a large area of the RFICs,

and their quality (Q) factors are critical for power management and RF signal processing.

Therefore, compact and high Q factor integrated inductors are required.

Microelectromechanical systems (MEMS) and multi-stacked inductors have previously

been developed [1−4]. A MEMS inductor, having a 15 μm thick, 2-turn coil and 40 μm

suspension over a silicon (Si) substrate, showed a high Q factor of 27 at 2.6 GHz due to low

substrate and ohmic losses. However, the inductor showed a small Inductance (L) density of 11.2

nH/mm2 (L of 1.8 nH) [2]. On the other hand, a two-staked inductor with 9.5 total coil turns

exhibited a high L-density of 283.7 nH/mm2 but low Q factor of 9.8 at 2.3 GHz [3]. The low Q

factor resulted from the large proximity effect and parasitic capacitance of the multi-stacked coil

structure.

In addition, integrated magnetic inductors have been widely investigated. A high

permeability material increases the L and Q factor of the inductors by increasing the magnetic

flux and decreasing coil resistance and parasitic capacitance [5, 6]. A 2 μm thick CoZrTa

magnetic film inductor with relative permeability of about 1000 showed a high L-density of 1.3

9
This work is to be submitted as “Gigahertz (GHz) ferrite inductor for radio frequency integrated
circuits,” in IEEE Transactions on Magnetics by Jaejin Lee, Yang-Ki Hong, et al.

133
μH/mm2 and a 28 times increase in L compared to that of the air-core inductor [6]. However, low

resistivity (ρ = 99 μΩ∙cm) of the CoZrTa film decreased the Q factor below 6 at 1 GHz.

In an effort to address the low Q factor, multilayered magnetic film [7−9] or high-

resistivity ferrite (ρ > kΩ∙cm) [10−13] were used in integrated inductors to suppress GHz eddy

current loss. A FeGaB/Al2O3 multilayer solenoid inductor showed an L-density of 15 nH/mm2

and Q factor of 14 at 1.2 GHz [8]. A Ni0.4 Zn0.4Cu0.2Fe2O4 ferrite spiral inductor had 12 nH/mm2

(L of 1.94 nH) and a Q factor of 17.2 at 4 GHz, which were increases of about 3.9 % and 6.3 %,

respectively, compared to those of the air-core inductor [13]. It was found that the high-

resistivity ferrite is more effective in increasing both the L and Q factor above 2 GHz compared

to the conventional MEMS and low-resistivity magnetic inductors [2, 3, 6, 8]. However, the L-

density and Q factor of the ferrite inductor still need to be improved.

In this paper, we have designed and fabricated integrated planar inductors based on high

magnetic anisotropy Zn0.13Co0.04Ni0.63Fe2.2O4 (Zn-Co-Ni) ferrite film. The experimental and

simulated inductor characteristics of the Zn-Co-Ni ferrite film inductor demonstrate high L-

density and Q factor at 2 GHz in comparison to integrated MEMS, low-resistivity magnetic, and

low-anisotropy ferrite inductors. In addition, the role of dynamic magnetic properties in inductor

L and Q is discussed based on Ba3Co2Fe24O41 (Co2Z) ferrites.

10.2. Experiment

10.2.1 Ferrite inductor design and fabrication

A one-turn spiral integrated inductor was designed as shown in Fig. 10.1(a). The inductor

has a coil area of 0.5 × 0.5 mm2, width of 15 μm, space of 50 μm, and thickness of 7 μm. A

bonding wire was used between the center and outermost pads for an electrical connection. The

134
(a)

(b)

FIG. 10.1. (a) Designed one-turn spiral inductor and (b) optical microscope image of the
fabricated air-core inductor.

designed coil structure was used in simulating inductor performance (using ANSYS HFSS ver.

11) and fabricating the air-core and ferrite inductors.

Figure 10.1(b) shows the air-core inductor fabricated on 0.6 μm thick silicon oxide

(SiO2)/600 μm thick Si substrate. The conductivity (σ) of the Si substrate is 13 S/m. The

fabricated ferrite inductors in Fig. 10.2(a) and (b) have two different magnetic configurations.

One is the double-side ferrite inductor, where the inductor coil is positioned between the bottom

Zn-Co-Ni ferrite film and top Co2Z ferrite particles as shown in Fig. 10.2(a). On the other hand,

135
(a)

(b) (c) (d)

FIG. 10.2. Photo-images of (a) double-side ferrite inductor (bottom Zn-Co-Ni ferrite film + top
Co2Z ferrite particles), (b) bottom-type Zn-Co-Ni ferrite film inductor, (c) short de-embedding,
and (d) open de-embedding structures.

Fig. 10.2(b) shows the bottom-type ferrite inductor, having the coil formed on the Zn-Co-Ni

ferrite film.

Ferrite film deposition and microfabrication processes were used to fabricate the ferrite

inductors. For the bottom-type ferrite inductor, a 2.45 μm thick Zn-Co-Ni ferrite film was

deposited on the SiO2/Si substrate, followed by Au (50 nm)/Ti (20 nm) seed layers for use in

copper (Cu) coil electroplating. A 10 μm thick photoresist (PR) mold (Microchem KMPR1005)

was then photolithographically produced on the Au/Ti layers. A 7 μm thick Cu coil was

subsequently deposited by electroplating, followed by PR stripping and ion beam etching (IBE)

of the Au/Ti layers as shown in Fig. 10.2(b). For the double-side ferrite inductor, a mixture of

136
Co2Z ferrite particles and ethanol was applied on the top of the bottom-type Zn-Co-Ni ferrite

inductor and dried in air as presented in Fig. 10.2(a).

