dynamic modulation of material properties by solid state

221
Dynamic Modulation of Material Properties by Solid State Proton Gating by Aik Jun Tan Submitted to the Department of Materials Science and Engineering in partial fulfilment of the requirements for the degree of Doctor of Philosophy in Materials Science and Engineering at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY June 2019 © Massachusetts Institute of Technology 2019. All rights reserved Author…………………………………………………………………………………………........ Department of Materials Science and Engineering May 8th,2019 Certified by……………………………………………………………………………………........ Geoffrey S. D. Beach Professor of Materials Science and Engineering Thesis Supervisor Accepted by…...………………………………………………………………………………........ Donald R. Sadoway Chairman, Department Committee on Graduate Studies

Upload: others

Post on 16-May-2022

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Dynamic Modulation of Material Properties by Solid State

Dynamic Modulation of Material Properties by Solid

State Proton Gating

by Aik Jun Tan

Submitted to the Department of Materials Science and Engineering in partial fulfilment of the

requirements for the degree of

Doctor of Philosophy in Materials Science and Engineering

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

June 2019

© Massachusetts Institute of Technology 2019. All rights reserved

Author…………………………………………………………………………………………........

Department of Materials Science and Engineering

May 8th,2019

Certified by……………………………………………………………………………………........

Geoffrey S. D. Beach

Professor of Materials Science and Engineering

Thesis Supervisor

Accepted by…...………………………………………………………………………………........

Donald R. Sadoway

Chairman, Department Committee on Graduate Studies

Page 2: Dynamic Modulation of Material Properties by Solid State

Dynamic Modulation of Material Properties by Solid

State Proton Gating

by Aik Jun Tan

Submitted to the Department of Materials Science and Engineering on May 8th, 2019 in partial

fulfilment of the requirements for the degree of

Doctor of Philosophy in Materials Science and Engineering

Abstract

As functionalities become more abundant in solid state devices, one key capability which

remains lacking is an effective means to dynamically tune material properties. In this thesis, we

establish a pathway towards this capability by utilizing the simplest ion known to mankind: the

proton. We demonstrate for the first time dynamic control of magnetic properties in an all-solid-

state heterostructures using solid state proton gating in a metal/oxide heterostructure. We also

demonstrate dynamic modulation of magnetic anisotropy at a metal-metal interface through

hydrogen insertion in a heavy metal adjacent to a ferromagnet. Besides magnetic properties,

solid state proton gating also enables dynamic modulation of optical properties in a thin film

oxide. We observe fast gating of optical reflectivity by ~10% at timescale down to ~20ms in a

metal/oxide/metal heterostructure. Finally, we also demonstrate a room temperature reversible

solid oxide fuel cell based on hydrogen storage. The cell has a small form factor which is

suitable for energy storage in solid state microelectronics application. Our work hence provides a

platform for complete control of material properties through solid state proton gating.

Thesis Supervisor: Geoffrey S.D. Beach

Title: Professor of Materials Science and Engineering

Page 3: Dynamic Modulation of Material Properties by Solid State
Page 4: Dynamic Modulation of Material Properties by Solid State

Acknowledgements

First and foremost, I would like to thank my advisor, Prof. Geoffrey Beach. From him, I learned

how to properly conduct a research, and how to effectively present the findings of the research.

He sets a good example in terms of his hardwork and his relentless pursuit of perfection. And he

gave me a lot of autonomy when it comes to my research. He is exactly the advisor I have hoped

for and I am grateful to him.

I would also like to thank my thesis committee Prof. Harry Tuller and Prof. Eugene Fitzgerald.

Prof. Tuller has been very kind to me, and I have learned so much about solid state ionics from

him. Prof. Fitzgerald has given me very useful inputs on my thesis.

I would like to thank all my teammates in the Beach group: Max, Lucas, Mantao, Felix, Ivan,

Sara, Jason, Can, Kohei, Uwe, Satoru, Liz, Parnika, Minae, Shwoo, Chi Feng, Usama, Daniel,

and Siying. They are wonderful people to work with, and my life in MIT was made colorful by

them.

I would like to thank the staff in DMSE and MRL who have made this thesis possible: David

Bono, Charlie, Libby, Mike, Tara, Angelita, Elissa, Jessie, Dominique, John, and Jessie. They

are some of the nicest people I know, and they have given me help whenever asked.

Finally, I would like to thank my family. My parents are a source of wisdom and strength, and

they are the most important people in my life. My elder brother, who always looks after his

younger siblings, is someone I can always have a frank conversation with. And my younger

sister, who is the most intelligent among us, is someone I can always poke fun at.

Page 5: Dynamic Modulation of Material Properties by Solid State

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

1.1 Motivation ............................................................................................................................. 2

1.2 Thesis Outline ....................................................................................................................... 3

Chapter 2: Background ................................................................................................................... 6

2.0 Magnetic Hysteresis Loop .................................................................................................... 7

2.1 Magnetic Anisotropy ............................................................................................................ 9

2.2 Magnetization Dynamics and Spin Current ........................................................................ 20

2.3 Magneto-Electric and Magneto-Ionic Effects ..................................................................... 24

2.4 Solid Oxide Fuel and Electrolyzer cell ............................................................................... 31

2.5 Solid Oxide Proton Conductors .......................................................................................... 36

2.6 Water Electro-Catalysis ...................................................................................................... 44

2.7 Electrodes for Solid Oxide Cells......................................................................................... 47

2.8 Electrochemical Impedance Spectroscopy ......................................................................... 51

Chapter 3: Experimental Methods ................................................................................................ 54

3.1 Sputter Deposition .............................................................................................................. 55

3.2 Sample Structure and Patterning ......................................................................................... 61

3.3 Magneto-Optical Kerr Effect .............................................................................................. 67

3.4 Anomalous and Planar Hall Effect ..................................................................................... 72

3.5 Time Resolved Hall Magnetometry under different Atmospheric Conditions ................... 74

3.6 Spin-torque Ferromagnetic Resonance ............................................................................... 77

3.7 Solid Oxide Cell Characterization ...................................................................................... 79

Chapter 4: Effect of H2O on Voltage-induced Co Oxidation in a Pt/Co/GdOx Heterostructure .. 81

4.1: Experimental Methods ....................................................................................................... 84

4.2: Probing Water Uptake in GdOx ......................................................................................... 86

4.3: Voltage-induced Co Oxidation in Hydrated and Non-hydrated Pt/Co/GdOx Devices ...... 90

4.4: H2 evolution during Voltage-induced Co Oxidation in Pt/Co/GdOx ................................. 94

4.5: In-situ XAS probe of Co during Voltage-induced Co Oxidation in Pt/Co/GdOx ............. 98

Chapter 5: Magneto-ionic Control of Magnetism using a Solid-state Proton Pump .................. 100

5.1: Experimental Methods ..................................................................................................... 103

5.2: Co Redox through Water Electrolysis ............................................................................. 105

5.3: Modulation of Magnetic Anisotropy through Proton Injection ....................................... 112

Page 6: Dynamic Modulation of Material Properties by Solid State

5.4: Magnetic Response under Short Circuit and Open Circuit.............................................. 117

5.5: Electrical Gating of Magnetic Anisotropy at a Heavy-metal/ferromagnet Interface ....... 120

5.6 Comparison between Au and Pt Top Electrodes .............................................................. 124

Chapter 6: Voltage Gating of Magnetic Damping and Spin-Orbit Torques using Proton .......... 126

6.1 Experimental Methods ...................................................................................................... 129

6.2 Spin Torque Ferromagnetic Resonance to Probe Voltage Gating of Spin Orbit Torque and

Magnetic Damping.................................................................................................................. 131

Chapter 7: Voltage-induced Magneto-Ionic Effect in Pt/Co/MOx Heterostructure (M= Gd, Y, Zr,

and Ta) ........................................................................................................................................ 138

7.1 Experimental Methods ...................................................................................................... 141

7.2 Rate of Voltage-Induced Magnetic Modulation at Positive Bias ..................................... 142

7.3 Rate of Voltage-Induced Magnetic Modulation at Negative Bias .................................... 145

Chapter 8: Room Temperature Reversible Solid Oxide Fuel Cell ............................................. 148

8.1 Experimental Methods ...................................................................................................... 150

8.2 Proton Conductivity of GdOx............................................................................................ 152

8.3 Cell Performance and Scalability...................................................................................... 154

8.4 Gating of Magnetism using Built-in Voltage .................................................................... 158

Chapter 9: Voltage Gating of Optical Properties ........................................................................ 161

9.1 Experimental Methods ...................................................................................................... 163

9.2 Voltage Gating of Optical Reflectivity in Pt/GdOx/Au Heterostructure .......................... 164

9.3 Source of Irreversible Optical Change in Pt/GdOx/Au Heterostructure ........................... 169

9.4 Voltage Gating of GdOx Heterostructures with different Top and Bottom Electrodes .... 172

9.5 Optical Modulation outside Active Region due to Hydrogen Diffusion .......................... 175

Chapter 10: Electrical Properties of GdOx .................................................................................. 178

Chapter 11: Summary and Outlook ............................................................................................ 186

11.1 Summary ......................................................................................................................... 187

11.2 Outlook ........................................................................................................................... 189

11.2.1 Integration of Hydrogen Storage in Proton Magneto-Ionic Device ......................... 189

11.2.2 Proton Magneto-Ionics for Memory and Logic Devices ......................................... 190

11.2.3 Proton Magneto-Ionics to Quantify Proton Conductivity in Thin Film Oxides ...... 193

References ................................................................................................................................... 194

Page 7: Dynamic Modulation of Material Properties by Solid State

1

Chapter 1:

Introduction

Page 8: Dynamic Modulation of Material Properties by Solid State

2

1.1 Motivation

As solid state devices continue its path towards miniaturization, there are two important trends to

note: (1) interfaces play an increasingly important role in determining material properties1, and

(2) the electric field, which is the primary tool by which we control device functions, becomes

progressively larger. Simultaneously, the need to dynamically toggle material properties have

become more crucial due to limited functionalities and performance imposed by static devices.

For instance, in magnetic memory, it is extremely difficult to achieve both thermal stability and

low writing power simultaneously because thermal stability implies large energy barrier to

magnetic switching, while low writing power implies low energy barrier to magnetic

switching2,3. These are two conflicting requirements which are very difficult to optimize in a

static device.

In this thesis, we take advantage of functional interfaces and large electric field in nanoscale

devices in order to provide a mechanism by which we can dynamically induce large changes in

material properties. We demonstrate dynamic modulation of material properties in thin film

devices through solid state proton gating. Proton is used because it has high mobility at room

temperature which allows for fast device operation. At the same time, it can induce very large

changes in device properties because it disrupts chemical bonding at functional interfaces. In this

sense, it captures the best of both worlds in terms of speed of electronic modulation and

magnitude of ionic modulation. The term “magneto-ionic” is used to refer to magnetic-ionic

coupling where ions are used to modulate magnetic properties.

Page 9: Dynamic Modulation of Material Properties by Solid State

3

1.2 Thesis Outline

This thesis is written with the aim to enable a wide audience with little knowledge of magnetism,

ionics, or electrochemistry to interpret the new findings.

In chapter 1, we give a preliminary introduction to dynamic modulation of material properties

using solid state proton gating.

In chapter 2, we provide some scientific background which is necessary to understanding the

results presented in this thesis. The chapter starts off with introduction to magnetism since a

large part of the thesis is focused on modulation of magnetic properties. The chapter eventually

transitions into introduction to proton conductors and water electro-catalysis because we source

protons from water and transport them through an oxide electrolyte to modulate material

properties.

In chapter 3, we provide a detailed overview of the primary experimental techniques used to

perform the research in this thesis. The chapter starts off with description of the sample

structures and fabrication steps. This is followed by characterization techniques used to probe the

magnetic and electrochemical properties of the samples.

In chapter 4, we present the first experimental evidence that voltage-induced Co oxidation in a

Pt/Co/GdOx/Au heterostructure is dominated by water oxidation instead of oxidation by oxygen

ions. The findings represent an important breakthrough in understanding of voltage induced

redox in ferromagnet/oxide systems and show that water can play a crucial role in previously

observed magneto-ionic effect.

Page 10: Dynamic Modulation of Material Properties by Solid State

4

In chapter 5, we show for the first time solid state gating of magnetic anisotropy using proton.

This discovery allows for 90° toggling of the magnetization in a Pt/Co/GdOx system at

unprecedented speed and cyclability. Proton magneto-ionics also allow for gating of magnetic

anisotropy at a metal-metal interface through hydrogen insertion in a heavy metal adjacent to a

ferromagnet.

In chapter 6, we demonstrate voltage gating of magnetic damping and spin torques probed using

spin-torque ferromagnetic resonance. The results show that a wide range of fundamental

magnetic properties can be gated using ions.

In chapter 7, we compare the rates of voltage-induced magnetic changes in different Pt/Co/MOx

heterostructures, where M is Gd, Y, Zr, or Ta. The results show that speed of magnetic

modulation can change significantly depending on the choice of MOx as the proton conducting

electrolyte.

In chapter 8, we introduce a room temperature reversible solid oxide cell based on hydrogen

storage in a thin film oxide. The cell is a miniaturized version of a conventional solid oxide fuel

cell, and can be operated like a battery for microelectronics. The finding is important because it

shows the applicability of solid-oxide fuel cells for energy storage in microelectronics.

In chapter 9, we demonstrate voltage-induced modulation of optical properties in a thin film

metal/oxide/metal heterostructure using proton. Fast optical response down to 20ms was

achieved with good cyclability. This establishes the wide applicability of voltage-induced

protonics to modulate material properties in thin film heterostructures.

In chapter 10, we revisit some electrical properties of GdOx since it is used as the proton

conducting oxide in most of our gated devices.

Page 11: Dynamic Modulation of Material Properties by Solid State

5

In chapter 11, we summarize the findings of the thesis and discuss their impacts. This is followed

by a list of recommended future work to fill the remaining gap in our understanding proton

gating of material properties. We end by providing a brief outlook for this new field of research.

Page 12: Dynamic Modulation of Material Properties by Solid State

6

Chapter 2:

Background

Page 13: Dynamic Modulation of Material Properties by Solid State

7

2.0 Magnetic Hysteresis Loop The simplest and densest way to describe a magnetic sample is a magnetic hysteresis loop4. A

magnetic hysteresis loop is a plot of magnetization vs magnetic field as shown in Figure 2.1(a).

In magnetic hysteresis loops, the same magnetic field which is applied does not always produce

the same magnetization. The magnetization, M depends not only on the magnetic field which is

currently applied, it also depends on its previous state (This is why it is called a “hysteresis”

loop).

Figure 2.1. Magnetic hysteresis loop. (a) Easy-axis magnetic hysteresis loop. The coercive

field, HC is typically half the width of the hysteresis loop, while the saturation magnetization, MS

is half the height of the hysteresis loop. (b) A general profile of the applied magnetic field. (c)

The measured magnetization corresponding to the magnetic field in (b). (d) The direction of the

applied magnetic field and the magnetization. Both are pointing along the y-axis.

To generate a magnetic hysteresis loop, one sweeps the magnetic field and measures the

magnetization at each field according to the profile shown in figure 2.1b. For a sample which is

(1) uniformly magnetized in the direction of the magnetic field, and (2) where the magnetization

rotates coherently as a single domain (Stoner-Wohlfarth model), the resulting magnetization is

shown in figure 2.1c. When we first increase the magnetic field in the +y direction, the magnet

Page 14: Dynamic Modulation of Material Properties by Solid State

8

starts out pointing in a –y direction (hysteresis from its previous state). When the magnetic field

exceeds a critical field, the magnet “abruptly” flips 180º to the positive direction. This critical

field is called the coercive field, HC and it quantifies the energy barrier to switch the

magnetization 180º (assuming Stoner-Wohlfarth model). As one increases the magnetic field

further in the positive direction, the measured magnetization remains flat. The value at which this

magnetization plateaus is called the saturation magnetization, MS. When we next increase the

magnetic field in the –y direction, the magnetization again flips 180º at HC and plateaus at MS

but with a negative sign. If we plot the magnetization versus the magnetic field, we get the

magnetic hysteresis loop shown in figure 2.1a.

A magnetic hysteresis loop is measured using a variety of techniques such as magneto-optical

Kerr effect, and anomalous Hall effect. In these cases, M is measured indirectly as Kerr rotation

or Hall resistance respectively. These will be described in greater details in chapter 3. While we

can always extract HC and MS from a magnetic hysteresis loop, the magnetic information is not

limited to the two quantities. For instance, if the magnetic sample is not in a single domain state,

then the hysteresis loop can give us qualitative information on the domain configuration. It can

also give us information on the magnetic anisotropy depending on which direction the magnetic

field is applied in. The details will be discussed in subsequent sections.

In magnetism, there are two kinds of units which are frequently used: cgs (centimeter-gram-

second) and SI units. While SI units are more standardized, the magnitude is often too large for

thin film magnetism. For this thesis, we will mainly be using cgs units unless stated otherwise.

