passive wireless radiation dosimeters for radiation

145
Passive Wireless Radiation Dosimeters for Radiation Therapy Using Fully-Depleted Silicon-on-Insulator Devices A DISSERTATION SUBMITTED TO THE FACULTY OF THE UNIVERSITY OF MINNESOTA BY Yulong Li IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOTOR OF PHILOSOPHY Prof. Steven J. Koester, Advisor January 2017

Upload: others

Post on 11-Apr-2022

21 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Passive Wireless Radiation Dosimeters for Radiation

Passive Wireless Radiation Dosimeters for Radiation

Therapy Using Fully-Depleted Silicon-on-Insulator Devices

A DISSERTATION

SUBMITTED TO THE FACULTY OF THE

UNIVERSITY OF MINNESOTA

BY

Yulong Li

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS

FOR THE DEGREE OF DOTOR OF PHILOSOPHY

Prof. Steven J. Koester, Advisor

January 2017

Page 2: Passive Wireless Radiation Dosimeters for Radiation

© Yulong Li 2017

Page 3: Passive Wireless Radiation Dosimeters for Radiation

i

Acknowledgement

I would like to thank my advisor Prof. Steven Koester for his continuous support,

guidance and encouragement throughout the time toward my Ph. D. degree. It’s been an

incredible journey for me in Koester Nano Devices Laboratory. Started as someone

knowing little about transistors and now becoming a devoted researcher in semiconductor

devices, I have not only learnt knowledge and skills from Prof. Koester but also found my

passion about what to do in the next stage of my life.

I would like to thank our collaborator, Prof. Bruce Gerbi for his engagement in our

research project. My dissertation project would not go far without Prof. Gerbi’s support to

guarantee enough beam time and help for setting up the clinical environment before every

single experiment.

I would also like to thank Prof. Paul Ruden and Prof. Stephen Campbell for teaching

me the fundamental of semiconductors and serving on my committee.

I am very thankful to Warren Poter for his help while conducting the radiation

experiments.

I want to thank Dr. David Deen and Dr. Ankur Khare for being a role model for me in

graduate school and teaching me how to do good research.

I also want to thank all my colleagues in Koester Nano Devices Laboratory in KNDL:

Dr. Eric Olson, Dr. Jing Li, Dr. Seon Namgung, Dr. Yoska Anugrah, Dr. Chaitanya

Kshirsagar, Dr. Mona Ebrish, Yang Su, Nazila Haratipour, Jiaxi Hu, Matt Robbins, Yao

Page 4: Passive Wireless Radiation Dosimeters for Radiation

ii

Zhang, Rui Ma, Saran Chaganti, Qun Su and Andrew Stephan. I have spent pleasant five

years with them. The insightful discussions in the office, the long fabrication days in clean

room, as well as the happy hours on Fridays are all beautiful memories I will carry on in

my life.

I also want to thank staffs at Minnesota Nano Center: Mark Fisher, Paul Kimani, Glenn

Kuschke, Kevin Roberts, Tony Whipple and Lage von Dissen, for their dedicated work to

maintain the equipment so that I can graduate in time.

Last but not least, I would like to thank my parents, Xiaohong Li and Huiru Lyu. They

may not understand my research, but I know no matter who happened they will always be

there for me with love and care. And my wife Yaliu He, who is not only my partner, but

also my best friend and my soul mate.

It’s my great honor to start my career at IBM where Prof. Koester started twenty years

ago. I will carry on what I learnt during my Ph. D. and keep learning. This dissertation ends

one chapter of my life and opens a new one. I am very much looking forward to it.

Page 5: Passive Wireless Radiation Dosimeters for Radiation

iii

To My Wife: Yaliu He

And My Parents: Xiaohong Li, And Huiru Lyu

Page 6: Passive Wireless Radiation Dosimeters for Radiation

iv

Abstract

Radiation cancer therapy is one of the most important methods of cancer treatment.

The goal of radiation therapy is to apply maximum amount of dose to the tumor region

while applying minimum amount of dose to the normal tissues. Being able to measure

radiation dose accurately is critical for radiation therapy because it will make sure the

radiation is properly delivered to the tumor region as planned. In this dissertation, the

development of a passive wireless radiation dosimeter for radiation cancer therapy based

on fully-depleted silicon-on-insulator (FDSOI) technology was presented through the

following steps.

In Chapter 1, the radiation cancer therapy and existing dosimetry technology was

introduced. Because of the high demand of modern radiation therapy and the deficiency of

existing dosimetry. An ultra-small, passive radiation dosimeter will greatly help the dose

verification in today’s radiation therapy. In Chapter 2, the building block of proposed

dosimeter technology, FDSOI technology, was introduced and cutting edge development

of FDSOI dosimeter was summarized. In Chapter 3, the radiation response of FDSOI

metal–oxide–semiconductor field-effect transistors (MOSFETs) over a wide range of

therapeutic X-ray beam energies, angles, dose ranges and applied substrate voltages were

characterized. The charge collection efficiency was calculated from both technology

computer aided design (TCAD) simulation and an analytical model to evaluate the

radiation effect quantitatively. Based upon the experimental results, the design of passive

wireless sensors using FDSOI varactors was proposed. In Chapter 4, the fabrication process

of FDSOI varactors was developed at University of Minnesota. Capacitance vs. voltage

Page 7: Passive Wireless Radiation Dosimeters for Radiation

v

(CV), current vs. voltage (IV), and TCAD modeling were used to fully understand the

device characterization. The capacitance based sensing of Cobalt-60 (Co-60) gamma

radiation was demonstrated using FDSOI varactors. An analytical model was developed to

further understand the device response under irradiation. Based on the analytical model,

the optimal working condition at different frequencies was discussed and a guide line of

improving the device sensitivity was provided. TCAD simulation was also conducted to

optimize the design space of FDSOI varactors for wireless sensing. In Chapter 5. 120-

finger varactors with 0 V threshold voltage were fabricated using electron beam

lithography with small device variability. An impedance measurement setup was built in

clinical environment. Completely passive wireless radiation sensing was demonstrated

using FDSOI varactor and the device showed good linearity of resonant frequency change

under radiation. The possibilities to improve the performance of device were discussed.

Chapter 6 summarized the progress of FDSOI dosimeter. The future development towards

more reliable dosimeter with smaller sized was proposed. A novel device structure with

two-dimensional material that could expend the application of solid-state radiation

dosimeter was also proposed.

Page 8: Passive Wireless Radiation Dosimeters for Radiation

vi

Table of Contents

List of figures ............................................................................................................... x

List of tables............................................................................................................... xv

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

1.1 Radiation therapy .......................................................................................... 1

1.1.1 Brief history of radiation therapy .......................................................... 1

1.1.2 Classification of radiation therapy ......................................................... 3

1.1.3 Generation of high-energy photons ....................................................... 5

1.2 Interactions between radiation and matter .................................................... 7

1.2.1 Interaction of photons and matter .......................................................... 7

1.2.2 Interaction of photons and SiO2........................................................... 10

1.3 Radiation dosimetry .................................................................................... 12

1.3.1 Optically stimulated luminescence dosimeter (OSLD) ....................... 12

1.3.2 Radiation-Sensing Field-Effect-Transistor (RADFET) ....................... 14

1.3.3 MOS capacitor ..................................................................................... 17

1.3.4 MEMS technology ............................................................................... 18

1.4 Dissertation goal ......................................................................................... 20

Chapter 2: FDSOI approach ................................................................................ 22

Page 9: Passive Wireless Radiation Dosimeters for Radiation

vii

2.1 SOI technology overview ........................................................................... 22

2.2 Advantage of FDSOI for CMOS ................................................................ 25

2.3 Total ionizing dose effects on FDSOI devices ........................................... 26

2.4 FDSOI radiation dosimeter ......................................................................... 27

2.5 Existing literatures about SOI radiation dosimeter ..................................... 29

Chapter 3: Radiation response of FDSOI MOSFETs ....................................... 31

3.1 Device structure and characterization ......................................................... 31

3.2 Experimental setup...................................................................................... 33

3.3 Experimental results.................................................................................... 36

3.4 Charge trapping efficiency .......................................................................... 41

Chapter 4: Radiation response of FDSOI varactors .......................................... 46

4.1 Device fabrication ....................................................................................... 46

4.1.1 Overview ............................................................................................. 46

4.1.2 Thinning of SOI layer .......................................................................... 48

4.1.3 Ion implantation of source/drain.......................................................... 50

4.1.4 Activation ............................................................................................ 51

4.1.5 Growth of gate oxide ........................................................................... 53

4.2 Device characterization ............................................................................... 55

4.3 Experimental setup...................................................................................... 58

Page 10: Passive Wireless Radiation Dosimeters for Radiation

viii

4.4 Experimental results.................................................................................... 61

4.5 Analytical model of FDSOI varactors ........................................................ 67

4.6 Simulation and optimization of FDSOI varactors ...................................... 76

4.6.1 Requirements of FDSOI varactor for wireless sensing ....................... 76

4.6.2 Device optimization ............................................................................. 77

Chapter 5: Passive wireless sensing using FDSOI devices ................................ 82

5.1 Theory of wireless sensing.......................................................................... 82

5.2 Device design and fabrication ..................................................................... 84

5.2.1 Overview ............................................................................................. 84

5.2.2 Electron beam lithography ................................................................... 85

5.2.3 Threshold voltage adjustment .............................................................. 87

5.3 Device characterization ............................................................................... 88

5.4 Experimental results.................................................................................... 92

5.5 Device integration with small size .............................................................. 98

Chapter 6: Conclusion and future work ........................................................... 100

6.1 Conclusion ................................................................................................ 100

6.2 Proposed future work ................................................................................ 101

6.2.1 Further development of FDOSI radiation dosimeter ......................... 101

6.2.2 Radiation dosimeter based on two-dimensional materials ................ 104

Page 11: Passive Wireless Radiation Dosimeters for Radiation

ix

Bibliography ............................................................................................................ 109

Appendix A: ............................................................................................................. 124

Appendix B: ............................................................................................................. 126

Page 12: Passive Wireless Radiation Dosimeters for Radiation

x

List of figures

Figure 1-1. Print of Wilhelm Röntgen's first "medical" X-ray, of his wife's hand . ..... 2

Figure 1-2. Spectral distribution of X rays calculated for a thick tungsten target. ....... 5

Figure 1-3. Energy level diagram for the decay of Co-60 nucleus. .............................. 6

Figure 1-4. Mass attenuation coefficient for soft tissues .............................................. 9

Figure 1-5. Major physical processes underlying radiation response of a Si-SiO2-Si

structure............................................................................................................................. 10

Figure 1-6. Schematic diagram of the energy levels of a crystalline material that

sustains optical luminescence ........................................................................................... 13

Figure 1-7. The structure of a typical RADFET and its working machenism ............ 15

Figure 1-8. Typical Id vs. Vg characteristic of a RADFET before and after radiation. 16

Figure 1-9. Structure of a MOS varactor dosimeter ................................................... 18

Figure 1-10. Structure of the MEMS microdosimeter. ............................................... 19

Figure 1-11. (a) Image of the MEMS dosimeter in a hypodermic needle. (b) typical

wireless sensing results of the MEMS dosimeter with different capacitor size. .............. 20

Figure 2-1. The Smart-Cut ® process. ........................................................................ 24

Figure 2-3. Working machenism of FDSOI wireless dosimeter ................................ 28

Figure 3-1 Optical image of IBM FDSOI MOSFETs ................................................ 31

Figure 3-2 Output characteristic of a FDSOI MOSFET ............................................. 32

Figure 3-3 Transfer characteristic of a FDSOI MOSFET ......................................... 33

Figure 3-4 Experimental setup for radiation dosimetry experiments. ........................ 34

Page 13: Passive Wireless Radiation Dosimeters for Radiation

xi

Figure 3-5. Id vs. Vgs curve for an FDSOI MOSFET after different dose fractions of

radiation ............................................................................................................................ 35

Figure 3-6. Threshold voltage shift under step irradiation ......................................... 36

Figure 3-7. Threshold shift for FDSOI MOSFETs at Vsub = +5 V ............................. 37

Figure 3-8. Threshold voltage shift under step irradiation with different energy X rays.

........................................................................................................................................... 38

Figure 3-9. Threshold voltage shift under step irradiation versus different angles of

beam incidence.................................................................................................................. 39

Figure 3-10. Sensitivity vs. absolute threshold voltage for all data ............................ 40

Figure 3-11. Points: simulated Id – Vgs curve with different amount of positive charge

at the Si/BOX interface. Line: experimental data. ............................................................ 43

Figure 4-1. Fabrication process of FDSOI varactor. ................................................... 47

Figure 4-2. Thicknesses of SOI wafers measured from spectroscopic ellipsometry .. 49

Figure 4-3. Simulation of implantation condition....................................................... 51

Figure 4-4. Optical image of a TLM .......................................................................... 52

Figure 4-5. TLM measurement result. ........................................................................ 53

Figure 4-6. Profilometer measurement of silicon layer after gate oxidation. ............. 54

Figure 4-7. Optical image of a typical FDSOI varactor.............................................. 55

Figure 4-8. (a) Cg vs. Vg characteristics of FDSOI varactor ....................................... 56

Figure 4-9. (a) Band diagram simulation of FDSOI varactor ..................................... 57

Figure 4-10 Operational region for radiation experiments. ........................................ 58

Figure 4-11 Setup for capacitance based radiation experiment.. ................................ 59

Page 14: Passive Wireless Radiation Dosimeters for Radiation

xii

Figure 4-12 Capacitance and dose vs. time for FDSOI varactors with stepped dose

profile ................................................................................................................................ 61

Figure 4-13. Change in capacitance vs. dose for two different delay times. .............. 62

Figure 4-14. Measurement error as a function of dose ............................................... 63

Figure 4-15. Capacitance change vs. dose under different substrate bias .................. 65

Figure 4-16. Comparison of capacitance vs. dose for one device when it was first

irradiated and when it was irradiated for 230 Gy ............................................................. 66

Figure 4-17. (a) Equivalent circuit model for FDSOI varactor. (b) Diagram showing

physical location of circuit elements................................................................................. 68

Figure 4-18. (a) Modeled and experimentally extracted inversion capacitance vs. gate

voltage, (b) predicted Cinv vs dose at 100 kHz compared to radiation experiment ........... 72

Figure 4-19. Maximum predicted sensitivity vs. frequency for an FDSOI varactor in

the subthreshold and strong inversion limits .................................................................... 74

Figure 4-20. Illustration of multi-finger design of FDSOI varactors. ......................... 77

Figure 4-21. (a) Cg vs. Vg. (b) The intrinsic resistance vs. Vg. (c) Calculated Q-factor

for devices with different Lg. ............................................................................................ 79

Figure 4-22. (a) Percent capacitance change vs. Dose. (b) Resistance vs. Dose. (c) Q-

factor vs. Dose for devices with different Tsi .................................................................... 79

Figure 4-23. (a) Percent capacitance change vs. Dose. (b) Resistance vs. Dose. (c) Q-

factor vs. Dose for device with different metal gate ......................................................... 80

Figure 4-24. Estimated Q-factor as a function of radiation dose at 1 GHz with optimized

design parameters.............................................................................................................. 81

Page 15: Passive Wireless Radiation Dosimeters for Radiation

xiii

Figure 5-1. Equivalent network of the inductively couple sensor system. ................. 82

Figure 5-2. (a) Etching of different layers. (b) Profilometer measurement indicating a

700 nm deep trench was successfully created. .................................................................. 86

Figure 5-3. (a) Optical image of a 120-finger FDSOI varactor. (b) Cg vs. Vg

characteristic of 4 identical devices. ................................................................................. 87

Figure 5-4. (a) “Partially wireless” measurement setup. (b) “Fully wireless”

measurement setup ............................................................................................................ 88

Figure 5-5 (a) Cg vs. Vg characteristics at Vsub = 0, +2, +4, +6 V. (b) equivalent series

resistance vs. Vg at same Vsub. ........................................................................................... 89

Figure 5-6 (a) Read inductor resistance vs. frequency. (b) Resonant frequency vs. the

capacitance at Vg = 0 V and Vsub = 0 to 6 V. ..................................................................... 90

Figure 5-7. Estimated Q-factor at different Vsub. ........................................................ 91

Figure 5-8. Picture of experimental device ................................................................. 93

Figure 5-9. Cg vs. Vg characteristic of the experimental device .................................. 94

Figure 5-10. (a) Read inductor resistance vs. frequency at different radiation dose. (b)

Measured resonant frequency vs. dose compared with model. ........................................ 95

Figure 5-11. (a) Read inductor resistance vs. frequency at different radiation dose for

supercapacitor. (b) Resonant frequency vs. radiation dose ............................................... 96

Figure 5-12 (a) Integrated device with commercial inductor and capacitor. (b) Wireless

readout result of the device at left. .................................................................................... 99

Figure 6-1. SEM image of the gate-level metal deposition of a 1000-finger varactor

with Lg = 200 nm ............................................................................................................ 102

Page 16: Passive Wireless Radiation Dosimeters for Radiation

xiv

Figure 6-2. Proposed structure of final dosimeter. ................................................... 103

Figure 6-3. (a) Proposed new device structure of radiation dosimeter. (b) Band

alignment of the charge trapping layer ........................................................................... 107

Page 17: Passive Wireless Radiation Dosimeters for Radiation

xv

List of tables

Table 1-1. Summary of existing radiation dosimeter technologies. ........................... 21

Table 3-1 Extracted charge trapping efficiency for different substrate bias voltages and

simulation conditions ........................................................................................................ 45

Table 4-1. Nominal FDSOI varactor simulation parameters ...................................... 78

Page 18: Passive Wireless Radiation Dosimeters for Radiation

1

Chapter 1: Introduction

1.1 Radiation therapy

Radiation therapy is one of the most common and successful methods used for the

treatment of cancer. It utilizes high-energy radiation (e.g. X rays, gamma rays or charged

particles), originating either from sources external to the body or from implanted

radioactive material to kill malignant cells [1]. According to the National Cancer Institute,

about half of all cancer patients receive radiation therapy during the course of their

treatment [2]. Annually, over one million people are treated with radiation therapy and the

number is increasing each year [3].

1.1.1 Brief history of radiation therapy

The history of radiation therapy can be tracked back to the days of the discovery of X

rays. In 1895, in the study of cathode rays (streams of electrons) in a gas discharge tube,

German physics professor Wilhelm Röntgen observed that another type of radiation was

produced and could be detected outside of the tube. This radiation could penetrate opaque

substance, produce fluorescence, blacken a photographic plate, and ionize a gas. He named

the new radiation X rays. Figure 1-1 shows the print of Wilhelm Röntgen's first "medical"

X-ray of his wife's hand [4].

The first therapeutic use of X rays in cancer quickly followed this initial discovery.

Only 7 months after Roentgen’s discovery, a 1896 issue of the Medical Record described

a patient with gastric carcinoma who had benefited from radiotherapy [5]. After the

Page 19: Passive Wireless Radiation Dosimeters for Radiation

2

discovery of natural radioactivity by Antoine-Henri Becquerel in 1896, by 1904, patients

were undergoing implantation of radium tubes directly into tumors, representing some of

the first interstitial brachytherapy treatments. Early radiotherapy consisted of a single

massive dose of radiation, typically lasting an hour. Side effects were severe. In the early

1920s French radiation oncologist Henri Coutard pioneered the use of fractionated

radiotherapy in a wide variety of tumors without disastrous side effects [6]. This approach,

known as fractionation, is one of the most important underlying principles in radiation

therapy.

Although promising as a therapeutic modality, an important limitation of the early X

ray machines was their inability to produce high-energy, deeply penetrating beams.

Figure 1-1. Print of Wilhelm Röntgen's first "medical" X-ray, of his wife's hand [4].

Page 20: Passive Wireless Radiation Dosimeters for Radiation

3

Therefore, it was difficult to treat deep-seated tumors without excessive skin reactions. In

the 1960s, high-energy (megavoltage) treatment machines, known as linear accelerators or

linacs, were introduced. Linacs were capable of producing high-energy, deeply penetrating

beams, allowing for the very first time treatment of tumors deep inside the body without

excessive damage to the overlying skin and other normal tissues [7]. Another major

development at this time was the production of teletherapy units that used cobalt-60 as their

source, which produces 1.17 and 1. 33 MeV radiation. The relatively low cost and

availability of these units made them the favored source of high-energy beams, and this

popularity was a major stimulus for the subsequent growth of the field.

Advances in radiation physics and computer technology during the last quarter of the

20th century made it possible to aim radiation more precisely. For example, conformal

radiation therapy (CRT) uses computed tomography (CT) images and special computers to

precisely map the location of a cancer in 3 dimensions [8]. The radiation beams are matched

to the shape of the tumor and delivered to the tumor from several directions. Intensity-

modulated radiation therapy (IMRT) is similar to CRT, but along with aiming photon

beams from several directions, the intensity (strength) of the beams can be adjusted [9].

This gives even more control in decreasing the radiation reaching normal tissue while

delivering a high dose to the tumor.

1.1.2 Classification of radiation therapy

Depending on the position of the radiation source, there are three main divisions of

radiation therapy: systemic radioisotope therapy, brachytherapy, and external beam

radiation therapy.

Page 21: Passive Wireless Radiation Dosimeters for Radiation

4

In systemic radiation therapy, a patient swallows or receives an injection of a

radioactive substance, such as radioactive iodine or a radioactive substance bound to a

monoclonal antibody. Radioactive iodine (I-131) is commonly used to help treat certain

types of thyroid cancer [10]. Thyroid cells naturally take up radioactive iodine. For some

other types of cancer, a monoclonal antibody helps deliver the radioactive substance to the

target [11]. The antibody joins to the radioactive substance and travels through the blood,

locating and killing tumor cells.

Brachytherapy uses sealed radioactive sources placed precisely in the area under

treatment. In brachytherapy, radioactive isotopes are sealed in tiny pellets or “seeds.” These

seeds are placed in patients using delivery devices, such as needles, catheters, or some other

type of carrier. As the isotopes decay naturally, they give off radiation that damages nearby

cancer cells [12].

External beam radiotherapy is the most common form of radiotherapy. During the

therapy, the patient sits or lies on a couch and an external source of radiation is pointed at

a particular part of the body. External beam radiation therapy is most often delivered in the

form of photon beams (either X rays or gamma rays) because of their high energy and large

penetration depth [13]. Electron beams are also used as radiation source to irradiate

superficial tumors, such as skin cancer [14]. Proton therapy is another external beam

radiotherapy under development. Compared to photons, protons deposit much of their

energy at the end of their path therefore reducing the exposure of normal tissue to radiation

[15].

Page 22: Passive Wireless Radiation Dosimeters for Radiation

5

For the interest of this thesis, the radiation source used here are high-energy photos,

either X rays generated from a linac or gamma rays generated from Co-60 [13].