10.2.2 Preparation of high magnetic anisotropy Zn-Co-Ni ferrite film and Co2Z ferrite particles

The high-anisotropy Zn-Co-Ni ferrite film was prepared by the DC magnetron sputtering

of nickel (Ni), iron (Fe), zinc (Zn), and cobalt (Co) metal targets in a mixture of 75 % Ar and 25 %

O2. During the sputtering deposition, the substrate temperature was held at 850 oC, and the

working pressure and DC power for each target were 5 mTorr and 250 W (200 W for Zn),

respectively. The deposited multilayered oxide films were post-annealed at 850 oC for 4 hours in

a mixture of 67 % Ar and 33 % O2 in order to crystallize the films into the Zn-Co-Ni ferrite.

For preparation of the Co2Z ferrite particles, the solid-state reaction process was used,

and the detailed process is reported elsewhere [14, 15]. The Co 2Z ferrite particles with and

without 10-hour ball-milling were used for the fabrication of the double-side ferrite inductors.

10.2.3. Measurement

Magnetic properties and crystalline phases of the Zn-Co-Ni ferrite film and Co2Z ferrite

particles were characterized by a vibrating sample magnetometer (MicroSense EV9) and X-ray

diffractometer, respectively. A scanning electron microscope (SEM: JEOL 7000) was used to

observe the cross-sectional image of the Zn-Co-Ni ferrite film.

To characterize the inductor performances, we used a vector network analyzer (Agilent

N5260A) and GSG probes (Cascade Microtech ACP40-GSG-200) to measure scattering (S)-

parameters for the air-core and ferrite inductors. Then, the L and Q factor were calculated from

the measured complex two-port S-parameters (S11, S12, S21, and S22) [16]. The experimental

137
setup for the S-parameter measurement of the inductor is shown in Fig. 2(a). The probes were

calibrated by the short-open-load-through (SOLT) technique with an impedance standard

substrate (Cascade Microtech 101-190). The de-embedding technique [17] was also used to

eliminate the parasitic effects of the probe pads and ground structure. The fabricated short and

open de-embedding structures for the ferrite inductors are shown in Fig. 10.2(c) and (d).

10.3. Results and discussion

10.3.1. Characteristics of the Zn-Co-Ni ferrite film and Co2Z ferrite particles

The measured X-ray diffraction patterns confirm that the Zn-Co-Ni ferrite film (in Fig.

10.3) and Co2Z ferrite particles (not shown here) were well crystallized. In addition, the cross-

sectional SEM image in Fig. 10.4 shows that the 2.45 μm thick Zn-Co-Ni ferrite film had a

smooth and clean surface.


Si (400)
Spinel (311)
Intensity (a.u)

Spinel (440)
Spinel (220)

Spinel (333)
Spinel (422)

Spinel (533)
Spinel (400)

Spinel (622)
Spinel (222)
Si (200)

20 25 30 35 40 45 50 55 60 65 70 75 80
2 (degree)
FIG. 10.3. X-ray diffraction pattern of the fabricated Zn-Co-Ni ferrite film on silicon substrate.

138
FIG. 10.4. Scanning electron microscope (SEM) cross-sectional image of the Zn-Co-Ni ferrite
film.

400
Zn-Co-Ni ferrite film
300
Magnetization, M (emu/cm )
3

200

100

-100

-200

-300

-400
-10 -8 -6 -4 -2 0 2 4 6 8 10
Applied field (kOe)
FIG. 10.5. Measured magnetic hysteresis loop of the Zn-Co-Ni ferrite film.

The measured magnetic hysteresis loop of the Zn-Co-Ni ferrite film in Fig. 10.5 shows

saturation magnetization (Ms) of 334emu/cm3 and coercivity (Hc) of 178 Oe. The in-plane

magnetic anisotropy field (Hk) of the Zn-Co-Ni ferrite film is 420 Oe, which was obtained by the

method in [18]. It was found that the fabricated Zn-Co-Ni ferrite film has higher Ms and Hk than

those of the low magnetic anisotropy Ni-Zn-Cu and YIG ferrites [13]. Accordingly, the

ferromagnetic resonance frequency (fFMR) of the Zn-Co-Ni ferrite film increases with the Ms and

Hk according to the relation fFMR = (γ/2π)(4πMsHk)0.5.


139
50 non-milled Co2Z
40 10-hour milled Co2Z

Magnetization,  (emu/g)
30
20
10
0
-10
-20
-30
-40
-50
-10 -8 -6 -4 -2 0 2 4 6 8 10
Applied field (kOe)
FIG. 10.6. Measured magnetic hysteresis loops of the non-milled and 10-hour milled Co2Z
ferrite particles.

Figure 10.6 shows two magnetic hysteresis loops for the Co 2Z ferrites. One is for the

non-milled Co2Z, and the other is for the 10-hour milled Co2Z ferrite particles. The 10-hour

milled Co2Z particles exhibit saturation magnetization (σs) of 49 emu/g and Hc of 168 Oe, while

the non-milled particles show σs of 50 emu/g and Hc of 23 Oe. The larger Hc of the milled Co2Z

indicates a higher magnetic anisotropy field (Hk) than that of the non-milled Co2Z [14]. This

suggests the improved GHz magnetic properties.