Page 15: Dynamic Modulation of Material Properties by Solid State

9

2.1 Magnetic Anisotropy

In an anisotropic system, the magnetization vector tends to point in certain preferred axis

because it minimizes its energy. This is called magnetic anisotropy4. The minimum energy axis

is called an easy magnetization axis whereas the maximum energy axis is called the hard

magnetization axis. In thin film magnetism, the term “magnetic anisotropy” is typically used in a

more specific sense; it represents an energy density, KU needed to rotate the magnetization from

the easy-axis to the hard-axis direction. This value is extremely important because it quantifies

the energy barrier which the magnetic moment needs to overcome in order switch180º. In other

words, the moment goes through the hard axis as it switches from one direction to the opposite

direction4,5.

Page 16: Dynamic Modulation of Material Properties by Solid State

10

Figure 2.2. Magnetic anisotropy. (a) Energy coordinate of M with the magnetic field applied

along the easy axis in –y direction. (b) Easy axis magnetic hysteresis loop. (c) Energy coordinate

of M with the magnetic field applied along the hard axis in +z direction. (d) Hard-axis magnetic

hysteresis loop. Adapted from reference5

In figure 2.2, we assumed the easy-axis is the y-axis while the hard axis is the z-axis and we

start out with the magnetization pointing in the +y direction. When we apply a magnetic field, H

along the easy axis in the -y direction (figure 2.2(a),), we decrease the energy of the –y state

while increasing the energy of the +y state. When the energy due to the magnetic field (Zeeman

energy) equals the energy barrier arising from the magnetic anisotropy, KU, the magnetization

switches from the +y direction to the -y direction. If one looks at the easy-axis hysteresis loop

which shows the y-projection of M (figure 2.2(b)), this happens at the coercive field, HC. If we

instead apply the magnetic field along the hard axis in the +z-direction (figure 2.2(c)), we reduce

the energy of the +z state while keeping the ±y state the same. As a result, there is no 180º

switching of the magnetization; the magnet simply rotates 90º from +y to +z. If one looks at the

hard-axis hysteresis loop (figure 2.2(d)), the z-projection of M gradually increases and saturates

at the anisotropy field, HK. Both HC and HK are equal to 2𝐾𝑈

𝑀𝑆 , hence we can see why magnetic

Page 17: Dynamic Modulation of Material Properties by Solid State

11

anisotropy is very important in determining the threshold magnetic field to switch or rotate the

magnetization.

Magnetic anisotropy arises from two microscopic origins: dipolar interactions and spin-orbit

coupling4–7. Dipolar interactions refers to the interaction between magnetic dipoles, and

intuitively, one can think of the dipolar energy being minimized when the one magnetic dipole is

aligned along the magnetic field generated by a second magnetic dipole. Dipolar interactions

lead to magnetic anisotropy because moments lower their energy when they are aligned along a

common axis. In bulk systems, there will be a very large number of these microscopic dipoles;

hence to find the stable magnetic configuration, one has to integrate all the individual dipoles to

find the minimum dipolar energy of the system. The second microscopic origin of magnetic

anisotropy is spin-orbit coupling (SOC). Spin-orbit coupling refers to the coupling between the

spin and orbital moments of each atom.

Figure 2.3. Spin orbit coupling. (a) Trajectory of electron from reference frame of the nucleus.

(b) Trajectory of the nucleus from reference frame of the electron

When electron spin orbits the nucleus of an atom, it experiences an effective magnetic field from

the nucleus because from the electron’s reference frame, it “sees” the positively charged nucleus

orbiting itself (a moving charge generates a magnetic field) (figure 2.3). The magnetic field

Page 18: Dynamic Modulation of Material Properties by Solid State

12

“generated” from the orbiting nucleus is represented by the orbital moment, L while the electron

spin is represented by the spin moment, S. As a result, the energy of spin-orbit coupling is

expressed as 𝐸𝑆𝑂 = −𝜉𝐿. 𝑆 where 𝜉 is the coefficient which represents the strength of the spin-

orbit coupling4–7. SOC leads to magnetic anisotropy because it couples the spin moment to the

orbital moment. And because the orbital moment is in turn coupled to the crystal lattice, a form

of magnetic anisotropy arises where the magnetic moments are stabilized along certain high

symmetry axes of the crystal lattice. This form of anisotropy is called magnetocrystalline

anisotropy.

In this thesis, we will be focusing on thin film 3d ferromagnet systems where the thickness of the

magnetic film is on the order of nm. While the microscopic origins (the two stated above) of

magnetic anisotropy are the same for all magnetic systems, in these thin film systems, we

classify the magnetic anisotropy into two contributions: a volumetric and a surface contribution.

The volumetric contribution, KV (energy per unit volume) is mainly due to shape anisotropy.

This anisotropy refers to the dependence of the magnetization easy axis on the shape of the

magnet and it arises due to dipolar interactions. For each shape and magnetization direction,

there is a corresponding total dipolar energy, or magnetostatic energy, 𝐸𝑚𝑎𝑔𝑛𝑒𝑡𝑜𝑠𝑡𝑎𝑡𝑖𝑐 which is

given by:

𝐸𝑚𝑎𝑔𝑛𝑒𝑡𝑜𝑠𝑡𝑎𝑡𝑖𝑐 =1

2𝐻𝐷⃗⃗ ⃗⃗ ⃗. �⃗⃗� (Equation 2.1)

𝐻𝐷⃗⃗ ⃗⃗ ⃗ = −𝑁𝐷�⃗⃗� (Equation 2.2)

Here, 𝐻𝐷⃗⃗ ⃗⃗ ⃗ is the demagnetizing field, 𝑁𝐷 is a geometry-dependent tensor, and �⃗⃗� is the

magnetization vector. The demagnetizing field 𝐻𝐷⃗⃗ ⃗⃗ ⃗, is an effective field experienced by each

magnetic dipole due to all other dipoles in the magnet. Equation 2.1 and 2.2 show that shape

Page 19: Dynamic Modulation of Material Properties by Solid State

13

anisotropy arises because there is certain magnetization direction relative to the magnet shape

where 𝐸𝑚𝑎𝑔𝑛𝑒𝑡𝑜𝑠𝑡𝑎𝑡𝑖𝑐 is minimum. One can further simplify the picture and think of a stable

magnetic configuration as one that results in the minimum area of “free” magnetic poles (figure

2.4).

Figure 2.4. Shape anisotropy. The magnetic configuration in (b) results in smaller area of

“free” magnetic poles (shaded in red) in the magnet compared to (a); as a result, (b) has smaller

Emagnetostatic.

In an ideal thin film which (1) stretches infinitely in the in-plane directions (x and y-axis) and

where (2) the magnetization points out-of-plane (z-axis), 𝐻𝐷 = 4πMS (MS is the saturation

magnetization) (figure 2.5) 4–7. This field points opposite to the out-of-plane magnetization,

resulting in 𝐸𝑚𝑎𝑔𝑛𝑒𝑡𝑜𝑠𝑡𝑎𝑡𝑖𝑐 of 2πMS2. If the magnetization points in any in-plane direction, the

resulting 𝐸𝑚𝑎𝑔𝑛𝑒𝑡𝑜𝑠𝑡𝑎𝑡𝑖𝑐 is 0. Hence, for a thin film magnet, assuming a negative convention for

in-plane anisotropy (magnet prefers to point in-plane), the volumetric magnetic anisotropy, KV =

-2πMS2.

Page 20: Dynamic Modulation of Material Properties by Solid State

14

Figure 2.5. Demagnetizing Field. Cross section of a thin film magnet showing the directions of

the magnetic moment and the demagnetizing field.

For practical applications, it is the out-of-plane magnetic configuration which is desired in order

to achieve the highest memory density2,8–11. Another commonly used term for out-of-plane

magnetic anisotropy is perpendicular magnetic anisotropy (PMA). Fortunately, in a thin film

system, the surface magnetic anisotropy, KS (energy per unit area) tends to favor PMA.

Generally speaking, in thin film ferromagnet with broken inversion symmetry, there is splitting

in degeneracy of the different orbitals (figure 2.6). Due to electrostatic interactions, these lower

energy orbitals tend to have an out-of-plane anisotropy.

Figure 2.6. Surface magnetic anisotropy. Schematic illustration of splitting of degeneracy in

thin film magnet with broken inversion symmetry. Adapted with permission from reference12.

Page 21: Dynamic Modulation of Material Properties by Solid State

15

Because PMA is generated by filling bands with out-of-plane orbital anisotropy, the position of

the fermi level is very important. One way to tune the fermi level is through hybridization with

heavy metals at the ferromagnetic interface. For instance, it has been known for many years that

large PMA exists in in Co/Pt and Co/Pd multilayers13,14 due to hybridization of the Co 3d orbitals

with the 5d and 4d orbitals of Pt and Pd respectively15–19. This hybridization tunes the fermi level

such that the out-of-plane orbitals are maximally filled and provides large spin-orbit coupling

from the Pt and Pd atoms. This hybridization is localized at the Co/Pt or Co/Pd interface and, for

a Co thickness of at least a few monolayers, this interface constitutes the majority of the PMA

across the entire Co film (figure 2.7)17.

Figure 2.7. Orbital moment near the Co/Pt interface. The perpendicular magnetic anisotropy

is largest closest to the Co/Pt interface. Adapted with permission from reference17.

More recently, it has been shown that hybridization between the s-p orbitals of the oxygen atom

and dz-orbitals of a 3d ferromagnetic metals such as Co and Fe can also stabilize PMA20,21. This

is remarkable given the low spin-orbit coupling strength in oxygen. It was also found that the

Page 22: Dynamic Modulation of Material Properties by Solid State

16

position of the oxidation front is very important in determining the strength of the PMA. By

starting out with a Pt/Co/Al trilayer structure and subjecting the top Al layer to different duration

of plasma oxidation front, Manchon et al found that the largest PMA is obtained when the

oxidation front is right at the Co/Al interface, where the structure is exactly Pt/Co/AlOx(figure

2.8)22,23. When the Al is underoxidized, the Co magnetization points in-plane whereas if it is

overoxidized (Pt/Co/CoO/AlOx), the PMA starts to decrease again.

Figure 2.8. PMA as function of interfacial oxidation. (a) Anisotropy field, Han as a function of

plasma oxidation time of the Pt/Co/Al trilayer structure. (b) Hall effect magnetic hysteresis loops

corresponding to different plasma oxidation times. Largest PMA is obtained when the oxidation

front is exactly at the Co/Al interface. Adapted with permission from reference22.

The total magnetic anisotropy, KU is given by the sum of the surface (KS) and volumetric

magnetic anisotropy (KV), according to the equation16,18:

Page 23: Dynamic Modulation of Material Properties by Solid State

17

𝐾𝑈 = 𝐾𝑉 +𝐾𝑆

𝑡 Equation 2.3

Where 𝑡 is the thickness of the ferromagnetic film. Due to the competition between KS and KV,

there is a critical thickness below which one gets PMA and above which one gets an in-plane

magnetic state. This is shown in figure 2.9, where the authors plotted Keff *tCo against tCo (the

symbol Keff is used in place of KU, Co is the magnetic material). Note that the intercept is 2KS

because there are two surfaces where KS is present.

Figure 2.9. Keff*t versus t for a Pd/Co multilayers. Keff*t corresponds to in-plane magnetic

anisotropy while Keff*t > 0 corresponds to perpendicular magnetic anisotropy. Adapted with

permission from reference16.

While we have so far focused on surface anisotropy arising from interfacial hybridization in a

static structure, it has also been known that gas adsorption on ultra-thin magnetic films can alter

magnetic anisotropy. One of the most studied gas is hydrogen, H2 which can either physiosorb as

molecules or chemisorb on metallic thin film after splitting into individual H atom24–31. Mankey

et al first found that hydrogen chemisorption reduces the surface magnetization of Co while

Page 24: Dynamic Modulation of Material Properties by Solid State

18

enhancing that of Fe24. According to the authors, this is because the majority sub-band is nearly

filled in Co whereas it is only partially filled in Fe; as a result, when electrons are added due to

hybridization with hydrogen, they are added to the minority sub-band in Co and majority sub-

band in Fe. This changes the Co and Fe magnetizations in opposite directions. Valvidares et al

have found through magneto-optical Kerr imaging that adsorption of hydrogen on a Pt(111)

substrate before the deposition of a thin Co layer (t < 1.3nm) decreases the PMA significantly25.

Using ab-initio calculations, they found a large reduction in magnet moments of the Co, which in

turn reduces the magnetic moment in Pt induced by proximity effect. As Pt atoms constitute the

largest source of spin-orbit-coupling, this results in a significant reduction in the PMA. Similarly,

Munbodh et al have found that hydrogen absorption in Co/Pd multilayers decreases both the

magnetic moment and PMA in the structure due to decreased induced moments in the Pd layer26.

Unlike Co heterostructures, Sander et al instead found that exposure of Ni/Cu(001) to hydrogen

induces PMA, and by cycling the hydrogen partial pressure they were able to cycle the

magnetization between in-plane and out-of-plane state reversibly (figure 2.10)27. In the case of

Ni, the authors attribute the change in magnetic anisotropy mainly to tetragonal distortion in the

Ni crystal structure.

Page 25: Dynamic Modulation of Material Properties by Solid State

19

Figure 2.10. Reversible hydrogen gas induced magnetization switching. Polar magneto-

optical Kerr signal (proportional to out of plane magnetization) of a Ni/Cu(001) thin film

structure as a function of partial pressure of hydrogen. Inset shows the direction of the

magnetization. Adapted with permission from reference27.

Page 26: Dynamic Modulation of Material Properties by Solid State

20

2.2 Magnetization Dynamics and Spin

Current

So far, we have pictured magnetization as moments which are aligned in a static configuration

along a magnetic field. However, magnetic moments undergo a precessional motion when

subjected to the slightest torque exerted by either a magnetic field or a spin current.

Magnetization dynamics is usually modelled by the Landau-Lifshiftz-Gilbert (LLG) equation

which is given by:

𝜕𝑚

𝜕𝑡= −𝛾𝑚 × 𝐻𝑒𝑓𝑓 + 𝛼𝑚 ×

𝜕𝑚

𝜕𝑡 Equation 2.4

Where 𝑚 is the magnetic moment vector, 𝜕𝑚

𝜕𝑡 is evolution of the magnetic moment as a function

of time, and 𝐻𝑒𝑓𝑓 is an effective field which includes contribution from the demagnetizing field,

interfacial anisotropy, and the applied field, H. 𝛾 is the gyromagnetic ratio and 𝛼 is the Gilbert

damping parameter. The first term on the right (−𝛾𝑚 × 𝐻𝑒𝑓𝑓) represents the field-like torque

which causes the magnetization to precess around 𝐻𝑒𝑓𝑓 while the second term on the right is the

damping-like torque (𝛼𝑚 ×𝜕𝑚

𝜕𝑡) which causes the magnetic moment to gradually “dampens”

towards the direction of 𝐻𝑒𝑓𝑓. Pictorially, this can be represented as:

Page 27: Dynamic Modulation of Material Properties by Solid State

21

Figure 2.11. Magnetization dynamics according to the LLG equation. Adapted from

reference32.

The original LLG equation was later modified to include spin-current induced torques

𝜕𝑚

𝜕𝑡= −𝛾𝑚 × 𝐻𝑒𝑓𝑓 + 𝛼𝑚 ×

𝜕𝑚

𝜕𝑡+ 𝜏𝐹𝐿

𝑠×𝑚

|(𝑠×𝑚)|+ 𝜏𝐷𝐿

𝑚×(𝑠×𝑚)

|𝑚×(𝑠×𝑚)| Equation 2.5

Where 𝜏𝐹𝐿 and 𝜏𝐷𝐿 are magnitudes of spin-current induced field-like and damping-like torques,

and s is the polarization of the spin current. Here, the third 𝜏𝐹𝐿𝑠×𝑚

|(𝑠×𝑚)| and the fourth

(𝜏𝐷𝐿𝑚×(𝑠×𝑚)

|𝑚×(𝑠×𝑚)|) terms are called the spin current induced field-like and damping-like

(Slonczewski) torques respectively because they have the same symmetry as the corresponding

field induced field-like and damping like torques. In this case, the spin current provides an

effective exchange field along the direction of its spin polarization. The third term drives the

moment to precess around this spin polarization direction while the fourth term drives the

moment to “dampen” towards the spin polarization direction. One way to generate spin current is

using the spin Hall effect where spin-orbit coupling in heavy metals drives spin current to flow in

tranverse direction to the charge curren (figure 2.12)33–35 .

Page 28: Dynamic Modulation of Material Properties by Solid State

22

Figure 2.12. Spin current induced by the spin Hall effect in Pt layer. The spin current

injected into the CoFe ferromagnetic layer generates a torque to drive magnetization dynamics or

switching. In this schematic, the charge current is along the x-axis, the spin current is along the

z-axis, and the spin polarization is along the y-axis. There is an exchange field due to the spin

current along the spin polarization direction. Adapted with permission from reference36.