1.1.3 Generation of high-energy photons

The high-energy X rays used for radiation therapy generated from a linac. A linac uses

microwaves to accelerate electrons and allows them to collide with a heavy metal target to

produce high-energy X rays. There are two different types of X rays produced:

bremsstrahlung X rays and characteristic X rays.

The bremsstrahlung process is the result of radiative interactions between a high speed

electron and a nucleus. As the electron passes in the vicinity of a nucleus, due to coulomb

forces of attraction, it suffers a sudden deflection and acceleration. As a result, part or all

of its energy is dissociated from the electron and propagates in space in the form of photons.

Since the bremsstrahlung interaction may result in partial or complete loss of electron

Figure 1-2. Spectral distribution of X rays calculated for a thick tungsten target.

Page 23: Passive Wireless Radiation Dosimeters for Radiation

6

energy, the resulting bremsstrahlung photon may have any energy up to the initial energy

of the electron.

Characteristic X rays are produced when electrons interact with the atom and an

electron is ejected from an inner orbital. When a vacancy is created in an orbit, an outer

orbital electron will fall down to fill that vacancy. In doing so, the energy is radiated in the

form of characteristic X rays. Because of the discrete energy level of the orbitals, the

characteristic X rays are emitted at discrete energies. Figure 1-2 shows the spectral

distribution of X rays calculated for a thick tungsten target, along with the discrete energy

characteristic X rays [13].

The gamma rays used in radiation therapy are produced from the decay of, Co-60,

which has a half time of 5.27 years. Co-60 decays by beta minus decay to the stable isotope

Figure 1-3. Energy level diagram for the decay of Co-60 nucleus.

Page 24: Passive Wireless Radiation Dosimeters for Radiation

7

Ni-60. The activated nickel nucleus emits two gamma rays with energies of 1.17 MeV and

1.33 MeV. Figure 1-3 shows the energy level diagram for the decay of a Co-60 nucleus.

1.2 Interactions between radiation and matter

When X rays or gamma-rays passes through a medium, energy will be transferred to

the medium due to the interactions between photons and matter. This energy transfer

process first causes the ejection of electrons from the atoms of the absorbing medium. Then

these high-speed electrons transfer their energy by ionizing or exciting the atoms along

their paths. Most of the absorbed energy is converted into heat. If the absorbing medium

consists of body tissues, sufficient energy may be deposited within the cells, destroying

their reproductive capacity [16].

1.2.1 Interaction of photons and matter

An X ray or gamma-ray beams usually consist of large number of photons. Once they

pass through matter, the numbers of the photons will decrease due to the interactions with

matter. If the intensity of photons is I and the thickness of absorber is dx, then:

IdxdI , 1-1

where µ is the constant of proportionality, called the attenuation coefficient. The minus

sign indicated that the number of photons decreases as the absorber thickness increases.

The differential equation can be solved to yield the following equation:

uxeIxI 0)( , 1-2

where I(x) is the intensity transmitted by a thickness x and I0 is the incident intensity on the

absorber. The attenuation produced by a thickness x depends on the number of electrons

Page 25: Passive Wireless Radiation Dosimeters for Radiation

8

presented in that thickness. Therefore, µ depends on the density of the material. Usually

the µ is divided by the density ρ, resulting int more fundamental coefficient (µ/ ρ), called

the mass attenuation coefficient.

The attenuation of photon beam by an absorbing material is caused by four major types

of interactions. The first interaction is called Rayleigh Scattering. This process can be

visualized as a photon scattered by an electron elastically. Photon’s energy is preserved

during the scattering. Only the incident angle of the photon is slightly changed. This

process is most-probable in high atomic number materials with low-energy photons.

The second effect is called the photoelectric effect. When a photon interacts with an

atom, the entire energy of the photon can be absorbed by the atom which then ejects an

orbital electron. After the electron has been ejected from the atom, a vacancy is created in

the shell, thus leaving the atom in an excited state. The vacancy can be filled by an outer

orbital electron with the emission of characteristic X rays or Auger electrons.

The third interaction is known as the Compton effect, where the photon interacts with

an atomic electron as though it were a free electron. In this interaction, the electron receives

some energy from the photon and is emitted from the atom. The photon, with reduced

energy is scattered to a different angle. The overall energy of the incident photon is shared

between the photon and emitted electron after the interaction.

Finally, the fourth interaction is production of an electron/positron pair. In this process,

the photon interacts strongly with the electromagnetic field of an atomic nucleus and

produces a electron and positron pair. Because the rest mass energy of the electron is

Page 26: Passive Wireless Radiation Dosimeters for Radiation

9

equivalent to 0.51 MeV, a minimum photon energy of 1.02 MeV is required to create the

pair production.

Figure 1-4 shows the mass attenuation coefficient for soft tissues as a function of

photon energy, for different interactive process [17]. For the energy range of radiation

therapy, the photoelectric effect and Compton scattering dominate the interaction between

photons and matter.

Figure 1-4. Mass attenuation coefficient for soft tissues as a function of photon energy,

for different effects. © 2011, Lippincott Williams

Page 27: Passive Wireless Radiation Dosimeters for Radiation

10

1.2.2 Interaction of photons and SiO2

As will be shown in the next section, SiO2 is an important material to measure radiation

dose. The interaction of photons with SiO2 builds the foundation of this thesis. Therefore,

the interaction of photons with SiO2 will be briefly introduced in this section and more

details can be found in [18]. Figure 1-5 shows the schematic energy band diagram of a Si-

SiO2-Si system, where a positive voltage is applied to the Si on the left side. Upon

irradiation, a sequence fo four processes will contribute to the total ionizing effect.

Firstly, when radiation passes through SiO2, electron-hole pairs are created due to the

photoelectric effect and Compton scattering. It takes about 18 eV to create one electron-

hole pair in SiO2 [19]. Based on this value, one can calculate the charge pair volume density

per unit dose, g0 = 8.1 × 1014 pairs/cm3-Gy, where Gy is the unit of radiation dose, defined

Figure 1-5. Major physical processes underlying radiation response of a Si-SiO2-Si

structure.

Page 28: Passive Wireless Radiation Dosimeters for Radiation

11

as 1 J of radiation been absorbed by 1 kg of matter. Because the electrons in the SiO2 are

much more mobile than holes [20], the electrons will be swept out of the oxide rapidly,

usually on the order of picoseconds. However, some of the electrons will recombine with

the holes. The fractions of holes escaping recombination, fy(Eox), is determined mainly by

two factors: the magnitude of electrical field which separates the electron-hole pairs, and

the initial line density of charge pairs created by the incident radiation [21], [22]. This

process is labeled as A in Figure 1-5. The holes that did not recombine will “slowly” move

to the SiO2/Si interface through a hopping process [23], as labeled as B in Figure 1-5. The

time for this process can vary substantially but normally takes less than 1 second at room

temperature, where the precise time is determined by the applied field, temperature, oxide

thickness, and oxide processing history [24]–[26].

When the holes reach the SiO2/Si interface, some will fall into relatively deep long-

lived trap states, labeled as C in Figure 1-5. The deep hole traps near the interface originate

from the transition region where the oxidation is not complete. This region contains oxygen

vacancies: when there is one oxygen atom missing from the SiO2 lattice configuration, a

weak Si-Si bond will form. The incident holes can break this Si-Si bond, and leading to

trapping of the holes [27]. The trapped holes are relatively stable, but they do undergo a

long-term annealing process due to tunneling or thermal excitation [28], [29]. This process

can extend for hours or even years.

Radiation can also induce interface traps within the Si band gap, as labeled D in Figure

1-5. In the Si lattice, each Si atom is bonded with four other Si atoms. On the Si side of the

SiO2/Si interface, due to incomplete oxidation, some Si atoms are only bonded with three

Page 29: Passive Wireless Radiation Dosimeters for Radiation

12

other Si atoms, with a dangling bond extending into the oxide. In subsequent processing,

almost all these centers are passivated by reacting with hydrogen and forming Si-H bonds.

When holes transport through the oxide, they free hydrogen and form protons. When the

protons reach the interface, they react, breaking the Si-H bonds and form H2 and a trivalent

Si defect [30]–[32]. These defects are amphoteric, negatively charged above mid-gap,

neutral near mid-gap and positively charged below mid-gap.

1.3 Radiation dosimetry

The primary goal of radiation therapy is to maximize the dose to the tumor, while

minimizing the dose to the normal adjacent tissues. Therefore, it is important to be able to

measure the radiation dose accurately and make sure the proper amount of radiation is

delivered. The advancement of modern radiation therapy technology, such as intensity

modulated radiotherapy [33] and image-guided radiation therapy [34] make dose

verification a more and more complicated problem. The requirements of clinical radiation

dosimetry for modern technology are as follows: the accuracy of the dose measurement

has to be predictable up to 100 Gy, and the dose resolution should be as small as 1 cGy and

position resolution within 1 mm. In this section the existing radiation dosimetry technology

will be briefly summarized and their advantages and disadvantages will be discussed.

1.3.1 Optically stimulated luminescence dosimeter (OSLD)

Optically stimulated luminescence dosimeters (OSLDs) use the optically stimulated

luminescence technique to detect radiation. Figure 1-6 shows the schematic diagram of the

energy levels of a crystalline material that sustains optical luminescence. Pure crystalline

dielectric materials either contain or have added trace amounts of contaminations that form

Page 30: Passive Wireless Radiation Dosimeters for Radiation

13

crystal-lattice imperfections. These imperfections act as traps for electrons or holes and

also can act as luminescence centers, which emit light when electrons and holes recombine

near them. After radiation, free electrons and holes can be generated and trapped in the

forbidden band (1-5 in Figure 1-6). Figure 1-6 contains one hole trap (2) as recombination

center and three types of electron traps representing shallow traps (3), dosimetric traps (4)

and deep traps (5). Shallow traps are unstable at room temperature and can only hold charge

for very short periods of time. Deep traps can only release charge at very high temperature

or stimulated with ultraviolet light. Dosimetric traps are energetically deep enough to hold

charge at room temperature for long periods of time, but are not so deep that the charge

can be released by illumination with visible light. Once the trapped charges escape,

electron–hole recombination is possible with resulting in luminescence. The total

Figure 1-6. Schematic diagram of the energy levels of a crystalline material that sustains

optical luminescence

Page 31: Passive Wireless Radiation Dosimeters for Radiation

14

luminescence associated with a particular trapping level is proportional to the trapped

charge concentration and, ideally, to the absorbed dose of radiation [35]–[37].

Therefore, in OSLDs, the trapped charge concentration provides a record of the total

dose absorbed by the crystal. This record can be “read” by stimulating the trapped charges

using a light source to cause recombination of electron and hole and measuring the

luminescence using a photomultiplier tube [38]. The advantage of OSLD is that it provides

an accurate measurement for low radiation dose < 20 Gy with small or no dependence on

beam quality, temperature, and angle of irradiation [39]. It also has a low radiation

detectable threshold of 0.1 mGy [40] and is re-usable.

However, since OSLDs requires a photomultiplier tube to read the radiation dose, it

cannot be used for real time measurements. They are also sensitive to light and therefore

require extra packaging. Even though some groups have demonstrated a prototype OSLD

dosimeter system for in vivo measurements [38][41], an optical fiber was needed to deliver

the dosimeter into the body, making the measurement inconvenient and complicated.

1.3.2 Radiation-Sensing Field-Effect-Transistor (RADFET)

Radiation-Sensing Field-Effect-Transistors (RADFETs) are another technology that

has been used in radiation detection. Figure 1-7(a) shows the structure of a typical

RADFET which is simply a p-MOSFET with a thick gate oxide. The oxide thickness varies

from a few hundred nano-meters to a few micro-meters [42]. The absorption of radiation

(Figure 1-7(b)) will create electron-hole pairs in the thick SiO2 oxide due to. Under a

positive gate bias, the electrons will be swept out of gate oxide while the holes will transfer

Page 32: Passive Wireless Radiation Dosimeters for Radiation

15

towards the channel and be trapped at the SiO2 and Si channel interface. The trapped holes

will induce electrons in the channel and turn the device toward the “off” state.

Figure 1-8 shows the typical Id vs. Vg characteristic curve on a semi-log scale. Radiation

will cause the curve to shift leftward. Therefore, if the device can be biased using a constant

current source, the radiation dose can be characterized as a shift of gate bias (threshold

voltage) [43]. The RADFET has been used in space [44]–[46] as well as clinical

applications [47]–[49]. The thick gate oxide in RADFET makes it very sensitive to

radiation and since the threshold voltage is anti-proportional to oxide thickness [50], it is

able to detect small radiation doses. Also since the fabrication process is compatible with

standard complementary metal–oxide–semiconductor (CMOS) technology, new structure

and functionality could be added to the RADFET device, as shown later.

One disadvantage of RADFETs is that the gate oxide is used for both radiation charge

collection and read out. During the radiation, a constant gate bias needs to be applied to

separate the electron-hole pairs while in the reading period, a different gate bias need to be

Figure 1-7. (a) The structure of a typical RADFET. (b) After irradiation, the trapped holes in

SiO2 and Si interface will turn the device towards “off” state.

Page 33: Passive Wireless Radiation Dosimeters for Radiation

16

applied to measure the threshold voltage. As a result, the device needs to switch between

“write” and “read” mode during and after radiation. This makes the measurement

complicated. Some groups adopted flash memory technology into radiation sensing and

used a floating gate structure at the gate. Before radiation, some charges are pre-stored in

the floating gate to create a built-in electric field to help separate the radiation generated

electron-hole pairs [51]–[55]. Therefore, no external bias is needed during radiation and

the measurement is simplified.

However, this voltage measurement technique all requires a “wired” configuration,

which makes the radiation measurement inconvenient especially for in vivo dosimetry.

Some groups have combined RADFETs with a RF transmitter and were able to make

wireless measurement [56], [57]. Particularly, Sicel Technology developed an implantable

Figure 1-8. Typical Id vs. Vg characteristic of a RADFET before and after

radiation.

Page 34: Passive Wireless Radiation Dosimeters for Radiation

17

wireless radiation dosimeter that has good dose reproducibility and accuracy [58], [59].

However, due to the active electronic component in their device, the size of device was

still too large for convenient implantation.

1.3.3 MOS capacitor

As an alternative to building the dosimeter using active electronic components, it is

also possible to sense radiation passively. One approach is to use MOS capacitors as a

radiation sensitive variable capacitor (varactor). Figure 1-9 shows the structure and

working mechanism of a MOS varactor. The varactor consists of a gate metal electrode, a

thick SiO2 layer for absorption of radiation and a lightly-doped n-type Si substrate. The

total capacitance, Ctot measured from the gate and substrate is equal to the series

combination of oxide capacitance, Cox and silicon layer capacitance, Csi,

siox

siox

totCC

CCC

, 1-3

where Cox, is a fixed value which depends on the oxide thickness, while Csi depends on the

depletion layer thickness and can be modulated by gate bias or radiation. Just like RADFET,

radiation generated holes could be trapped at the SiO2/Si interface and decrease the

depletion layer thickness in silicon. As a result, Ctot will change as a function of radiation.

If connected with an inductor to form a resonant circuit, the resonant frequency will change

as a function of radiation and can be potentially measured wirelessly. The varactor and

wireless sensing concept is similar to the FDSOI approach in my dissertation and will be

discussed in more details in next chapter.

Page 35: Passive Wireless Radiation Dosimeters for Radiation

18

Even though the MOS varactor enables the possibility of passive wireless sensing [60],

[61], no wireless sensing has been demonstrated in the literature using a MOS varactor to

date. One disadvantage of this structure is that it requires a thick SiO2 layer to be able to

detect low doses of radiation. Since the capacitance is anti-proportional to SiO2 layer

thickness, the capacitance per unit area of a MOS varactor is too low and this makes the

device less scalable. Also since the capacitance is modulated by the depletion capacitance,

the tuning range of the total capacitance is small. This could limit the detection range of

radiation.

1.3.4 MEMS technology

Another way to realize a passive varactor is to use MEMS technology. C. Son et al.

demonstrated a microdosimeter composed of a radiation sensitive parallel plate capacitor

Figure 1-9. Structure of a MOS varactor and its working mechanism for radiation

sensing.

Page 36: Passive Wireless Radiation Dosimeters for Radiation

19

and an inductor as shown in Figure 1-10 [62]. In that work, the bottom plate of a capacitor

is designed to be deflectable and an electret layer with pre-stored charge is placed under

the top metal electrode. Radiation induced electron-hole pairs in the air gap will be

collected from the electrode and reduce the surface charge density. As a result, the force

between two parallel plates will decrease and the air gap will increase. In this way, the

capacitance of the parallel plate capacitor will decrease as a function of radiation dose and

the radiation can be detected as a changing of resonant frequency.

The dosimeter can be measured wirelessly and implanted with a hypodermic needle.

Figure 1-11 shows the image of the dosimeter in a hypodermic needle (a) and a typical

wireless measurement result with different capacitor sizes (b) [62]. The overall size of the

dosimeter was 2.5 mm in diameter and 2.8 cm in length. Thus, is still too large to be

Figure 1-10. Structure of the MEMS microdosimeter.

Page 37: Passive Wireless Radiation Dosimeters for Radiation

20

implanted easily and since the air gap between two plates is usually in micro-meters, the

low capacitance per unit area makes further scaling difficult.

1.4 Dissertation goal

Overall, for each of the existing technologies, some drawbacks exist. OSLDs have high

accuracy and detectability but they are difficult for in vivo measurement. RADFETs are

based on CMOS technology and have high reliability but their active electronic component

is not preferred for in vivo radiation sensing; MOS capacitors and MEMS capacitors enable

passive wireless sensing and greatly simplify the measurement but their sizes are too large

for easy implantation. Table 1-1 summarizes the main advantage and disadvantage for

existing dosimeter technologies.

Figure 1-11. (a) Image of the MEMS dosimeter in a hypodermic needle. (b) typical wireless

sensing results of the MEMS dosimeter with different capacitor size. © 2008 IEEE

Page 38: Passive Wireless Radiation Dosimeters for Radiation

21

The goal of this dissertation is to combine the reliability of CMOS and convenience of

passive wireless sensing to realize a passive wireless radiation dosimeter for in vivo

radiation therapy measurements. This will be accomplished using fully-depleted silicon-

on-insulator (FDSOI) technology, which will be described in the next chapter.

Table 1-1. Summary of existing radiation dosimeter technologies.

Page 39: Passive Wireless Radiation Dosimeters for Radiation

22

Chapter 2: FDSOI approach

2.1 SOI technology overview

Silicon-on-insulator (SOI) wafers consist of a bulk silicon substrate, on top of which is

a SiO2 layer and then a top silicon thin film which is used to build active devices and

circuits. If the SOI wafer is used in a MOSFET process, and the thickness of the top silicon

layer is larger than its depletion depth, it is called partially-depleted SOI (PDSOI). If the

thickness of the top silicon layer is smaller than its depletion depth, is the wafer is known

as fully-depleted SOI (FDSOI).

One of the early motivations in the development of SOI technology is from the

radiation hard properties of the SOI devices. When bulk MOSFETs are exposed to a

radiation environment, in the silicon region where radiation passes, electron-hole pairs are

generated. If small electric field exists, either from external bias or internal built-in

potential difference, to separate the electrons and holes, a large amount of current will be

produced. The current may flow to the active silicon region and seriously affect the device

operation and reliability [63]. For SOI devices, the existence of the insulating SiO2 layer

isolates the active thin-film region therefore, radiation induced current in the substrate has

little influence on device performance. For this reason, undesired radiation effects, such as

single event effects, can be greatly reduced [64], [65].

Initially sapphire was used as the insulating layer [66], [67]. Since the 1980s, SiO2

replaced sapphire as the buried oxide layer due to the success of SIMOX (separation by

implantation of oxygen) technology [68], [69]. Many other technologies have also been

Page 40: Passive Wireless Radiation Dosimeters for Radiation

23

developed to produce the single-crystal silicon film on top of the buried oxide insulator.

Some of them are based on the epitaxial growth of silicon on a crystalline insulator

(heteroepitaxial techniques) [70], [71]. Some techniques are based on the crystallization of

a thin silicon layer from the melt (laser recrystallization [72], e-beam recrystallization [73]

and zone-melting recrystallization [74]). The silicon thin film could also be produced from

a bulk silicon wafer by isolating a thin silicon layer from the substrate through the

formation and oxidation of porous silicon (FIPOS) [75] or through the ion beam synthesis

of a buried insulator layer [76]. Finally, the top silicon layer can be obtained by bonding

the silicon wafer to an insulator or a mechanical substrate followed by thinning the top

silicon layer (wafer bonding [77], BESOI [78]).

State of the art technology to produce the SOI layer combines ion implantation (Smart-

Cut® process [79], [80]) and wafer bonding technique to transfer a thin surface layer from

a wafer onto another wafer. The basic steps of this technique are shown in Figure 2-1. The

first step is ion implantation of hydrogen ions into an oxidized silicon wafer. The implanted

dose is on the order of 5 × 1016 cm-2. After implantation, micro-cavities and micro-bubbles

are formed at the depth of implantation. The top oxide layer will become the buried oxide

of SOI structure at the end of the process (Figure 2-1 a). Next, bonding of wafer A to a

handle wafer (B) is performed (Figure 2-1 b). Wafer B can be either bare or oxidized

depending on the desired thickness of the buried oxide layer. Then a two phase annealing

of the bonded wafer is performed. The first annealing step takes place at low temperature,

about 500 °C, during which, the implanted wafer A splits into two parts: a thin layer of

monocrystalline silicon which remains bonded onto wafer B and the remainder of wafer A

that can be recycled for later use (Figure 2-1 c). The second annealing take place at higher

Page 41: Passive Wireless Radiation Dosimeters for Radiation

24

temperature, about 1100 °C to strengthen the bond between the handle wafer and the SOI

film. Finally, chemo-mechanical polishing is performed on the SOI film to reduce the

surface roughness (Figure 2-1 d). The Smart-Cut® process produces good quality SOI

wafers and is also economical. It requires only N+1 wafers to produces N SOI wafers,

while the BESOI process require 2N wafers to produce N SOI wafers.

Figure 2-1. The Smart-Cut ® process (a) Hydrogen implantation. (b) Wafer bonding. (c)

Annealing and splitting of wafer A. (d) Polishing of both wafers.

Page 42: Passive Wireless Radiation Dosimeters for Radiation

25

2.2 Advantage of FDSOI for CMOS

Besides radiation hardening for single event effects, FDSOI technology has other

advantages over conventional bulk Si technology, making it attractive for CMOS

applications.