We have calculated the frequency-dependent complex permeability for the Zn-Co-Ni

ferrite film using Eq. (10.1) and measured the complex permeability for the Co2Z ferrite particles

in order to estimate permeability, magnetic loss, and fFMR. The complex permeability of the

ferrite film can be given by Eq. (10.1) [19]

140
 4M s
    1 
H k  j
, (10.1)
 2 
 1  2
 H k   4M s  j H k  j    

where ω is the angular driving frequency, γ is the gyromagnetic constant (1.76 × 107 rad/Oe∙s),

and α is the damping constant. Two different sets of magnetic and damping parameters were

used in Eq. (10.1) to exploit the effect of magnetic loss on the inductor performance. The one set

includes the experimental Ms of 334 emu/cm3 (in Fig. 10.5), in-plane Hk of 420 Oe, and α of 0.15

of the fabricated Zn-Co-Ni ferrite film. For magnetic property comparison, the other set has Ms

of 430 emu/cm3, Hk of 520 Oe, and α of 0.05. Fig. 10.7 shows the calculated complex

permeability of the ferrite films with the different magnetic and damping parameters. The real

part of permeability (μ′) is about 13 at 2 GHz for both sets, while the magnetic loss tan δμ (=μ′′/μ′)

60
' (Ms 334 emu/cc, Hk 420 Oe,  = 0.15)
50 '' (Ms 334 emu/cc, Hk 420 Oe,  = 0.15)
' (Ms 430 emu/cc, Hk 520 Oe,  = 0.05)
40
Complex permeability

" (Ms 430 emu/cc, Hk 520 Oe,  = 0.05)

30
20
10
0
-10
-20
-30
0.1 1 10
Frequency (GHz)
FIG. 10.7. Calculated complex permeability of Zn-Co-Ni ferrite film using the experimental
parameters Ms = 334 emu/cm3, Hk = 420 Oe, α = 0.15, and designed parameters Ms = 430
emu/cm3, Hk = 520 Oe, α = 0.05.

141
for the second set is 78 % lower than that of the first set. The loss tan δμ of the first and second

sets at 2 GHz is 0.33 (0.12 at 1 GHz) and 0.07 (0.03 at 1 GHz), respectively. The fFMR of the

second set is 4.92 GHz, which is much higher than 3.75 GHz of the first set. It is noted that the

higher Hk of the second set contributes to an improvement in GHz magnetic loss properties as

compared to the first set. The fFMR was determined by the maximum imaginary part of

permeability (μ′′) in Fig. 10.7. Both calculated complex permeabilities were used in inductor

performance simulation.

To measure the frequency-dependent permeability of the Co 2Z ferrites, the non-milled

and 10-hour milled Co2Z ferrite powders were formed into toroidal rings (inner diameter: 3.2

mm, outer diameter: 7.8 mm). Then, the ferrite rings were characterized for complex

permeability with an impedance/material analyzer (Agilent E4991). Fig. 10.8 shows the

measured complex permeabilities of the non-milled and 10-hour milled Co2Z ferrites. The μ′ at 2

GHz is 3.4 (4.3 at 1 GHz) for the non-milled Co2Z and 2.1 (2.0 at 1 GHz) for the milled Co 2Z.

6
' '' (non-milled Co2Z)
5 ' '' (10-hour milled Co2Z)
Complex permeability

0
0.01 0.1 1
Frequency (GHz)
FIG. 10.8. Measured complex permeability of the non-milled and 10-hour milled Co2Z ferrites.

142
The corresponding loss tan δμ are 0.64 (0.3 at 1 GHz) and 0.14 (0.06 at 1 GHz), respectively. The

lower magnetic loss of the 10-hour milled Co2Z ferrite is attributed to higher Hk than that of the

non-milled Co2Z. As previously described, this also leads to a higher fFMR (beyond the measured

frequency range in Fig. 10.8). These two different Co2Z ferrites were later used to investigate the

dynamic magnetic property dependence of the L and Q factor.

10.3.2. Performance of high anisotropy ferrite film inductor

The fabricated bottom-type and double-side ferrite inductors were characterized for the L

and Q factor. Fig. 10.9(a) and (b) show the measured and simulated L and Q factor of the

bottom-type Zn-Co-Ni ferrite film inductor compared to those of the air-core inductor. For ferrite

inductor simulation, we used the calculated frequency-dependent values of μ′ and loss tan δμ of

the Zn-Co-Ni ferrite film in Fig. 10.7. The measured L and Q factor of the ferrite and air-core

inductors are in close agreement with the simulation data. The measured L of the Zn-Co-Ni

ferrite and air-core inductors are 4.5 and 4.2 nH at 2 GHz, respectively. The corresponding L-

densities are 18.0 and 16.8 nH/mm2. It is noted that the Zn-Co-Ni ferrite inductor shows a 7.1 %

increase in L compared to that of the air-core inductor. On the other hand, the Q factor of the Zn-

Co-Ni ferrite inductor at 2 GHz is 4.8 (8.5 at 1 GHz), which is lower than 6.7 (12.5 at 1 GHz) of

the air-core inductor. The maximum Q factors are 12.3 at 0.5 GHz for the Zn-Co-Ni ferrite and

14.1 at 0.55 GHz for the air-core inductors. The relatively low Q factor of the ferrite inductor is

mainly attributed to the GHz loss tan δμ of the Zn-Co-Ni ferrite. In addition, the low GHz Q

factors of both the air-core and Zn-Co-Ni ferrite inductors are attributed to the large substrate

loss of the high-conductivity Si substrate (σ = 13 S/m) used in the inductor fabrications.