Magnetization dynamics can be characterized using ferromagnetic resonance (FMR). In

conventional FMR, a magnetic field at high frequency is applied to a magnetic sample to provide

a torque (equation 2.4) which drives the moments to precess. Simultaneously, the absorption of

microwave radiation by the magnetic sample is measured in a microwave cavity. At resonance,

there will maximum absorption of radiation by the sample. The resonance frequency (or field)

then provides quantitative information about the 𝐻𝑒𝑓𝑓 experienced by the sample while the width

of the absorption peak is proportional to its damping parameters.

In this thesis, we instead rely on a variation of conventional FMR called spin-torque FMR (ST-

FMR) to characterize the magnetic properties of the magnetic sample37–40. A heavy metal such as

Ta or Pt is used to generate spin current using the spin Hall effect39. The generated spin current is

then used to drive magnetization precession in an adjacent ferromagnetic layer. The precession

of the magnetic layer manifests itself electrically as a change in DC resistance through

Page 29: Dynamic Modulation of Material Properties by Solid State

23

magnetoresistance, which can be measured as a mixing voltage(figure 2.13). A detailed

description of the method and analysis will be given in chapter 6.

Figure 2.13. Spin-torque ferromagnetic resonance (ST-FMR). (a) Schematic of the ST-FMR

measurement configuration. (b) Current induced change in magnetoresistance (ref. spin torque

diode effect). Adapted with permission from references38,39.

Page 30: Dynamic Modulation of Material Properties by Solid State

24

2.3 Magneto-Electric and Magneto-Ionic

Effects In magnetic memory, the ultimate goal is to have high memory density, low writing power, and

high thermal stability2. To achieve high memory density, the size of magnetic bit can be made

smaller. However smaller bits lead to degradation of the thermal stability because the energy

barrier to flip the magnetization is given by ∆𝐸 = 𝐾𝑢𝑉, where V is the volume of the magnetic

bit. ∆𝐸 has to be at least 60kT in order to have stable recording for more than 10 years (k is the

Boltzmann’s constant). One may compensate for this size reduction by increasing the magnetic

anisotropy, 𝐾𝑢 but as discussed in the previous section (chapter 2.1) the magnetic field, 𝐻𝑐

required to switch the magnetization is proportional to 𝐾𝑢. As a result, the total writing power

also increases. This challenge in achieving high memory density, low writing power, and high

thermal stability simultaneously in magnetic memory is called the trilemma.

One approach to overcoming the trilemma is to utilize a dynamic system where we reduce the

switching barrier temporarily during the writing operation. This can be done most effectively

using a gate voltage, and one of the mechanisms which allows for such modulation is the

magneto-electric effect. Magneto-electric effect refers to a wide range of phenomena which

allow for control of magnetization using an electric field. Magneto-electric phenomena can be

broadly classified into three main classes of materials: dilute magnetic semiconductors (DMS),

multiferroic materials, and ultra-thin metallic ferromagnet/oxide bilayers. The magneto-electric

coupling can be an intrinsic feature of the single phase material, or it can be coupled in

heterogeneous media through strain or interfacial interactions.

Page 31: Dynamic Modulation of Material Properties by Solid State

25

Dilute magnetic semiconductors (DMS) are semiconductors which exhibit magnetic properties

due to low level doping by ferromagnetic metals such as Mn41,42. The most studied of these DMS

is Mn-doped InGaAs which exhibits ferromagnetic properties due to hole-mediated interaction41–

44. With decreasing hole concentration, a super-exchange interaction which favors

antiferromagnetic configuration starts to dominate; as a result an electric field can reversibly

change the magnetic properties by changing the hole concentration through charge injection43–45.

This was first demonstrated by Chiba et al where they electrically gate the magnetic moment of

(In,Mn)As through modulation of its Curie temperature (figure 2.14)44. However, the major

problems with these systems are their extremely low Curie temperature (typically <50K) and the

complex fabrication steps to induce the ferromagnetism.

Multiferroics, on the other hand, are materials which exhibit magnetic and ferroelectric orders

simultaneously46. All of these materials have perovskite structures. In these materials, non-zero

magneto-electric coupling coefficient can arise from structural asymmetry; as a result, an electric

field can induce a change in magnetization and vice versa. Besides magneto-electric coupling in

a single phase system with non-zero magneto-electric coefficient, magneto-electric coupling can

also be engineered in heterogeneous ferroic media47–50. In this case, exchange or elastic coupling

between a ferroelectric and ferromagnetic material can allow for electric field control of

magnetization. Figure 2.15 shows an example of such system, where magnetic CoFe2O4

nanopillars are embedded with out-of-plane epitaxy in a ferroelectric BaTiO3 medium. When an

electric field is applied, the resulting strain in the ferroelectric BaTiO3 is imparted to the

CoFe2O4 nanopillars, causing a change in its magnetization48,49. While these systems have very

interesting physics, the main problem remains their complex fabrication process and the stringent

Page 32: Dynamic Modulation of Material Properties by Solid State

26

growth requirements (substrate epitaxy, growth temperature etc.) which are incompatible with

current CMOS processes.

Figure 2.14. Electric field control of ferromagnetism in a dilute magnetic semiconductor. (a) Schematic of device operation of a magnetic (In,Mn)As during electrical gating. (b) Hall

hysteresis loops of the (In,Mn)As at different gate voltages. Adapted with permission from

reference44.

Figure 2.15. Magneto-electric coupling in a heterogeneous multiferroic system. (a)

Schematic of magnetic CoFe2O4 (CFO) nanopillars embedded in a ferroelectric BaTiO3

(BTO)matrix. (b)-(c) Magnetic force microscope images of the CFO-BTO composite before and

after electrical poling at +12V. The region in red circle represents magnetization reversal upon

electrical poling. The green circle represents multi-domain formation. Adapted with permission

from references48,49.

Page 33: Dynamic Modulation of Material Properties by Solid State

27

The third class of materials where magneto-electric coupling has been observed is ultra-thin

metallic ferromagnet/ oxide bilayer systems12,51,52. In these devices, the ferromagnet needs to be

thin (<1nm) due to Coulombic screening in metals. Electric field-induced magnetic changes in

these materials has been attributed to a few factors: (1) spin-dependent screening of the electric

field can induce a net surface magnetization in an ultra-thin ferromagnet53. (2) electric field can

also lead to different occupation of the 3d orbitals at the surface layer of the ferromagnet; this in

turn changes the magnetic anisotropy54. (3) Besides changing the orbital occupation, it was also

proposed that an electric field directly changes the band structure of the ferromagnet and hence

the magnetic anisotropy55. The magneto-electric efficiency that has been predicted and

demonstrated in these ultra-thin ferromagnet is on the order of ~10 fJ/Vm. This value however is

too small for practical device applications. To put the value into context, an electric field of

~1MV/cm only induces a magnetic anisotropy change of a few percent in thermally stable

ferromagnetic devices. Figure 2.16 shows an example of an ultra-thin film ferromagnet/oxide

magnetoelectric device and its operation.

Figure 2.16. Magneto-electric effect in ultra-thin film ferromagnet/oxide device. (a)

Magneto-optical Kerr effect hysteresis loops of the Fe80Co20 magnetic film in (a) at V = ±200V.

(c) Simulated change in orbital magnetic anisotropy and magnetic anisotropy energy (MAE) as a

function of electric field for a free standing 15 monolayer thick Fe (001) film. Adapted with

permission from reference56,57.

Page 34: Dynamic Modulation of Material Properties by Solid State

28

Besides magneto-electric effect, another promising route to effective voltage control of

magnetism is through ionic modulation of magnetic interfaces58–65. This approach relies on

voltage-induced ionic migration and electrochemistry to modulate magnetic properties ranging

from magnetic anisotropy to spin-orbit torques. One example of magneto-ionic control of

magnetism is reversible oxidation and reduction of the Co ferromagnet layer in a Pt/Co/GdOx/Au

heterostructure (figure 2.17). The redox of the Co layer is confirmed by electron energy loss

spectroscocpy (EELS) and x-ray magnetic circular dichroism (XMCD) analyses60,61. While the

voltage-induced redox process is originally confined to the electrode edge, further optimizations

enabled uniform redox across the entire device region (figure 2.14). Oxygen-based magneto-

ionics were also demonstrated in Co/SrCoOx66,67, Co/HfOx

68, Co/ZnO69, and CoFeB/MgOx

systems70.

Figure 2.17. Magneto-ionic control of magnetism through voltage-induced redox of Co. (a)

Schematic of a Pt/Co(0.9nm)/GdOx/Au device. (b) EELS spectrum of the normalized O-K edge

count as a function gate voltage (Ubias). (c) Polar magneto-optical Kerr effect hysteresis loop of

the Pt/Co/GdOx/Au device under different gate voltages at 100ºC. Adapted from reference61.

Page 35: Dynamic Modulation of Material Properties by Solid State

29

Besides oxygen-ion, Group I ions like lithium ions and protons have also been used to induce

changes in magnetic properties67,71–73. For instance, Zhang et al have demonstrated large changes

in magnetic moments by reversible intercalation of lithium ions in spinel structures like

Fe2O3(figure 2.18)73. These experiments are done in a battery-like configuration with the spinel

structure acting as the cathode inside a liquid electrolyte. Similarly, proton-induced modulation

of magnetic properties have also been demonstrated by Nan et al in an acidic solution through

reversible absorption and desorption of hydrogen on an ultra-thin Co film74. More recently, Lu et

al have demonstrated that hydrogen and oxygen-induced phase transformation in SrCoOx can

lead to reversible transition between a paramagnetic, ferromagnetic and antiferromagnetic state

(figure 2.18b)67.

Magneto-ionic gating of magnetism has garnered great interest in recent years due to the

extremely large magneto-electric efficiency that can achieved, which is on the order of

~5000fJ/Vm61. This allows its implementation in devices to be more practical in terms of power

saving. However, there are a few major issues which remain to be addressed. For oxygen-ion

gating, it has been shown that irreversible structural and chemical degradation of the target

ferromagnet always accompany the magnetic property changes62. In addition, while oxidation of

the ferromagnet has indeed been observed through TEM-EELS studies, there is a lack of

understanding on the mechanism of oxidation and the source of oxidant. For lithium ion gating,

the major problem remains the incompatibility of most Group I ions with CMOS processing due

to formation of traps and defects in Si or SiO275. The exception to this is proton, where it is

relatively innocuous in its standard state. While proton gating is promising, it has only been

demonstrated using liquid electrolyte. This thesis henceforth aims to fill the gap in understanding

Page 36: Dynamic Modulation of Material Properties by Solid State

30

of oxygen-ion gating of magnetic properties, and to demonstrate for the first time proton gating

of magnetic properties in an all-solid-state device.

Figure 2.18: Magneto-ionic control of magnetism using lithium ion and proton. (a)

Reversible change in magnetization of Fe through intercalation of Li in Fe2O3 spinel structure.

(b) Reversible insertion of protons and oxygen ions in SrCoO2.5 to change the magnetization

between antiferromagnetic, paramagnetic, and ferromagnetic states. Adapted with permission

from references67,73.

Page 37: Dynamic Modulation of Material Properties by Solid State

31

2.4 Solid Oxide Fuel and Electrolyzer cell

The simplest fuel cell produces power from reaction of H2 and O2 to produce H2O, while an

electrolyzer cell produces H2 fuel and O2 from electrolysis of H2O76–78. Both cells share the same

device structure; whether it is a fuel or electrolyzer cell depends on the mode of operation. A

solid oxide cell, unlike conventional liquid electrolyte cells, uses a solid oxide electrolyte such as

Yttria-stabilized Zirconia (YSZ); as a result they are typically operated at high temperature (>

700C). A solid oxide cell can be further classified into a proton conducting oxide cell or an

oxygen-ion conducting oxide cell79,80. The main difference lies in the type of ions conducted

across the oxide; in the former case, it is a proton which gets transported, while in the latter case,

it is an oxygen ion. Figure 2.16 shows schematics of a proton-conducting and oxygen-ion

conducting solid oxide cell respectively run in electrolysis mode. The cells consist of two

electrodes sandwiching a solid oxide electrolyte layer. One electrode is the anode, where

oxidation takes place, while the other is the cathode where reduction takes place. Because the

anode and cathode assignment can change depending on operations modes, the electrodes facing

O2 and H2 are also known as the air and hydrogen electrode respectively79,80.

Page 38: Dynamic Modulation of Material Properties by Solid State

32

Figure 2.19. Solid oxide cell in electrolysis mode. (a) Proton conducting solid oxide cell. (b)

Oxygen-ion conducting solid oxide cell.

During operation of a proton conducting electrolyzer cell, a positive bias larger than the

thermodynamic potential of water splitting is applied to the air electrode (anode), which splits

water to produce proton and oxygen gas. The proton, 𝐻+ then gets transported across the oxide

due to the applied electric field, where it gets reduced at the hydrogen electrode (cathode) to

produce hydrogen gas. The reactions are shown below:

Proton conducting oxide electrolyzer cell

Anode: 2𝐻2𝑂 → 𝑂2 + 4𝐻+ + 4𝑒− Equation 2.6

Cathode: 4𝐻+ + 4𝑒− → 2𝐻2 Equation 2.7

Page 39: Dynamic Modulation of Material Properties by Solid State

33

For an oxygen-ion conducting solid oxide electrolyzer cell, the reaction starts at the hydrogen

electrode (cathode), where a negative bias splits water to produce oxygen ions and hydrogen gas.

The oxygen ions are driven by the electric field to the air electrode (anode), where it gets

oxidized to oxygen gas.

Oxygen-ion conducting oxide electrolyzer cell

Cathode: 2𝐻2𝑂 + 4𝑒− → 2𝐻2 + 2𝑂2− Equation 2.8

Anode: 2𝑂2− → 𝑂2 + 4𝑒− Equation 2.9

Equation 2.6 and 2.9 are known as the oxygen evolution reaction (OER) while equation 2.7 and

2.8 are known as the hydrogen evolution reaction (HER) (More discussion in Chapter 2.6). The

operation of a fuel cell is essentially an electrolyzer cell run in reverse mode; for instance, in

figure 2.19(a), hydrogen gas would be oxidized to H+, which gets transported to the cathode

where it recombines with O2 gas to form H2O.

Some examples of well-known proton conducting oxide electrolyte include barium cerate,

BaCeO3 (BCO) and barium zirconate, BaZrO3 (BZO) while examples of oxygen-ion conducting

oxide electrolyte include yttria-stabilized zirconia, YxZr1-xO2 (YSZ) and gadolinium-doped ceria,

Ce1-GdxO2 (GDC) 81–83. Common air electrodes include lanthanum strontium manganite, La1-

xSrxMnO3 (LSMO) while hydrogen electrodes include Ni-ceramic composites such as Ni-

YSZ84,85. A more extensive discussion of solid oxide cell electrodes will be given in Chapter 2.6

to 2.7.

One figure of merit to characterize performance in a solid oxide fuel cell is power density. It is

measured by sourcing current from a fuel cell and measuring the resulting voltage at a specific

temperature and partial pressure of H2 fuel. This is repeated for increasing current values to get

Page 40: Dynamic Modulation of Material Properties by Solid State

34

the power density curve. Power density depends on a wide range of factors such as overpotential

(discussed below), number of active sites on the electrode catalyst, and gas transport of the

reactants to the active sites. Figure 2.20(a) shows an example of a power density curve for a solid

oxide fuel cell with 0.3mm thick YSZ. An analogous plot can also be generated for an

electrolyzer cell (Figure 2.20(b)); in this case, low voltage and power are desired for operation of

the cell.

Figure 2.20. Performance of a solid oxide fuel cell and electrolyzer cell. (a) Exemplary power

density curve of a fuel cell. (b) Analogous performance plot for an electrolyzer cell. Adapted

with permission from reference78,86

In an electrolyzer cell, the goal is to maximize the rate of H2 production from water splitting at

minimum potential. For a fuel cell, the goal is to achieve highest power output at maximum

potential from hydrogen-oxygen recombination to form water. In both cases, the maximum (fuel

cell) and minimum (electrolyzer cell) achievable voltage is the thermodynamic potential, which

is 1.23V for H2O at standard conditions. To characterize deviations from this thermodynamic

potential, a parameter called overpotential, 𝜂 is used, which essentially tells us how many

Page 41: Dynamic Modulation of Material Properties by Solid State

35

additional volts above the thermodynamic potential is required to split water in an electolyzer

cell, or how many volts less than the thermodynamic potential that we can extract from water

generation in a fuel cell. Studies of electrolyte and electrode materials are intended to minimize

this overpotential, as it represents a loss. There are in general three types of overpotentials:

activation overpotential, ohmic overpotential, and concentration overpotential. Activation

overpotential arises because, the kinetics of charge transfer at thermodynamic potential is very

slow; as a result an additional voltage is required to drive this reaction in the anodic or cathodic

direction. Ohmic overpotential arises from ohmic loss and is mitigated by reducing the overall

resistance of the cell. Concentration overpotential arises from limited kinetics of the mass

transport of either the reactants to the active sites or products from the active sites. For solid

oxide cells, activation and concentration overpotential are primarily minimized through

optimization of the electrodes. For ohmic overpotential, the overall resistance is reduced through

optimization of the electrodes, electrolyte, and also the electrode/electrolyte interface. The net

voltage, 𝑉 that can be extracted from a fuel cell after accounting for all sources of overpotential

is given in equation 2.10. Similarly, the voltage that is needed in an electrolyzer cell after

accounting for all sources of overpotential is given in equation 2.11. Another term which is

commonly used in place of overpotential is polarization loss.