As the scaling of device continues, short channel effects (SCEs) become more and more

significant and detrimental for device performance. SCEs occur as the channel length

shrinks, and the controllability of the gate over the channel depletion region reduces due to

the increased charge sharing from source/drain. This causes a lack of pinch off and a

threshold voltage roll-off as well as drain-induced barrier lowing (DIBL). Moreover, SCEs

will also degrade the subthreshold slope and increase the subthreshold current. In bulk

silicon MOSFETs, SCEs are controlled by a number of techniques, including forming

shallow junctions and increasing the body doping. However excessively large doping

degrades the carrier mobility and junction capacitance and gives rise to band to band

tunneling leakage current. In FDSOI structure, the silicon film is completely depleted

compared to PDSOI and the gate voltage has excellent control of silicon channel potential.

Therefore the SOI film thickness can entirely govern device scaling and SCEs can be

improved by scaling the film thickness together with the channel length [81]–[83].

The inverse subthreshold slope of long channel FDSOI MOSFETs can be near the ideal

60 mV/decade at 300 K [84]. The steeper subthreshold slope permits a lower threshold

voltage for the same off-current, which in turn allows the device to be used at a lower

supply voltage, a great advantage for low power operation.

Page 43: Passive Wireless Radiation Dosimeters for Radiation

26

For SOI MOSFETs, the source and drain regions extend down to the buried oxide,

yielding reduction of junction capacitance. This offers the opportunity to fabricate CMOS

circuits with lower power dissipation and improved speed. Since each single-transistor

islands is dielectrically isolated from each other, this isolation protects the thin active

silicon layer from a lot of parasitic effects induced from bulk silicon substrate, such as

leakage current and latch up. One the other hand, the lateral isolation makes interdevice

separation much easier. The overall technology and circuit design are highly simplified

[85], [86].

2.3 Total ionizing dose effects on FDSOI devices

The total ionizing dose (TID) effects on FDSOI MOSFETs have been studied

extensively in past decades. The main mechanism attributes the TID effects is the positive

Figure 2-2. Id vs. Vg characteristic of a typical of FDSOI N-MOSFET under different

levels of radiation dose.

Page 44: Passive Wireless Radiation Dosimeters for Radiation

27

charge buildup in the buried oxide layer causing the band diagram in the silicon body to

change [87]. Figure 2-2 shows the Id vs. Vg characteristic of a typical of FDSOI N-

MOSFET under different levels of radiation dose [88].

At low radiation dose, the front gate threshold voltage of the device shifted leftward

due to the trapped holes in the buried oxide and the coupling effect between front and back

gate [89]. The subthreshold slope degraded due to the build-up of radiation induced

interface traps [90]. At moderate radiation and negative gate bias, the leakage current

increased. This is because when the gate-to-source voltage becomes increasingly negative,

a high electric field is created at the surface of gate-to-drain overlap region. This results in

a band-to-band tunneling process that increases the gate induced grain leakage (GIDL)

current [91], [92]. At large dose, the leakage is high and the device cannot be turned off by

front gate voltage. This leakage current is a combination of two effects. One is the inversion

of the back channel interface due to the trapped holes. The other is the “total dose latch”

effect. This is caused by the charge trapping in buried oxide modulating the body potential.

As the body to source barrier height is lowered, excess electrons can be injected into the

body region and be collected at the drain. If the electric field near the drain is high enough

to cause impact ionization, this can lead to significant increase of leakage current [93], [94].

2.4 FDSOI radiation dosimeter

Even through the initial motivation of SOI technology was for radiation hardness, the

buried oxide layer in SOI structure could be utilized as a good medium for absorbing

radiation, just like RADFETs. Moreover, the special structure of FDSOI devices provides

unique advantages to overcome some of the limitations of existing dosimeter technologies.

Page 45: Passive Wireless Radiation Dosimeters for Radiation

28

The basic mechanism of FDSOI radiation dosimeter is shown in Figure 2-3. The

FDSOI device could be utilized as a radiation sensitive variable capacitor (varactor). The

device must be designed to initially be in the “OFF” state. That is, before radiation, no

inversion layer is present, and the capacitance measured between the top gate and the

source/drain electrodes is dominated by the fringing and overlap capacitances (Figure 2-3

(a)). As shown in Figure 2-3(b), when radiation is applied and holes are trapped in the

BOX layer, the trapped positive charge induces electrons in the SOI and turn the device

from “OFF” to “ON” state, allowing the gate to couple the entire channel. Therefore, at a

fixed gate bias, the measured gate to source/drain capacitance will change as a function of

Figure 2-3. (a) Before irradiation, the capacitance measured between the top gate and the

source/drain electrodes is dominated by the fringing and overlap capacitances. (b) When

radiation is applied and holes are trapped in the BOX layer, the trapped positive charge induces

electrons in the SOI and turn device from “OFF” to “ON” state. (c) At a fixed gate bias, the

measured gate to source/drain capacitance will change as a function of radiation. (d) and (f) If

this varactor is then integrated with an inductor, the frequency will change with dose and can

be measured wirelessly.

Page 46: Passive Wireless Radiation Dosimeters for Radiation

29

radiation (Figure 2-3(c)). If this varactor is then integrated with an inductor (Figure 2-3(d)),

the frequency will change with dose (Figure 2-3(e)) and can be measured wirelessly [95].

Compared to OSLDs, FDSOI wireless radiation dosimeter is more suitable for in situ

measurement. Compared to RADFETs, FDSOI devices collect radiation-induced charge in

the thick buried oxide layer, but read out the dose using the top-gated capacitor. By

separating the charge collection and read out components of the sensor, real-time

measurement can be easily realized without switching between “write” and “read” mode.

Since it operates at higher frequency, it has potential to overcome the 1/f noise limit, which

limits the sensitivity of RADFET dosimeters [96]. Moreover, few-nanometer gate oxide

thickness in the FDSOI varactors provides high capacitance per unit area as well as

capacitance tuning range [97], [98]. Therefore, the devices are much more scalable than

the MOS capacitors and MEMS based dosimeters. The separation of charge collection and

read out also make it possible to achieve high sensitivity and scalability simultaneously,

which does not exist in other technologies.

2.5 Existing literatures about SOI radiation dosimeter

A pioneering study using SOI devices as radiation dosimeters has been performed by

M. R. Shaneyfelt et al. [99]. In their work, the feasibility of developing an embeddable SOI

buried oxide MOS dosimeter has been demonstrated and fully-functional read-out circuitry

has been fabricated. Discrete SOI dosimeters have been irradiated under various dose rates

and bias conditions. The output sensitivity is ~ 125 µV/cGy. Data show only a small dose

rate dependence. However, instead of separating the charge collection and readout using

Page 47: Passive Wireless Radiation Dosimeters for Radiation

30

the BOX and top gate oxide respectively, the BOX layer was used for both purposes, which

is similar to conventional RADFETs.

In [100] the response of FDSOI MOSFETs to 8 keV X rays as well as to proton and to

-particle radiation was reported. In that work, it was shown that the threshold voltage of

an FDSOI MOSFET could be shifted by ~ 0.3 mV/Gy using 1 MeV protons and ~ 1 mV/Gy

using 8 keV X rays. The devices had a good charge retention time up to 90 days. The results

shown are promising to utilize a FDSOI device as a radiation dosimeter but the response

in clinically-related radiation conditions was not demonstrated and only a wired

configuration was used to characterize the radiation response.

The possibility of realizing FDSOI varactors with high quality factor (Q-factor) has

been studied using TCAD simulation [101]. A multi-finger design was used and results

show Q > 100 at f = 1 GHz can be achieved for Lg = 300 nm and tSOI = 20 nm. However,

the varactor character and radiation was not correlated and the frequency output under

radiation was not estimated.

Page 48: Passive Wireless Radiation Dosimeters for Radiation

31

Chapter 3: Radiation response of FDSOI MOSFETs

3.1 Device structure and characterization

FDSOI MOSFETs with nominally the same dimensions and layer thickness as those

reported in [100] were first studied. Simple MOSFETs were utilized for this experimental

study first in order to study the basic response of FDSOI structures to medically-relevant

radiation conditions.

The MOSFETs were fabricated at the IBM T. J. Watson research center using standard

Si process.

Figure 3-1 shows an optical image of an IBM FDSOI n-MOSFET. The device has a

40-nm-thick intrinsic SOI body, a 145-nm-thick BOX layer, a 3-nm-thick SiO2 gate oxide,

Figure 3-1 Optical image of IBM FDSOI MOSFETs

Page 49: Passive Wireless Radiation Dosimeters for Radiation

32

and an n-type polysilicon gate electrode. The devices were fabricated in a ring-shape

geometry and had channel lengths from100 to 200 μm.

Similar to bulk MOSFETs, FDSOI MOSFETs can operate as a 4-terminal device.

Figure 3-2 shows the output characteristic of a FDSOI MOSFET at different gate to source

voltage, Vgs from -0.25 V to 1 V and substrate voltage Vsub = 0 V. Drain to source voltage,

Vds was swept from 0 V to 1 V. The device has a threshold voltage Vth about 0.25 V and

displays typical long-channel device performance.

Figure 3-2 Output characteristic of a FDSOI MOSFET at Vgs from -0.25 V to 1 V and Vsub

= 0 V.

Page 50: Passive Wireless Radiation Dosimeters for Radiation

33

Figure 3-3 shows a typical transfer characteristic of the device at Vds = 0.5 V and

different substrate voltage, Vsub from -4 V to 2 V. Vsub will determine the substrate potential

of the device, and because of the coupling between front gate and substrate, the Id vs. Vgs

characteristic was shifted at different Vsub. As a result, the threshold voltage of the device

was changed because of the changing of the substrate potential. This also mimics the

radiation response. Since the accumulated holes in the BOX layer will also change the

substrate potential of the device, the threshold voltage will change due to radiation.

3.2 Experimental setup

The radiation response of the device was tested at the University of Minnesota,

Fariview Clinic. The testing was conducted using a setup as shown in Figure 3-4. The

device was exposed to an X-ray beam, with 10 × 10 cm2 cross-sectional area, generated

Figure 3-3 Typical transfer characteristic of the device at Vds = 0.5 V and Vsub from -4 to 2

V.

Page 51: Passive Wireless Radiation Dosimeters for Radiation

34

using a Varian 2300 CD linear accelerator. Polystyrene, a tissue equivalent material, was

placed above the device to a total thickness of 1.5 cm. High-energy beams have a dose

build-up region due to the forward momentum of the scattered electrons. This build-up

material is necessary so that the measurement device is placed at the maximum dose region

and not in a high-dose-gradient region [102]. The source-to-surface distance measured to

the top of these slabs was 100 cm. While the X-ray beam was on, the linac provided a

constant radiation dose rate of 4 Gy/min, which is within the typical range (3-6 Gy/min)

for cancer therapy. The device characterization was performed using an HP4145B

semiconductor parameter analyzer located in a control room, roughly 12 m from the

treatment room. The drain current, Id, was measured vs. gate-to-source voltage, Vgs, at a

fixed drain-to-source voltage, Vds of + 0.5 V. The measurement sequence was as follows.

Before irradiation, the Id vs. Vgs characteristic was measured repeatedly to quantify any

drift in the device characteristic. Next, a positive substrate voltage, Vsub, between +1 V and

Figure 3-4 Experimental setup for radiation dosimetry experiments.

Device

Control room Treatment room

DEVICE

POLYSTYRENE

POSITION-ADJUSTABLETREATMENT TABLE

TREATMENT ROOMCONTROL ROOM

HP4145BSEMICONDUCTOR

PARAMETERANALYZER

Linac Head

Adjustable table

Polystryrenephantom

Parameter analyzer

Electrical connection box

Si

GS

SiO2

Si

D

n+n+

Page 52: Passive Wireless Radiation Dosimeters for Radiation

35

+ 5 V was applied and this voltage remained on continuously throughout the entire

measurement sequence. The substrate voltage helps to minimize recombination of the

generated electron-hole pairs, as well as to drive the holes to the top surface of the BOX.

The Id vs. Vgs characteristic was measured with the substrate bias applied. Next, the device

was irradiated for a fixed dose, and then Id vs. Vgs measured again immediately afterward.

A fixed time was then allowed to elapse before the device was measured again, but before

the next dose of radiation. This sequence allowed us to plot the device response as a

function of time as well as dose.

Figure 3-5. Typical Id vs. Vgs curve for an FDSOI MOSFET after different dose fractions

of radiation from a linac with 6-MV accelerating voltage. The applied substrate bias,

Vsub is + 5 V.

Page 53: Passive Wireless Radiation Dosimeters for Radiation

36

3.3 Experimental results

Figure 3-5 shows a typical response of one device to a series of exposures, consisting

of 4 Gy per fraction, up to a total dose of 40 Gy. For each fraction, the irradiation time was

1 min and a 3-min delay was utilized between fractions. Therefore, for each 4-Gy fraction,

the Id vs. Vgs sweeps were separated by 4 min, and were taken immediately after the beam

was turned off. The results show a clear, consistent leftward shift in the Id vs. Vgs

characteristic, consistent with the trapping of positive charge in the BOX layer.

Figure 3-6 shows the threshold voltage shift, ∆Vth, plotted vs. time for radiation dose

up to 40 Gy using the same experimental setup as for the data depicted in Figure 3-5. Here

the threshold voltage, Vth, is defined as the voltage at Id = 2 μA. The plot in Figure 3-6

Figure 3-6. Threshold voltage shift under step irradiation with substrate voltage Vsub =

0, +1 V, +3 V and +5 V. A clear response to each irradiation is observed that closely

tracks the accumulated dose.

Page 54: Passive Wireless Radiation Dosimeters for Radiation

37

shows the evolution of the ∆Vth vs. time plot for Vsub = 0, +1 V, +3 V and +5 V, while the

shaded area shows the accumulated dose vs. time. The threshold voltage shows a clear step

response to each interval of radiation, and tracks the applied dose extremely closely. By

applying a substrate bias, the sensitivity is improved by a factor of 20 for a Vsub = +5 V

compared to Vsub = 0 V. The increased sensitivity with increasing Vsub is primary due to the

increased fractional yield of charge collected in the BOX layer [18].

For therapeutic radiation treatments, it is important that the device response can be

predicted under a wide range of conditions, and therefore, the response was measured over

a wide dynamic range, as well as for different energies and incident angles. Figure 3-7

shows the sensitivity of an FDSOI MOSFET for radiation doses up to 8 Gy, 40 Gy and 160

Gy. (In the 8-Gy experiment, the dose delivery rate was reduced to 1 Gy/min.) As shown

in the figure, linear response was observed for the lower dose ranges, while for very heavy

accumulated doses, the response became sublinear. In addition, we also observed slightly

different sensitivities for the three dose ranges shown in Figure 3-7, and this effect is related

Figure 3-7. Measured threshold shift for FDSOI MOSFETs at Vsub = +5 V for experiments with

maximum dose of (a) 8 Gy, (b) 40 Gy and (c) 160 Gy.

Page 55: Passive Wireless Radiation Dosimeters for Radiation

38

to the radiation history of the device as will be discussed later in this thesis. Though the

saturation and history effects are not entirely understood at this time, a possible explanation

could be the accumulation of interface traps which degrade the subthreshold slope. Such

degradation was observed in these samples, but a quantitative extraction of the interface vs.

bulk charge trapping could not be performed on our current samples due to the relatively

high off current in our large ring-shaped MOSFETs. Additional experiments on MOSFETs

with more regular geometries will help in separating out bulk vs. interface charge trapping

effects.

The dependence of radiation response with different X-ray beam energies and incident

angles was also determined experimentally. Figure 3-8 shows ∆Vth plotted vs. time for

irradiation using different X-ray beam energies. These experiments were performed in the

Figure 3-8. Threshold voltage shift under step irradiation with different energy X rays.

Blue triangles: energy vs. time increasing. Red squares: energy vs. time, decreasing.

Page 56: Passive Wireless Radiation Dosimeters for Radiation

39

following way. In experiment 1 (blue triangles), the device was first irradiated using X rays

generated from a linac beam created using a 6 MV accelerating voltage in 4-Gy fractions

up to accumulated dose of 40 Gy. Then, on the same device, the dose sequence was applied

using X rays generated using a 10 MV and an 18 MV accelerating voltage. An additional

delay of a few minutes occurred between energies to monitor the stability of device. This

test was followed by experiment 2 (red squares in Figure 3-8), once again using the same

device, where the sequence of applied radiation beam energies was reversed. The “zero

point” threshold voltage was reset between experiments 1 and 2. The plot shows that the

threshold voltage tracks the dose closely with only a small change in the slope at higher

doses, consistent with the trend observed in Figure 3-7(c). The fact that this sublinear

behavior occurs at high doses, independent of the accelerating energy, provides strong

evidence that the device response is independent of X-ray beam energy.

Figure 3-9. Threshold voltage shift under step irradiation versus different angles of

beam incidence.

Page 57: Passive Wireless Radiation Dosimeters for Radiation

40

The plot in Figure 3-9 shows the dependence of the device threshold shift under

different incident beam angles. The figure shows ∆Vth plotted vs. time for 6 MV X rays for

incident beam angles of 75˚, 60˚, 45˚, 30˚, and 0˚ relative to normal incidence. Similar to

the energy-dependence experiment, no systematic angular dependence was observed.

These results are consistent with [59].

In both of the experiments described in Figure 3-8 and Figure 3-9, a small reduction of

the slope was observed with increasing accumulated radiation dose, similar to the trend

observed in Figure 3-7(c). Figure 3-10 shows the sensitivity (defined as the change in

threshold voltage per unit dose) vs. the absolute threshold voltage for all data sets at Vsub =

+5 V. Since the same device was “recycled” for multiple experiments, for any given dataset,

Figure 3-10. Plot of sensitivity vs. absolute threshold voltage for all data for Vsub = +5

V.

Page 58: Passive Wireless Radiation Dosimeters for Radiation

41

the device already had some degree of trapped charge. This plot indicates that when

adjusted for the actual threshold voltage, the sensitivity characteristics collapse onto a

single curve determined by the prior dose history. This result explains the slight differences

in sensitivity observed in previous data and suggest that a high degree of accuracy in the

dosimeter readout can be obtained when this nonlinearity is taken into account. However,

additional statistical analysis is needed to determine the limits to the stability and accuracy

of this sensing mechanism. While the origin of this trend is still under investigation, this

behavior could be related to accumulation of interface traps as well as “saturation” of the

hole traps due to the high quality of the buried oxide in the commercial SOI wafers.

3.4 Charge trapping efficiency

As mentioned in Chapter 1, after electron-hole pairs are generated because of

irradiation, some of them will recombine immediately and eventually only a small portion

of holes will be trapped at the SiO2 and Si interface. The percentage of holes trapped at the

oxide compared to the total generated electron-hole pairs, or, the charge trapping efficiency

is an important parameter for dosimetry application. The higher the charge trapping

efficiency, the more holes are trapped in the oxide, therefore the higher sensitivity the

device will have.

The charge trapping efficiency, f, is determined by two factors. First, the percentage of

holes survived from the initial recombination process, which is called the fractional yield

of holes. The fractional yield is a factor of radiation source, electrical field. Second, the

percentage of hole eventually trapped in the oxide compared to the total holes reach to the

Page 59: Passive Wireless Radiation Dosimeters for Radiation

42

SiO2 and Si interface. It can be influenced by factors such as oxide quality, temperature

[103].

In order to extract f, based on the radiation response results in Figure 3-6, it is first

important to consider the fact that the conducting channel in the SOI can form either at the

top or bottom Si/SiO2 interface. In either case the threshold voltage shift can be expressed

as:

topsith CQV / , 3-1

where Ctop is the capacitance of the top gate oxide to the silicon channel and ∆Qsi is the

charge per unit area in the SOI layer that results from the radiation-induced trapped charge

in the BOX, QBOX. If the inversion channel is located at the bottom interface, then the

threshold shift is determined as:

oxoxSiSiBOXth ttQV bot , 3-2

where ox and Si are the dielectric constant of SiO2 and Si. tox and tSi are the thickness of

top gate oxide and thickness of SOI layer. This equation comes about in the back-channel

conduction, since in this case, QSi is essentially equal to QBOX, but the top gate capacitance

must also consider the thickness of the SOI layer. In the opposite extreme, if the inversion

layer is at the top interface, then the top gate capacitance only needs to consider tox, but a

portion of QBOX will be mirrored at the back (substrate) interface of the BOX layer, and

therefore the expression for Vth becomes

ox

ox

BOXBOXSiSi

SiSiBOXth

t

tt

tQV

top , 3-3

Page 60: Passive Wireless Radiation Dosimeters for Radiation

43

where tBOX is the BOX thickness as shown in Figure 2-3 (a). Since in reality, the channel

could be distributed between top and bottom interfaces, the actual threshold voltage shift

is somewhere between the values in 3-2 and 3-3:

bottop ththth VVV . 3-4

From 3-2 and 3-3, it is clear that if the channel location is known, then the measured

∆Vth value can be used to determine QBOX. Then, to determine the charge trapping

efficiency the following relation can be used:

DftgqQ BOXBOX 0 , 3-5

where D is the dose, g0 = 8.1×1014 cm-3/Gy is the density of electron-hole pairs created per

unit dose in SiO2, and q is the electronic charge.

Figure 3-11. Points: simulated Id – Vgs curve with different amount of positive charge

at the Si/BOX interface. Line: experimental data.

Page 61: Passive Wireless Radiation Dosimeters for Radiation

44

In our current devices, the relations in 3-2 and 3-3 give a broad range of possible

fractional yields due to the fact that the Si layer thickness (tSi = 40 nm) is large compared

to the gate oxide (tox = 3 nm). Therefore, we have fit the experimental data to TCAD

simulations using Synopsys Sentaurus DeviceTM to obtain an accurate estimate of the

channel location in our devices. The device in simulation has the same dimensions as the

experimental device except a 2 µm channel length and 1 µm width. Figure 3-11 shows the

simulated Id-Vgs characteristic normalized to experimental device dimensions for different

values of positive charge at the SOI / BOX interface along with an experimental curve from

a device before irradiation. As expected, the Id – Vgs shifts toward more negative values as

more oxide charge is added.

The Vth vs. QBOX characteristics obtained from TCAD were then compared to the

analytical models in 3-2 and 3-3 oxoxSiSiBOXth ttQV bot , 3-2 and the fractional

yield calculated in all three situations. These results are summarized in . The table clearly

shows very close agreement between the TCAD results and the analytical back-channel

model. The possibility of back-channel formation, even at relatively low doses, was

neglected in [100] but is consistent with [104] and [92], showing results with very large

doses. The small discrepancy between the TCAD and analytical models is likely due to

uncertainty in the precise thicknesses of the layers used in the experimental devices. This

result is important for the design of optimized dosimeters, particularly the wireless sensors

to be described later, since the location of the channel formation can affect the device

tuning range and sensitivity. Additional details of the optimization for wireless sensing are

provided in the following sections.