However, when we used the low-conductivity Si substrate (σ = 0.01 S/m) in inductor

143
9
Meas. (air-core, Si  = 13 S/m)
8 Meas. (ferrite, Si  = 13 S/m)
Sim. (air-core, Si  = 13 S/m)
7 Sim. (ferrite, Si  = 13 S/m)
Sim. (air-core, Si  = 0.01 S/m)
Inductance (nH) 6 Sim. (ferrite, Si  = 0.01 S/m)
Sim. (low loss ferrite, Si  = 0.01 S/m)
5
4
3
2
1
0
0.1 1 10
Frequency (GHz)
(a)
50
Meas. (air-core, Si  = 13 S/m)
Meas. (ferrite, Si  = 13 S/m)
Sim. (air-core, Si  = 13 S/m)
40 Sim. (ferrite, Si  = 13 S/m)
Sim. (air-core, Si  = 0.01 S/m)
Sim. (ferrite, Si  = 0.01 S/m)
Quality factor

30 Sim. (low loss ferrite,


Si  = 0.01 S/m)

20

10

0
0.1 1 10
Frequency (GHz)
(b)

FIG. 10.9. Measured and simulated (a) L and (b) Q factor of the air-core and bottom-type ferrite
film inductors.

144
performance simulation, the Q factors of the Zn-Co-Ni ferrite and air-core inductors significantly

increased to 20.0 and 31.9 at 2 GHz, respectively. As shown in Fig. 10.9, the Q factor of the Zn-

Co-Ni ferrite inductor reaches its maximum of 22.6 at 1.4 GHz, and the maximum Q factor of

the air-core inductor is 42.2 at 5.5 GHz. The simulated ferrite inductor characteristics correspond

well to the frequency-dependent loss tan δμ of the Zn-Ni-Co ferrite (shown in Fig. 10.7).

Accordingly, these results indicate that GHz magnetic loss of the ferrite plays a critical role in

inductor Q factor.

We have, therefore, used the frequency-dependent μ′ and tan δμ of the high-anisotropy,

low-loss ferrite in Fig. 10.7, with Ms of 430 emu/cm3, Hk of 520 Oe, and α of 0.05, for the L and

Q factor simulation of the bottom-type ferrite inductor. The simulated results are presented in Fig.

10.9. The L of 4.5 nH for the high-anisotropy ferrite inductor at 2 GHz is the same as that of the

Zn-Co-Ni ferrite inductor due to the similar μ′. The maximum Q factor of the high-anisotropy

ferrite inductor significantly increases to 34.3 at 2.1 GHz from 22.6 at 1.4 GHz for the Zn-Co-Ni

ferrite inductor (at Si σ = 0.01 S/m). Importantly, the Q factor of the high-anisotropy ferrite

inductor is greater than the Q factor of the air-core inductor (at Si σ = 0.01 S/m) up to 2.35 GHz

as depicted in Fig. 10.9. This simulation result shows that high magnetic anisotropy GHz ferrites

can lead to high-Q integrated inductors above 2 GHz.

10.3.3. Dependence of dynamic magnetic properties of ferrites on L and Q factor

To demonstrate the dynamic magnetic property dependence of the L and Q factor, the

double-side ferrite inductors were fabricated by applying the 10-hour milled (i.e. high anisotropy)

or non-milled (i.e., low anisotropy) Co2Z ferrite particles on the top of the fabricated Zn-Co-Ni

ferrite film inductor as shown in Fig. 10.2(a). Fig. 10.10 shows the measured L and Q factor of

145
the double-side ferrite inductors in comparison to the measured characteristics of the air-core and

bottom-type Zn-Co-Ni ferrite inductors. The double-side ferrite inductors show L of 5.6 nH

(non-milled Co2Z) and 4.8 nH (10-hour milled Co2Z) at 2 GHz, which are greater than 4.2 nH of

the air-core and 4.5 nH of the Zn-Co-Ni ferrite film inductors. The corresponding L-densities are

22.4 nH/mm2 with the non-milled Co2Z and 19.2 nH/mm2 with the 10-hour milled Co2Z. The L

of the double-side ferrite inductors are 133 % and 114 % of the L of the air-core inductor,

respectively. The remarkable increase in L occurs because the top Co2Z ferrite enhances

magnetic flux density and also reduces a leakage magnetic flux. On the other hand, the additional

magnetic loss from the Co2Z ferrite caused a reduction in the Q factor as shown in Fig. 10.10(b).

However, it is noted that the high anisotropy, low-loss Co2Z (i.e., 10-hour milled) ferrite inductor

shows a higher Q factor of 4.0 (7.2 at 1 GHz) than 2.8 (5.3 at 1 GHz) for the low anisotropy,

high-loss Co2Z (i.e., non-milled Co2Z) inductor at 2 GHz.

The measured characteristics of the double-side ferrite inductors well explain the

dependence of the dynamic properties of the ferrites on the inductor L and Q factor. The

magnetic properties for the two different Co2Z ferrites can be understood from the experimental

complex permeability in Fig. 10.8. The measured characteristics of the fabricated air-core and

ferrite inductors are summarized in Table 10.1.