𝑉 = 𝐸𝑜 − 𝜂𝑜ℎ𝑚𝑖𝑐 − 𝜂𝑎𝑐𝑡𝑖𝑣 − 𝜂𝑐𝑜𝑛𝑐 Equation 2.10

𝑉 = 𝐸𝑜 + 𝜂𝑜ℎ𝑚𝑖𝑐 + 𝜂𝑎𝑐𝑡𝑖𝑣 + 𝜂𝑐𝑜𝑛𝑐 Equation 2.11

Here, 𝜂𝑜ℎ𝑚𝑖𝑐 , 𝜂𝑎𝑐𝑡𝑖𝑣, and 𝜂𝑐𝑜𝑛𝑐 are the ohmic, activation and concentration overpotentials

respectively.

Page 42: Dynamic Modulation of Material Properties by Solid State

36

2.5 Solid Oxide Proton Conductors

Solid oxide proton conductors have mainly been developed as electrolytes for fuel cell

applications. The main motivation for choosing an electrolyte with the largest proton

conductivity is to reduce the ohmic overpotential and to maximize the power density that can be

extracted from the fuel cell.

Proton conduction can in general be classified into two broad categories: one based on the

Grotthuss mechanism and another based on the vehicle mechanism81,82,87. In Grothuss

mechanism, proton migration is achieved by proton “hopping” between oxygen host lattice sites,

and the net transfer of proton depends crucially on factors like lattice dynamics and distance

between the oxygen host atoms. If proton migration only depends on proton “hopping” between

static host oxygen atoms, the conductivity should increase with decreasing distance between

neighboring sites. However, in many cases, this is completely opposite; structures with larger

oxygen-oxygen distance exhibits larger proton conductivity. The reason for this is the transfer of

proton involves cooperative movement of neighboring oxygen atoms which get closer/further

apart momentarily due to lattice vibrations. This allows proton to break and reform hydrogen

bonds with neighboring atoms, resulting in a net migration of proton. The lattice dynamics of

oxygen atoms is crucial for hydrogen bond breaking; and for short stiff bonds where the oxygen

atoms approximate a static lattice, the proton remains essentially “trapped” in a symmetrical

bond. In general, bulk proton conductivity (non-grain boundary) of solid-oxide proton

conductors can be attributed to the Grotthuss mechanism.

Page 43: Dynamic Modulation of Material Properties by Solid State

37

Figure 2.21. Coordinates of proton conduction through Grotthuss mechanism. (a)

Conduction through proton “hopping” between static oxygen lattice atoms. (b) Conduction

involving both proton “hopping” and oxygen atom lattice vibration. (c) Conduction involving

only oxygen atom lattice vibration to break hydrogen bonds. Adapted with permission from

reference88.

The second type of proton conduction mechanism is the vehicle mechanism, where proton

migration through the electrolyte is a mediated by a vehicle, such as a H2O molecule. In this

case, the rate of diffusion of the vehicle can play a crucial role in determining the overall proton

conductivity (in addition to the rate hydrogen bond breaking). The vehicle mechanism dominates

in a liquid-like environment where molecules can easily diffuse around, unlike a solid oxide

lattice where the atoms are relatively rigid.

The most studied solid oxide proton conductors are the perovskite-structure oxides such as doped

BaCeO3 and BaZrO379,89–91. Other well-known proton conductors include phosphates like LaPO4,

Page 44: Dynamic Modulation of Material Properties by Solid State

38

rare earth oxides such as Gd2O3 , orthoniobates and orthotantalates such as LaNbO492–94. A

compilation of their conductivity values are shown in figure 2.22.

Figure 2.22. Proton conductivities of different oxides. Adapted with permission from

reference82.

In almost all proton conducting oxides, acceptor doping is required to generate oxygen

vacancies, and these vacancies in turn absorb water to form hydroxide defect according to the

reaction:

𝐻2𝑂 + 𝑂𝑂𝑥 + 𝑉𝑂

∙∙ → 2𝑂𝐻𝑂∙ Equation 2.12

Here 𝑂𝑂𝑥 represents an oxygen ion on a normal oxygen site, 𝑉𝑂

∙∙ an oxygen vacancy with net

double positive charge relative to the normally occupied lattice site, and 𝑂𝐻𝑂∙ a singly positively

Page 45: Dynamic Modulation of Material Properties by Solid State

39

charged proton localized around an oxygen ion sitting on a normal oxygen site. These localized

protons form the charge carriers, and because proton conductivity (𝜎) is a function of both the

concentration of carriers (𝑛), charge (𝑞) and mobility (𝜇), according to the reaction 𝜎 = 𝑛𝑞𝜇, the

larger the amount of dissolved water, the larger is the proton conductivity. While large acceptor

doping can increase the concentration of carriers, excess doping will lead to lattice reordering

and strain effects, which in turn reduce the mobility, 𝜇. In addition, not all oxygen vacancies that

are created by doping will react with water to form hydroxide defects according to equation 2.12.

This will depend on the enthalpy of dissolution of water in the oxide. In general, the more

exothermic is the enthalpy, the more water can be dissolved in the oxide. For rare earth oxides,

higher stability of oxide leads to larger exothermic enthalpy and hence dissolution of water in the

oxide according to equation 2.12. This is counterintuitive because it is the least stable rare earth

oxides (La2O3) that is most reactive with water to form hydroxide. On the other hand, for

perovskites such as BaZrO3 and BaCeO3, it is the larger basicity that has been attributed to larger

dissolution of water. In these perovskites, their stability is inversely related to their capacity for

dissolution of water. For instance, in the case of BaCeO3, it decomposes to Ba(OH)2 and CeO2 at

high partial pressure of water95.

Page 46: Dynamic Modulation of Material Properties by Solid State

40

Figure 2.23. Rotation and proton transfer in a perovskite structure. Adapted with

permission from reference82

For the mobility of proton carriers, the trend is more complicated, as it depends on a series of

steps such as rotation around the oxygen host atom and proton transfer (involving hydrogen-

bond breaking). There are however some general trends: (1) the reduction in symmetry of the

lattice reduces the mobility, such as in SrCeO3, because dissimilar environment for proton

hopping reduces net permutations of ways to get from one site to another96. (2) For rare earth

oxides, mobility decreases as the lattice parameter of the lattice decreases. So far, the best known

solid oxide proton conductors are BaCeO3 and BaZrO3, with proton conductivities between 10-

1S/cm and 10-3S/cm at ~700C. However, as mentioned, BaCeO3 is not stable under high water

partial pressure. BaZrO3 on the other hand, cannot be sintered well which leads to large volume

of grain boundaries and higher resistance. A lot of research effort is hence focused on improving

the stability of BCO and sinterability of BZO through doping and special fabrication techniques.

Besides proton conductivity through grains, there has also been evidence of proton conductivity

through surface and grain boundaries in some nanocrystalline materials. In these systems, there

is water uptake in the grain boundaries and surface, and because the volumetric ratio of both

Page 47: Dynamic Modulation of Material Properties by Solid State

41

components to grain is very large, this form of proton conduction can dominate97–99. One

disadvantage of such proton conductors is increasing temperature will decrease the proton

conductivity as water vaporizes, as shown for nanocrystalline ceria in figure 2.24(a) and

nanocrystalline YSZ in figure 2.24(b). At higher temperature, the ionic conductivity is

dominated by another charge carrier, such as oxygen vacancies. In addition, the difference in

ionic conductivities in wet and dry atmospheres can be a few orders of magnitude at room

temperature.

Figure 2.24. Proton conductivity of nanocrystalline oxide. (a) Conductivity plot of

nanocrystalline ceria. (b) Conductivity plot of nanocrystalline YSZ. Ionic conduction in both

oxides are dominated by proton conduction through water. Adapter with permission from

references98,99.

For this thesis, there are some differences between the devices studied and majority of solid

oxide cells in literature. While there are many proton conducting oxides that have been

Page 48: Dynamic Modulation of Material Properties by Solid State

42

investigated at intermediate temperatures (~500C to 700C), almost none has been studied at

room temperature (25C). For voltage-gating of functional interfaces using protons, we are mainly

only interested in operation at low temperatures (<100ºC). From extrapolation of the

intermediate temperature data, we can already observe significantly different trends and

candidates from what were generally known as “good” proton conductors. Another significant

difference between conventional studies and this thesis is the dimension of the device. The

thinnest proton conducting electrolyte that has been studied for fuel and electrolyzer cells is a

few hundred nm; this thickness is constrained by mechanical strength of the electrolyte

membrane when subjected to high partial pressure of reactant gases. However, the thickness of

the proton conductor used in this thesis is down to 4nm, which is >100x smaller. At such

dimension, the electric field is extremely large (1-10MV/cm) and can have drastic, non-linear

effect on the motion of ions, as seen in many memristive devices100. Figure 2.25 shows some

driving forces which may be present under such large fields.

Figure 2.25. Driving forces for ionic migration in memristive devices under large fields. (a)

Driving force due to drift. (b) Driving force due to electromigration (c) Driving force due to

concentration gradient. (d) Driving force due to temperature gradient. Adapted with permission

from reference100.

Page 49: Dynamic Modulation of Material Properties by Solid State

43

Finally, for voltage gating of functional interfaces using protons, the primary figure of merit for

fast device response is proton mobility rather than proton conductivity as long as there is

sufficient concentration of proton to alter the interfaces. While oxides with large proton

conductivities generally have large proton mobilities, this is not always the case. One example of

this is BaTiO3, where the proton conductivity is very low due to low solubility of hydroxide

defects, but the proton mobility can be very high101.

Page 50: Dynamic Modulation of Material Properties by Solid State

44

2.6 Water Electro-Catalysis

While electrolyte conducts ions across a cell, water splitting/recombination take place at the

anode and the cathode. In water electro-catalysis, there are four types of reactions: oxygen

evolution reactions (OER), oxygen reduction reactions (ORR), hydrogen evolution reaction

(HER), and hydrogen oxidation reaction (HOR)96,102–109. The first two reactions: the OER and

HER are involved during electrolysis of water (electrolyzer mode) to produce O2 and H2

respectively, while the latter two reactions: the ORR and HOR are involved during consumption

of O2 and H2 to produce H2O (fuel cell mode).

The net ORR and HOR are just the reverse reactions of the OER and HER respectively.

Typically, the oxygen reactions (OER and ORR) constitute the largest source of overpotential

because strong oxygen bonds need to be broken in both processes110–113. This can be understood

by looking at the intermediate species which are produced during the reactions. Rossmeisl et al.

broke down the OER reaction into four distinct elementary steps in an acidic electrolyte111,112,

each involving the transfer of one electron, according to equation 2.13. The elementary steps for

HER is shown in equation 2.14. The corresponding reaction coordinates for OER are shown in

figure 2.26.

2𝐻+ + 2𝑒− → 𝐻+ + 𝐻∗ + 𝑒− Equation 2.14

→ 𝐻2

2𝐻2𝑂 → 𝐻𝑂∗ + 𝐻2𝑂 + 𝐻+ + 𝑒− Equation 2.13

→ 𝑂∗ + 𝐻2𝑂 + 2𝐻+ + 2𝑒−

→ 𝐻𝑂𝑂∗ + 3𝐻+ + 3𝑒−

→ 𝑂2 + 4𝐻+ + 4𝑒−

Page 51: Dynamic Modulation of Material Properties by Solid State

45

Figure 2.26. Reaction coordinates of the OER in acidic electrolyte solution. Adapted with

permission from reference112.

From the figure, one can see the Gibbs free energy of H2O and O2 species are exactly equal at an

applied bias of 1.23V (the thermodynamic potential), however the intermediate products are at

different energy levels at this bias. Because the reactions happen in series, the kinetics are

essentially limited by the largest energy barrier between the intermediate products. The reaction

kinetic is then quantified by the current, 𝑖 according to the equation:

𝑖 = 𝑖𝑘exp (−∆𝐺𝑟

𝑘𝑇) Equation 2.15

Where 𝑖𝑘 is a constant, ∆𝐺𝑟 is the Gibbs free energy barrier of the rate limiting step, k is the

Boltzmann constant, and T is the absolute temperature. An overpotential, or more specifically

the activation overpotential, hence serves to reduce the ∆𝐺𝑟 of the rate-limiting step.

For water electro-catalysis, the best catalysts usually neither have the highest nor the lowest

binding energy to the reactants, but rather have intermediate values, according to the Sabatier’s

principle. As a result, when one plots the catalytic activity against the reactant binding energy,

Page 52: Dynamic Modulation of Material Properties by Solid State

46

one typically gets a “volcano” plot. Some of the plots for metal and binary oxide OER catalysts

are shown in figure 2.27. Generally speaking, this “volvano” trend exists because too weak of

binding energy leads to low conversion rate of reactants to intermediates, whereas too strong of

binding energy leads to low conversion rate of intermediates to final products. The ideal catalyst

is one where the energy splitting between all the intermediates are the same, as shown in figure

2.24(c) for the case of an OER catalyst. Pt and RuOx are the best metal and oxide (non-

perovskite) OER catalysts because their binding energies to the intermediates are closest to the

ideal catalyst (equi-energy splitting). Note that because ORR is just the reverse of OER, one

should expect the best OER catalyst to also be the best ORR catalyst.

Figure 2.27. Volcano plots. Volcano plots for (a) metallic and (b) binary oxide OER catalysts.

(c) Energy profile of an ideal OER catalyst. Adapted with permission from references104,113.

Page 53: Dynamic Modulation of Material Properties by Solid State

47

2.7 Electrodes for Solid Oxide Cells

So far we have focused on the electronic structure of catalyst which primarily affects charge

transfer at the electrode/electrolyte interface. However, mass transport can also be a significant

source of overpotential; in fact it often dominates in systems like solid oxide cell where gas

phase diffusion and ionic transfer across the electrode/electrolyte interface are involved114,115.

Fundamentally, most losses in solid oxide cell electrodes boil down to the fact that a single phase

cannot conduct all three species involved in a gas phase reaction: an electron, an ion, and a gas

molecule. Only an electronically conductive material can transport electrons, only an ionically

conductive material can transport ions, and only a gas phase can transport a gas molecule. As a

result, a complete reaction can only occur at the intersection of these three phases; this

intersection is called a triple phase boundary (TPB). Figure 2.28 shows an example of a TPB at

the cathode of an oxygen-ion conducting oxide fuel cell.

Figure 2.28. Triple phase boundary. Schematic illustration of the triple phase boundary at the

cathode of an oxygen-ion conducting solid oxide fuel cell. Adapted with permission from

reference114.

Page 54: Dynamic Modulation of Material Properties by Solid State

48

The necessity of three phases for a complete reaction not only reduces the total number of

reaction sites, it also increases the overall complexity due to added intermediate steps and

species. For instance, it has been proposed that the oxygen reduction reaction happening at the

cathode of a proton conducting oxide fuel cell consists of eight elementary steps, each with its

own reaction order resulting in a net reaction of 𝑂2 + 4𝐻+ + 4𝑒− → 2𝐻2𝑂 (table 2.1). Because

these steps happen in series, the slowest step will be the rate-limiting step and will determine the

total resistance from the electrode115. For electrodes in solid oxide cells, they are mainly

characterized by their area specific resistance (ASR), with 𝑅𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑑𝑒 =𝐴𝑆𝑅

𝐴𝑟𝑒𝑎 . Here, 𝑅𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑑𝑒 is

the total electrode resistance. ASR is not normalized to the thickness of the electrode because in

most cases, the resistance of the electrode does not scale with its thickness; it is a function of

extrinsic properties such as microstructure and porosity.

Table 2.1. Elementary reaction steps of ORR at the cathode of a proton conducting solid oxide

fuel cell. Reproduced with permission from reference115.

In the past, to increase the area of the triple phase boundary, porous Pt was commonly used as

both the anodes and cathodes. However due to the prohibitive cost of Pt and the limited added

TPB from porosity alone, new methods have been employed. These methods have mainly

revolved around using materials with mixed ionic and electronic conductivities (MIEC). By

Page 55: Dynamic Modulation of Material Properties by Solid State

49

using MIEC, one only needs the intersection of two phases for reactions to take place because

one of the phases conduct both electrons and ions. As a result, the overall activity increases due

to larger area where electrons, ions, and gas molecules can coexist. MIEC can exist as a one-

phase material such as (La,Sr)(Co,Fe)O3-δ (LSCF) or two-phase material consisting of a

composite of electronic conductor and an ionic conductor, such as Ni-YSZ and LSM-

YSZ84,85,116,117. Schematics of operation of an MIEC cathode in a proton conducting fuel cell is

shown in figure 2.29.