Page 62: Passive Wireless Radiation Dosimeters for Radiation

45

Table 3-1 Extracted charge trapping efficiency for different substrate bias voltages and

simulation conditions

Page 63: Passive Wireless Radiation Dosimeters for Radiation

46

Chapter 4: Radiation response of FDSOI varactors

4.1 Device fabrication

4.1.1 Overview

The experiments based on IBM MOSFETs demonstrated the radiation response of such

devices under clinically relevant irradiation conditions. The final objective of this project

is to realize a wireless radiation dosimeter, which requires the use of variable capacitors

instead of MOSFETs. However, the IBM devices are not suitable for further studies for the

following reasons. First, since the devices require wire bonding to be tested in a clinical

environment, and due to their ring-shaped design, the wire for the gate electrode must be

bonded directly onto gate, which may induce extra leakage for the device. Second, the wire

directly on top of the gate region causes extra scattering of radiation which can generate

secondary electrons which could cause errors in dose measurement. Finally, the Q-factor

of the IBM devices is small due to the long gate length, which results in high resistance.

Due to these reasons, it is necessary to design and fabricate our own FDSOI varactors

that are suitable for radiation detection. Therefore, a custom fabrication process for FDSOI

varactors was developed at the University of Minnesota and a brief description of the

process flow is as follows:

As shown in Figure 4-1, the process starts with commercial 4-inch SOI wafers

purchased from SOITEC with about 100-nm SOI layer and 200-nm BOX layer. The SOI

and silicon substrate were lightly P-type doped. It is followed by thermal oxidation to

partially oxidize the SOI layer and then etch the top SiO2 away so the remaining SOI is

Page 64: Passive Wireless Radiation Dosimeters for Radiation

47

about 30 nm thick. Next a layer of 7-nm thermal screening oxide is grown on top of SOI.

It is followed by patterning mesa and etching away excessive silicon. Then source/drain

(S/D) region is then doped by ion implantation and activation. After that, the screening

oxide is etched away and a 7-nm thermal gate oxide is grown. Finally gate and S/D are

patterned and metal layers are deposited. A more detailed process flow can be found in

appendix A.

Compared with a standard CMOS process in industry, our custom process is greatly

simplified. For example, it utilizes a non-self-aligned process with significant overlap

between gate and S/D electrodes. This creates extra parasitic capacitance in the device.

Figure 4-1. (a) Structure of commercial SOI wafer. (b) Thin the SOI layer to ~ 30 nm.

(c) Grow screen oxide and pattern mesa. (d) Ion implantation, activation and growth of

gate oxid. (e) Pattern and deposit gate electrode. (f) Pattern and deposit S/D electrode.

Page 65: Passive Wireless Radiation Dosimeters for Radiation

48

Also, there is no raised S/D, and as a result, the device has relatively high resistance due to

the very thin silicon layer at S/D region. The lack of a raised source/drain also limits how

thin the SOI layer can be made, since sufficient thickness must be allowed for the

source/drain implants. Nevertheless, the varactor is good enough to demonstrate radiation

sensing and those issues will be address later for wireless sensing applications.

In the next a few subsections, more details about a few critical steps in fabrication will

be discussed. They are thinning of the SOI layer, ion implantation, and gate oxide growth.

4.1.2 Thinning of SOI layer

The as-purchased commercial SOI wafers usually have a thick SOI layer. In order to

ensure the SOI is fully-depleted, the thickness of the SOI layer must be made smaller than

its maximum depletion width. As will be discussed in the following chapter, the thinner

the SOI layer at the channel region, the higher sensitivity the dosimeter will have. For these

reasons, the final SOI layer needs to be less than 30 nm.

One approach to thinning the SOI layer is to directly etch the silicon. One approach is

to use wet etching, such as KOH solution [105], while another way is to use dry plasma

etch, such as XeF2 plasma [106]. However, such techniques suffer from poor controllability.

Therefore, the preferred way to etch the SOI layer with high precision is to use a two-step

process, i.e. oxidation and etching. This process starts with thermally oxidizing a part of

the SOI layer to form a layer of SiO2 on top, then using buffered oxide etch (BOE) to etch

away the top SiO2 layer. Because the thermal oxidation is a very well understood process,

the thickness of SiO2 can be controlled precisely, on the scale of nanometers. Also, since

the BOE solution has high selectivity to etch SiO2 from Si, the remaining SOI layer can be

Page 66: Passive Wireless Radiation Dosimeters for Radiation

49

well preserved. Moreover, this two-step etching process can be repeated a few times until

the desired SOI thickness is achieved.

The equipment used to characterize the thickness of silicon or SiO2 layers is called an

ellipsometer. An ellipsometer utilizes the interaction between an incident laser and the

material of interest. After reflection, the changing of polarization of the incident laser will

depend on the thickness of material and its dielectric constant [107]. Since the dielectric

constant of silicon and SiO2 are well known, the thickness can be found by fitting the signal

with an existing model. In a stacked structure where the thicknesses of more than one layer

needs to be measured, the laser sources with multiple wavelength can be used. Figure 4-2

shows an example of measuring a commercial SOI wafer (before thinning) using a

spectroscopic ellipsometer and comparison with the model. In this sample, the native oxide

Figure 4-2 Thicknesses of SOI wafers measured from spectroscopic ellipsometry. By

fitting the signal with existing model, the thickness of each layer was determined.

Page 67: Passive Wireless Radiation Dosimeters for Radiation

50

layer was 2.08 nm ± 0.21nm, the SOI layer was 101.02 nm ± 0.07 nm and the BOX layer

was 197.21 nm ± 0.27 nm, consistent with the thicknesses provided from the vendor.

4.1.3 Ion implantation of source/drain

In order to make the silicon in the S/D regions conductive and create a potential barrier

between channel and S/D, it is necessary to heavily dope the S/D region. This process is

done by ion implantation. Because the silicon layer is sandwiched between the screen oxide

and BOX, it is important to optimize the implantation parameters to make sure sufficient

dose is incorporated without fully amorphizing the SOI layer. The important parameters

include the energy of incident ions and number of ions per unit area. The energy of the

incident ions will determine the stopping range of implantation. If the energy is too low,

the ions will not be able to penetrate the screen oxide and reach the silicon layer. On the

other hand, if the energy is too high, most of the ions will stop at the BOX layer instead of

silicon. This will make implantation less efficient and more importantly, it could

completely amorphize the silicon layer and make it unable to recrystallize at the activation

step [108]. The number of ions per unit area, will determine the doping density. Ideally the

doping should be as high as possible without completely amorphizing the silicon layer.

In order to obtain the best implantation condition, simulation was conducted using The

Stopping and Range of Ions in Matter (SRIM) before implantation. Figure 4-3(a) shows

the simulation result of 9 keV As ions implanted into a SiO2-Si-SiO2 structure with a 7-

degree incident angle. This 7-degree incident angle is needed to prevent the channeling

effect of ion implantation in silicon [109]. The top SiO2 layer was 7 nm and middle silicon

layer 34 nm. Most of the ions stopped in the silicon layer. Figure 4-3(b) shows the

Page 68: Passive Wireless Radiation Dosimeters for Radiation

51

calculated doping density as a function of depth for As with energy of 7 keV, 9 keV, 11

keV and 13 keV. The number of As was 2 × 1015/ cm2. Based on the simulation, 9 keV was

found to be the optimal energy because the peak of doping density resides near the surface

of the silicon layer while the bottom of the silicon layer still has sufficiently low dose to

prevent amorphization.

4.1.4 Activation

The implanted ions (As) will reside randomly within the silicon lattice after

implantation. In order to make them function properly as donors of electrons, they need to

take the place of silicon atoms within the crystal lattice [110]. This process is called

activation and it is done by high temperature annealing. After implantation, the SOI

samples were annealed at 850 °C for 2 mins to activate the dopants. Right after activation,

the screen oxide was etched away and the sheet resistance of SOI layer was measured using

Figure 4-3. (a) Illustration of 9 KeV As stopping in SiO2-Si-SiO2 stacked layers with 7 nm top

SiO2 and 34 nm silicon. (b) Calculated doping density of 7 KeV, 9 KeV, 11 KeV and 13 KeV

As implanted in the structure same as (a). Number of implants: 2 × 1015 /cm2.

Page 69: Passive Wireless Radiation Dosimeters for Radiation

52

a four-point probe to check the effectiveness of implantation and activation (this step was

done using a control sample without a patterning mesa). For implantation conditions

mentioned in the previous section, the typical sheet resistance of SOI layer was about 560

Ω/□. This equals to 3 × 1019 cm-3 doping density in 30-nm silicon, which is consistent with

STRIM simulation.

The doping density was further calculated using a transfer length measurement (TLM)

after the device was successfully fabricated. Figure 4-4 shows the optical image of the

TLM structure fabricated together with the FDSOI varactors. The silicon width was 20 μm

and the distance of each contacts varies from 5 µm to 100 μm. Figure 4-5(a) shows the IV

measurement of one set of TLM structure and Figure 4-5(b) shows the extracted resistance

as a function of distance between two electrodes. The sheet resistance of SOI layer can be

Figure 4-4. Optical image of a TLM structure where the distance between each two

electrodes varies from 5 µm to 100 μm

Page 70: Passive Wireless Radiation Dosimeters for Radiation

53

calculated from the slope of resistance vs. distance curve and a typical number is about

1500 Ω/□, which equals to a doping density of 1 × 1019 /cm3.

The slightly lower number of doping density calculated from TLM measurement

compared to four-point probe is expected. Because the TLM measurement was conducted

after gate oxide growth and etching away at S/D region, some dopants could be lost during

these steps.

4.1.5 Growth of gate oxide

After activation of the dopants, the screen oxide was etched away immediately and was

followed by the growth of the gate oxide. Ideally, the gate oxide would be as thin as

possible because the thinner the gate oxide, the higher density of capacitance per unit area,

and the sensitivity would be higher. However, the gate oxide cannot be too thin otherwise

there will be too much tunneling leakage. For this custom fabrication process, the thickness

Figure 4-5. (a) IV characteristic of a TLM structure. (b) Extracted resistance vs. electrode

distance for the same TLM structure.

Page 71: Passive Wireless Radiation Dosimeters for Radiation

54

of the gate oxide was limited by the facilities at University of Minnesota. The thinnest

oxide that can be grown is about 6 nm at an oxidation condition of 850 °C for 1 min. During

the growth step, the silicon at S/D region will also be oxidized. It is important to realize

that the oxidation rate of implanted silicon is higher than undoped silicon [111]. Figure 4-6

shows the profilometer measurement of the silicon layer after gate oxidation. Clearly, the

S/D region is higher than the channel region, indicating a faster oxidation rate in the S/D

region. As a result, this step will consume more silicon in the S/D region than the channel

region. It will also increase the series resistance and decrease the Q-factor for wireless

sensing. However this issue only matters for the the custom device fabrication process,

since the S/D could be raised by epitaxially growing more silicon which is a standard

technique in production processing [112].

Figure 4-6. Profilometer measurement of silicon layer after gate oxidation.

Page 72: Passive Wireless Radiation Dosimeters for Radiation

55

4.2 Device characterization

Figure 4-7 shows the optical image of an FDSOI varactor, which have 20 µm gate length

and 100 µm width. The device has 7 nm gate oxide and about 20 nm SOI layer with a

200 nm BOX.

Cg vs. Vg (Figure 4-8(a)) characteristics were measured to verify the operation of the

device under different substrate bias, Vsub from -30 V to + 50 V. Depending upon at which

SOI interface the inversion will occur first while sweeping Vg, the device can operate at

either front channel or back channel mode. As shown in Figure 4-8(a), when Vsub changes

from -30 V to +50 V, the Cg vs. Vg curve shifts leftward as expected. At large negative Vsub,

the band bending of the SOI layer keeps the back channel from formation, and only the

front channel can be turned on by increasing Vg, as shown in Figure 4-9(a). However, as

Figure 4-7 Optical image of a typical FDSOI varactor. Lg = 20 µm, W = 100 µm.

Page 73: Passive Wireless Radiation Dosimeters for Radiation

56

Vsub is increased, the band bending of the SOI layer will gradually switch from front-

channel to back-channel mode (Figure 4-9(b)). In this mode, as Vg increases, the back

channel turns on first, and then after further increasing Vg, eventually the front channel will

also be turned on. For instance, the “hump” in Figure 4-8(a) at Vsub = 0 V and +5 V show

this transition of the front channel from the “off” to “on” state. In Figure 4-8(a), the

capacitance variations at Vg = +2 V for all values of Vsub are due to slight differencees in

the probe pad parasitic capacitance induced by the substrate bias. Similar variations are

observed for Vsub = -30 V and +5 V at at Vg = -3.5 V.

These capacitance variations arise due to the substrate below the pads switching

between depletion and accumulation as Vsub is varied. At very large positive Vsub, the back

channel will be completely turned on at all values of Vg (Figure 4-9(c)), so that the

capacitance level will be high even at large negative Vg.

Figure 4-8. (a) Cg vs. Vg characteristics of FDSOI varactor at different Vsub from -30 V to +50

V. (b) Id vs. Vg characteristics of the same device at different Vsub from -30 V to +50 V.

Page 74: Passive Wireless Radiation Dosimeters for Radiation

57

This change of the location of the channel can be also seen in Id vs. Vg characteristics

at different Vsub as shown in Figure 4-8(b). At large negative Vsub, the back channel is kept

off and the front channel can be turned “off” and “on” by sweeping Vg. At large positive

Vsub, the back channel is always on therefore the current is always a large value regardless

of Vg. In both Figure 4-8(a) and Figure 4-8(b) the shift of characteristics when Vsub changed

from 0 V to +5 V is larger than the shift when Vsub changed from -30 V to -25 V. Again,

this is due to the different locations of the channel. At the front channel mode, the

equivalent oxide thickness from the top gate is just the gate oxide thickness, tox. At back

channel mode, the equivalent oxide thickness is the combination of gate oxide and silicon

channel, tox + 1/3 tsi. A thicker equivalent oxide thickness results smaller equivalent

capacitance and since Q = CV, it will require more voltage to induce same amount of charge

in the channel. As a result, the shift of characteristics in back channel mode is larger than

the shift in front channel mode.

Figure 4-9 (a) Band diagram simulation of FDSOI varactor at Vsub = -30 V, Vg = 0 V. (b) Band

diagram simulation of FDSOI varactor at Vsub = + 5 V, Vg = 0 V. Band diagram simulation of

FDSOI varactor at Vsub = +50 V V, Vg = 0 V.

Page 75: Passive Wireless Radiation Dosimeters for Radiation

58

In all radiation experiments, the proper Vg and Vsub value were selected such that the

devices operated in the back channel mode, as shown in Figure 4-10.

4.3 Experimental setup

The devices were then irradiated and tested simultaneously using a commercial

Gamma-Knife Co-60 clinical therapy unit (Figure 4-11), where a special mount was

fabricated to allow the varactors to reside at the center of the focused radiation beams. The

accuracy of the device positioning in the beam was confirmed using radiochromic film

(Gafchromic™ EBT2, Ashland, Inc.). The applied dose was verified using an optically

stimulated luminescent dosimeter (nonoDotTM, Landauer, Inc.). The device was wire

bonded to a header that was connected to the capacitance meter. When the device was

Figure 4-10 Operational region for radiation experiments.

Page 76: Passive Wireless Radiation Dosimeters for Radiation

59

irradiated, the device in the mounting assembly was moved into the focus of the 201 beams

by the movement of the patient couch. The capacitance was measured at 100 kHz with

amplitude of 50 mV for all radiation experiments.

During the experiment, Cg was measured with a constant positive Vsub. The substrate

voltage helps to minimize recombination of the generated electron-hole pairs, as well as to

drive the holes to the top surface of the BOX. The measurement sequence was as follows.

Before irradiation, the Cg vs. Vg characteristic at different Vsub values was measured. Based

on this characteristic, the proper Vg value was selected to make sure the device was operated

at the transition region between “off” and “on” states, where the capacitance change is most

sensitive to irradiation, as labeled in Figure 4-10. Next, Cg was monitored as a function of

time (without irradiation) at fixed Vg and Vsub to ensure that there was no drifting in the

device. Then the device was irradiated. After the accumulated dose was reached, the device

was moved out of the radiation beam for a fixed time before being irradiated again. For

Figure 4-11 Setup for capacitance based radiation experiment. A special mount was fabricated

to allow the varactors to reside at the center of the focused radiation beams. The device was

connected with a capacitance meter and capacitance was measured in real time during

irradiation.

Page 77: Passive Wireless Radiation Dosimeters for Radiation

60

some measurements, one continuous dose sequence was utilized, and for others, a series of

step stress doses (i.e. several shorter dose sequences separated by short delay times where

no radiation was incident) was applied to allow observation of room temperature annealing

on the device response.

Page 78: Passive Wireless Radiation Dosimeters for Radiation

61

4.4 Experimental results

Figure 4-12(a) shows a typical response of one device to a series of exposures,

consisting of 8 Gy per fraction, up to a total dose of 64 Gy, where a 5 minute delay was

allowed to elapse between each fraction. The capacitance was measured in real time with

Figure 4-12 Capacitance and dose vs. time for FDSOI varactors with stepped dose

profile of 8 Gy per fraction up to 64 Gy. (b) Cg vs. Vg before and after irradiation. The

points A and B on the capacitance vs. voltage curve correspond to the same points on

the capacitance vs. dose graph in (a). Device dimension: Lg = 7 m, Wg = 40 m.

Page 79: Passive Wireless Radiation Dosimeters for Radiation

62

a constant substrate bias of Vsub = +5 V. The shaded area shows the accumulated dose vs.

time. The capacitance shows a clear step response to each interval which tracks the applied

dose extremely closely. The “spikes” in the capacitance between each fraction are due to

mechanical and electromagnetic interference produced when the patient couch is moved to

bring the device in and out of the radiation beams. Nearly all of these occur during the

transition period. After the device arrives at the radiation beam center, the capacitance

response was found to be very stable and linear under in-situ radiation conditions. I believe

this noise is not a fundamental limitation of the sensor technology but rather reflects non-

idealities in our measurement setup. Figure 4-12(b) shows the C-V characteristics of the

device before and after irradiation. It is clear that the capacitance curve shifts leftward upon

irradiation, as expected for positive charge trapping in the BOX. Since the device was

Figure 4-13. Change in capacitance vs. dose for two different delay times. Device

dimension: Lg = 7 m, Wg = 40 m.

Page 80: Passive Wireless Radiation Dosimeters for Radiation

63

biased at a constant gate voltage of -1.12 V during irradiation, the capacitance increases

from point A to point B on the graph. The parallel shift of the C-V curve also shows that

most of the holes were trapped in SiO2 instead of creating traps at the Si/SiO2 interface.

In radiation dosimetry, particularly in medical applications, it is desirable that the

device response is independent of the time delay between dose fractions. In our

experiments, we tested this behavior by comparing the device response to continuous

radiation, and radiation delivered in a step-wise fashion, with short (~5 minute) intervening

times where no dose was applied. In our setup, due to the natural decay of Co-60, the dose

rate was fixed at a value of 2.96 Gy/min. Figure 4-13 shows the capacitance vs. dose for

FDSOI varactors with 0 and 5 minute delay times between each fraction. The figure shows

excellent overlay of sensor response indicating that, for the short delay times utilized in

Figure 4-14. Measurement error as a function of dose for intra-device and inter-device

experiments.

Page 81: Passive Wireless Radiation Dosimeters for Radiation

64

this experiment, no room temperature annealing effects were observed. Additional

experiments are needed, however, to investigate annealing effects over longer delay times.

Figure 4-14 shows the measurement error vs. dose at Vsub = +5 V, where the data was

generated from two sets of experiments. In the first intra-device experiment, a linear dose

calibration curve was established based upon an initial capacitance vs. dose measurement.

Then, on the same device, the gate voltage was changed to reset the capacitance to the

initial value before calibration. Then, the device was re-irradiated and the percent error

(defined as 100% × |measured dose – actual dose| / actual dose) in the measured dose

relative to the actual dose was determined. This experiment was performed for two separate

devices and the results are shown by the black squares in Figure 4-14.

In the inter-device experiment, a dose calibration curve was established on one device

and then a second (nominally identical) device was used as the dosimeter. In this

experiment, the gate bias was adjusted to match the initial capacitances of both devices

before irradiation. The percent error from these measurements was shown by red dots and

circles in Figure 4-14 and two pairs of devices were utilized. For all experiments,

measurements were performed under continuous radiation conditions up to 64 Gy. The

measurement error was found to be < 7% after 10 Gy and reduced to 1 % for accumulated

dose of 60 Gy. In the single device measurements, the residual error is likely due to

reduction of trapping efficiency with increasing dose. The slight change in gate voltage (<

0.1 V) between calibration and measurement is not expected to have a significant effect on

the measurement error [113]. In the device-pair experiments, geometric variations as well

as measurement noise could account for the observed sensitivity errors.

Page 82: Passive Wireless Radiation Dosimeters for Radiation

65

Figure 4-15(a) shows the capacitance change vs. radiation dose up to 64 Gy with

substrate voltage Vsub = +1, +5 and +10 V. The increased sensitivity with increasing Vsub is

due to the higher charge fractional yield of holes in the BOX layer with increasing electric

Figure 4-15. (a) Capacitance change vs. dose under different substrate bias voltages.

(b). Capacitance vs. dose for Vsub = +10 V with low radiation dose. Device dimension:

Lg = 30 m, Wg = 40 m.

Page 83: Passive Wireless Radiation Dosimeters for Radiation

66

field [18]. Assuming all electrons were swept out of the oxide, the percentages of holes

(relative to the number initially generated by the radiation) trapped in the BOX layer are

1.9%, 7.4%, 9.7% for Vsub = +1, +5, +10 V, respectively [114]. The low value could be due

to the high quality of BOX layer in the commercial SOI wafer, where only very limited

oxygen vacancies exist near the Si/SiO2 interface. However, this also indicates a large

space to improve the sensitivity by engineering the vacancy concentration in the BOX to

increase hole trapping efficiency [115]. Another possibility is that, in addition to trapping

holes, the BOX layer also traps a significant amount of electrons which compensates for

the positive trapped charge [116] [117]. More comprehensive experiments are needed to

distinguish between these two mechanisms. Figure 4-15 (b) shows a detailed portion of

device response at Vsub = +10 V for low dose levels from 0 to 8 Gy. The results show that

Figure 4-16. Comparison of capacitance vs. dose for one device when it was first

irradiated and when it was irradiated for 230 Gy. Device dimension: Lg = 30 m, Wg =

40 m.

Page 84: Passive Wireless Radiation Dosimeters for Radiation

67

the current sensors have minimum detectable dose of ~ 0.1 Gy. However, this value is

limited by non-idealities in current measurement setup and we believe that significant

improvements in sensitivity are possible, as will be described in the next section.