10.3.4. Comparison of high anisotropy ferrite inductor to previous integrated inductors

For performance comparison, the electrical characteristics of the fabricated Zn-Co-Ni

ferrite film inductor are compared to various integrated inductors in Table 10.2. It is found that

the high magnetic anisotropy Zn-Co-Ni ferrite film inductor holds relatively high L-density of

18.3 nH/mm2 (measured) and Q factor of 20.0 at 2 GHz (simulated at Si σ = 0.01 S/m) compared

146
6

5
Inductance (nH)
4

3 aircore
ZnCoNi ferrite (bottom)
2 ZnCoNi ferrite (bottom)
+ Co2Z ferrite (top)
ZnCoNi ferrite (bottom)
1 + 10 h milled Co2Z ferrite (top)

0
0.1 1 10
Frequency (GHz)
(a)

14

12

10
Quality factor

6
aircore
4 ZnCoNi ferrite (bottom)
ZnCoNi ferrite (bottom)
+ Co2Z ferrite (top)
2 ZnCoNi ferrite (bottom)
+ 10 h milled Co2Z ferrite (top)
0
0.1 1 10
Frequency (GHz)
(b)

FIG. 10.10. Measured (a) L and (b) Q factor of the air-core, Zn-Co-Ni ferrite (bottom), Zn-Co-Ni
ferrite (bottom) + non-milled Co2Z ferrite (top), and Zn-Co-Ni ferrite (bottom) + 10-hour milled
Co2Z ferrite (top) inductors.

147
Table 10.1. Measured characteristics of the fabricated air-core and ferrite inductors at 2 GHz.
At 2 GHz L (nH) L increase (%) Q Max. Q
Air-core 4.2 - 6.7 14.1 at 0.55 GHz
Zn-Co-Ni ferrite film (bottom) 4.5 7.1 4.8 12.3 at 0.50 GHz
Zn-Co-Ni ferrite film (bottom) 5.6 33.3 2.8 12.6 at 0.35 GHz
+ non-milled Co2Z ferrite (top)
Zn-Co-Ni ferrite film (bottom) 4.8 14.3 4.0 11.7 at 0.43 GHz
+ 10 h-milled Co2Z ferrite (top)

Table 10.2. Performance comparison of various integrated inductors.


Inductor types L (nH) L/area (nH/mm2) Q factor at 2 GHz
MEMS 1.8 11.2 27.5
(spiral) [2]
Multi-stacked 8.2 283.7 9.6
(spiral) [3]
CoZrTa 100 1300 <1
(spiral) [6]
FeGaB/Al2O3 2.3 15 9.2
(solenoid) [8]
Ni-Zn-Cu ferrite 1.94 12 12.3
(spiral) [13]
Zn-Co-Ni ferrite 4.5 18.0 20.0
(spiral) [This work] (Simulated at Si σ = 0.01
S/m)

to those of the MEMS [2] and low-resistivity magnetic [6], [8] inductors. In addition, the high

saturation magnetization of 334 emu/cm3 and large in-plane Hk of 420 Oe of the fabricated Zn-

Co-Ni ferrite film result in higher L-density and Q factor than those of the low anisotropy Ni-Zn-

Cu ferrite inductor [13]. It is, therefore, suggested that high anisotropy ferrite inductors can

realize high L-density and Q GHz integrated inductors.

10.4. Conclusion

The GHz integrated Zn-Ni-Co ferrite film inductor was characterized for L and Q factor

in comparison to the air-core inductor. The Zn-Ni-Co ferrite film and air-core inductors showed

148
L-densities of 18.0 and 16.8 nH/mm2, respectively. The Q factors at 2 GHz were 4.8 for the Zn-

Ni-Co ferrite and 6.7 for the air-core inductors with high-conductivity Si substrate (σ = 13 S/m).

The Q factors of the ferrite and air-core inductors significantly increased to 20.0 and 31.9 at 2

GHz with the low-conductivity Si substrate (σ = 0.01 S/m). In addition, the high magnetic

anisotropy ferrite inductor showed a simulated Q of 34.3 at 2.1 GHz. The measured and

simulated results suggest that the combined technology of high anisotropy, low-loss GHz ferrites

and high Q-inductor design will give rise to high L-density and Q factor integrated inductors.

149
References

1. X. N. Wang, X. L. Zhao, Y. Zhou, X. H. Dai, and B. C. Cai, Microelectron. J., 36, 737
(2005).

2. E. C. Park, Y. S. Choi, J. B. Yoon, S. Hong, and E. Yoon, IEEE Trans. Microwave Theory
Tech., 51, 289 (2003).

3. X. Xu, P. Li, M. Cai, and B. Han, IEEE Trans. Electron Devices, 59, 2011 (2012).

4. J. Zhan, C. Yang, X. Wang, Q. Fang, Z. T. Shi, Y. Yang, T. L. Ren, A. Wang, Y. H. Cheng,


L. T. Liu, Sens. Actuators, A, 195, 231 (2013).

5. M. Yamaguchi, M. Baba, and K. I. Arai, IEEE Trans. Microwave Theory Tech., 49, 2331
(2001).

6. D. S. Gardner, G. Schorom, P. Hazucha, F. Paillet, T. Karnik, S. Borkar, R. Hallstein, T.


Dambrauskas, C. Hill, C. Linde, W. Worwang, R. Baresel, and S. Muthukumar, J. Appl.
Phys., 103, 07E927 (2008).