LSCF, a one-phase MIEC, is a perovskite oxide, where the La and Sr atoms are at the A-site, and

the Co and Fe atoms are at the B-site. Sr doping induces a change in oxidation state of both the

Co and Fe from +3 to +4 resulting in p-type electronic conductivity. Simultaneously, there is also

charge compensation through formation of oxygen vacancies, resulting in increased oxygen-ion

conductivity. As a result, LSCF has both electronic and ionic conductivities at high temperature.

It is most commonly used with a ceria electrolyte. LSM ( La1-xSrxMn O3-δ ), which is typically

the electronically conductive component of a two-phase MIEC, is a perovskite where the A-site

La3+ is doped with Sr with an oxidation state of +2. This acceptor doping is mainly compensated

by change in oxidation state of the Mn from +3 to +4 which results in p-type electronic

conductivity. However, in this case, there is minimal change in the oxygen vacancy

concentration. LSM is usually mixed with YSZ, (oxygen ion conductor) to form a two-phase

composite MIEC structure. This cathode is most commonly used on a YSZ electrolyte due to the

excellent match in thermal expansion coefficients. For the anode, the most commonly used

material is Ni-YSZ cermet (cermet is short for ceramic-metal), which is a two-phase composite

MIEC. Ni is used due to three reasons. (1) it is one of the best HOR catalyst, (2) it is cheap, and

(3) it provides mechanical support to the fuel cell, especially in cases where the electrolyte is

Page 56: Dynamic Modulation of Material Properties by Solid State

50

very thin. Other examples of anode include Ni-SDC (Samarium-doped ceria, SmxCe1-xO2-δ) and

Ni-GDC (Gadolinium-doped ceria, GdxCe1-xO2-δ). While the primary ionic carrier in most MIEC

is oxygen-ion, there are also MIEC such as BaCe1-xFexO3 whose main ionic conducting species

is the proton. These proton-based MIEC are typically doped BCO (BaCeO2-δ) and BZO (BaZrO2-

δ)80,115.

Figure 2.29. Active sites in different types of cathodes for a proton conducting oxide fuel

cell. (a) Porous single phase electronic conductor such as Pt. (b) Two phase mixed-electronic-

ionic conductor (MIEC). (c) Single phase MIEC such as BaCe1-xFexO3.

Page 57: Dynamic Modulation of Material Properties by Solid State

51

2.8 Electrochemical Impedance

Spectroscopy

In solid state ionics, one of the most used techniques to characterize ionic conductivity is

electrochemical impedance spectroscopy (EIS). It is a simple technique where a small sinusoidal

AC voltage (~10mV to 100mV) is applied to the sample and the resulting impedance is

measured. To get a full impedance spectrum, the frequency of the AC voltage is swept (~0.1Hz

to 1MHz) to probe different processes which contribute to the total conductivity. The spectrum is

then fitted with one or more RC components (figure 2.30) to get the corresponding resistance and

capacitance associated with each process118–121.

Figure 2.30. EIS spectrum and its equivalent circuits. Adapted with permission from

reference121.

1

𝑍𝑎=

1

𝑅𝑎+

1

𝐶𝑎𝑗𝜔 Equation 2.16

𝑍𝑡𝑜𝑡𝑎𝑙 = 𝑍𝑎 + 𝑍𝑏 + 𝑍𝑐 Equation 2.17

Page 58: Dynamic Modulation of Material Properties by Solid State

52

In real systems, the capacitor is usually replaced with a constant phase element (CPE) to model

non-ideal capacitor behavior. In this case, the capacitance of CPE can be calculated according to

equation 2.19.

𝑍𝐶𝑃𝐸 =1

𝑄(𝑗𝜔)𝑛 Equation 2.18

𝐶𝐶𝑃𝐸 = (𝑅1−𝑛𝑄)𝑛 Equation 2.19

Here, Q and n are the CPE constants. In solid oxide electrolyte, a “brick layer” model is typically

used to represent a polycrystalline material with grain boundaries (figure 2.31)122. Due to the

presence of ionic conduction in both the grains and grain boundaries, the impedance spectrum

for the oxide electrolyte is fitted with two RC circuits. Figure 2.32 shows an example spectrum

for YSZ where the grain and grain boundary contributions to ionic conduction can be clearly

resolved. The third semicircle corresponds to reactions at the electrodes. In this case, the figure

also shows increasing grain boundary resistance with decreasing grain size.

Figure 2.31. “Brick layer” model for polycrystalline solid oxide. Grains have size L and grain

boundaries have width 2b. Reproduced with permission from reference123.

Page 59: Dynamic Modulation of Material Properties by Solid State

53

Figure 2.32. Impedance spectra of YSZ. There are 3 RC components corresponding to oxygen-

ion conduction in the bulk (grain) and grain boundaries, and the reactions at the electrode. The

data shows increasing resistance of the grain boundary with decreasing grain size, dg.

Reproduced with permission from reference123.

Page 60: Dynamic Modulation of Material Properties by Solid State

54

Chapter 3:

Experimental Methods

Page 61: Dynamic Modulation of Material Properties by Solid State

55

3.1 Sputter Deposition

Sputtering is a physical vapor deposition technique where a target material is bombarded with

energetic particles from a plasma which results in ejection of the material from the target surface

and its eventual deposition on a substrate124. Figure 3.1 shows a schematic of the sputtering

process. Sputtering is done at high or ultra-high vacuum (<10-5Torr) to sustain the sputtering

plasma and to ensure high purity of deposited films. Ar gas is usually used as the sputtering gas.

Sputtering is a powerful technique for both laboratory and industrial scale production because

the deposited thin films can be made uniform and smooth across a large area. In addition, a large

variety of materials ranging from metals to metal oxides can be deposited.

Figure 3.1. Sputtering of Au on a substrate using Ar as the sputtering gas.

In sputtering, microstructure control is one of the most important things to consider. A very

useful guide for this is the zone structure model, which gives a qualitative description of the

expected microstructure for different sputtering parameters125. Figure 3.2(a) shows a structure

zone model developed by Thornton et al to describe the thin film microstructure as a function of

the substrate temperature, TS and Ar sputtering gas pressure126. For generality, TS is usually

Page 62: Dynamic Modulation of Material Properties by Solid State

56

normalized by the melting temperature of the target material, TM and the ratio is called the

homologous temperature. One can divide the different microstructures into three zones: Zone I,

Zone II, and Zone III. A fourth zone, Zone T is latter added between Zone I and II. Qualitatively,

as one goes from zone I to zone III, the kinetic energy of the sputtered atoms increase. As a

result, the atoms have higher rate of diffusion and can coalesce to form larger grains. In zone I,

the thin film is fine-grained and porous, in zone II, the thin film is columnar, and in zone III, the

thin film has large equiaxed grains. An example cross section of the different zones (including

zone T) is also shown in figure 3.2(b)127.

Figure 3.2. Structure zone model. (a) Structure zone model for different argon sputtering

pressure and substrate temperature. (b) Cross sections of microstructures in different zones.

Adapted with permission from reference126,127.

Besides the substrate temperature and sputtering gas pressure, other factors that affect the

microstructure of sputtered films include background pressure and the impurity level. If one uses

oxygen for sputtering, partial oxygen pressure can also affect the overall grain size, as shown in

figure 3.3(e) and (f) for the case of ITO128,129.

Page 63: Dynamic Modulation of Material Properties by Solid State

57

Figure 3.3. Microstructure as a function of oxygen partial pressure. (a)-(b) SEM images of

ITO grains deposited at 4x10-5mbarr (e) and 4.75x10-5mbarr (f) of oxygen partial pressure.

Adapted with permission from reference128.

In this thesis, all the samples are fabricated using sputtering system shown in figure 3.4. The

base pressure of the chamber reach down to 10-7Torr using a Pfeiffer turbomolecular pump

backed by an Alcatel 2008A roughing pump. The system consists of four sputtering sources

which are mounted on the floor of the sputter chamber. Each sputtering source has its own

chimney and shutter for better control of the sputtering area and timing. For depositing metallic

layers, metal targets are sputtered using DC power sources. For depositing oxide layers, there are

two approaches: (1) a metal target is reactively sputtered using a DC power source under a

partial pressure of oxygen, or (2) an oxide target is directly RF sputtered (also under oxygen

partial pressure to control microstructure of the film). The argon sputtering gas pressure is

between 2-5mTorr in all cases, whereas the oxygen partial pressure can range between

0.01mTorr to 0.9mTorr depending on the materials.

In the sputtering chamber, there is a rotating substrate table with slots for four substrate holders,

and a masking table with slots for 6 masks. The masking table allows one to sputter selectively

on different substrate slots. Typically, sputtering is done while rotating the substrate and mask

Page 64: Dynamic Modulation of Material Properties by Solid State

58

table in order to achieve better uniformity. The rotation rate of the tables is ~36/min. All

subtrates are kept at room temperature during sputtering.

Figure 3.4. Schematic of sputtering system. (a) Substrate table. (b) Substrate holder slot. (c)

Liquid nitrogen reservoir. (d) Gears for rotating the substrate table. (e) Mask table. (f) Mask slot.

(g) Pins to align substrate and mask tables. (h) Chamber floor (i) Jack and bellows below the

chamber floor. (j) Chimney for sputter source. Adapted from reference130

Page 65: Dynamic Modulation of Material Properties by Solid State

59

Figure 3.5. Image of substrate holder.

The thicknesses of the sputtered layers are calibrated using x-ray reflectivity. For metallic layers,

the deposition rates are between 1 to 5nm/min at 0.4A of sputtering current. The rates are very

similar for oxide layers deposited using reactive sputtering of a metal target at low oxygen partial

pressure (<0.1mT). For reactive sputtering at high oxygen partial pressure, target poisoning takes

place where the surface of the metal target becomes oxidized before it is sputtered. As a result,

the deposition rate goes down significantly (~1 order of magnitude). Figure 3.6 shows the

deposition rate of a Gd target before and after target poisoning. For RF sputtering of oxide

layers, the deposition rates are typically <0.5nm/min at 100W of sputtering power. The rates

quoted are all for rotating sputter deposition; for stationary deposition the rate can be up to 7

times larger but the uniformity is poorer. Stationary sputtering is only done for contact pads or

the top electrodes.

Page 66: Dynamic Modulation of Material Properties by Solid State

60

Figure 3.6. Deposition rates. (a)Deposition rates of an S-gun Gd target at 0.4A current, 3mTorr

of Ar sputtering gas, and at different O2 partial pressure, PO2. (b) Oxygen flow rates required to

achieve PO2.

Page 67: Dynamic Modulation of Material Properties by Solid State

61

3.2 Sample Structure and Patterning

In this thesis, the general sample structure used for the magneto-ionic devices is Ta/Pt/Co/MOx

/Au , where M is Gd, Y, Zr, Ta or Mg. Ta is the adhesion layer, Pt is the bottom electrode, Co is

the magnetic layer, MOx is the ionic-conducting oxide, and Au is the top electrode. Another

magneto-ionic device structure is Ta/Pd/Co/Pd/MOx/Au where the Ta/Pd/Co/Pd form the bottom

electrode layers and Au is the top electrode. All magneto-ionic devices are sputter deposited on

p-doped Si substrate with 50nm of thermal oxide.

To probe dynamic modulation of magnetic properties in these structures, magneto-optical Kerr

effect (MOKE) polarimetry (Chapter 3.3) and Hall magnetometry (Chapter 3.4) were used. For

devices used in the MOKE experiments, the Ta/Pt/Co/MOx layers (or Ta/Pd/Co/Pd/MOx layers)

were deposited on Si substrate as continuous films, with the bottom metallic layers uncovered to

provide electrical connection to the ground (figure 3.7). 200µm diameter top Au electrodes were

then patterned on these continuous films. In total, three sputtering runs with two vacuum breaks

are required to make the gated MOKE devices (ie Ta/Pt, then Co/MOx, then Au). In cases where

the top Au electrodes need to be deposited in situ directly after the MOx layer, an in situ mask

aligner is used to pattern the Au electrodes.

Page 68: Dynamic Modulation of Material Properties by Solid State

62

Figure 3.7. Sample structure for MOKE magneto-ionic device. (a) Schematic of a magneto-

ionic device used for time-resolved MOKE measurement during voltage gating. (b) Layout of the

mask used for patterning the top Au electrodes. The diameter of the electrodes is 200µm.

For the devices used in Hall measurement, the Ta/Pt/Co/MOx layers are patterned into Hall bar

geometry as shown in figure 3.8. A thicker MOx layer is then deposited in the form of a

rectangular patch with its area larger than the active region to ensure no parasitic current leakage

path between the top and bottom electrodes. The top Au is patterned into a square, where there is

a strip protruding out at 45º to both the Hall bar arms to allow easy probe access to the top Au

gate. In total, four sputtering runs with three vacuum breaks are required to make the gated Hall

devices.

Page 69: Dynamic Modulation of Material Properties by Solid State

63

Figure 3.8. Sample structure for gated hall bar device. (a) Schematic of a gated hall bar used

for time-resolved Hall measurement during voltage gating. (b) Detailed layouts of the four masks

used to pattern the complete device. The active region is 500µm x 500µm.

For the reversible solid oxide cells studied in this thesis, Ta/Pt/GdOx/Au structure was used. In

this case, there is no magnetic layer and the GdOx serves as the charge storage layer. The

structure has a cross bar geometry with the bottom Ta/Pt bottom electrode on the horizontal arm

and the Au top electrode on the vertical arm. The GdOx is deposited as a rectangular patch

between the Ta/Pt and the Au layers. A simplified device schematic is shown in figure 3.9(a) and

the mask designs are shown in figure 3.9(b). In total, four sputtering runs with three vacuum

breaks are required to make the reversible solid oxide cells.

Page 70: Dynamic Modulation of Material Properties by Solid State

64

Figure 3.9. Sample structure for cross bars. (a) Schematic of cross bar used for reversible

solid oxide fuel cells. (b) Detailed layouts of the four masks used to pattern the complete device.

The active areas can range between ~10-3cm2 and 4x10-1cm2 depending on the width of the arms.

For the gated spin-torque ferromagnetic resonance (ST-FMR) device used in chapter 6, device

layout as shown in figure 3.10 was used. The overall device consists of a Ta/Pt/Co/GdOx FMR

planar waveguide on top of which we deposit an overlayer of GdOx which covers the entire

active region. The active region is then gated by a top Au electrode which is connected by an

extended arm to a contact pad. The waveguides used have dimensions of 10 µm x5µm, 10 µm

x10µm, 20 µm x5µm and 20 µm x10µm. The mask layout is also included in figure 3.10.

Page 71: Dynamic Modulation of Material Properties by Solid State

65

Figure 3.10. Sample structure for gated ST-FMR device. (a) Schematic of gated ST-FMR

device. (b) Detailed layouts of the four masks used to pattern a complete device.

For patterning most devices, shadow mask lithography is the primary technique used. For this, a

flexible 0.01” thick PEEK sheet serves as the mask and a laser cutter is used to define features

down to ~10µm (figure 3.11). During sample preparation for sputtering, the PEEK sheets are

pressed tight again the Si substrates and taped using a Cu tape to minimize any shadowing effect.

For patterning sub-micron to micron scale structures such as the gated ST-FMR devices, optical

lithography was used. Si substrates were coated with positive photoresist, Megaposit SPR700,

exposed, baked, and developed with Microposit MF-CD26 developer. This is repeated n number

of times where n is the number of mask layers.

Page 72: Dynamic Modulation of Material Properties by Solid State

66

Figure 3.11. Shadow mask lithography. (a) Image of a PEEK mask and top electrode patterns

defined using a laser cutter. (b) Backside of the PEEK mask where a substrate is attached tightly

using Cu tape.

Page 73: Dynamic Modulation of Material Properties by Solid State

67

3.3 Magneto-Optical Kerr Effect

Magneto-optical Kerr effect (MOKE) polarimetry is a powerful technique to detect

magnetization in ultra-thin films. With a magnetic field, it allows one to obtain magnetic

hysteresis loops. The basic idea behind MOKE is a linearly polarized light which is incident on a

magnetic sample is rotated by an angle, 𝜃𝑆 (Kerr rotation) and gains a slight ellipticity, 𝜖𝑆 (Kerr

ellipticity) (figure 3.11)131,132.

Figure 3.12. Magneto-optical Kerr effect (MOKE). Adapted from reference131.

A MOKE setup for detecting magnetization consists of a laser source, an incident polarizer, an

analyzer, and a photodiode detector. The analyzer is a polarizer which is set at (90+ δ)º to the

incident polarizer. The setup is shown in figure 3.13. The incident light first passes through the

incident polarizer and becomes linearly polarized. When the polarized light is reflected off a

magnetic surface, its polarization is rotated by 𝜃𝑆 . The photodiode detects this rotation as a

change in intensity, 𝐼 which can be expressed as equation 3.1.