The effect of dose history on the sensor response was also characterized. Figure 4-16

shows the capacitance vs. dose for a device irradiated under two conditions. The black

squares in Figure 4-16 show the capacitance vs. dose response of a device with no prior

radiation history. This device displays a sensitivity of 0.24 %/Gy, where we define

sensitivity as the percent capacitance change per unit dose. The red circles show the

response of the same device after it had accumulated a dose of 230 Gy. For this

measurement, the gate voltage was adjusted so that the starting capacitance was the same

for both measurements. In this case, the sensitivity was reduced to 0.20 %/Gy. This change

in sensitivity is likely due to saturation of hole traps in the BOX [118], which is consistent

with our previously published results [114].

4.5 Analytical model of FDSOI varactors

We have developed an analytical formalism to quantify the sensitivity limits of FDSOI

radiation dosimeters. Figure 4-17 shows the equivalent circuit model for FDSOI varactors,

along with a schematic diagram of the device structure showing the critical circuit elements.

In this model, Cov is the overlap capacitance between the gate and source/drain

electrodes, Rinv is the channel resistance, Cinv is the capacitance between the top gate and

the channel region between the source and drain and R0 is an additional extrinsic resistance.

If the total impedance of the circuit in Fig. 4-17 is Z, then for a given angular frequency,

,

Page 85: Passive Wireless Radiation Dosimeters for Radiation

68

1)/1

1(0

ov

invinv

CjCjR

RZ

. 4-1

Since in a series model, CjRZ /1 , therefore ZC Im/1 , then we have:

1

22 1

invovinv

ovinv

invovCCC

CCRCC . 4-2

The value of Rinv is related to the charge density in the channel, Qinv, via

invg

eff

invQW

LR

1

4 , 4-3

where Leff is the effective gate length, and is the electron mobility. The factor 4 in

equation 4-3 comes from the length between gate and source/drain equals to 1/2 Leff, and

Figure 4-17. (a) Equivalent circuit model for FDSOI varactor. (b) Diagram showing

physical location of circuit elements.

Page 86: Passive Wireless Radiation Dosimeters for Radiation

69

when source and drain electrodes are shorted together, the equivalent resistance reduced to

half of the gate to source/drain resistance.

The value of Cinv, by definition, can be expressed as

geff

g

invinv WL

dV

dQC , 4-4

where Vg is the top gate voltage. Qinv is a function of both Vg and the radiation generated

charge, Qrad, and can be expressed as either

kTCQVVq

iinvinvgradgenqtQ

// '0

, 4-5

or

radthgginv QVVCQ ' , 4-6

where, q is the electron charge, tinv is the thickness of the inversion layer, ni is the intrinsic

charge density of silicon, the V0 is the gate voltage at which the Fermi level in the channel

equals the intrinsic Fermi level. In addition, Cg’ = ox / (tox + tSi / 3) is the equivalent gate

capacitance per unit area in the back channel mode, and Cg = LeffWgCg’. The factor of 3

accounts for the approximate relative permittivity difference between Si and SiO2. When

Vg < Vth, the device is in the subthreshold regime, and then equation 4-5 applies. Conversely,

when Vg > Vth, the device is in inversion regime and the inversion charge is determined by

4-6. The different relations for Qinv come about due to the fact that, for Vg < Vth, the carrier

concentration in the channel is very low and the electric field lines from Qrad will terminate

on the top gate electrode. For Vg > Vth, the carrier concentration in the channel is high

enough to screen the electrical field lines from Qrad. Therefore, from equations 4-3 through

4-6 the total capacitance can be expressed as

Page 87: Passive Wireless Radiation Dosimeters for Radiation

70

inv

ovinv

oveff

ov CCC

C

kT

qLCC

12

2

41

, 4-7

1

22 1

govg

ovg

invovCCC

CCRCC , 4-8

where, once again, equation 4-7 and 4-8 correspond to Vg < Vth and Vg > Vth, respectively.

The relations in equation 4-7 and 4-8 provide interesting insight into how the

capacitance change is occurring. In the subthreshold regime, the capacitance tuning occurs

entirely due to the change of Cinv, and in the limit of low frequencies defined by

(Leff2q/4kT)2 << 1, then C Cov + Cinv. In the superthreshold regime, since Cg is a

constant, the total capacitance C is modulated by the change in the channel resistance Rinv.

Finally, we can relate Qrad to the dose [114] in an FDSOI device using

ftqGycmQ BOXrad /101.8 314 , 4-9

where f is the percentages of holes, relative to the number initially generated by the

radiation, trapped in the BOX layer and tBOX is the BOX layer thickness. If we define

sensitivity as the percent capacitance change per unit dose, and we calculate the optimum

bias point for the maximum sensitivity. The optimum bias point was found at 0/22 invCC

and 0/22 invRC at Vg < Vth and Vg > Vth, respectively. Since C Cov + Cinv when Vg < Vth,

the optimum bias point is independent of capacitance. When Vg > Vth, the optimum bias

point occurs at

gov CCC

4

3 , 4-10

and at the optimum bias point,

Page 88: Passive Wireless Radiation Dosimeters for Radiation

71

ov

ovg

g

invC

CC

CR

3

1 . 4-11

When Vg < Vth, the maximum sensitivity is calculated by:

Dose

Q

Q

Q

Q

C

C

C

CDose

C

C

rad

rad

inv

inv

inv

inv

11 , 4-12

where

1

invC

C , 4-13

kT

WLq

Q

C geff

inv

inv

, 4-14

g

inv

rad

inv

CkT

Qq

Q

Q

, 4-15

ftcm

Dose

QBox

rad

314101.8 , 4-16

as a result:

GycmftkT

ttq

CC

C

Dose

C

CBox

ox

siox

ovinv

inv /101.8)3/(

|1 314

2

max

. 4-17

When Vg > Vth, the maximum sensitivity is calculated by

Dose

Q

Q

Q

Q

R

R

C

CDose

C

C

rad

rad

inv

inv

inv

inv

11 , 4-18

where

ovg

ovg

inv

govg

ovg

inv

inv CC

CCR

CCC

CCR

R

C

2

2

222

1

, 4-19

eff

invg

inv

inv

L

RW

Q

R2

4

, 4-20

1

rad

inv

Q

Q , 4-21

Page 89: Passive Wireless Radiation Dosimeters for Radiation

72

finally, as a result:

Gycmft

L

L

L

ttq

CC

C

Dose

C

CBox

ov

eff

effox

Siox

gov

g/101.81

2

3/3

4

3|

1 314

2max

, 4-22

Figure 4-18. (a) Modeled and experimentally extracted inversion capacitance vs. gate

voltage at frequencies of 100 kHz and 100 MHz. (b) predicted Cinv vs dose at 100 kHz

compared to radiation experiment showing our model fit in the low capacitance regime.

Page 90: Passive Wireless Radiation Dosimeters for Radiation

73

where Lov is the overlap distance between the gate or source/drain electrode, Cov = oxLovWg

/ (tox + tSi / 3).

Figure 4-18(a) shows a comparison of the modeled and experimentally extracted values

of Cinv vs. Vg at frequencies of 100 kHz and 100 MHz. At 100 kHz, the experimental data

is fit to the subthreshold model in 4-17. Excellent agreement is achieved at lower

capacitance values, though some deviation occurs at higher capacitance, due to the

transition between subthreshold and superthreshold operation where our simple analytical

formalism is less accurate. Figure 4-18(b) shows a detail of Cinv vs. dose on a semi-log

plot which show the initial exponential increase in capacitance vs. dose, which is in

agreement with the subthreshold model. At 100 MHz, we find that the model in 4-22 fits

the experimental data, indicating that at higher frequencies, the capacitance switching

occurs above threshold. This switching regime is likely to be useful for wireless sensors

due to the lower overall channel resistance. However, the sensitivity is lower compared to

the subthreshold operation.

The analytical formalism is particularly useful for evaluating the maximum sensitivity

of FDSOI varactor dosimeters. Figure 4-19 shows a plot of the maximum theoretical

sensitivity vs. frequency where the models in equation 4-17 and 4-22 have been used. For

these calculation the geometric parameters of the devices were the same as those shown in

Figure 4-12. At low frequencies, the maximum sensitivity occurs in the subthreshold mode

of operation and is independent of frequency as shown by the black dashed line. However,

as the frequency increases, at some frequency, the maximum sensitivity transitions to

occurring in the superthreshold regime. When this occurs, the maximum sensitivity

Page 91: Passive Wireless Radiation Dosimeters for Radiation

74

decreases with further increase in frequency. The point at which this transition occurs,

depends upon many parameters, including the gate length, overlap capacitance, gate oxide

thickness and others.

The plot also shows the experimental sensitivity measured at 100 kHz. The

experimental sensitivity for this device (0.18 %/Gy) is much less than the theoretical

prediction. This difference can be attributed to two main effects. The first is the low f value

in 4-9 and if this value could be increased to near 100%, the sensitivity could increase by

a factor of 10x. The second effect is the substantial overlap and parasitic capacitance in our

Figure 4-19. Maximum predicted sensitivity (in percent capacitance change per unit

dose) vs. frequency for an FDSOI varactor in the subthreshold (black) and strong

inversion (red) limits. The solid blue symbol shows the current device performance

and the open symbols show expected improvements with improved f value and parasitic

capacitance optimization. Further improvement is possible with increased BOX

thickness and decreased gate length.

Page 92: Passive Wireless Radiation Dosimeters for Radiation

75

devices. Once again, optimizing this value, for instance by utilizing a self-aligned process

and an insulating substrate, another factor of 10x in sensitivity could be achieved. This

sensitivity metric only applies to capacitance-based dosimeters. In order to compare with

other dosimeters, additional studies are needed to determine the minimum capacitance

change that can be measured. Assuming 1 fF capacitance resolution with 10 pF total

capacitance and f = 50%, the minimum detectable dose is 5 × 10-4 Gy, which is in the same

order of RADFET [119].

Finally, the plot also shows that the sensitivity can be increased by increasing tBOX,

provided that the substrate bias can be scaled sufficiently to maintain constant f. The ability

to increase sensitivity by increasing the radiation collection layer thickness is a unique

advantage of the FDSOI approach. For instance, increasing the oxide thickness does not

increase sensitivity in conventional MOS-based capacitor dosimeters, due to the trade-off

between overall capacitance and charge collection volume. Furthermore, for high-

frequency operation, the sensitivity can be increased by decreasing the lateral dimensions

of the device. Such optimization would be important for wireless dosimeters where

maintaining high Q-factor as well as high sensitivity would be important.

The effect of substrate was ignored in our analytical model. The substrate will add

another parallel parasitic capacitance, and including this effect will provide a more

comprehensive analysis. Finally, the frequency dependent response of radiation generated

traps was not included in this study. The behavior of switching states near the Si/SiO2

interface [120], especially border traps, on different time scales and bias points may change

Page 93: Passive Wireless Radiation Dosimeters for Radiation

76

the shape of the C-V characteristics, changing the optimum bias point and sensitivity of

the dosimeter. This effect should also be taken into account for practical designs.

4.6 Simulation and optimization of FDSOI varactors

4.6.1 Requirements of FDSOI varactor for wireless sensing

In order to utilize FDSOI varactor as part of the LC resonant circuit for wireless sensing,

a few requirements need to be achieved.

First, the threshold voltage of the device needs to be about 0 V. Since no top gate bias

will be applied in wireless operation and the device will work at the threshold region, where

the capacitance is sensitive to bias, 0 V threshold voltage will make sure the capacitance

changes immediately upon irradiation. In order to achieve this, a gate metal with proper

work function needs to be selected for a specific device design. Second, the device needs

to have relatively large overall capacitance. A large capacitance will increase the signal to

noise ratio and capacitance tuning range of the device. Both of them will increase the

sensitivity, which is critical for small dose detection. This could be achieved by fabricating

devices with a large gate area. Third, the device needs to have a high Q-factor. The Q-

factor plays an important role for wireless sensing as the higher the Q-factor, the more

accurate the measurement [95]. The Q-factor can be calculated by RCL // . Since the

changing of inductance and capacitance in the circuit will change the resonant frequency,

which is usually tuned in a specific range, the best way to increase the Q-factor is to reduce

the resistance in the circuit. The intrinsic resistance is usually high at the threshold region,

which will dominate the total resistance in the circuit. In order to reduce the intrinsic

Page 94: Passive Wireless Radiation Dosimeters for Radiation

77

resistance of the FDSOI varactors, a multi-finger design needs to be utilized, as shown in

Figure 4-20. The key point of this structure is that the dimension of each fingers is small

while the number of finger is large. Since the fingers are parallel to each other, the total

capacitance is preserved while the overall resistance is reduced.

4.6.2 Device optimization

In order to determine the design space for the FDSOI varactors, Synopsys Sentaurus

DeviceTM was utilized to calculate the total input impedance from the gate to the S/D

electrodes (which were shorted together) for a single finger device. The default device

parameters are listed in Table 4-1. The SOI thickness, tSi, was 6 nm and the equivalent

gate oxide thickness was 1.6 nm. These values are achievable using state-of-the-art thin

SOI technology [121]. Additional parameters are listed in Table 4-1, which includes

Figure 4-20. Illustration of multi-finger design of FDSOI varactors.

Page 95: Passive Wireless Radiation Dosimeters for Radiation

78

estimations of the parasitic resistances associated with the narrow gate fingers and the

metal-semiconductor contact resistance. The gate electrode was assumed to be a metal

gate where the work function was used as an adjustable parameter.

Symbol Quantity Default Value

EOT Equivalent oxide thickness 1.6 nm

NA Channel p-doping 1015 cm-3

ND Source/drain n-doping 1020 cm-3

tBOX BOX thickness 200 nm

Rc S/D contact resistance 100 Ω-m

Rg Sheet resistance at gate 10 Ω/sq.

Wg Gate finger length 3 m

N Number of gate fingers 1000

f Nominal frequency 109 Hz

Table 4-1. Nominal FDSOI varactor simulation parameters

The extrinsic resistances associated with the contacts, Rcontact, and the gate metal, Rg-

metal, were added afterward. For an intrinsic channel resistance from the TCAD, Rintr, then

the total resistance, Rtot, can be calculated as

metalgcontactintrtot RRRR , 4-23

where

NW

RR

g

ccontact

2, 4-24

and,

NL

WRR

g

gg

metalg

. 4-25

Figure 4-21 shows the design trade-off of the finger length for a device with tsi = 20

nm. Figure 4-21(a) shows the Cg vs. Vg for a single finger device with finger length equals

Page 96: Passive Wireless Radiation Dosimeters for Radiation

79

to 100 nm, 200 nm and 400 nm. Figure 4-21(b) shows the intrinsic resistance vs. Vg of such

devices. Figure 4-21(c) shows the calculated Q-factor for such devices with 1000 fingers.

Q-factor of a capacitor is calculated by 1/2πfRC, where R is the total resistance and C is

the total capacitance at a certain gate voltage respectively. Longer finger length results in

higher a capacitance tuning range while having higher resistance. As a result, the Q-factor

is smaller for longer finger length devices.

Figure 4-22 shows the percent capacitance change vs. Dose (Figure 4-22-a), resistance

vs. Dose (Figure 4-22-b) and Q-factor vs. Dose (Figure 4-22-c) for a 200-nm finger length

Figure 4-21. (a) Cg vs. Vg for a single finger device with finger length equals to 100 nm, 200 nm

and 400 nm. (b) The intrinsic resistance vs. Vg of such devices. (c) Calculated Q-factor for such

devices with 1000 fingers.

Figure 4-22. (a) Percent capacitance change vs. Dose. (b) Resistance vs. Dose. (c) Q-factor vs.

Dose for device with finger length = 200 nm and Tsi = 10 nm, 20 nm and 40 nm with 15%

charge collection efficiency and optimized threshold voltage.

Page 97: Passive Wireless Radiation Dosimeters for Radiation

80

device with tsi equals to 10 nm, 20 nm and 40 nm, assuming 15% charge collection

efficiency with optimized threshold voltage. A thinner silicon layer provides a higher

capacitance tuning range due to the higher silicon layer capacitance. However, the

resistance is higher and the Q-factor is therefore smaller.

Figure 4-23 shows the optimization of gate metal work function of a device with a 200

nm finger length and 20 nm Tsi. Using a different work function metal will change the

threshold voltage of the device (a). As a result, the capacitance change under irradiation

will also be different (b). The change of capacitance will be measured as a frequency

change for wireless sensing (c).

Based on optimized structure, the Q-factor of the varactor can be calculated as 1/2πfRC,

where R is the resistance, C is capacitance of the varactor and f is the working frequency.

Figure 4-23 shows the Q-factor as a function of radiation dose. Overall, a trade-off exists

for the optimization of finger length and SOI thickness. However, with proper design, a

large frequency tuning (0.46%/Gy) and a large Q-factor (>20 at 1 GHz) could be achieved,

making FDSOI device promising for wireless radiation sensing applications.

Figure 4-23. (a) Percent capacitance change vs. Dose. (b) Resistance vs. Dose. (c) Q-factor vs.

Dose for device with finger length = 200 nm and Tsi = 10 nm, 20 nm and 40 nm with 15%

charge collection efficiency and optimized threshold voltage.

Page 98: Passive Wireless Radiation Dosimeters for Radiation

81

It is worth mentioning that the calculated Q-factor is for the varactor only instead of

LC circuit. Since the inductor will also have its own Q-factor, the overall Q-factor will not

exceed the smaller value of those two Q-factors separately. The overall Q-factor for

wireless sensing needs to be estimated when the wireless radiation dosimeter is fabricated,

as discussed in later sections.

Figure 4-24. Estimated Q-factor as a function of radiation dose at 1 GHz with optimized

design parameters.

Page 99: Passive Wireless Radiation Dosimeters for Radiation

82

Chapter 5: Passive wireless sensing using FDSOI devices

5.1 Theory of wireless sensing

As mentioned in the second chapter, once a passive LC resonant circuit is formed with

a FDSOI varactor and an inductor, the radiation will change the resonant frequency of this

circuit. In order to obtain the radiation information, a wireless reading technique needs to

be developed. The theory of wireless readout for a passive LC circuit was explained in

detail in [95]. For the integrity of this thesis, it is briefly described again here.

Figure 5-1 shows the equivalent network of an inductively coupled sensor system. The

left part of the system is the reading circuit, which consists of a reading inductor with

inductance L1 and an extrinsic resistance R1. The right part is the sensing circuit which has

a varactor C2 and a sensing inductor L2 with extrinsic resistance R2. The sensing inductor

and reading inductor are inductively coupled with a mutual inductance M. If the voltage

and current in the reading circuit are V1 and I1, andthe voltage and current in the sensing

circuit are L2 and R2, and the AC signal in the circuit has a frequency f, then:

Figure 5-1. Equivalent network of the inductively couple sensor system.

Page 100: Passive Wireless Radiation Dosimeters for Radiation

83

211111 22 fMIjIfLjIRV , 5-1

122222 22 fMIjIfLjIRV , 5-2

and

21LLkM , 5-3

where k is the geometry-dependent coupling coefficient with a value between 0 and 1. Also

in the sensing circuit, we have

222 2 VfCjI . 5-4

Therefore the equivalent input impedance Z1 measured from the reading circuit can be

derived from 5-1 to 5-4 as:

2

22

22

11

1

11

2

12(

)2(2

fCfLjR

MffLjR

I

VZ

. 5-5

If we define

22

22

1

CLf

, 5-6

2

2

2

2

1

C

L

RQ , 5-7

where Q2 is the Q-factor and f2 is the resonant frequency of the sensor circuit. The real part

of the total impedance can be derived from 5-5 as

2

22

2

2

22

2

111)//(1

/2Re

ffffQ

ffQkfLRZ

. 5-8

The resonant frequency fmax can be obtained from the maximum of Re{Z1} with respect to

f2,

Page 101: Passive Wireless Radiation Dosimeters for Radiation

84

22

2

2

max

24

2f

Q

Qf

. 5-9

If Q2 is large, then fmax ≈ f2. So if by using an impedance analyzer to measure the resistance

vs. frequency from the reading sensor, the resonant frequency can be found at the peak

resistance.

5.2 Device design and fabrication

5.2.1 Overview

The Q-factor plays an important role for wireless sensing as the higher the Q-factor,

the more accurate the measurement [95]. The Q-factor can be calculated from

2

2

2

2

1

C

L

RQ , 5-7. Since changing the inductance and capacitance in the circuit will

change the resonant frequency, which is usually tuned in a specific range, the best way to

increase the Q-factor is to reduce the resistance in the circuit. For radiation sensing, the

FDSOI device operates at the transition region between “off” and “on” states, where the

resistance value is usually high, therefore the intrinsic resistance of the device will

dominate the total resistance in the circuit. In order to reduce the intrinsic resistance of the

FDSOI varactors, a multi-finger design was utilized, as shown in Figure 4-20. The key

point of this structure is that the dimension of each finger is small while the number of

fingers is large. Since the fingers are parallel with each other, the total capacitance value is

large while the overall resistance is reduced.

The fabrication process for multi-finger varactors is similar to the process of single-

finger varactors and can be found in chapter 3. However, since the dimensions of each

Page 102: Passive Wireless Radiation Dosimeters for Radiation

85

finger in the multi-finger design is much smaller than single-finger devices and the overlap

region between gate and S/D needs to be reduced to reduce the parasitic capacitance,

electron beam lithography (EBL) was utilized in the fabrication. Also, the threshold voltage

of the device needs to be adjusted to around 0 V, since for wireless sensing, no external

voltage can be applied from the top gate. A different gate metal was used to adjust the

threshold voltage.

5.2.2 Electron beam lithography

One of the first challenges to fabricate small FDSOI varactors using EBL is to create

alignment marks, which is the first layer of device fabrication. Since all the subsequent

layers will be aligned to the alignment mark, it is important to make the marks with sharp

edges and high contrast. The easiest way is to deposit a metal layer as alignment marks, for

example, Au. However, this will not work for FDSOI device fabrication since the

subsequent steps involves high temperature process, for example, activation and gate oxide

growth. At high temperature, metal tends to melt or diffuse into other materials, which will

degrade the device performance and contaminate fabrication facilities.

Another way is to create deep trenches as alignment marks. Since the sample has a

stacked layer of silicon and SiO2, the trench needs to be deep to have enough contrast. A

typical thickness is about 1 µm. Therefore, an etching process needs to be developed to

etch through a few layers of silicon and SiO2. The etching process used in this work is as

follows. To begin with, a thick layer (~ 800 nm) of PMMA is spin-coated to act as a soft

mask and pattern the alignment marks using EBL. The PMMA layer needs to be thick

enough to survive the subsequent etching steps. Then, the sample is dipped into 10:1 BOE

Page 103: Passive Wireless Radiation Dosimeters for Radiation

86

solution for about 15s to etch away the top screen oxide. Then a combination of C4F8, SF6

and O2 plasma gas was used to etch away the exposed 30 nm SOI layer. This etching has a

good selectivity for silicon over SiO2, and it will stop at the BOX layer. After that, CF3

plasma gas can be used to completely etch away the exposed 200 nm BOX layer. Last, the

same method that etches the SOI layer to etch another at 800 nm into the Si substrate. These

will result in about a 1 µm deep trench which can be used for alignment marks. Figure

5-2(a) illustrates the etching of different layers and Figure 5-2(b) shows the profilometer

measurement indicating a 700 nm deep trench was successfully created.