7. N. Sato, Y. Endo, and M. Yamaguchi, J. Appl. Phys., 111, 07A501 (2012).

8. Y. Gao, S. Zare, X. J. Yang, T. X. Nan, Z. Y. Zhou, M. Onabajo, K. P. O’Brien, U. Jalan, M.


El-tanani, P. Fisher, M. Liu, A. Aronow, K. Mahalingam, B. M. Howe, G. J. Brown, and N.
X. Sun, J. Appl. Phys., 115, 17E714 (2014).

9. R. P. Davies, C. Cheng, N. Sturcken, W. E. Bailey, and K. L. Shepard, IEEE Trans. Magn.,


49, 4009 (2013).

10. K. Kaneko, N. Inoue, N. Furutake, and Y. Hayashi, Jpn. J. Appl. Phys., 49, 04DB15 (2010).

11. W. Qu, X. H. Wang, and L. Li, J. Magn. Magn. Mater., 257, 284 (2003).

12. H. L. Cai, J. Zhan, C. Yang, X. Chen, Y. Yang, B. Y. Chi, A. Wang, and T. L. Ren, J.
Nanomater., 2013, 1 (2013).

13. C. Yang, F. Liu, X. Wang, J. Zhan, A. Wang, T. L. Ren, L. T. Liu, H. Long, Z. Wu, and X.
Li, IEEE Trans. Electron Devices, 56, 3141 ( 2009).

14. J. Lee, Y. K. Hong, S. Bae, J. Jalli, G. S. Abo, J. Park, W. M. Seong, S. H. Park, and W. K.
Ahn, J. Appl. Phys., 109, 07E530 (2011).

15. S. Bae, Y. K. Hong, J. J. Lee, J. Jalli, G. S. Abo, A. Lyle, I. T. Nam, W. M. Seong, J. S. Kim,
and S. H. Park, IEEE Trans. Magn., 45, 2557 (2009).

16. J. M. Yook, J. Hyun Ko, M. L. Ha, and Y. S. Kwon, IEEE Trans. Microwave Theory Tech.,
53, 2230 (2005).

150
17. L. Li, D. W. Lee, K. P. Hwang, Y. Min, T. Hizume, M. Tanaka, M. Mao, T. Schneider, R.
Bubber, and S. X. Wang, IEEE Trans. Adv. Packag., 32, 780 (2009).

18. K. Kondo, T. Chiba, H. Ono, and S. Yoshida, Y. Shimada, N. Matsushita, and M. Abe, J.
Appl. Phys., 93, 7130 (2003).

19. T. Tanaka, Y. K. Hong, S. H. Gee, M. H. Park, D. W. Erickson, and C. Byun, IEEE Trans.
Magn., 40, 2005 (2004).

151
CHAPTER 11

CONCLUSION

This dissertation provides an understanding of the roles of ferrites in RF antennas and

inductors. Also, the knowledge of the dynamic properties of ferrites, attained by analysis of the

fundamental characteristics of RF devices, reveals the ferrites’ importance in RF applications

and elucidates ferrite development techniques and device design methods. Ferrites have been

shown to possess the following characteristics: (i) the gyromagnetic nature of the ferrites gives

asymmetrical permeability tensor, leading to non-reciprocal wave propagation, (ii) the magnetic

permeability and dielectric permittivity of ferrites causes electromagnetic scaling and a decrease

in wavelength of an incident wave in a ferrite medium, (iii) the possession of both relative

permeability and permittivity higher than unity of ferrites results in improved impedance

matching at the interface of ferrite medium and free-space, and (iv) the high permeability and

electrical resistivity of ferrites provide a high inductive characteristic and low eddy current loss

at high frequencies, thereby counteracting a strong capacitive property and improving device

efficiency. In this dissertation research, the above characteristics of ferrites gave rise to

development of miniature RF antennas and inductors. The research accomplishments are

described below.

In ferrites, permeability becomes smaller than 2.5 in the low end of the gigahertz (GHz)

range according to the fundamental relation between frequency and permeability. Antenna

performance simulation results showed that a volume of the small 1.97-permeability ferrite

antenna was decreased by 30 % compared to the high 9.4-permittivity dielectric antenna. In

addition, the measured radiation efficiency of a fabricated 1.57 GHz small permeability (μr =

1.97) ferrite antenna was 66 % at 1.57 GHz, which was much higher than the simulated radiation

152
efficiency of 55.4 % for the high-permittivity (εr = 11) dielectric antenna. Basedon simulation and

experimental results, even small permeability of GHz ferrites greatly contributes to

miniaturization, bandwidth, and gain. The significance of this work is that it is the first attempt to

understand the effect of small permeability on GHz ferrite antennas.

Conventional material processing approaches, i.e., control of microstructure and

optimization of annealing process, have shown a decrease in magnetic loss of ferrites but were

still high for antenna applications. In this research, a significant reduction in magnetic loss

tangent of ferrites was demonstrated with a small DC magnetic field, thereby improving antenna

gain. Magnetic loss of the M-type BaFe9.6Co1.2Ti1.2O19 ferrite at 200 MHz decreased

significantly from 11.8 % to less than 1 % as the applied DC magnetic field increased from 0 to

400 Oe. Based on the loss reduction mechanism, small and high gain Ni0.5 Mn0.2Co0.07Fe2.23O4

ferrite antenna was developed with compact Nd-Fe-B permanent magnets. The volume of the

ferrite antenna is 0.346 cm3. The antenna gain increased by 54 % (from -10.58 to -4.81 dBi) with

two pairs of the magnet, which equal to a magnetic field of 1,600 Oe. Both experimental

dynamic magnetic properties of the ferrite and ferrite antenna performance results demonstrate

that the ferrite antenna gain can be signifincantly improved with small DC magnetic fields, while

mainintaining high miniaturization factor. These results suggest that ferrite antenna with

permanent magnet layers in a form of thick films can lead to future development of low-profile

and high-gain RF antennas.