Page 74: Dynamic Modulation of Material Properties by Solid State

68

𝐼 = 𝐼𝑜(1 +𝜃𝑆

𝛿) Equation 3.1

If one inserts a quarter wave plate before the analyzer, the detected intensity becomes

proportional to 𝜖𝑆 (Kerr ellipticity), which can be expressed as equation 3.2:

𝐼 = 𝐼𝑜(1 +𝜖𝑆

𝛿) Equation 3.2

Figure 3.13. MOKE configurations. (a) MOKE setup consisting of a laser source, incident

polarizer, analyzer, and a photodiode detector. (b) Polar MOKE configuration for detecting out-

of-plane magnetization. (c) Longitudinal MOKE configuration for detecting in-plane

magnetization which is parallel to the incidence plane. (d) Transverse MOKE configuration for

detecting in-plane magnetization which is perpendicular to the incidence plane. Adapted from

reference133.

There are three basic types of MOKE configurations for detecting magnetization in thin films: a

polar MOKE, a transverse MOKE, and a longitudinal MOKE. Their setups are shown in figure

3.13. A polar MOKE is used to detect magnetization which is out-of-plane, a longitudinal

MOKE is used to detect magnetization which is in-plane and parallel to the incidence plane,

Page 75: Dynamic Modulation of Material Properties by Solid State

69

while a transverse MOKE is used to detect magnetization which is in-plane but perpendicular to

the incidence plane.

For this thesis, we primarily use the polar MOKE configuration to measure out-of-plane

magnetic hysteresis loops. A simplified measurement setup is shown in figure 3.14. A laser

source with wavelength of ~655nm is used, and the laser spot which is incident on the sample

can be focused down to <10µm using an objective lens. A CCD aligned along the vertical axis

allows us to image the sample. A sample stage with an out-of-plane magnetic coil is set up which

can produce 1000Oe of out-of-plane field. To obtain MOKE hysteresis loops, we sweep the

magnetic field and measure the reflected intensity from a thin film sample. The amplitude of this

MOKE hysteresis loop is proportional to the out-of-plane magnetization (MZ) of the sample

(figure 3.15).

Figure 3.14. Polar MOKE setup with electrical probes and out-of-plane field for time-resolved

MOKE measurement during voltage gating.

Page 76: Dynamic Modulation of Material Properties by Solid State

70

For time-resolved measurements during voltage gating, a series of MOKE magnetic hysteresis

loops are acquired continually, and parameters like coercivities (HC) and MZ can be plotted as a

function of time (figure 3.15 and 3.16). For voltage gating, mechanically compliant CuBe probes

with radius of 25µm were used. One probe is landed on the top electrode to apply a gate voltage,

while a second probe is landed on the uncovered bottom electrode to make the ground

connection. The laser spot is focused on the middle of the top electrode during measurements

(figure 3.16).

Figure 3.15. Time resolved MOKE. (a) Magnetic field profile and measured Kerr signal

(intensity) as a function of time. Both values are continuously measured and MOKE hysteresis

loops are generated for each time period (1s). (b) Exemplary MOKE hysteresis loop generated

from time period 1.

Page 77: Dynamic Modulation of Material Properties by Solid State

71

Figure 3.16. Time series data. (a) Time series of MZ and HC extracted from MOKE hysteresis

loops during each cycle. (b) Optical micrograph of an Au top electrode with a CuBe probe

landed. Also shown is the focused laser spot.

Page 78: Dynamic Modulation of Material Properties by Solid State

72

3.4 Anomalous and Planar Hall Effect

In ordinary Hall effect, a potential difference (known as Hall voltage, VHall) develops in the

transverse direction when an electrical current flows along a strip of conductor under an applied

perpendicular magnetic field. The ordinary Hall voltage (𝑉𝑂𝐻𝐸)is generated due to the Lorentz

force from the magnetic field, and is hence proportional to the product 𝐼 × �⃗⃗� where 𝐼 is the

current vector and �⃗⃗� is the field vector.

Figure 3.17: Hall effects. (a) Coordinate system. (b) Ordinary Hall effect. (c) Anomalous Hall

effect. (d) Planar Hall effect. In all three cases, the current flow is in the +x direction and the

sample plane is the x-y plane. “+” and “-“ signs represent polarity of measured Hall voltage.

Adapted from reference134.

Page 79: Dynamic Modulation of Material Properties by Solid State

73

The ordinary Hall effect however, does not allow one to detect magnetization because the

measured Hall voltage is proportional to the applied magnetic field, not the magnetization of the

sample. To detect magnetization in a thin film sample, we instead rely on the anomalous and

planar Hall effect (figure 3.17)134,135. The anomalous Hall effect is used to probe an out-of-plane

magnetized sample because its Hall resistance (𝑅𝐴𝐻𝐸) is proportional to the out-of-plane

magnetization (𝑀𝑍). The Hall resistance is simply the ratio of transverse voltage to the current in

the longitudinal direction. The planar Hall effect is used to probe an in-plane magnetized sample

because its Hall resistance (𝑅𝑃𝐻𝐸)is proportional to the product of the magnetization projections

along the two in-plane orthogonal directions (𝑀𝑥and 𝑀𝑌) . Using the convention in figure 3.17,

𝑅𝐴𝐻𝐸 and 𝑅𝑃𝐻𝐸 are proportional to 𝑀𝑍and 𝑀𝑥𝑀𝑌 respectively, and their sums can be expressed

more generally as

𝑅𝐻𝑎𝑙𝑙 = 𝑅𝐴𝐻𝐸 + 𝑅𝑃𝐻𝐸

= 𝑅𝐴𝐻𝐸° 𝑀𝑐𝑜𝑠𝜃 + 𝑅𝑃𝐻𝐸

° 𝑀𝑠𝑖𝑛2𝜃𝑠𝑖𝑛2𝜑 Equation 3.3

Here 𝑅𝐴𝐻𝐸° and 𝑅𝑃𝐻𝐸

° are constants which depend on the dimensions of the sample and the

intrinsic material properties (Berry curvature etc). As mentioned earlier, if we have an out-of-

plane sample, the second term vanishes and 𝑅𝐻𝑎𝑙𝑙 = 𝑅𝐴𝐻𝐸 , whereas if we have an in-plane

magnetized sample, the first term vanishes and 𝑅𝐻𝑎𝑙𝑙 = 𝑅𝑃𝐻𝐸. We have ignored 𝑅𝑂𝐻𝐸 in the sum

for 𝑅𝐻𝑎𝑙𝑙. Hence, to measure the magnetizations, we just have to source a longitudinal current (𝐼)

and measure the resulting Hall voltage, 𝑉𝐻𝑎𝑙𝑙.

𝑉𝐻𝑎𝑙𝑙 = 𝐼𝑅𝐻𝑎𝑙𝑙 Equation 3.4

Page 80: Dynamic Modulation of Material Properties by Solid State

74

3.5 Time Resolved Hall Magnetometry

under different Atmospheric Conditions

Anomalous and planar Hall effects can be used to measure magnetic hysteresis loops of thin film

samples (figure 3.17). In this thesis, the anomalous and planar Hall measurement system is

shown in figure 3.18. A lock-in amplifier is used to both source the longitudinal current and

measure the transverse voltage. A lock-in frequency of 2kHz is typically used with an integration

time of 30ms while the source current used depends on the dimension and material of the

samples. An out-of-plane magnetic coil provides out-of-plane field for measurement of

anomalous Hall hysteresis loops while an in-plane magnetic coil set at 45º to the current injection

direction provides an in-plane field for obtaining planar Hall hysteresis loops. Similar to time

resolved MOKE, anomalous and planar Hall magnetic hysteresis loops are alternately acquired

continually during voltage gating experiment. This allows us to probe both the out-of-plane and

in-plane magnetizations in real time under a gate bias. A Keithley sourcemeter unit was used to

provide the gate voltage.

Figure 3.18. Hall hysteresis loops. (a) Anomalous Hall hysteresis loop. (b) Planar Hall

hysteresis loop.

Page 81: Dynamic Modulation of Material Properties by Solid State

75

Figure 3.19. Hall measurement system. (a) Top-down and (b) side views of the anomalous and

planar Hall measurement system.

The Hall measurement system is set up in a Lakeshore CPX-VF probe station which allows us to

probe magnetization in situ when applying a top gate voltage under different atmospheric

conditions. Hall devices can be subjected to vacuum condition down to 10-4mbarr using a

turbomolecular pump. For dry conditions, gases such as O2 and N2 (purity of >99%) can be

introduced directly into the chamber using a venting valve. For wet conditions, gases are bubbled

through water in a Fisher Scientific bubbler before being introduced into the chamber.

Page 82: Dynamic Modulation of Material Properties by Solid State

76

Figure 3.20. Schematic of the CPX-VF probe station with Hall measurement system and

atmospheric control.

Page 83: Dynamic Modulation of Material Properties by Solid State

77

3.6 Spin-torque Ferromagnetic

Resonance

For spin-torque ferromagnetic resonance (ST-FMR) experiments in chapter 6, Ta/Pt/Co/GdOx

structure was patterned into ST-FMR planar waveguide with a top GdOx/Au gate as shown in

figure 3.1039,136. A coaxial GS microwave probe (GGB Industries Picoprobe 40A series) with

pitch of 150µm to 250µm was used for electrical contact (one tip was landed on A, another on B

as shown in figure 3.21) while a signal generator was used to inject microwave frequency power

between 5.36GHz to 12 GHz into the planar waveguide. This microwave power acts a source of

current-induced torque which drives ferromagnetic resonance in the planar waveguide. For most

measurements, the power output of the signal generator was set to 15dbm. The output of the

signal generator was modulated by a lock-in amplifier in order to allow low frequency detection

of the mixing voltage (Vmix) which results from the change in DC resistance. Since the technique

involves signal input and detection at high and low frequencies respectively, a bias-tee was used

to separate the two components. To apply a gate voltage, a Keithley 2400 sourcemeter and CuBe

DC probe were used.

In all measurements, an in-plane magnetic field, 𝐻 is applied at 45° to the waveguide and the

measurement protocol involves sweeping 𝐻 while measuring 𝑉𝑚𝑖𝑥 at a fixed frequency, f. The

𝑉𝑚𝑖𝑥 can then be fitted according to equation below:

𝑉𝑚𝑖𝑥 = 𝑆𝑊2

(𝐻− 𝐻𝐹𝑀𝑅)2+𝑊2 + 𝐴𝑊(𝐻− 𝐻𝐹𝑀𝑅)

(𝐻− 𝐻𝐹𝑀𝑅)2+𝑊2 Equation 3.5

Page 84: Dynamic Modulation of Material Properties by Solid State

78

Where 𝑆 and 𝐴 are the amplitudes of the symmetric and antisymmetric Lorentzian components

of the resonance peak respectively, and 𝑊 and 𝐻𝐹𝑀𝑅 are the width and position of both these

Lorentzian components. Figure 3.22 shows an example of a ST-FMR spectrum measured at

8GHz.

Figure 3.21. Spin-torque ferromagnetic resonance system to probe voltage gating of magnetic

properties.

Figure 3.22. ST-FMR spectrum of a Pt(3nm)/Co(6nm)/GdOx(30nm) device at f=8GHz.

Page 85: Dynamic Modulation of Material Properties by Solid State

79

3.7 Solid Oxide Cell Characterization

For studying the performance of a reversible solid oxide fuel cell, the main metrics are the

capacity, power density and cyclability. For this, the first plot that is usually generated is the

discharge curve, which is a plot of cell voltage versus total charge. The measurement is

performed by repeatedly sourcing a constant current from the cell and measuring the resulting

voltage. When the cell voltage drops to a threshold value, the measurement is stopped, and the

capacity of the cell is given by the total integrated charge. The discharge curve provides

qualitative information regarding chemical reaction steps and the cell overpotential, besides the

capacity. The second plot which is generated is the power density curve. The measurement is

very similar to the discharge curve; however in this case, the sourced current is gradually ramped

up while measuring the voltage. And instead of plotting the cell voltage, it is the power at each

source current that is plotted; the peak power density from the cell can then be obtained from the

curve. To study cyclability, the cell is repeatedly charged and discharged, and the capacity

during each cycle is plotted as a function of cycle number. Exemplary plots of discharge curve,

power density curve, and cyclability are shown in figure 3.23.

Page 86: Dynamic Modulation of Material Properties by Solid State

80

Figure 3.23. Performance plots of a reversible solid oxide fuel cell. (a) Discharge curves at

different discharge currents. (b) Power density curve. (c) Plot of charge versus cycle number to

study the cyclability of a device.

All the measurements are done using a Keithley 6430 source meter in a Lakeshore CPX-VF

probe station. The cells studied have a cross bar geometry where one arm is connected to the top

electrode and the other arm is connected to the bottom electrode. Active areas range from ~10-

3cm2 to 0.4cm2. For charging and discharging cells under different atmospheric conditions,

different gases can be introduced into the chamber as described in the previous section. In this

case, all the magnets are removed for measurement.

Page 87: Dynamic Modulation of Material Properties by Solid State

81

Chapter 4: Effect of

H2O on Voltage-

induced Co Oxidation

in a Pt/Co/GdOx

Heterostructure

Page 88: Dynamic Modulation of Material Properties by Solid State

82

Magneto-ionic effect in a Pt/Co/GdOx/Au system was first observed when a negative gate

voltage (VG) applied to the top Au causes a significant reduction in the magnetic anisotropy at

the electrode edge, which traps domain wall propagation58. Because the electrode edge is the

triple phase boundary where O2 gas, Au and O2- ions meet, the authors proposed that the

magnetic anisotropy is reduced due to voltage-driven O2- motion from GdOx into the Co layer

under a negative gate bias. By then optimizing the geometry of the devices, the authors were

subsequently able to gate the magnetic anisotropy across the entire electrode region. EELS

studies subsequently confirmed that a negative gate voltage does indeed oxidize the Co layer in

the Pt/Co/GdOx device61. However, voltage-driven O2- transport from GdOx to Co layer has still

not been directly observed eventhough it was generally assumed to be the operative mechanism

in voltage-induced Co oxidation.

While O2- can be the active oxidant, it is also well-known that many oxides readily absorb

water from the atmosphere, incorporated as hydroxide defects situated at oxygen ion vacancies

through the following defect reaction:137

𝐻2𝑂 + 𝑂𝑂𝑥 + 𝑉𝑂

∙∙ → 2𝑂𝐻𝑂∙ Equation 4.1

Here 𝑂𝑂𝑥 represents an oxygen ion on a normal oxygen site, 𝑉𝑂

∙∙ an oxygen vacancy with net

double positive charge relative to the normally occupied lattice site, and 𝑂𝐻𝑂∙ a singly positively

charged proton localized around an oxygen ion sitting on a normal oxygen site. For memristors,

it is already well established that humidity can alter the resistive switching behavior of oxides

such as TaOx138,139, HfOx

139, SiOx140 and SrTiOx

141,142 memristive cells due to its effect on bulk

oxide properties and interface reactions at the anode and cathode. The high basicity of rare earth

oxides makes them particularly hygroscopic143, and Gd2O3 is known to react with moisture to

Page 89: Dynamic Modulation of Material Properties by Solid State

83

form hydroxides144–148 such as Gd(OH)3 with consequent changes to electrical147 and ionic

properties92.

In this chapter, we show that it is H2O stored in GdOx as Gd(OH)3 that oxidizes Co under a

negative gate bias, and that oxygen migration plays an insignificant role. We further show that

hydrogen-induced CoO reduction leads to water uptake back into the GdOx matrix, allowing for

closed-system electrochemical and magnetic property switching without the need for

atmospheric exchange. These results provide a mechanistic understanding of magneto-ionic

switching in metal/oxide heterostructures and essential insights to enable magneto-ionic device

engineering.

Page 90: Dynamic Modulation of Material Properties by Solid State

84

4.1: Experimental Methods

Sample preparation. Ta(4 nm)/Pt(3 nm)/Co(0.9 nm)/GdOx(tGdOx nm) films were fabricated on

thermally oxidized Si (100) substrates using magnetron sputtering at room temperature and

3mTorr Ar pressure. The metal layers were grown by DC sputtering. All GdOx layers were

deposited using DC reactive sputtering with PO2 of 0.07mTorr except for the XAS samples,

where the deposition was done using RF sputtering with PO2 of 0.7mTorr O2. For MOKE

measurements, 200 μm diameter Au(3 nm) electrodes were patterned on top of the

Ta/Pt/Co/GdOx continuous film, with the Ta (4 nm)/Pt(3 nm) underlayer was uncovered by

GdOx at the sample edge to allow electrical contact to the back. For in situ XAS measurements,

the Ta/Pt/Co/GdOx/Au structure was patterned into a cross-bar geometry with 1mm arm width.

X-ray reflectivity (XRR) measurements. XRR was carried out using a Bruker D8 Discover

HRXRD instrument with Cu K-α radiation at wavelength of 1.54Å.

X-ray Photoelectron Spectroscopy (XPS) measurements. XPS was carried out using a

Physical Electronics Versaprobe II X-ray Photoelectron Spectrometer at a base pressure of 5x10-

9 Torr.

Polar magneto-optical Kerr effect (MOKE) measurements. MOKE measurements were

performed using a 1 mW laser with a wavelength of 655 nm focused to spot size of ~10µm.