In order to have good lift-off for small features, P(MMA/MAA) (poly methyl

methacrylate/methacrylicacid) / PMMA (poly methyl methacrylate) bi-layer resist was

used in the lithography process for metal depositions. Because of the lower molecular

weight of P(MMA/MAA), it has a lower clearing dose compared to PMMA. As a result,

an undercut profile will form after developing which prevents deposition of metal on the

side walls of the resist and gives good lift-off [122].

Figure 5-2. (a) Etching of different layers. (b) Profilometer measurement indicating a 700 nm

deep trench was successfully created.

Page 104: Passive Wireless Radiation Dosimeters for Radiation

87

5.2.3 Threshold voltage adjustment

As can be seen in Figure 4-8, the threshold voltage of single-finger FDSOI varactors

was about -1 V. In order to detect radiation efficiently the device has to operate around a

threshold voltage where the capacitance changes abruptly as a function of gate potential.

However, in wireless sensing applications, no external gate bias can be applied, therefore

the threshold voltage of the device need to be adjusted to around 0 V. The threshold voltage

of the MOSFET can be adjusted by either changing the gate oxide thickness, changing the

doping level in the channel, or changing the gate metal work function. For FDSOI varactors,

the channel is lightly doped to obtain fully-depleted condition while the gate oxide is kept

as thin as possible to get high sensitivity. Therefore, the only practical way to change the

threshold voltage is to adjust the gate work function. Therefore, for multi-finger varactors,

a high work function metal, Pd, was used as the gate metal. Figure 5-3(a) shows the optical

image of a 120-finger varactor with 4 µm gate length and 3 µm width for each finger.

Figure 5-3. (a) Optical image of a 120-finger FDSOI varactor with Lg = 4 µm and W = 3 µm

for each finger. (b) Cg vs. Vg characteristic of 4 identical devices indicating the threshold

voltage was around 0 V with small device to device variability.

Page 105: Passive Wireless Radiation Dosimeters for Radiation

88

Figure 5-3(b) shows the Cg vs. Vg characteristic for 4 identical devices, and as can be seen,

the threshold voltage was near 0 V with small device to device variability.

5.3 Device characterization

A 3-µm gate length device was used to characterize the wireless sensing capability.

The FDSOI varactor was mounted on a header and wire-bonded to a hand-wound inductor

to form an LC resonant circuit. A second hand-wound inductor was then closely positioned

to the sensing inductor and connected to an impedance analyzer. The impedance of the

readout inductor was then monitored vs. frequency to determine the resonant frequency,

fres, of the LC circuit. A key challenge for wireless radiation dosimeters is the fact that very

little radiation-induced charge is trapped when the substrate bias, Vsub = 0. Therefore, a

positive substrate bias needs to be applied in order to increase the fractional yield and

achieve reasonable sensitivity. This requirement is not inconsistent with passive sensing,

since the substrate bias does not dissipate any power, and so could be achieved using a

Figure 5-4. (a) “Partially wireless” measurement setup, where the substrate bias is supplied

using an external power supply. (b) “Fully wireless” measurement setup there the substrate bias

is supplied using a supercapacitor. In this configuration, the supercapacitor is mounted locally,

in close proximity to the varactor.

Page 106: Passive Wireless Radiation Dosimeters for Radiation

89

simple charge storage device such as a supercapacitor or electret. In our initial

measurements, the wireless dosimeters were tested using a “partially wireless”

configuration (Figure 5-4(a)) where a voltage source was used to provide the substrate bias.

However, in later experiments, a supercapacitor was utilized to create a “fully wireless”

sensor (Figure 5-4(b)), and results from both configurations are provided in this thesis.

Initially, the stand-alone varactors were characterized using the impedance analyzer.

The total impedance and phase angle was measured from the impedance analyzer and the

varactor capacitance and resistance can be determined from the imaginary and real part of

the impedance. As shown in Figure 5-5(a), the gate capacitance, Cg, vs. the gate voltage Vg,

(relative to the source/drain electrodes) was measured under different substrate bias Vsub

values from 0 V to +6 V (for clarity, curves with Vsub = 0, +2, +4, +6 V are shown in the

figure). As Vsub increased, the Cg vs. Vg curve shifted leftward parallel, as expected. The

substrate bias mimics the effect of radiation, since the substrate bias has the same effect on

Figure 5-5 (a) Cg vs. Vg characteristics at Vsub = 0, +2, +4, +6 V. (b) equivalent series resistance

vs. Vg at same Vsub. Measurement frequency: 60 MHz.

Page 107: Passive Wireless Radiation Dosimeters for Radiation

90

the device capacitance as trapped charge in the BOX layer. The data in the figure shows

that the device has been correctly tuned so that the capacitance at Vg = 0 (zero top-gate bias)

will increase if charge is trapped in the BOX upon exposure to radiation. Figure 5-5(b)

shows the varactor resistance of the device vs. Vg at different Vsub. The resistance is used

to analyze the Q-factor of the device, as discussed later.

The device was then connected to the inductor and measured using the “partially

wireless” setup. In order to verify the basic wireless sensor operation, the measurements

were performed for different substrate bias values before any radiation was applied. Figure

5-6(a) shows the read inductor resistance vs. frequency for Vsub = 0 to +6 V. The resonant

frequency was extracted by fitting the read inductor resistance vs. frequency data using a

parabolic function, and the frequency that gives the highest resistance was defined as the

resonant frequency. It is observed that the resonant frequency decreases, consistent with

expectations, since LCf res 2/1 , and the capacitance, C, increases with Vsub at Vg = 0 V,

as shown in Figure 5-5(a). Figure 5-6(b) shows the resonant frequency (extracted from

Figure 5-6 (a) Read inductor resistance vs. frequency for Vsub = 0 to 6 V. (b) Resonant frequency

(extracted from Fig. 4(a)) vs. the capacitance at Vg = 0 V and Vsub = 0 to 6 V.

Page 108: Passive Wireless Radiation Dosimeters for Radiation

91

Figure 5-6(a)) vs. the capacitance (at Vg = 0 V) from Vsub equals 0 to +6 V, further

confirming this trend. Taking into account the parasitic component of the sensor, the

resonant frequency should be determined by )(2/1 var parres CCLf , where Cvar is the

variable capacitance, which is a function of Vsub (in this set of measurements) or radiation

dose (in actual radiation experiment), Cpar is the parasitic capacitance and L is the

inductance of the sensor inductor. Cpar and L have fixed values and can be extracted by

fitting the frequency vs. capacitance data in Figure 5-6(b). For this device, it is found that

Cpar was about 5.3 pF and L was about 830 nH.

The Q-factor can be estimated in two ways. In an LCR circuit, RCLQ // , where

R is the total resistance in the circuit. The Q-factor can also be measured from the read

inductor resistance vs. fres shown in Figure 5-6(a). FWHMres ffQ / , where ∆fFWHM is the half

Figure 5-7. Estimated Q-factor at different Vsub.

Page 109: Passive Wireless Radiation Dosimeters for Radiation

92

maximum bandwidth [95]. Figure 5-7 shows the Q-factor at different Vsub, where Qmeasure

was extracted from Figure 5-6(a) and Qcalculate was calculated by assuming the varactor

resistance dominates the total resistance and total capacitance is the sum of Cvar and Cpar.

The overall trend of Q-factor vs. is consistent with previous simulation results [101], even

though the Q is much smaller due to larger gate length and less fingers compared to

simulation. The difference of between Qcalculation and Qmeasure could be due to the

measurement error of varactor resistance. At the measurement frequency (60 MHz) the

Im{impedance} is much larger and Re{impedance}, small error of phase angle could lead

to noticeable change of R.

5.4 Experimental results

For all radiation experiments, the devices were irradiated and measured in real time

using a commercial GammaKnife Co-60 clinical therapy unit, where a special mount was

fabricated to allow the varactors to reside at the center of the focused radiation spot. The

accuracy of the device positioning in the beam was confirmed using radiochromic film

(Gafchromic™ EBT2, Ashland, Inc.). When the device was irradiated, the device in the

mounting assembly was moved into focus of the beams by the movement of the patient

couch. The device was first measured as a varactor to confirm the capacitance change under

radiation and then connected with sensor inductor and measured wirelessly. During the

experiment, the real part of the read inductor impedance (i.e. the resistance) was measured

using a continuous frequency scan with a repetition rate of 10 seconds.

Page 110: Passive Wireless Radiation Dosimeters for Radiation

93

Figure 5-8 shows a picture of the experimental device with read inductor closely

coupled with the sense inductor. The experimental varactor has 4-µm gate length and 120-

fingers. Figure 5-9 shows the Cg vs. Vg characteristic of the device at different Vsub from 0

V to +10 V. Similar to the 3- µm gate length varactor, the capacitance at Vg = 0 V was

modulated by Vsub. This indicates that the capacitance will also be modulated by irradiation

and was further confirmed by measuring the FDSOI device as a single varactor under Co-

60 irradiation. It was found that the capacitance changing rate was about 1.4 fF/Gy at a

substrate bias of Vsub = +5 V. This sensitivity was used to compare with the wireless

measurement result, as will be discussed later.

Figure 5-8. Picture of experimental device with the read inductors closely coupled with

sense inductors

Page 111: Passive Wireless Radiation Dosimeters for Radiation

94

Next, the “partially wireless” test setup in Figure 5-4(a) was used to determine the

resonant frequency shift under Co-60 irradiation at Vsub = +5 V. Figure 5-10(a) shows the

read inductor resistance vs. frequency at different radiation doses up to 80 Gy. Clearly a

leftward shift of the peak resistance is observed with increasing dose, consistent with

increased capacitance, as expected. Figure 5-10(b) shows the extracted resonant frequency

vs. dose during radiation, while the inset shows the resonant frequency before and after

radiation. Only a minimal amount of drift is observed without radiation, while a linear

change in resonant frequency is observed with radiation with a measured sensitivity of -

2.53 kHz/Gy. The change in the resonant frequency is consistent with the capacitance

change described earlier, as shown in Figure 5-10(b). The blue line shows the predicted

Figure 5-9. Cg vs. Vg characteristic of the experimental device under different Vsub.

Page 112: Passive Wireless Radiation Dosimeters for Radiation

95

resonant frequency vs. dose. Overall, the measured resonant frequency change matches

with the prediction reasonably well.

Figure 5-10. (a) Read inductor resistance vs. frequency at different radiation dose. (b)

Measured resonant frequency vs. dose compared with model using a substrate power

supply bias of Vsub = +5 V. The sensitivity is 2.53 kHz/Gy. Inset: Resonant frequency

vs. time before and after irradiation showing good stability without radiation.

Page 113: Passive Wireless Radiation Dosimeters for Radiation

96

Using an external power supply to provide the substrate voltage is not preferred in

practical applications because of its large size and also it will diminish the advantage of

wireless sensing. However, as mentioned previously, since the substrate bias is

dissipationless, the power supply can be replaced by a simple charge storage device, such

Figure 5-11. (a) Read inductor resistance vs. frequency at different radiation dose for

supercapacitor substrate bias of Vsub = +2.6 V. (b) Resonant frequency vs. radiation dose

showing sensitivity of 2.02 kHz/Gy.

Page 114: Passive Wireless Radiation Dosimeters for Radiation

97

as a supercapacitor. We therefore used a 2 mm × 3 mm × 1 mm size supercapacitor (11

mF) as shown in Figure 5-4(b), and repeated the measurement sequence. The

supercapacitor provided a constant substrate voltage of +2.6 V, and this bias value was

verified before and after irradiation. The results using this “fully wireless” configuration

are shown in Figure 5-11. Similar to the results in Figure 5-10, the read inductor resistance

vs. frequency curve shifted leftward as irradiation was increased (Figure 5-11(a)), resulting

in a linear decrease of resonant frequency vs. radiation dose as shown in Figure 5-11(b).

The slightly lower sensitivity of -2.02 kHz/Gy, is consistent with the lower substrate bias

provided by the supercapacitor (Vsub = +2.6 V) compared to the power supply (Vsub = +5.0

V).

The non-optimized design of our current sensors offers considerable opportunity for

sensitivity improvement through design optimization. The sensitivity in our devices is

limited by two main factors: (1) the high ratio of parasitic capacitance to variable

capacitance and (2) the low charge collection efficiency of the BOX layer. Regarding (1),

utilizing a self-aligned process can eliminate the gate overlap capacitance with the

source/drain implants. Furthermore, by moving to a monolithic design on an insulating

substrate, where the inductor and varactor are integrated onto the same chip, the parasitic

capacitance associated with the large bond pads and chip header could be eliminated. This

on-chip component integration would also greatly reduce the final size of the sensor, which

could be beneficial for clinical use. Concerning issue (2), we showed previously that our

devices have relatively low hole collection efficiency (~ 7%) at Vsub = 5 V [114]. Therefore,

the possibility exists to increase the charge collection efficiency of the BOX layer [117].

Page 115: Passive Wireless Radiation Dosimeters for Radiation

98

Our calculations indicate that the combined effect of each of these optimizations could

increase the sensitivity by as much as 50×.

An additional method of enhancing the sensor performance is to increase the Q-factor,

since the minimum detectable dose change can be improved with higher Q. For the “fully

wireless” configuration, Q was only ~ 8 and we believe this is due to several non-idealities,

including the large gate length of our devices which increases their intrinsic resistance, as

well as extrinsic effects such as solder resistances. Our previous calculations [114] have

shown that much higher Q values can be achieved by utilizing multi-finger devices with

sub-micron gate lengths, and monolithic integration should minimize resistance associated

with the resonator assembly. The accuracy of the device could be increased by

implementing more sophisticated wireless readout techniques.

5.5 Device integration with small size

The demonstrated passive wireless sensing was conducted using an FDSOI varactor

connected with a hand-wound inductor. Since commercial inductors and capacitors with

small size are available, it is possible to integrate the radiation dosimeter with smaller

components and make it more compact.

Figure 5-12(a) shows the integrated device. Two small 560 nH inductors were used as

sense and read inductors and closely placed to each other. A 6.2 pF capacitor was connected

with the sense inductor as a parasitic capacitance to adjust the resonant frequency. The

FDSOI varactor, sitting on a supercapacitor was also connected with the parasitic capacitor

and sense inductor to complete the LC resonant circuit. The connections were made using

re-flow soldering and wire bonding. The whole device was positioned on a quartz substrate

Page 116: Passive Wireless Radiation Dosimeters for Radiation

99

to minimize unwanted parasitic elements. The total size of the device was 2.8 mm × 6 mm

× 2.5 mm. Figure 5-12(b) shows the resistance and phase angle of the read inductor as a

function of frequency. A resonant peak can clearly be seen in the figure and the resonant

frequency was about 80 MHz, as expected (the varactor capacitance was about 0.75 pF and

with 6.2 pF capacitor and 560 nH inductor, the expected frequency was around 80 MHz).

To date, a radiation experiment using this compact device has not been conducted but is

planned in the near future.

Figure 5-12 (a) Integrated device with commercial inductor and capacitor. (b) Wireless readout

result of the device at left.

Page 117: Passive Wireless Radiation Dosimeters for Radiation

100

Chapter 6: Conclusion and future work

6.1 Conclusion

Radiation cancer therapy is one of the most important methods of cancer treatment.

The goal of radiation therapy is to apply maximum amount of dose to the tumor region

while applying minimum amount of dose to the normal tissues. Being able to measure

radiation dose accurately is critical for radiation therapy because it will ensure that

radiation is properly delivered to the tumor region as planned. In my dissertation, a passive

wireless radiation dosimeter for cancer therapy based on FDSOI technology was proposed

and realized through the following steps.

First the radiation response of FDSOI MOSFETs over a wide range of therapeutic X-

ray beam energies, angles, dose ranges and applied substrate voltages were characterized.

Results show that the threshold voltage shift under radiation was independent of X-ray

energies and incident angles. The response was reliable over large dose ranges for radiation

therapy and applied substrate voltage could improve the sensitivity of the device. The

charge collection efficiency was calculated from both TCAD simulation and an analytical

model to evaluate the radiation effect quantitatively. Based upon the experimental results,

the design of passive wireless sensors using FDSOI varactors was proposed.

Then the fabrication process of FDSOI varactors was developed at the University of

Minnesota. CV and IV and TCAD modeling were used to fully understand the device

characterization. The capacitance based sensing of Co-60 gamma radiation was

demonstrated using FDSOI varactors. The devices showed good linear response of

Page 118: Passive Wireless Radiation Dosimeters for Radiation

101

capacitance change upon irradiation with small variability and encouraging accuracy. An

analytical model was developed to further understand the device response under irradiation.

Based on the analytical model, the optimal working condition at different frequencies was

discussed and a guideline of improving the device sensitivity was provided. TCAD

simulation was also conducted to optimize the design space of FDSOI varactors for

wireless sensing.

Last a multi-finger design of FDSOI varactors with improved Q-factors and threshold

voltage adjustment method for wireless sensing were proposed. 120-finger varactors with

0 V threshold voltage were fabricated using electron beam lithography with small device

variability. An impedance measurement setup was built in a clinical environment.

Completely passive wireless radiation sensing was demonstrated using FDSOI varactor

and the device showed good linearity of resonant frequency change under radiation with a

sensitivity of 2.02 kHz/Gy. The possibilities to improve the performance of device were

discussed. Overall, the results show that FDSOI technology has the potential to realize an

ultra-small, passive wireless radiation dosimeter for in vivo radiation cancer therapy with

high sensitivity and accuracy.

6.2 Proposed future work

6.2.1 Further development of FDOSI radiation dosimeter

Even though passive wireless sensing was demonstrated in this dissertation, a

significant amount of work remains to make the device suitable for clinical application. I

would like to propose my ideas for short term improvement and long term development.

Page 119: Passive Wireless Radiation Dosimeters for Radiation

102

In the short terms, the performance of the device needs to be improved from at least

three aspects, i.e. the sensitivity, the minimum detectable dose and the size of the device.

The current sensitivity of the device is limited by (1) its low ratio of varactor

capacitance to overall capacitance and (2) low charge collection efficiency. The first one

could be addressed by increasing the number of fingers in one device while decreasing the

parasitic capacitance of the device. As mentioned in the last chapter, the parasitic

capacitance could be reduced by using a thicker BOX, insulating substrate and monolithic

design.

Figure 6-1. SEM image of the gate-level metal deposition of a 1000-finger varactor

with Lg = 200 nm

Page 120: Passive Wireless Radiation Dosimeters for Radiation

103

The minimum detectable dose is limited by a low Q-factor and noise in the

measurement due to non-ideal reading technique. Q-factor can be improved by fabricating

a device with smaller gate length and more fingers and more sophisticated reading

technique, for example phase lock method could be utilized in wireless sensing. Figure 6-1

shows the SEM image of the gate-level metal deposition of a 1000-finger varactor with Lg

= 200 nm. In order to fabricate varactor with large number of fingers and small gate length,

some lithography difficulty needs to be overcome, to prevent shorting from each finger.

The final dosimeter will consist of three parts: the FDSOI varactor, the inductor and

the supercapacitor. The size of the varactor can be very small due to high capacitance per

unit area of FDSOI devices and modern CMOS fabrication technique. The size of the

device is mostly likely limited by the inductor and supercapacitor. Currently the inductor

was hand-wounded and has a large size. However, on chip inductors could be fabricated

Figure 6-2. Proposed structure of final dosimeter. Size: 2 mm3 Graph by: S. Koester.

Page 121: Passive Wireless Radiation Dosimeters for Radiation

104

and monolithic integration of the device could be utilized to significantly reduce the size.

The size of the supercapacitor used in the experiments was about 2 mm * 3mm * 1mm.

Commercial solid state batteries with much smaller size are available and can be used to

provide the substrate bias [123]. The final size of the integrated device should be able to

achieve sub 2 mm3. Figure 6-2 shows a proposed structure of final dosimeter.

In the long term, given the size of the device can be reduced, proper packaging method

needs to be developed to protect the device form environmental factors, such as moisture.

It is well known that radiation could cause degradation for CMOS [124], even though the

dose for radiation therapy is small, it is important to test the reliability of the device

throughout the whole radiation therapy. Also silicon based device are known to have some

stability issues due to temperature variation [125] and bias [126]. Therefore, the stability

of the device over time of a course of radiation therapy also needs to be tested. After that,

the device will be ready for in vivo test and animal studies.

6.2.2 Radiation dosimeter based on two-dimensional materials

Even though FDSOI structure provide unique advantages to measure radiation

wirelessly, some intrinsic design trade-offs and disadvantages exist and may limit its

application for radiation sensing. For example, fundamental trade-off exists between

sensitivity and Q-factor for the dosimeter. In order to get high sensitivity, the silicon

channel thickness needs to be as thin as possible. On the other side, the thinner channel,

the higher resistance in the device, as a result, the Q-factor will be lower. Also, SiO2/Si

interface can trap charge but it under goes some slow annealing process and the charge will

be lost as time goes by [127], [128]. The rate of this annealing process dependents on

Page 122: Passive Wireless Radiation Dosimeters for Radiation

105

temperature, stress, bias, process and so on and is difficult to control. This may not

significantly influence the operation for radiation therapy because of the short time range

but makes it difficult to use FDSOI device to detect low radiation dose at relative large

time scale, for example, in occupational dosimetry applications[129]. Moreover, since the

SOI wafers are purchased from vendors, less flexibility exists to engineer the wafer and

change its property, except doing some high temperature annealing or implantation. It

would be great if engineers have the freedom to create customized substrate and make them

radiation sensitive.

The discovery of 2-dimentional materials, for example grapheme [130], [131] and

transition metal dichalcogenides (TMD), such as MoS2 [132] and WSe2 [133] has expended

people’s understanding about materials and provided unique electrical properties. For

example, grapheme has semi-metallic characteristics and extremely high mobility [134].

TMDs’ have tunable band gap depending on their number of layers [135]. Numerous

potential applications exist for those material such as high frequency RF devices [136],[137]

and flexible electronics [138].

Two-dimensional materials also provide new opportunities for radiation sensing and

can overcome some of the disadvantages of FDSOI devices. For example, because of the

semi-metal property and quantum capacitance of grapheme, it is possible to realize a

grapheme varactor for wireless sensing with extremely high Q factors. Wireless vapor

sensing using grapheme varactors has been demonstrated [139], and it is possible to apply

the same technique to radiation sensing. Also, MoS2 has a band gap of about 1.8 eV and

can be utilized as a channel material for varactors. One unique advantage of such 2-

Page 123: Passive Wireless Radiation Dosimeters for Radiation

106

dimesional materials is that their flexibility. Since they can be made through chemical

vapor deposition (CVD) [140], they can be transferred on almost any substrates. This

provides a new dimension for engineering and customizing the substrate to make it more

sensitive to radiation.