Long-term evolution (LTE) is the latest standard in mobile network technology and

requires a multiple-input multiple-output (MIMO) antenna system. However, the limited space

of a mobile handheld device is a major challenge for development and implementation of high-

isolation LTE MIMO antennas. To address this issue, a miniature ferrite antenna array was

153
designed and developed for LTE MIMO applications. The ferrite MIMO antenna showed good

isolation of 16.4 dB at 720 MHz without conventional suspended line, bandstop filter, or

decoupling network. Accordingly, the developed miniature and high-isolation ferrite antenna

provides simple MIMO antenna system and also has about 11 times smaller volume than the

suspended line or bandstop filter LTE MIMO antennas. Thus, the novel ferrite antenna approach

to the latest technology’s challenge will make a significant impact on the development of next

generation MIMO systems.

Planar ferrite film and air-core inductor 3 × 3 arrays were fabricated on 4 inch silicon

wafer to characterize inductor performance. The ferrite inductor showed inductance (L) of 50 nH

and maximum Q-factor of 59 at 2.87 MHz. These values are a 22.5 % increase in L and 154 %

increase in Q-factor compared to those of an air-core inductor with the same structure. Based on

the studied ferrite inductor, a DC-DC converter was modeled, and the power efficiency of the

ferrite inductor DC-DC converter was calculated. The calculated power efficiency was about

91.7 % at a load current of 0.647 A. This performance was competitive to conventional spiral

inductors. The experimental and modeling results indicate that the developed ferrite inductors are

applicable to miniature and high-efficiency on-chip power module for mobile device

applications.

A high magnetic anisotropy Zn0.13Co0.04Ni0.63Fe2.2O4 (Zn-Co-Ni) ferrite film inductor

showed L of 4.5 nH at 2 GHz. The corresponding L-density of the ferrite inductor is 18.0

nH/mm2. An air-core inductor with the same coil structure has L of 4.2 nH and L-density of 16.8

nH/mm2. The Q factors at 2 GHz were 4.8 (8.5 at 1 GHz) for the Zn-Co-Ni ferrite film and 6.7

(12.5 at 1 GHz) for the air-core inductors. The low GHz Q factors of both the inductors are

attributed to the high substrate loss of the Si substrate (conductivity (σ) of Si = 13 S/m) used in

154
the inductor fabrication. The Q factors noticeably increased to 20.0 for the ferrite and 31.9 for

the air-core with the low-conductivity Si substrate (σ = 0.01 S/m). Furthermore, the high

anisotropy ferrite inductor showed a simulated Q of 34.3 at 2.1 GHz. The measured and

simulated results suggest that high-resistivity and large magnetic-anisotropy ferrites in

combination with high Q-inductor design can lead to the development of high L-density Q-

inductors up to 2 GHz for radio frequency integrated circuits (RFICs).

In this dissertation, the versatile properties of ferrites make a significant impact on a wide

range of RF passive devices. The device form factor and efficiency are closely related to the

frequency-dependent permeability and magnetic loss of ferrites. In addition, in emerging

technologies of wireless power transfer and near-field communications, ferrites’ dynamic

magnetic and electrical properties increase wireless signal and power transmission efficiency.

The development of low-loss and high-permeability ferrites is, therefore, indispensible in an

effort to meet needs of established and rapidly growing technologies.

155
CHAPTER 12

FUTURE WORK

Future work will need to address a low permeability and large magnetic loss of ferrite at

GHz frequency. One way to increase permeability is to crystallographically orient grains of the

ferrite, and the magnetic loss could be reduced when the grain size is in the range of single

magnetic domain size. Hexagonal ferrites have both in-plane (H ) and out-of-plane (Hθ)

anisotropies. Since H is smaller than Hθ, oriented hexagonal ferrites can have high permeability

compared to randomly oriented ferrites. For example, the permeability of Co2Z hexaferrite

increases from 11 to 32 as the degree of orientation increases from 0 (i.e. randomly oriented

grains) to 0.9 as shown in Fig. 12.1. Therefore, the orientation can increase the Snoek limit,

which is the product of ferromagnetic resonance frequency and permeability (μr).

50 Co2Z
CoZnZ
40
Permeability

30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
Degree of orientation

FIG. 12.1. Permeability of samples of Co2Z and CoZnZ with different degrees of crystal
orientation [1].

As discussed in the chapters 2 and 6, magnetizing process involves domain wall motion

and spin rotation. The domain wall motion needs to be depressed to minimize the domain wall

resonance, and therefore, magnetic loss tangent decreases. To depress the domain wall motion,

156
there are two approaches: (i) control of grain size below a critical radius and (ii) application of a

DC magnetic field.

In summary, future research in the area of RF ferrite devices will focus on development

of low-loss and high-permeability ferrites and low-temperature fabrication technique compatible

with complementary metal–oxide–semiconductor (CMOS) processing temperature. In addition,

future research will need to address a critical scientific challenge for understanding of magnetic

loss mechanisms.