Experiments were performed in polar geometry and hence sensitive to the out-of-plane

magnetization component. To apply gate voltage VG to the circular electrodes, a CuBe probe

was landed near the edge of the electrode and the Ta/Pt back electrode was grounded. The laser

spot was focused at the middle of the electrode. All experiments were performed at room

Page 91: Dynamic Modulation of Material Properties by Solid State

85

temperature. For experiments where different atmospheric conditions are required, VG was

applied ex-situ in a CPX-VF probe station. The MOKE hysteresis loop was measured before and

after VG was applied. Experiments under controlled gas environments were performed by

backfilling the chamber with either O2 gas (99.999% purity) or N2 gas (99.999% purity).

Humidity was introduced into the N2 gas flow by bubbling through water. Wet N2 and ambient

condition at 25C corresponds to ~20mT and 12mT of H2O partial pressure respectively.

Vacuum condition corresponds to a base pressure of 10-4mbarr. All experiments were performed

at room temperature.

X-Ray absorption spectroscopy (XAS). In situ XAS data were taken at the In-situ and Operando

Soft X-ray Spectroscopy (IOS, 23-ID-2) beamline at the National Synchrotron Light Source II,

Brookhaven National Laboratory. Partial fluorescence yield (PFY) spectra were acquired using a

Vortex EM silicon drift detector. The incident soft x-ray beam has a footprint of ~100 x 20 µm

and is directed at 30° relative to the sample normal, while the PFY detector is positioned at 40° to

the sample normal (Supplementary Information V). The sample used for the measurement has

crossbar geometry with sample structure Ta (4nm)/Mg(30 nm)/Pd(10 nm)/Co(0.9 nm)/GdOx (30

nm) and a 3 nm Au top gate. The XAS incident beam spot was located on the sample by first

scanning the scanning stage to locate the crossbars through the total electron yield of the top Au

electrode and chemical signature of the bottom electrode (Mg K-edge) (More details in

Supplementary Information V). Measurement is done with the VG applied in situ. For experiments

which require humidity, H2O vapor is introduced into the chamber through a leak valve and the

flow rate is adjusted to maintain PH2O of 10 Torr. At vacuum condition, the main chamber pressure

is ~2 x 10-7 Torr after H2O evacuation, and the sample is kept at room temperature throughout the

measurement.

Page 92: Dynamic Modulation of Material Properties by Solid State

86

4.2: Probing Water Uptake in GdOx

Figure 4.1(a) shows the effect of hydrating a GdOx thin film deposited on a SiO2/Si substrate.

The hydration treatment involves placing the sample at 90ºC under wet nitrogen gas at ambient

pressure with PH2O = 525 Torr, for up to 168 hours. X-ray reflectivity (XRR) spectra were

obtained periodically during the hydration process to follow the evolution of the film thickness

and density. During hydration, Gd2O3 is expected to react with H2O to form Gd(OH)3 according

to the reaction:146

𝐺𝑑2𝑂3 + 3𝐻2𝑂 → 2𝐺𝑑(𝑂𝐻)3 Equation 4.2

The XRR spectra were fitted by modeling the film as a bilayer of Gd2O3 and Gd(OH)3, with

variable thicknesses, roughnesses (structural and/or chemical), and mass densities (converted to

x-ray scattering length densities (SLDs). Figure 4.1(b) shows SLD profiles corresponding to fits

of the XRR spectra (Fig. 4.2), where two distinct layers are clearly resolved. With increasing

hydration time, we observe a gradual progression of the Gd(OH)3 layer deeper into the film.

Note that there is a diffuse gradient between a completely dry Gd2O3 and a fully hydrated

Gd(OH)3. In fact, the fitted roughness between the Gd2O3 and Gd(OH)3 layers arises from a

density gradient between the two layers, not from a structural roughness. During the hydration

process, the dry Gd2O3 first dissolves water molecules in the form of proton defects according to

Eq. 4.1. When the layer is completely hydrated, it forms Gd(OH)3 according to net reaction

depicted in Eq. 4.2. The transition region hence is expected to consist of a mixed phase of

hydrated Gd2O3 with dissolved water and Gd(OH)3.

Page 93: Dynamic Modulation of Material Properties by Solid State

87

Figure 4.1(c) shows the fitted thickness of the Gd(OH)3 layer as a function of hydration time

while figure 1(d) shows the XRR spectra of the non-hydrated (0 h) and hydrated (144 h) GdOx

films respectively. From the data, we see that a 22.8 nm as-prepared GdOx film takes

approximately 144 h to fully transform to Gd(OH)3. The fully transformed hydroxide shows an

increase in thickness of 50%, expanding from 22.8 nm to 34.1 nm and a decrease in density of

28%, from 8.3g/cm3 to 6.0g/cm3. This corresponds very well to the transformation of monoclinic

Gd2O3 to Gd(OH)3, with bulk densities of 8.3g/cm3 and 5.6g/cm3 respectively149,150.

In order to confirm the chemical state of the GdOx layer, we also performed x-ray

photoelectron spectroscopy (XPS) on the surface of a 3 nm GdOx thin film which was exposed to

ambient atmosphere for >2 weeks. Figure 4.1(e) shows the O1s spectrum of the thin film, where

the data is best fitted by two peaks at 531.8eV and 529.2eV, which correspond to the O-H bond

in Gd(OH)3 and the O-O bond in Gd2O3151,152 respectively. These results show that GdOx

readily uptakes water even in ambient conditions to form a hydroxide phase.

Page 94: Dynamic Modulation of Material Properties by Solid State

88

Figure 4.1. Probing water uptake in GdOx. (a) Schematic of a non-hydrated, partially

hydrated, and fully hydrated GdOx thin film on a SiO2/Si substrate. (b) Scattering length density

of the GdOx thin film as a function of hydration time. The fitted mass densities of Gd2O3 and

Gd(OH)3 are 8.3g/cc and 6.0 g/cc respectively. c)Fitted thickness of Gd(OH)3 as a function of

hydration time. d) X-ray reflectivity (XRR) spectra of a non-hydrated and hydrated GdOx thin

film (144h of hydration). The solid and dashed lines are the raw data and fits respectively. e) X-

ray Photoelectron Spectroscopy (XPS) data of a 3nm GdOx thin film surface which has been

exposed to ambient for > 2 weeks.

Page 95: Dynamic Modulation of Material Properties by Solid State

89

Figure 4.2. X-ray reflectivity (XRR) spectra. [(a)-(h)] XRR spectra of a GdOx thin film on a

SiO2/Si substrate as a function of hydration time. The solid and dashed lines are the raw data and

fits respectively. Hydration treatment involves placing the sample at 90C under 525Torr of PH2O.

Page 96: Dynamic Modulation of Material Properties by Solid State

90

4.3: Voltage-induced Co Oxidation in

Hydrated and Non-hydrated Pt/Co/GdOx

Devices

Figure 4.3 shows the comparison of voltage-induced Co oxidation between a non-hydrated

and hydrated Pt (3 nm)/Co(0.9 nm)/GdOx(10 nm)/Au(3 nm) device. Here, we probe the magnetic

state by measuring hysteresis loops probed locally using a polar magneto-optical Kerr effect

(MOKE) polarimeter. In its metallic state, the film exhibits a perpendicular magnetic anisotropy,

whereas in the oxidized state there is no magnetic signal, which provides a convenient means to

probe interfacial chemical state changes61.

For the non-hydrated device, the top Au electrode was deposited using an in situ shadow

mask immediately after the deposition of the Pt/Co/GdOx layers without vacuum break, so as to

serve as a capping layer to minimize water uptake upon exposure to ambient. Characterizations

of the non-hydrated devices were then done immediately after fabrication (within a day) in order

to preserve the non-hydrated state. For the hydrated device, the Pt/Co/GdOx structure was first

placed at 90ºC under PH2O = 525 Torr for 72 hours before the deposition of the top Au electrode.

All gate voltages (VG) were applied to the top Au electrode while the bottom Pt was grounded.

Figure 4.3(a) shows a MOKE hysteresis loop of a virgin non-hydrated device while figures

4.3(b)-(e) show MOKE hysteresis loops of the non-hydrated device after VG = -3V has been

applied for 600s in ambient, vacuum, wet N2 and dry O2 environments. No oxidation of Co is

observed under a negative bias even after 600s in dry O2 and wet N2 for the non-hydrated device.

Figure 4.3(f) shows the MOKE hysteresis loop of a virgin hydrated device while figure 4.3(g)-(j)

Page 97: Dynamic Modulation of Material Properties by Solid State

91

show MOKE hysteresis loops of the hydrated device after VG = -3V has been applied for 600s in

ambient, vacuum, wet N2 and dry O2. The results show that the Co layer is completely oxidized

even in vacuum, indicating that (1) the oxidant is present in the GdOx layer, and (2) O2 gas is not

required for the oxidation process. These combined results can be explained using the schematic

in figure 4.3(k)-(j). When VG =-3V is applied to the top Au gate of a non-hydrated device, the Co

layer remains metallic due to the absence of any oxidant in the GdOx film. When VG =-3V is

applied to the top Au electrode of a hydrated device, H2O stored in the oxide film in the form of

Gd(OH)3 oxidizes Co to CoO (Eq. 4.3). The proton, H+, produced from the reaction is then

driven by the electric field through the GdOx layer to the top Au electrode, where it is reduced by

electrons, e- (flowing through the external circuit) to form hydrogen gas (Eq. 4.4). The net

reactions, depicted in figure 4.3(l), are shown below:

The reaction described by Eq. 4.4 is also known as the hydrogen evolution reaction (HER) 153,154.

With Co and Gd(OH)3 densities of 8.9g/cc and 6.0g/cc respectively, the oxidation of 0.9 nm of

Co would require the decomposition of only ~ 3 nm of Gd(OH)3. It is likely that Gd(OH)3 does

not completely transform to dry Gd2O3 during the Co oxidation process. Rather, the Gd(OH)3

should instead transform to a semi-hydrated GdOx with dissolved water, leading to a gradient in

the water content adjacent to the Co interface. We note that the interface reaction corresponding

to the case in which the GdOx adjacent to Co is not the fully transformed hydroxide phase but

rather a hydrated oxide phase would be described by

Anode: 2𝑂𝐻𝑂+ + 𝐶𝑜 → 𝐶𝑜𝑂 + 2𝐻+ + 𝑂𝑂

𝑥 + 𝑉𝑂2+ + 2𝑒− Equation 4.5

Anode: 2𝐺𝑑(𝑂𝐻)3 + 3𝐶𝑜 → 𝐺𝑑2𝑂3 + 3𝐻2𝑂 + 3𝐶𝑜

→ 𝐺𝑑2𝑂3 + 6𝐻+ + 3𝐶𝑜𝑂 + 6𝑒− Equation 4.3

Cathode: 6𝐻+ + 6𝑒− → 3𝐻2 Equation 4.4

Page 98: Dynamic Modulation of Material Properties by Solid State

92

In this case, the cathode reaction will be the same as Eq. 4.3. This reaction could be controlling

under conditions where the cell is exposed to humid environments, but not as high as in this

study where formation of hydroxide is observed. In both cases, the bottom Co acts as the anode

while the top Au acts as the cathode. Note that if oxygen were available at the cathode, the

hydrogen formed at the cathode would react with the oxygen to form water as described in

Chapter 5.

Page 99: Dynamic Modulation of Material Properties by Solid State

93

Figure 4.3. Effect of GdOx hydration on voltage induced Co oxidation. [(a)-(e)] MOKE

hysteresis loops of non-hydrated Pt(3nm)/Co(0.9nm)/GdOx(10nm)/Au(3nm) device in virgin

state (a) and after VG = -3V has been applied for 600s in ambient (b), vacuum (c), wet N2 (d),

and dry O2 (e). [(f)-(j)] MOKE hysteresis loops of hydrated

Pt(3nm)/Co(0.9nm)/GdOx(10nm)/Au(3nm) device in virgin state (f) and after VG = -3V has been

applied for 600s in ambient (g), vacuum (h), wet N2 (i), and dry O2 (j). [(k)-(l)] Schematic of

voltage-induced reaction in a non-hydrated(k) and hydrated (l) device.

Page 100: Dynamic Modulation of Material Properties by Solid State

94

4.4: H2 evolution during Voltage-induced

Co Oxidation in Pt/Co/GdOx

In order to further confirm the oxidation of Co by H2O in GdOx, we also fabricated

thicker Au (15 nm) electrodes on Pt/Co/GdOx to isolate the device from its surrounding

atmosphere. Figure 4.4(a) shows the hysteresis loop of a virgin device with thicker Au electrode

while figure 4.4(b)-(c) shows the hysteresis loops after VG = -3V has been applied for 600s in

ambient and in vacuum respectively. With thicker Au, we still observe complete oxidation of Co

in roughly the same time as the thinner 3 nm Au device (fig. 4.5). This further confirms that the

oxidant is stored in the GdOx layer.

Figures 4.4(d)-(e) show optical micrographs of a hydrated

Pt(3nm)/Co(0.9nm)/GdOx(10nm) device with 15nm thick Au top electrodes, before and after

applying VG = -3V for 600s to completely oxidize the Co layer . Gas bubble formation is clearly

observed in the device. In order to verify if the gas bubbles are formed at the top GdOx/Au

interface or bottom Co/GdOx interface, we also performed a similar experiment with a Pt(3

nm)/Co(0.9 nm)/GdOx(10 nm) device with 3 nm thick Au after VG = -3V is applied for 600s

(Figure 3f). In this case, no gas bubbles are observed because the electrode is porous and the

evolved hydrogen gas escapes to the surrounding61. This indicates that the bubbles seen in the

thicker Au electrode case are formed at the GdOx/Au interface, and this gas is necessarily H2,

since the Au acts as the cathode at VG = -3V. Note that gas bubble formation is also observed

when VG = -3V is applied for 600s to Pt/Co/GdOx/Au devices in vacuum. This hydrogen

evolution reaction was confirmed in the next chapter where Pd and Mg layers were inserted in a

Page 101: Dynamic Modulation of Material Properties by Solid State

95

substrate/Mg/Pd/GdOx/Au stack structure and the formation of PdHx and MgHx was confirmed

using XAS upon applying a gate bias. In this case, a positive bias (VG > 0) was applied to the top

Au electrode to insert hydrogen in the bottom Pd and Mg layers.

Figure 4.4. Hydrogen gas bubble evolution. MOKE hysteresis loops of hydrated Pt(3

nm)/Co(0.9 nm)/GdOx(10 nm)/Au(15 nm) device in virgin state (a) and after VG = -3V was

applied for 600 s in ambient (b) and vacuum (c). [(d)-(e)] Optical micrographs of Pt(3

nm)/Co(0.9 nm)/GdOx(10 nm)/Au (15 nm) devices before (d) and after (e) bias voltage

application (VG = -3V for 600 s) showing generation of hydrogen bubbles under the electrode. (f)

Optical micrograph of Pt(3 nm)/Co(0.9 nm)/GdOx(10 nm)/Au (3 nm) after applying VG = -3V for

600s. The scratch marks on the side of the Au electrodes are due to the CuBe probes.

Page 102: Dynamic Modulation of Material Properties by Solid State

96

Figure 4.5. Voltage-induced Co oxidation in Pt/Co/GdOx/Au heterostructures probed by

MOKE polarimetry. Number is parentheses represent thickness of Au top electrode.

Interestingly, some of the hydrogen which is produced at the top Au electrode under VG < 0

can be stored in the GdOx film. In order to demonstrate this, we compared the magneto-ionic

response of two Pt/Co(CoO)/GdOx/Au devices with a hydrated GdOx layer, where the Co layer

has been oxidized. In the first device, the Co layer is metallic in its as-deposited state, and is then

oxidized completely by first applying VG = -3V for 300s to the top 3nm Au (figure 4.6a). In the

second device, the Co layer is deposited in its oxidized state by reactive sputtering with oxygen

gas (figure 4.6d). A positive bias (VG = +3V or +2V) bias is then applied to both devices in

vacuum and in ambient in order to reduce the CoO layer to metallic Co. This process has

previously been shown to occur through injection of protons to the CoO layer, where they react

with CoO to reduce it to a metallic state155. Figure 4.6(b) and (c) show the results for the first

device in vacuum and ambient respectively, while figure 4.6(e) and (f) show the corresponding

results for the second device. For the first device, we can clearly see that some of the CoO is

reduced in vacuum. This implies that some hydrogen is stored in the device after the initial VG =

-3V is applied which allows for closed-system electrochemical and magnetic property switching

without the need for atmospheric exchange. The reduction of CoO to metallic Co by the stored

Page 103: Dynamic Modulation of Material Properties by Solid State

97

hydrogen occurs through the reverse reactions described in Eqs. 4.3 and 4.4. In this case, at VG >

0V, the stored hydrogen donates its electrons driving the reduction of CoO to metallic Co. On the

other hand, for the second device where the Co layer is oxidized during deposition, we do not

observe any CoO reduction at VG = +3V in vacuum (figure 4.6e). The CoO layer is only reduced

when the positive gate bias is applied in ambient, where humidity is present so that a water-

splitting reaction can occur to provide a source of protons. In both devices, a positive bias in

ambient initially results in Co with perpendicular magnetic anisotropy. The Co magnetization

then rotates in-plane as more hydrogen is accumulated at the bottom interface155.