I would like to propose one new device structure which can potentially expend the

application of such radiation dosimeters. Figure 6-3(a) shows the structure of the device.

Similar to FDSOI varactors, it also has two components, the top varactor and the charge

trapping substrate. Different from FDSOI varactors, the top varactor channel is made with

2-dimensional material, such as grapheme and MoS2. The charge trapping substrate has

stacked layers of 20-nm Al2O3, 20-nm Si3N4 and 300-nm SiO2 from top to bottom. Figure

6-3(b) shows the band alignment of the charge trapping substrate. As can be seen form the

figure, Al2O3 and MoS2 have large band gap, about 9 eV and 8.9 eV respectively while

Si3N4 has relative small band gap about 4.5 eV and is sandwiched between Al2O3 and MoS2.

Once radiation is absorbed, it creates electron-hole pairs in the thick SiO2 layer and if small

electrical field exists from SiO2 to Si3N4 layer, the holes will transport towards the Si3N4.

Once they reached the Si3N4, the holes will be permanently trapped in that energy well and

will change the characteristics of top varactor. Specifically, for grapheme device, the

trapped charge will electronically dope the grapheme and change its quantum capacitance;

for MoS2 device, the trapped charge would turn the device from “off” to “on” state and the

top gate capacitance will be changed. This change of capacitance can be measured directly

from CV measurement, or transferred as a frequency signal and measured wirelessly. The

advantage of such structure is that it creates physical traps for holes and can potentially

Page 124: Passive Wireless Radiation Dosimeters for Radiation

107

trap the holes for very long time [141]. Such device could be suitable for occupational

Figure 6-3. (a) Proposed new device structure of radiation dosimeter. The channel is

made with 2-dimensional materials. The substrate has a stacked layers of 20 nm Al2O3,

20 nm Si3N4 and 200 nm SiO2 as charge trapping layer. (b) Band alignment of the

charge trapping layer. Due to its energy well structure, radiation generated holes will

be trapped in the Si3N4 layer permanently

Page 125: Passive Wireless Radiation Dosimeters for Radiation

108

dosimetry where radiation dose is measured in a relatively long time range. If grapheme is

used as a channel material, it could provide extremely high Q-factor. Therefore, ultra-low

radiation dose would be possibly measured wirelessly.

Page 126: Passive Wireless Radiation Dosimeters for Radiation

109

Bibliography

[1] C. M. Washington and D. Leaver, “principles and Practice of Radiation Therapy,”

Mosby, 1999. [Online]. Available: http://www.abebooks.com/principles-Practice-

Radiation-Therapy-Washington-Charles/1339540181/bd.

[2] “Radiation therapy data 1.” [Online]. Available: http://www.cancer.gov/about-

cancer/treatment/types/radiation-therapy/radiation-fact-sheet.

[3] “Cancer Treatment Survivorship.” [Online]. Available:

http://www.cancer.org/acs/groups/content/@research/documents/document/acspc-

042801.pdf.

[4] W. C. Röntgen, “On a New Kind of Rays,” Science (80-. )., vol. 3, no. 59, pp. 227–

231, 1986.

[5] P. Hodges, The life and times of Emil H. Grubbe: The biography of a pioneering

chicago radiologist, First Edit. University of Chicago Press, 1964.

[6] B. R. Young, “Should the method of coutard be applied in all cases of cancer theated

by Goentgen Rays?,” Radiology, no. 1, pp. 186–189, 1937.

[7] H. Kaplan, “Historic milestones in radiobiology and radiation therapy,” Semin

Oncol., vol. 6, no. 4, pp. 479–489, 1979.

[8] M. J. Zelefsky, S. a Leibel, P. B. Gaudin, G. J. Kutcher, N. E. Fleshner, E. S.

Venkatramen, V. E. Reuter, W. R. Fair, C. C. Ling, and Z. Fuks, “Dose escalation

with three-dimensional conformal radiation therapy affects the outcome in prostate

cancer.,” Int. J. Radiat. Oncol. Biol. Phys., vol. 41, no. 3, pp. 491–500, 1998.

[9] N. Lee, P. Xia, J. M. Quivey, K. Sultanem, I. Poon, C. Akazawa, P. Akazawa, V.

Weinberg, and K. K. Fu, “Intensity-modulated radiotherapy in the treatment of

nasopharyngeal carcinoma: an update of the UCSF experience.,” Int. J. Radiat.

Oncol. Biol. Phys., vol. 53, no. 1, pp. 12–22, May 2002.

Page 127: Passive Wireless Radiation Dosimeters for Radiation

110

[10] W. H. Beierwaltes, “The Treatment of thyroid carcinoma with radioactive iodine,”

Semin. Nucl. Med., vol. 8, no. 1, pp. 79–94, 1978.

[11] G. Riethmuller, E. Holz, G. Schlimok, W. Schmiegel, R. Raab, K. Hoffken, R.

Gruber, I. Funke, H. Pichlmaier, H. Hirche, P. Buggisch, J. Witte, and R. Pichlmayr,

“Monoclonal antibody therapy for resected Dukes’C colorectal cancer: seven-year

outcome of a multicenter randomized trial,” J Clin Oncol, vol. 16, no. 5, pp. 1788–

1794, 1998.

[12] B. Pierquin, Brachytherapy. St. Louis: W. H. Green, 1978.

[13] F. M. Khan, The physics of radiation therapy, 4th ed. Lippincott Williams & Wilkins,

2010.

[14] S. C. Klevenagen, Physics of electron beam therapy.Taylor & Francis, 1985.

[15] S. Yajnik, Proton beam therapy how protons are revolutionizing cancer treatment.

Springer Science, 2013.

[16] F. P. Ellinger and R. Gross, The biologic fundamentals of radiation therapy. Elsevier

pubilishing company, inc., 1941.

[17] J. T. Bushberg, J. A. Seibert, E. Leidholdt, and J. Boone, The essential physics of

medical imaging. Wolters Kluwer Health/Lippincott Williams & Wilkins, 2012.

[18] T. R. Oldham and F. B. McLean, “Total ionizing dose effects in MOS oxides and

devices,” IEEE Trans. Nucl. Sci., vol. 50, no. 3, pp. 483–499, Jun. 2003.

[19] H. E. Boesch and J. M. Mcgarrity, “Charge yield and dose effects in MOS capacitors

at 80 K,” IEEE Trans. Nucl. Sci, vol. NS-23, no. 6, pp. 1520–1525, 1976.

[20] R. C. Hughes, “Charge-carrier transport phenomena in amorphous SiO2: direct

measurement of the drift mobility and lifetime,” Phys. Rev. Lett., vol. 30, no. 26, pp.

1333–1336, 1973.

[21] L. Onsager, “Initial recombination of ions,” Phys. Rev., vol. 54, no. 8, pp. 554–557,

Page 128: Passive Wireless Radiation Dosimeters for Radiation

111

1938.

[22] T. R. Oldham, “Recombination along the tracks of heavy charged particles in SiO2

films,” J. Appl. Phys., vol. 57, no. 8, pp. 2695–2702, 1985.

[23] H. Scher and M. Lax, “Stochastic transport in a disordered solid. II. Impurity

conduction,” Phys. Rev. B, vol. 7, no. 10, pp. 4502–4519, 1973.

[24] R. C. Hughes, “Hole mobility and transport in thin SiO2 films,” Appl. Phys. Lett.,

vol. 26, no. 8, pp. 436–438, 1975.

[25] O. L. Curtis and J. R. Srour, “The multiple-trapping model and hole transport in

SiO2,” J. Appl. Phys., vol. 48, no. 9, pp. 3819–3828, 1977.

[26] E. Boesch and F. B. McLean, “Hole transport and trapping in field oxides,” IEEE

Trans. Nucl. Sci., vol. 32, no. 6, pp. 3940–3945, 1985.

[27] F. J. Feigl, W. B. Fowler, and K. L. Yip, “Oxygen vacancy model for the E1’ center

in SiO2,” Solid State Commun., vol. 14, pp. 225–229, 1974.

[28] D. M. Fleetwood, “Revised model of thermally stimulated current in MOS

capacitors,” 1997.

[29] V. Lakshmanna and A. S. Vengurlekar, “Logarithmic detrapping response for holes

injected into SiO2 and the influence of thermal activation and electric fields,” J.

Appl. Phys., vol. 63, no. 9, pp. 4548–4554, 1988.

[30] H. E. Boesch and F. B. McLean, “Interface state generation associated with hole

transport in metal-oxide-semiconductor structures,” J. Appl. Phys., vol. 60, no. 1, pp.

448–449, 1986.

[31] F. B. Mclean, “A framework for understanding radiation-induced interface states in

SiO2 MOS structures,” IEEE Trans. Nucl. Sci., vol. NS-27, no. 6, pp. 1651–1657,

1980.

[32] N. S. Saks, R. B. Klein, R. E. Stahlbush, B. J. Mrstik, and R. W. Rendell, “Effects

Page 129: Passive Wireless Radiation Dosimeters for Radiation

112

of post-stress hydrogen annealing on MOS oxides after Co-60 irradiation or Fowler-

Nordheim injection,” IEEE Trans. Nucl. Sci, vol. 40, no. 6, pp. 1341–1349, 1993.

[33] Y. Nishimura and R. Komaki, Intensity-modulated radiation therapy : clinical

evidence and techniques. Springer, 2015.

[34] J. D. Bourland, Image-guided radiation therapy Imaging in medical diagnosis and

therapy. CRC Press, 2012.

[35] J. M. Edmund and C. E. Andersen, “Temperature dependence of the Al2O3:C

response in medical iuminescence dosimetry,” Radiat. Meas., vol. 42, no. 2, pp.

177–189, 2007.

[36] B. G. Markey, S. W. S. McKeever, M. S. Akselrod, L. Botter-Jensen, N. Agersnap

Larsen, and L. . Colyyott, “The temperature dependence of optically stimulated

luminescence from Alpha-Al2O3:C,” Radiat. Prot. Dosimetry, vol. 65, pp. 185–189,

1996.

[37] E. G. Yukihara, V. H. Whitley, J. C. Polf, D. M. Klein, S. W. S. McKeever, A. E.

Akselrod, and M. S. Akselrod, “The effects of deep trap population on the

thermoluminescence of Al 2O3:C,” Radiat. Meas., vol. 37, no. 6, pp. 627–638, 2003.

[38] E. G. Yukihara and S. W. S. McKeever, “Optically stimulated luminescence (OSL)

dosimetry in medicine.,” Phys. Med. Biol., vol. 53, pp. R351–R379, 2008.

[39] P. A. Jursinic, “Characterization of optically stimulated luminescent dosimeters,

OSLDs, for clinical dosimetric measurements.,” Med. Phys., 2007.

[40] C. A. Perks, C. Yahnke, and M. Million, “Medical dosimetry using Optically

Stimulated Luminescence dots and microStar readers,” in 12th International

Contress of the International Radiation Protection Association, 2008.

[41] C. E. Andersen, S. K. Nielsen, S. Greilich, J. Helt-Hansen, J. C. Lindegaard, and K.

Tanderup, “Characterization of a fiber-coupled Al2O3:C luminescence dosimetry

system for online in vivo dose verification during 192Ir brachytherapy.,” Med. Phys.,

Page 130: Passive Wireless Radiation Dosimeters for Radiation

113

vol. 36, no. 3, pp. 708–718, 2009.

[42] A. Holmes-Siedle and L. Adams, “RADFET: A review of the use of metal-oxide-

silicon devices as integrating dosimeters,” Int. J. Radiat. Appl. Instrumentation. Part

C. Radiat. Phys. Chem., vol. 28, no. 2, pp. 235–244, 1986.

[43] A. Kelleher, M. O’Sullivan, J. Ryan, B. O’Neill, and W. Lane, “Development of the

radiation sensitivity of PMOS dosimeters,” IEEE Trans. Nucl. Sci., vol. 39, no. 3,

pp. 342–346, 1992.

[44] A. G. Holmes-Siedle, J. O. Goldsten, R. H. Maurer, and P. N. Peplowski, “RADFET

dosimeters in the belt: The Van Allen Probes on day 365,” IEEE Trans. Nucl. Sci.,

vol. 61, no. 2, pp. 948–954, 2014.

[45] J. R. Schwank, S. B. Roeske, D. E. Beutler, D. J. Moreno, and M. R. Shaneyfelt, “A

dose rate independent pMOS dosimeter for space applications,” IEEE Trans. Nucl.

Sci., vol. 43, no. 6, pp. 2671–2678, 1996.

[46] Y. Kimoto, H. Koshiishi, H. Matsumoto, and T. Gota, “Total dose orbital data by

dosimeter onboard tsubase (MDS-1) satellite,” IEEE Trans. Nucl. Sci., vol. 50, no.

6, pp. 2301–2306, 2003.

[47] L. J. Asensio, M. a. Carvajal, J. a. López-Villanueva, M. Vilches, a. M. Lallena,

and a. J. Palma, “Evaluation of a low-cost commercial mosfet as radiation

dosimeter,” Sensors Actuators A Phys., vol. 125, no. 2, pp. 288–295, Jan. 2006.

[48] R. Ramani, S. Russell, and P. O’Brien, “Clinical dosimetry using MOSFETs,” Int.

J. Radiat. Oncol. Biol. Phys., vol. 37, no. 4, pp. 959–964, 1997.

[49] C. F. Chuang, L. J. Verhey, and P. Xia, “Investigation of the use of MOSFET for

clinical IMRT dosimetric verification,” Med. Phys., vol. 29, no. 6, pp. 1109–1115,

2002.

[50] S. M. Sze and K. K. Ng, Physics of semiconductor devices. Eiley-Interscience, 2007.

[51] N. G. Tarr, K. Shortt, Y. Wang, and I. Thomson, “A sensitive , temperature-

Page 131: Passive Wireless Radiation Dosimeters for Radiation

114

compensated , zero-bias floating gate MOSFET dosimeter,” IEEE Trans. Nucl. Sci.,

vol. 51, no. 3, pp. 1277–1282, 2004.

[52] N. G. Tarr, G. F. Mackay, K. Shortt, and I. Thomson, “A floating gate MOSFET

dosimeter requiring no external bias supply,” in Radiation and Its Effects on

Components and Systems, 1997, pp. 277–281.

[53] M. N. Martin, D. R. Roth, a. Garrison-Darrin, P. J. McNulty, and a. G. Andreou,

“FGMOS dosimetry: design and implementation,” IEEE Trans. Nucl. Sci., vol. 48,

no. 6, pp. 2050–2055, 2001.

[54] E. Isern, M. Roca, R. Picos, J. Font, J. Cesari, and A. Pineda, “Floating gate CMOS

dosimter with frequency output,” IEEE Trans. Nucl. Sci., vol. 59, no. 2, pp. 373–

378, 2012.

[55] M. Alvarez, C. Hernando, J. Cesari, A. Pineda, and E. Garcia-Moreno, “Total

ionizing dose characterization of a prototype floating gate mosfet dosimeter for

space applications,” IEEE Trans. Nucl. Sci., vol. 60, no. 6, pp. 4281–4288, 2013.

[56] E. G. Villani, M. Crepaldi, D. DeMarchi, A. Gabrielli, A. Khan, E. Pikhay, Y. Roizin,

A. Rosenfeld, and Z. Zhang, “A monolithic 180 nm CMOS dosimeter for wireless

In Vivo Dosimetry,” Radiat. Meas., vol. 84, pp. 55–64, 2016.

[57] C. W. Scarantino, D. M. Ruslander, C. J. Rini, G. G. Mann, H. T. Nagle, and R. D.

Black, “An implantable radiation dosimeter for use in external beam radiation

therapy,” Med. Phys., vol. 31, no. 9, pp. 2658–2671, 2004.

[58] J. M. Fagerstrom, J. a. Micka, and L. a. DeWerd, “Response of an implantable

MOSFET dosimeter to 192 Ir HDR radiation,” Med. Phys., vol. 35, no. 12, pp. 5729–

5737, 2008.

[59] G. P. Beyer, G. G. Mann, J. a. Pursley, E. T. Espenhahn, C. Fraisse, D. J. Godfrey,

M. Oldham, T. B. Carrea, N. Bolick, and C. W. Scarantino, “An implantable

MOSFET dosimeter for the measurement of radiation dose in tissue during cancer

therapy,” IEEE Sens. J., vol. 8, no. 1, pp. 38–51, Jan. 2008.

Page 132: Passive Wireless Radiation Dosimeters for Radiation

115

[60] “US20100219494A1,” 2010.

[61] M. Gopalan, “Experimental study of MOS capacitors as wireless radiation dose

sensors,” 2010.

[62] C. Son and B. Ziaie, “A wireless implantable passive microdosimeter for radiation

oncology.,” IEEE Trans. Biomed. Eng., vol. 55, no. 6, pp. 1772–1775, Jun. 2008.

[63] T. P. Ma and P. V. Dressendorfer, Ionizing radiation effects in MOS devices and

circuits. Wiley, 1989.

[64] P. E. Dodd and L. W. Massengill, “Basic mechanisms and modeling of single-event

upset in digital microelectronics,” IEEE Trans. Nucl. Sci., vol. 50, no. 3, pp. 583–

602, 2003.

[65] O. Musseau, “Single-event effects in SOI,” IEEE Trans. Nucl. Sci., vol. 43, no. 2,

pp. 603–613, 1996.

[66] H. M. Manasevit and W. I. Simpson, “Single-crystal silicon on a sapphire substrate,”

J. Appl. Phys., vol. 35, no. 4, pp. 1349–1351, 1964.

[67] J. E. Maurits, “SOS wafers-some comparisons to silicon wafers,” IEEE Trans.

Electron Devices, vol. ED-25, no. 8, pp. 859–863, 1978.

[68] H. W. Lam, “Silicon on insulating substrates - recent advances,” IEDM. Dig., pp.

348–351, 1983.

[69] S. L. Partridge, “The current status of dilicon on insulator technologies - A

comparison,” IEDM. Dig., pp. 428–430, 1986.

[70] H. M. Manasevit, I. Golecki, L. A. Moudy, J. J. Yang, and J. E. Mee, “Si on cubic

zirconia,” J. Electrochem. Soc. Solid-state Sci. Tecnol., vol. 130, pp. 1752–1758,

1983.

[71] S. S. Lau, S. Matteson, J. W. Mayer, P. Revesz, J. Gyulai, J. Roth, T. W. Sigmon,

and T. Cass, “Improvement of crystalline quality of epitaxial Si layers by ion-

Page 133: Passive Wireless Radiation Dosimeters for Radiation

116

implantation techniques,” Appl. Phys. Lett., vol. 34, no. 1, pp. 76–78, 1979.

[72] G. K. Celler, H. J. Leamy, L. E. Trimble, and T. T. Sheng, “Annular grain structures

in pulsed laser recrystallized Si on amorphous insulators,” Appl. Phys. Lett., vol. 39,

no. 5, pp. 425–427, 1981.

[73] J. A. Knapp and S. T. Picraux, “Growth of Si on insulator using electron beams,” J.

Cryst. Growth, vol. 63, pp. 445–452, 1983.

[74] B. Tillack, K. Hoeppner, H. H. Richter, and R. Banisch, “Seeding recrystallization

for producing thick silicon-on-insulator films on non-planar substrates,” Mater. Sci.

Eng., vol. B4, pp. 237–241, 1989.

[75] K. Imai, “A new dielectric isolation method using porous silicon,” Solid. State.

Electron., vol. 24, no. 2, pp. 159–164, 1981.

[76] G. K. Celler, P. L. F. Hemment, K. W. West, and J. M. Gibson, “High quality Si-

on-SiO2 films by large dose oxygen implantation and lamp annealing,” Appl. Phys.

Lett., vol. 48, no. 8, p. 532, 1986.

[77] W. P. Maszara, G. Goetz, A. Caviglia, and J. B. McKitterick, “Bonding of silicon

wafers for silicon-on-insulator,” J. Appl. Phys., vol. 64, no. 10, pp. 4943–4950, 1988.

[78] T. Yonehara, K. Sakaguchi, and N. Sato, “Epitaxial layer transfer by bond and etch

back of porous Si,” Appl. Phys. Lett., vol. 64, no. 16, pp. 2108–2110, 1994.

[79] M. Bruel, “Silicon on insulator material technology,” Electron. Lett., vol. 31, pp.

1201–1202, 1995.

[80] M. Bruel, “Application of hydrogen ion beams to silicon on insulator material

technology,” Nucl. Inst. Methods Phys. Res. B, vol. 108, pp. 313–319, 1996.

[81] A. Chaudhry and M. J. Kumar, “Controlling short channel effects in deep-submicron

SOI MOSFETs for improved reliability: A review,” IEEE Trans. Device Mater.

Reliab., vol. 4, no. 1, pp. 99–109, 2004.

Page 134: Passive Wireless Radiation Dosimeters for Radiation

117

[82] L. T. Su, J. B. Jacobs, E. James, and D. A. Antoniadis, “Deep-submicrometer

channel design in silicon-on-insulator (SOI) MOSFET’s,” IEEE Electron Device

Lett., vol. 15, no. 9, pp. 183–185, 1994.

[83] K. K. Young, “Short-Channel effect in fully depleted SOI MOSFET’s,” IEEE Trans.

Electron Devices, vol. 36, no. 2, pp. 399–402, 1989.

[84] Y. Taur and T. H. Ning, Fundermentals of modern VLSI devices, 2nd ed. Cambridge,

2013.

[85] C. Mazuré, R. Ferrant, B. Y. Nguyen, W. Schwarzenbach, and C. Moulin, “FDSOI:

From substrate to devices and circuit applications,” ESSCIRC 2010 - 36th Eur. Solid

State Circuits Conf., pp. 45–51, 2010.

[86] C. Mazure and B.-Y. Nguyen, “Overview of fully depleted SOI technology and

circuit benefits,” ECS Trans., vol. 49, no. 1, pp. 41–48, 2012.

[87] J. R. Schwank and M. R. Shaneyfelt, “Radiation Effects in SOI Technologies,” vol.

50, no. 3, pp. 522–538, 2003.

[88] F. E. Mamouni, S. K. Dixit, R. D. Schrimpf, P. C. Adell, I. S. Esqueda, M. L. Mclain,

H. J. Barnaby, S. Cristoloveanu, and W. Xiong, “Gate-length and drain-bias

dependence of band-to-band tunneling-induced drain leakage in irradiated fully

depleted SOI devices,” IEEE Trans. Nucl. Sci., vol. 55, no. 6, pp. 3259–3264, 2008.

[89] H.-K. Lim and J. G. Fossum, “Threshold Voltage of Thin-Film Silicon-on-Insulator

(Sol) MOSFET‘s,” IEEE Trans. Electron Devices, vol. ED-30, no. 10, pp. 1244–

1251, 1983.