References

1. J. Smit and H. P. J. Wijn, Ferrites, John Wiley & Sons, New York, p. 248 (1959)

157
APPENDIX

LIST OF PUBLICATIONS

FIRST-AUTHORED JOURNAL PAPERS

1. Jaejin Lee, Yang-Ki Hong, Changhan Yun, Woncheol Lee, Jihoon Park, and Byoung-Chul
Choi, “Magnetic parameters for ultra-high frequency (UHF) ferrite circulator design,”
Journal of Magnetics 19 (4), 399 (2014).

2. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, Jihoon Park, Terumitsu Tanaka, and Changhan
Yun, “Electrically small ferrite antenna gain with dc magnetic field for mobile device
application,” Microwave and Optical Technology Letters 56 (7), 1531 (2014).
3. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, and Jihoon Park, “Impedance matching of
electrically small antenna with Ni-Zn ferrite film,” Journal of Magnetics 18 (4), 428 (2013).
4. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, Gavin S. Abo, Jihoon Park, Won-Mo Seong, and
Won-Ki Ahn, “Control of magnetic loss tangent of hexaferrite for advanced RF antenna
applications,” Journal of Applied Physics 113, 073909 (2013).
5. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, Gavin S. Abo, Jihoon Park, Won-Mo Seong, and
Seok Bae, “Role of small permeability in gigahertz ferrite antenna performance,” IEEE
Magnetics Letters 4, 5000104 (2013).
6. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, Gavin S. Abo, Jihoon Park, Nicholas Neveu,
Won-Mo Seong, Sang-Hoon Park and Won-Ki Ahn, “Soft M-type hexaferrite for very high
frequency (VHF) miniature antenna applications,” Journal of Applied Physics 111, 07A520
(2012).
7. Jaejin Lee, Yang-Ki Hong, Woncheol Lee, Gavin S. Abo, Jihoon Park, Ryan Syslo, Won-
Mo Seong, Sang-Hoon Park and Won-Ki Ahn, “High ferromagnetic resonance and thermal
stability spinel Ni0.7 Mn0.3-xCoxFe2O4 ferrite for ultra high frequency devices,” Journal of
Applied Physics 111, 07A516 (2012).
8. Jaejin Lee, Yang-Ki Hong, Jihoon Park, Won-Mo Seong, Gi-Ho Kim, and Akimitsu
Morisako, “M-type hexaferrite for gigahertz chip antenna applications,” IEEE Magnetics
Letters 2, 5000204 (2011).

9. Jaejin Lee, Yang-Ki Hong, Seok Bae, Gavin S. Abo, Won-Mo Seong, and Gi-Ho Kim,
“Miniature Long-Term Evolution (LTE) MIMO ferrite antenna,” IEEE Antennas and
Wireless Propagation Letters, 10, 603 (2011).
10. Jaejin Lee, Yang-Ki Hong, Seok Bae, Gavin S. Abo, Jeevan Jalli, Jihoon Park, Won-Mo
Seong, Sang-Hoon Park, Won-Gi An, and Gi-Ho Kim, “Broadband Bluetooth antenna based
on Co2Z hexaferrite-glass composite,” Microwave and Optical Technology Letters 53, 1222
(2011).

158
11. Jaejin Lee, Yang-Ki Hong, Seok Bae, Jeevan Jalli, Gavin S. Abo, Jihoon Park, Won-Mo
Seong, Sang-Hoon Park, and Won-Ki Ahn, “Low loss Co2Z (Ba3Co2Fe24O41)-glass
composite for gigahertz antenna application,” Journal of Applied Physics 109, 07E530
(2011).

12. Jaejin Lee, Yang-Ki Hong, Seok Bae, Jeevan Jalli, Jihoon Park, Gavin S. Abo, Gregory W.
Donohoe, and Byoung-Chul Choi, “Integrated ferrite film inductor for power system-on-chip
(PowerSoC) smart phone applications,” IEEE Transactions on Magnetics 47, 304 (2011).
13. Jaejin Lee, Yang-Ki Hong, Seok Bae, Ji-Hoon Park, Jeevan Jalli, Gavin S. Abo, Ryan Syslo,
Byoung-Chul Choi, and Gregory W. Donohoe, “High-quality factor Ni-Zn ferrite planar
inductor,” IEEE Transactions on Magnetics 46 (6), 2417 (2010).
14. Jaejin Lee, Yang-Ki Hong, Seok Bae, Jeevan Jalli, Gavin S. Abo, Won-Mo Seong, Won-Gi
Ahn, Sang-Hoon Park, Chul-Jin Choi, and Jung-Goo Lee, “Broadband NixZn0.8-xCu0.2Fe2O4
electromagnetic absorber for 1 GHz application,” IEEE Transactions on Magnetics 45 (10),
4230 (2009).

15. Jaejin Lee, Seok Bae, Yang-Ki Hong, Jeevan Jalli, Gavin S. Abo, Won-Mo Seong, Sang-
Hoon Park, Chul-Jin Choi, and Jung-Gu Lee, “Novel Ni–Mn–Co ferrite for gigahertz chip
devices,” Journal of Applied Physics 105, 07A514 (2009).

BOOK CHAPTER

1. Yang-Ki Hong and Jaejin Lee, “Chapter 8 - Ferrites for RF Passive Devices,” Solid State
Physics, volume 64, pp. 237–329 (2013), ISBN: 978-0-12-408130-7.

159

You might also like