[90] D. A. Neamen, Semiconductor physics and devices : basic principles. McGraw-Hill,

2012.

[91] M. L. McLain, H. J. Barnaby, and P. C. Adell, “Analytical model of radiation

response in FDSOI MOSFETS,” 2008 IEEE Int. Reliab. Phys. Symp., pp. 643–644,

Apr. 2008.

Page 135: Passive Wireless Radiation Dosimeters for Radiation

118

[92] P. C. Adell, H. J. Barnaby, S. Member, R. D. Schrimpf, and B. Vermeire, “Band-to-

Band Tunneling ( BBT ) Induced Leakage Current Enhancement in Irradiated Fully

Depleted SOI Devices,” IEEE Trans. Nucl. Sci., vol. 54, no. 6, pp. 2174–2180, 2007.

[93] C.-E. Chen, M. Matloubian, R. Sundaresan, B.-Y. Mao, C. C. Wei, and G. Pollack,

“Single transistor latch in SOI MOSFETs,” IEEE Trans. Nucl. Sci., vol. 9, no. 12,

pp. 636–638, 1988.

[94] S. Quoizola, J. L. Leray, J. L. Pelloie, and C. Raynaud, “Total dose Induce in short

channel NMOS / SOI transistors,” IEEE Trans. Nucl. Sci., vol. 45, no. 6, pp. 2458–

2466, 1998.

[95] R. Nopper, R. Niekrawietz, L. Reindl, and A. A. Model, “Wireless Readout of

Passive LC Sensors,” IEEE Trans. Instrum. Meas., vol. 59, no. 9, pp. 2450–2457,

2010.

[96] G. Sarrabayrouse and S. Siskos, “Radiation dose measurement using MOSFETs,”

IEEE instrumentation&measurement, no. June, pp. 26–34, 1998.

[97] N. H. W. Fong, J.-O. Plouchart, N. Zamdmer, L. F. Wagner, C. Plett, and N. G. Tarr,

“A 1-V 3.8-5.7-GHz wide-band VCO with differentially tuned accumulation MOS

varactors for common-mode noise rejection in CMOS SOI technology,” IEEE Trans.

Microw. Theory Tech., vol. 51, no. 8, pp. 1952–1959, Aug. 2003.

[98] J. O. Plouchart, N. Zamdmer, J. Kim, R. Trzcinski, S. Narasimha, M. Khare, L. F.

Wagner, S. L. Sweeney, and S. Chaloux, “A 243-GHz Ft and 208-GHz Fmax, 90-

nm SOI CMOS SoC technology with low-power mm wave digital and RF circuit

capability,” IEEE Trans. Electron Devices, vol. 52, no. 7, pp. 1370–1375, 2005.

[99] M. R. Shaneyfelt, T. A. Hill, T. M. Gurrieri, J. R. Schwank, R. S. Flores, P. E. Dodd,

and S. M. Dalton, “An Embeddable SOI Radiation Sensor,” IEEE Trans. Nucl. Sci.,

vol. 56, no. 6, pp. 3372–3380, 2009.

[100] J. Yau, M. S. Gordon, K. P. Rodbell, S. J. Koester, P. W. Dehaven, and W. E.

Haensch, “FDSOI Radiation Dosimeters,” in 2011 International Symposium on

Page 136: Passive Wireless Radiation Dosimeters for Radiation

119

VLSI Technology, Systems and Applications, 2011.

[101] G. Vaidhyanathan and S. J. Koester, “High-Q FDSOI Varactors for Wireless

Radiation Sensing,” in IEEE International SOI Conference, 2011.

[102] P. Mayles, A. E. Nahum, and J.-C. Rosenwald, Handbook of radiotherapy physics :

theory and practice. Taylor & Francis, 2007.

[103] T. H. Ning, “Capture cross section and trap concentration of holes in silicon dioxide,”

J. Appl. Phys., vol. 47, no. 3, pp. 1079–1081, 1976.

[104] J. R. Schwank, M. R. Shaneyfelt, P. E. Dodd, A. Burns, and P. W. Wyatt, “New

Insights into Fully-Depleted SO1 Transistor Response After Total-Dose Irradiation,”

IEEE Trans. Nucl. Sci., pp. 299–307, 2000.

[105] K. Sato, M. Shikida, Y. Matsushima, T. Yamashiro, K. Asaumi, Y. Iriye, and M.

Yamamoto, “Characterization of orientation-dependent etching properties of single-

crystal silicon: effects of KOH concentration,” Sensors Actuators A Phys., vol. 64,

no. 1, pp. 87–93, 1998.

[106] H. F. Winters and J. W. Coburn, “The etching of silicon with XeF2 vapor,” Appl.

Phys. Lett., vol. 34, no. 1, pp. 70–73, 1979.

[107] D. E. Aspnes and A. A. Studna, “High precision scanning ellipsometer.,” Appl. Opt.,

vol. 14, no. 1, pp. 220–228, 1975.

[108] G. Huang, Z. Xi, and D. Yang, “Crystallization of amorphous silicon thin films: The

effect of rapid thermal processing pretreatment,” Vacuum, vol. 80, no. 5, pp. 415–

420, 2006.

[109] K. Cho, W. R. Allen, T. G. Finstad, W. K. Chu, J. Liu, and J. J. Wortman,

“Channeling effect for low energy ion implantation in Si,” Nucl. Instruments

Methods Phys. Res. Sect. B Beam Interact. with Mater. Atoms, vol. B7, pp. 265–272,

1985.

[110] S. Campbell, Fabrication engineering at the micro and nanoscale. Oxford

Page 137: Passive Wireless Radiation Dosimeters for Radiation

120

University Press, 2013.

[111] C. P. Ho, J. D. Plummer, J. D. Mdindl, and M. Sklar, “Thermal oxidation of heavily

phosphorus-doped silicon,” J. Electrochem. Soc., vol. 125, no. 4, pp. 665–671, 1978.

[112] Z. Krivokapic, W. Maszara, F. Arasnia, E. Paton, Y. Kim, L. Washington, E. Zhao,

J. Chan, J. Zhang, A. Marathe, and M.-R. Lin, “High performance 25nm FDSOI

devices with extremely thin silicon channel,” Symp. VLSI Technol. Dig. Tech. Pap.,

pp. 131–132, 2003.

[113] N. S. Saks, M. G. Ancona, and J. A. Modolo, “Radiation effects in MOS capacitors

with very thin oxides at 80 degK,” IEEE Trans. Nucl. Sci., vol. 31, no. 6, pp. 1249–

1255, 1984.

[114] Y. Li, W. M. Porter, C. Kshirsagar, I. Roth, Y. Su, M. a. Reynolds, B. J. Gerbi, and

S. J. Koester, “Fully-Depleted Silicon-on-Insulator Devices for Radiation

Dosimetry in Cancer Therapy,” IEEE Trans. Nucl. Sci., vol. 61, no. 6, pp. 3443–

3450, 2014.

[115] D. M. Fleetwood and S. J. H, “Evidence that similar point defects cause 1/f noise

and radiation-induced-hole trapping in metal-oxide-semiconductor transistors,”

Phys. Rev. Lett., vol. 64, no. 5, pp. 579–582, 1990.

[116] D. M. Fleetwood, L. C. Riewe, J. R. Schwank, S. N. Laboratories, S. C. Witczak,

and R. D. Schrimpf, “RADIATION EFFECTS AT LOW ELECTRIC FIELDS IN

THERMAL, SIMOX, AND BIPOLAR-BASE OXIDES," IEEE Trans. Nucl. Sci.,

vol. 43, no. 6, pp. 2537–2546, 1996.

[117] D. M. Fleetwood, S. L. Miller, R. A. Reber, P. J. McWhorter, P. S. Winokur, M. R.

Shaneyfelt, and J. R. Schwank, “New insights into radiation-induced oxide-trap

charge through thermally-stimulated-current measurement and analysis,” IEEE

Trans. Nucl. Sci., vol. 39, no. 6, pp. 2192–2203, 1992.

[118] H. J. Barnaby, M. L. McLain, I. S. Esqueda, and X. J. Chen, “Modeling ionizing

radiation effects in solid state materials and CMOS devices,” IEEE Trans. Circuits

Page 138: Passive Wireless Radiation Dosimeters for Radiation

121

Syst. I Regul. Pap., vol. 56, no. 8, pp. 1870–1883, 2009.

[119] B. O. Connell, C. Mccarthy, B. Lane, and A. Mohammadzadeh, “Optimised stacked

RADFETs for micro gray dose measurement,” Radiat. Its Eff. Components Syst., pp.

101–105, 2000.

[120] D. M. Fleetwood, P. S. Winokur, R. a. Reber, T. L. Meisenheimer, J. R. Schwank,

M. R. Shaneyfelt, and L. C. Riewe, “Effects of oxide traps, interface traps, and

‘border traps’ on metal-oxide-semiconductor devices,” J. Appl. Phys., vol. 73, no.

1993, pp. 5058–5074, 1993.

[121] T. Poiroux, O. Rozeau, S. Martinie, P. Scheer, S. Puget, M. A. Jaud, S. El Ghouli, J.

C. Barbé, A. Juge, O. Faynot, and M. Campus, “UTSOI2 : A Complete Physical

Compact Model for UTBB and Independent Double Gate MOSFETs,” Electron

Devices Meet. (IEDM), 2013 IEEE Int., p. 12.4.1-12.4.4, 2013.

[122] Y. Todokoro, “Double-layer resist films for submicrometer electronbeam

lithography,” IEEE J. Solid-State Circuits, vol. 15, no. 4, pp. 508–513, 1980.

[123] “Ultra-small solid state battery.” [Online]. Available:

http://www.cymbet.com/products/enerchip-solid-state-batteries.php.

[124] R. D. Schrimpf, K. M. Warren, R. A. Weller, R. A. Reed, L. W. Massengill, M. L.

Alles, D. M. Fleetwood, X. J. Zhou, L. Tsetseris, and S. T. Pantelides, “Reliability

and radiation effects in IC technologies,” in IEEE International Reliability Physics

Symposium, 2008, pp. 97–106.

[125] T. Cheung, M. J. Butson, and P. K. N. Yu, “Effects of temperature variation on

MOSFET dosimetry.,” Phys. Med. Biol., vol. 49, pp. 191–196, 2004.

[126] V. Huard, M. Denais, F. Perrier, N. Revil, C. Parthasarathy, A. Bravaix, and E.

Vincent, “A thorough investigation of MOSFETs NBTI degradation,”

Microelectron. Reliab., vol. 45, no. 1, pp. 83–98, 2005.

[127] A. J. Lelis, T. R. Oldham, H. E. Boesch, and F. B. McLean, “Nature of the trapped

Page 139: Passive Wireless Radiation Dosimeters for Radiation

122

hole annealing process,” IEEE Trans. Nucl. Sci., vol. 36, no. 6, pp. 1808–1815, 1989.

[128] J. M. Benedetto, H. E. Boesch, F. B. McLean, and J. P. Mize, “Hole removal in thin-

gate MOSFETs by tunneling,” IEEE Trans. Nucl. Sci., vol. 32, no. 6, pp. 3916–3920,

1985.

[129] L. T. Niklason, M. V. Marx, and H.-P. Chan, “Interventional radiologists:

occupational radiation doses and risks,” Radiology, vol. 187, no. 3, pp. 729–733,

1993.

[130] K. S. Novoselov, A. K. Geim, S. V Morozov, D. Jiang, Y. Zhang, S. V Dubonos, I.

V Grigorieva, and A. A. Firsov, “Electric field effect in atomically thin carbon films,”

Science (80-. )., vol. 306, pp. 666–668, 2004.

[131] A. K. Geim and K. S. Novoselov, “The rise of graphene.,” Nat. Mater., vol. 6, pp.

183–191, 2007.

[132] K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, “Atomically thin MoS2: A

new direct-gap semiconductor,” Phys. Rev. Lett., vol. 105, no. 13, pp. 2–5, 2010.

[133] W. Zhao, Z. Ghorannevis, L. Chu, M. Toh, C. Kloc, P.-H. Tan, and G. Eda,

“Evolution of electronic structure in atomically thin sheets of WS2 and WSe 2,”

ACS Nano, vol. 7, no. 1, pp. 791–797, 2013.

[134] A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim,

“The electronic properties of graphene,” Rev. Mod. Phys., vol. 81, no. 1, pp. 109–

162, 2009.

[135] R. Coehoorn, C. Haas, J. Dijkstra, C. Flipse, R. de Groot, and A. Wold, “Electronic

structure of MoSe$_{2}$, MoS$_{2}$, and WSe$_{2}$. I. Band-structure

calculations and photoelectron spectroscopy,” Phys. Rev. B, vol. 35, no. 12, pp.

6195–6202, 1987.

[136] J. S. Moon, D. Curtis, M. Hu, D. Wong, C. McGuire, P. M. Campbell, G. Jernigan,

J. L. Tedesco, B. VanMil, R. Myers-Ward, C. Eddy, and D. K. Gaskill, “Epitaxial-

Page 140: Passive Wireless Radiation Dosimeters for Radiation

123

graphene RF field-effect transistors on Si-face 6H-SiC substrates,” IEEE Electron

Device Lett., vol. 30, no. 6, pp. 650–652, 2009.

[137] Y. Wu, K. A. Jenkins, A. Valdes-garcia, D. B. Farmer, Y. Zhu, A. A. Bol, C.

Dimitrakopoulos, W. Zhu, F. Xia, P. Avouris, and Y. Lin, “State-of-the-art graphene

high-frequency electronics,” Nano Lett., vol. 12, pp. 3062–3067, 2012.

[138] J. A. Rogers, T. Someya, and Y. Huang, “Materials and mechanics for stretchable

electronics,” Science (80-. )., vol. 327, pp. 1603–1607, 2010.

[139] E. J. Olson, R. Ma, T. Sun, M. A. Ebrish, N. Haratipour, K. Min, N. R. Aluru, and

S. J. Koester, “Capacitive sensing of intercalated molecules using graphene,” ACS

Appl. Mater. Interfaces, vol. 7, pp. 25804–25812, 2015.

[140] Y. H. Lee, X. Q. Zhang, W. Zhang, M. T. Chang, C. Te Lin, K. Di Chang, Y. C. Yu,

J. T. W. Wang, C. S. Chang, L. J. Li, and T. W. Lin, “Synthesis of large-area MoS2

atomic layers with chemical vapor deposition,” Adv. Mater., vol. 24, no. 17, pp.

2320–2325, 2012.

[141] R. Bez, E. Camerlenghi, A. Modelli, and A. Visconti, “Introduction to flash

memory,” Proc. IEEE, vol. 91, no. 4, pp. 1–16, 2009.

Page 141: Passive Wireless Radiation Dosimeters for Radiation

124

Appendix A:

FDSOI MOSFET fabrication process flow

I. Thinning the initial SOI layer. 1. Clean the wafer through RCA process.

2. Oxidation: Tylan Tube 14, recipe: Dry 1000, time: 2 hours and 50 mins, get

about ~950A oxide growth.

3. BOE etches away the oxide. About 35~40nm Si remains.

II. Grow screening oxide. 1. Right after thinning the silicon, immediately process this step

2. Oxidation: Tylan Tube 14, recipe: Gox850-A, time: 1 min, get about 6nm oxide.

About 34nm Si remains.

III. Pattern MESA. 1. Clean the sample and spin coat resist 1813 at 4000 rpm for 30s. Soft bake for

1min.

2. Pattern MESA at ma6, 6s exposure.

3. Develop for 30s.

4. STS dry etching. Recipe: Pjsnitd 1. Time: 40s.

5. Clean off the resist

IV. Implant 1. Clean the sample and spin coat resist 1813 at 4000 rpm for 30s. Soft bake for

1min.

2. Pattern implant layer at ma6, 6s exposure.

3. Image reversal

4. Oriel 6min+6min

5. Develop for 2mins

6. Implant with As 2.33e15/cm2, 9Kev

V. Activation and gate oxide growth 1. Clean the sample before activation

a. Soak at acetone overnight

b. Sonicate with 1165 at 70c for 15mins

c. STS O2 clean for 2 mins

2. Before activation, clean the sample again with H2SiO4+H2O2 for 15mins at 120c.

3. Activation at tylan tube 15, recipe: anneal 850 for 2mins (NO RCA clean!)

4. Right after activation, etch away screen oxide with BOE for 15s.

5. Oxidation. Tylan tube 14, recipe: Gox850-A , time: 2min. (Be aware the

implanted SD grows faster than gate).

6. Forming gas annealing at 425c for 30mins, H2+N2 (Tube 16 Alloy425)

Page 142: Passive Wireless Radiation Dosimeters for Radiation

125

VI. Gate metal deposition 1. Clean the sample and spin coat resist 1813 at 4000 rpm for 30s. Soft bake for

1min.

2. Pattern implant layer at ma6, 6s exposure.

3. Image reversal.

4. Oriel 6min+6min

5. Develop for 2mins

6. Ozone clean for 1min

7. Temescal deposit 10 nm Ti + 190nm Al

8. Liftoff

VII. SD metal deposition 1. Clean the sample and spin coat resist 1813 at 4000 rpm for 30s. Soft bake for

1min.

2. Pattern implant layer at ma6, 6s exposure.

3. Image reversal.

4. Oriel 6min+6min

5. Develop for 2mins

6. Ozone clean for 1min

7. Etch away SD oxide in BOE for 20s (Be aware the oxide there is thicker than

gate oxide)

8. Temescal deposit 10 nm Ti + 190nm Al

9. Liftoff

Page 143: Passive Wireless Radiation Dosimeters for Radiation

126

Appendix B:

EBeam FDSOI varactor fabrication process:

Start with 100 nm Si + 200 nm BOX SOI wafer

I. Thinning the SOI layer. 1. Clean the wafer through RCA process.

2. Oxidation: Tylan Tube 14, recipe: Dry 1000, time: 2 hours and 50 mins, get

about ~950A oxide growth.

3. BOE etches away the oxide. About 35~40nm Si remains.

II. Grow screening oxide. 1. Right after thinning the silicon, immediately process this step.

2. Oxidation: Tylan Tube 14, recipe: Gox850-A, time: 30 seconds, get about 6nm

oxide. About 34nm Si remains.

III. Create alignment mark

1. Spin PMMA 950 C6 3000 rpm 1 min, slow ramp. End with about 800 nm

thickness.

2. Pattern with vistec. Beam step size: 4 nm, current 5 nA , dose 800 uC/cm2

3. Develop with 3:1 MIBK for 90s.

4. Etching:

i. BOE dip for 15s to remove screen oxide.

ii. Deep trench CNF-14-V 30s to remove SOI layer, about 30 nm

iii. Oxford, o-SiO2-etch-RIE-mode, 3mins to etch BOX (Check afterward)

iv. Deep trench CNF-14-V 7min to etch bottom Si layer. This will give ~1

um deep alignment mark and etch away ~550nm resist.

IV. Pattern MESA: 1. Spin ma-N2403 resist at 3000rpm-slow, get about 250 nm thick film.

2. Pattern with vistec, beam step: 30nm, current 100 nA, Dose: 400 uC/cm2

3. Develp

4. Bake at 100c for 15 mins.

5. BOE dip 15s to tech away screen oxide.

6. STS slowpoly for 20s to etch away Si layer and stop at BOX. (For bare Si no

native oxide 20s slowpoly etches 110nm Si)

7. Remove resist, acetone, 1165, O2 clean.

V. Implantation 1. Spin PMMA 950 C4, 3000rpm slow ramp, end with 350nm thick film

2. Pattern with vistec, beam step size: 5nm, current 10nA, dose 1100 uC/cm2

3. Develop: 3:1 MIBK for 90s

4. Implant with As 2.33e15/cm2, 9Kev

Page 144: Passive Wireless Radiation Dosimeters for Radiation

127

VI. Activation and gate oxide growth 1. Clean the sample before activation

i. Soak at acetone overnight

ii. Sonicate until all resist on figure patterns are gone

iii. STS O2 clean for 2 mins

2. Before activation, clean the sample again with H2SiO4+H2O2 for 15mins at 120c.

3. Activation at tylan tube 15, recipe: anneal 850 for 2mins (NO RCA clean!)

4. Right after activation, etch away screen oxide with BOE for 15s.

5. Oxidation. Tylan tube 14, recipe: Gox850-A , time: 2min. (Be aware the

implanted SD grows faster than gate).

6. Forming gas annealing at 425c for 30mins, H2+N2 (Tube 16 Alloy425)

VII. Gate contacts deposition 1. Spin P(MMA/MAA) EL9 using recipe PMMA3000rpm-slow ramp

2. Bake at 150C for 2 mins

3. Spin PMMA 950 C2 using the same recipe

4. Bake at 180c for 8min (in total: 400nm P(MMA/MAA) and 200nm PMMA)

5. Pattern with vistec, beam step size: 3nm, current 2nA, dose 590 uC/cm2

6. Develop: 3:1 MIBK for 90s

7. Ozone clean for 1min

8. Deposit 10nm Pd + 10nm Ti + 180 nm Al at CHA.

9. Liftoff in Acetone overnight

VIII. S/D contacts deposition 1. Spin P(MMA/MAA) EL9 using recipe PMMA3000rpm-slow ramp

2. Bake at 150C for 2 mins

3. Spin PMMA 950 C2 using the same recipe

4. Bake at 180c for 8min (in total: 400nm P(MMA/MAA) and 200nm PMMA)

5. Pattern with vistec, beam step size: 3nm, current 2nA, dose 630 uC/cm2

6. Develop: 3:1 MIBK for 90s

7. BOE dip for 15s to remove oxide at S/D region

8. Deposit 10 nm Ti + 190 nm Al at CHA

9. Liftoff in Acetone overnight

Note:

1. P(MMA/MAA) PMMA double layer resist process is sensitive to previous steps, so the

optimal dose may vary at each fabrication.

2. Optimal dose for ebeam depends on smallest feature size and existing patterns. If the

smallest feature and space is smaller, optimal dose may change and should be retested.

(Specially for Gate and S/D layers).

3. If lifting off small features, 1165 can be used instead of acetone, temperature 100C.

Page 145: Passive Wireless Radiation Dosimeters for Radiation

128

Layer Resist Thickness Liftoff Smallest

Feature/space

Beam step

size/current

Optimal

Dose

(uC/cm2)

Align

Mark

PMMA

950 C6

~800 nm No 10um/large 4nm/5nA 800

Mesa Ma-

N2403

No 2um/large 30nm/100nA 400

Implant PMMA

950 C4

~350 nm No 800nm/300nm 5nm/10nA 1100

Gate MMA

EL9/

PMMA

950 C2

~400nm +

200nm

Yes 1um/3um 3nm/2nA 5zo

Contact MMA

EL9/

PMMA

950 C2

~400nm +

200nm

Yes 1um/3um 3nm/2nA 630