Figure 4.6. Hydrogen storage in GdOx film. (a) MOKE hysteresis loop of hydrated Pt(3

nm)/Co(0.9 nm)/GdOx(10 nm)/Au(3 nm) after VG = -3V has been applied for 300s in ambient.

[(b)-(c)] MOKE hysteresis loops after VG = +3V is applied to device in (a) in vacuum (b) and in

ambient (c). (d) MOKE hysteresis loop of hydrated Pt(3 nm)/CoO(0.9 nm)/GdOx(10 nm)/Au(3

nm) in virgin state. [(e)-(f)] MOKE hysteresis loops after VG = +2V is applied to device in (d) in

vacuum (e) and in ambient (f).

Page 104: Dynamic Modulation of Material Properties by Solid State

98

4.5: In-situ XAS probe of Co during

Voltage-induced Co Oxidation in

Pt/Co/GdOx

To further verify the oxidation of Co through a direct chemical probe, we also performed in-situ

x-ray absorption spectroscopy on a hydrated Pd(10nm)/CoO(0.9nm)/GdOx(30nm)/Au(3nm)

device while applying gate biases under different atmospheric conditions. In this case, the

hydration treatment at 90oC and 525Torr is performed for only 24 hours in order to retain a non-

hydrated state at the CoO/GdOx interface. The first and second columns of figure 4.7 show the

data for Co L2, L3 edge22,23 and their corresponding first derivatives, while the third column

shows schematically the chemical state of the Co and the GdOx layer near the interface . In its

virgin state, the Co layer is initially oxidized (figure 4.7a). When VG = +3V is applied to the top

Au in vacuum, the CoO layer remains oxidized (figure 5b). However, when VG = +3V is applied

in 10Torr of PH2O, the CoO layer is reduced to metallic Co by H+ sourced from H2O (figure 4.7c)

155. The H2O that is produced from this reaction is reincorporated back into GdOx in the form of

Gd(OH)3. To confirm this, we next applied VG = -3V to the metallic Co device in vacuum. The

data shows partial reoxidation of the metallic Co back to CoO, consistent with the hypothesis.

Similarly, if VG = -3V is applied to the metallic Co in 10Torr of PH2O instead of vacuum, partial

reoxidation of the Co is also observed.

Page 105: Dynamic Modulation of Material Properties by Solid State

99

Figure 4.7. X-ray absorption spectra (XAS) of Co L2 and L3 edge. The first, second, and third

columns are the raw data, the first derivatives, and the chemical state of the Co/GdOx interface.

[(a)-(c)] XAS spectra of Pd(10 nm)/Co(0.9 nm)/GdOx (30 nm)/Au(3 nm) device in virgin state(a)

and after VG = +3V has been applied for 600 s in vacuum(b) and 10 Torr of PH2O (c)

respectively. The experiments from (a) to (c) are done sequentially. [(d)-(e)] XAS spectra of the

device in (c) after VG = -3V is applied for 600 s in vacuum and in 10 Torr of PH2O respectively.

Two different devices in (c) are used for experiment in (d) and (e).

Page 106: Dynamic Modulation of Material Properties by Solid State

100

Chapter 5: Magneto-

ionic Control of

Magnetism using a

Solid-state Proton

Pump

Page 107: Dynamic Modulation of Material Properties by Solid State

101

Most research efforts on magneto-ionics have focused on voltage-induced oxidation and

reduction of a ferromagnetic metal to reversibly modulate its magnetic anisotropy, exchange bias,

and magnetization58–65. The problem with oxidation-based magneto-ionics is that magnetic

property changes are accompanied by chemical and structural changes in the target ferromagnet.

This often leads to irreversibility62 and would be detrimental to devices such as magnetic tunnel

junctions whose performance depends critically on structure and electronic properties of the

ferromagnet.

Alternatively, Group I ions such as Li can be inserted into a target ferromagnet to alter

magnetic properties without changing the chemical phase or structure72,73,156. Small ion size and

the possibility of super-ionic conduction makes this a promising approach to achieving fast,

reversible magnetic property switching, but most Group I ions are incompatible with CMOS,

limiting their viability for practical applications. The exception is H+, which is relatively

innocuous, and is at the same time the simplest possible ion, making it ideal for inducing rapid

electric field driven property changes in solid-state structures.

Here we show that H2O hydrolysis in ambient atmosphere catalyzed by a rare-earth

oxide/noble metal interface can serve as a solid-state proton pump that enables non-destructive

magnetic property gating with a modest voltage. We demonstrate reversible 90o magnetization

switching in a thin Co film at room temperature by either inserting H+ at its interface with an oxide

or loading hydrogen into an adjacent heavy metal layer. The mechanism permits both unipolar

toggle switching and nonvolatile state retention, with no discernible irreversibility in magnetic

properties of the ferromagnet after >2000 cycles. Moreover, since heavy metals like Pt and Pd that

exhibit strong spin-orbit coupling are also well-known hydrogen storage materials157,158 that can

be driven between a metal and metal-hydride phase, a host of spin-orbit induced phenomena at

Page 108: Dynamic Modulation of Material Properties by Solid State

102

heavy-metal/ferromagnetic interfaces33,159,160 becomes accessible to voltage gating despite the fact

that electric fields cannot be applied directly.

Page 109: Dynamic Modulation of Material Properties by Solid State

103

5.1: Experimental Methods

Sample preparation: Ta(4 nm)/Pt(3 nm)/Co(0.9 nm)/GdOx(tGdOx nm)/Au(3 nm) layers were

fabricated on thermally oxidized Si (100) substrates using magnetron sputtering at room

temperature and 3mTorr Ar pressure. The metal layers were grown by DC sputtering. The GdOx

layer was deposited either using reactive sputtering with PO2 of 0.07mTorr or RF sputtering with

PO2 of 0.7mTorr O2. For the samples described in Fig. 5.1 with Co in the initially-oxidized state,

the Co layer was reactively sputtered with PO2 of 0.07mTorr O2 with a deposition time

corresponding to the time required to deposit 0.9 nm of metallic Co. For AHE and PHE

measurements, the structure is patterned into a Hall cross geometry with 500µm arm width and

with Au(3nm) deposited over the 0.25 mm2 active region to serve as a gate electrode. For MOKE

measurements, 200 µm diameter Au(3 nm) electrodes were patterned on top of the GdOx layer of

a continuous film, with the Ta(4nm)/Pt(3 nm) underlayer uncovered by GdOx at the sample edge

to allow electrical contact to the back. All patterning was done using shadow mask lithography.

Hall Effect measurements in different atmospheres: Anomalous Hall effect (AHE) and planar

Hall effect (PHE) measurements were performed using a lock-in amplifier with an ac injected

current of amplitude 2 mA and frequency 1kHz. For the AHE measurements, the field was swept

perpendicular to the plane; for the PHE measurements, the field was oriented in the sample plane,

at 45o to the current flow axis. AHE and PHE hysteresis loops were acquired using a 2s field sweep

time. The measurements were performed in a modified CVX-PF Lakeshore Probe Station with a

base vacuum pressure of ~10-4 mbar. Experiments under controlled gas environments were

performed by backfilling the chamber with either O2 gas (99.999% purity) or N2 gas (99.999%

purity). Humidity was introduced into the N2 gas flow by bubbling through water. Wet N2 and

Page 110: Dynamic Modulation of Material Properties by Solid State

104

ambient condition at 25C corresponds to ~20mT and 12mT of H2O partial pressure respectively.

All experiments were performed at room temperature.

Polar magneto-optical Kerr effect (MOKE) measurements: MOKE measurements were

performed using a 1 mW laser with a wavelength of 660 nm focused to spot size of ~10µm.

Experiments were performed in polar geometry and hence sensitive to the out-of-plane

magnetization component. To apply gate voltage VG to the circular electrodes, a CuBe probe was

landed near the edge of the electrode and the Ta/Pt back electrode was grounded. The laser spot

was focused at the middle of the electrode. All experiments were performed at room temperature.

X-Ray absorption spectroscopy (XAS): XAS data was taken at Coherent Soft X-ray Scattering

(CSX) beamline at the National Synchrotron Light Source II, Brookhaven National Laboratory

using fluorescent yield. The incident soft x-ray beam has a footprint of ~200µm and the sample is

tilted 15º relative to the incident beam. The sample used for the measurement has a hall bar

geometry with sample structure Ta(4nm)/Pd(3nm)/Co(0.6nm)/Pd(4.5nm)/GdOx (30nm) and a 3nm

Au top gate. The main chamber base pressure is ~2 x 10-9 torr, and the sample is kept at 100K

throughout the measurement.

Page 111: Dynamic Modulation of Material Properties by Solid State

105

5.2: Co Redox through Water Electrolysis

For a Pt/Co/GdOx/Au nominal structure, we have shown in chapter 4 that it is H2O stored in GdOx

in the form of Gd(OH)3 that oxidizes Co to CoO when VG < 0 is applied to the top Au. In this

section, we will show that for a Pt/CoO/GdOx/Au structure , a positive gate voltage (VG > 0) can

electrochemically split atmospheric water and pump protons through the GdOx, to both reduce

CoO to metallic Co and to modulate the magnetic anisotropy of a metallic Co thin film.

Figure 5.1 shows the effect of applying a gate bias to Pt/Co/GdOx/Au under several atmospheric

conditions, demonstrating the critical role of ambient moisture in magneto-ionic control of Co

oxidation state and magnetic properties. We used a sample structure

Ta(4nm)/Pt(3nm)/CoO(0.9nm)/GdOx(30nm)/Au(3nm) sputter deposited on thermally oxidized Si.

A positive gate voltage (VG > 0) was applied to the top Au while the out-of-plane and in-plane

magnetization was monitored electrically through the anomalous Hall effect (AHE) and planar

Hall effect (PHE), respectively, using a Hall bar geometry (Fig. 5.1c).

Figures 5.1d-h show out-of-plane hysteresis loops probed through the AHE resistance (RAHE)

for the initially-oxidized sample. Results are shown for the virgin state (Fig. 5.1d), and after

applying VG = +3V for 1000s at room temperature in various atmospheres (Figs. 5.1e-h).

Consistent with literature60,61, under ambient atmosphere, a positive bias results in the appearance

of out-of-plane magnetization, corresponding to the reduction of nonmagnetic CoOx to metallic

Co with perpendicular magnetic anisotropy (PMA) (Fig. 5.1e). VG has no effect under vacuum

(Fig. 5.1f), even though a lower oxygen partial pressure (PO2) environment should make oxygen

extraction more favorable. Likewise, in dry O2 (Fig. 5.1g), no magnetic changes are observed, but

Page 112: Dynamic Modulation of Material Properties by Solid State

106

remarkably, under wet N2 (Fig. 5.1h), positive VG leads to the appearance of PMA, implying water-

assisted reduction of CoOx to Co. These results confirm the findings in chapter 4 that moisture and

proton transport through GdOx play a crucial role in voltage-induced redox of Co and that oxygen

ion migration plays an insignificant role.

Figure 5.1. In-situ probing of magneto-ionic switching in different atmospheres. a-b. Active

region of a Pt/CoO/GdOx (a) and Pt/Co/GdOx (b) device. c, Hall cross geometry used for

anomalous Hall effect (AHE) measurement to probe out-of-plane magnetization. d, AHE

hysteresis loop of virgin structure in (a). e-h, Hysteresis loops after applying a gate voltage VG =

+3V for 1000s in ambient (e), vacuum (f), dry O2 (g), and wet N2 (h).

It is well-known that water splitting reaction can be catalyzed by noble metals with oxide

support112,161,162. Water splitting and hydrogen incorporation had been shown to significantly

impact the switching behavior and electronic properties in metal/oxide/metal memristors163–165.

During reduction of CoOx under positive VG in a Pt/Co/GdOx/Au heterostructure, H2O is

hydrolyzed at the top electrode producing H+ and O2 through the oxygen evolution reaction111,112.

Page 113: Dynamic Modulation of Material Properties by Solid State

107

The proton is transported to the bottom electrode via a Grotthuss-type mechanism81,82,137 in which

the H+ ion hops between adjacent lattice oxygen atoms. At the bottom electrode, the proton reacts

with CoO to form Co and H2O. The net reactions, depicted in Fig. 5.2a, are

The water evolved at the bottom is then incorporated into GdOx itself as hydroxide, namely through

the reaction146: 𝐺𝑑2𝑂3 + 3𝐻2𝑂 → 2𝐺𝑑(𝑂𝐻)3 . Co oxidation under negative VG occurs by the

reverse process (Fig. 5.2b), with Pt acting as the anode for Co oxidation and Au acting as the

cathode for H2O recombination. Note that this H2O recombination reaction only occurs under the

presence of O2. If O2 is absent, a hydrogen evolution as described in chapter 4 will take place.

Figure 5.2 Electrochemical reactions in a magneto-ionic cell . a, Schematic of CoO reduction

at gate voltage VG > 0 involving H2O hydrolysis. b, Schematic of Co oxidation at gate voltage VG

< 0 involving H2O recombination.

Anode (Au) : 2𝐻2𝑂 → 4𝐻+ + 𝑂2 + 4𝑒− Equation 5.1

Cathode (Pt) : 4𝐻+ + 2𝐶𝑜2+ + 2𝑂2− + 4𝑒− → 2𝐻2𝑂 + 2𝐶𝑜. Equation 5.2

Page 114: Dynamic Modulation of Material Properties by Solid State

108

To confirm the role of ambient water as the source of proton in the reduction of CoO to

Co, we also looked at the effect of Au electrode thickness on the rate of voltage-induced CoO

reduction. Figure 5.3a shows the transient magneto-optical Kerr effect (MOKE) magnetization

remanence ratio, Mr/Ms at the center of a device with structure Pt(3nm)/CoO(0.9nm)/GdOx

(20nm)/Au (t =3,4,6,10 nm) as VG = +3V is applied. We observe that the voltage induced CoO

reduction becomes slower with increasing thickness of Au, confirming that hydrogen is sourced

from H2O in ambient during the reduction process. The reduction activity is highest at the

electrode edge where the thickness of Au is thinner due to shadowing effect from shadow mask

lithography. This can be seen in the optical micrograph in Figure 5.3b for a t = 4nm Au

electrode. The bottom plot in Figure 5.3b shows a snapshot of Mr/Ms across the electrode after

VG = +3V for ~600s. One can see that CoO is reduced to Co by H injection at the edges (Fig.

5.3c). With increasing bias dwell time, this region of modified properties moves inward towards

the center. Figure 5.3d shows the transient (VG = +3V) for a t = 6nm Au electrode at the edge vs

the center, where a relatively constant Mr/Ms followed by a sharp increase further indicates a

circular “front” of changed properties that is slowly moving inward from the edge towards the

center.

Page 115: Dynamic Modulation of Material Properties by Solid State

109

Figure 5.3 Voltage-driven injection of hydrogen at edge of electrode. a, Magneto-optical

Kerr effect (MOKE) magnetization remanence ratio, Mr/Ms versus time as VG =+3V is applied to

a Pt(3nm)/CoO(0.9nm)/GdOx (20nm)/Au (t nm) device. b, Top: Optical micrograph of a t = 4nm

electrode. Bottom: Snapshot of Mr/Ms across the electrode at ~600s. c, Hysteresis loops

corresponding to center and edge of electrode in b at ~600s. d, Magnetization transient for t =

6nm electrode at the edge vs the center.

If positive Vg pumps H+ as proposed, sustained bias application should lead to hydrogen

accumulation and evolution at the bottom electrode, and this is indeed observed. Figure 5.4a shows

scanning electron microscope (SEM) of a Pt(3nm)/GdOx(100nm)/Au(3nm) film after applying VG

= +3V for 5 hours. The inset optical micrograph shows bubble formation, which after cross-

sectioning is revealed to arise from gas evolution and GdOx delamination at the bottom electrode,

Page 116: Dynamic Modulation of Material Properties by Solid State

110

which could only be H2 since O2 evolution could only occur at the anode. Figure 5.4b shows the

voltage threshold for the onset of PMA in a CoO structure to be ~1.5V, which agrees well with the

proposed reactions in equation 5.1 and 5.2111–113.

Figure 5.4 Cross section SEM to image hydrogen evolution. a, Cross-sectional scanning

electron microscope image of a Pt (3nm)/GdOx (100nm)/Au (3nm) device after VG = +3V has

been applied for 5 hours. Inset shows optical micrograph of the device. Hydrogen gas is

produced between the Pt and GdOx layer at the end of the experiment. b, Out-of-plane remanent

magnetization ratio, Mr/Ms of Pt(3nm)/Co(0.9nm)/GdOx(tGdOx)/Au(3nm) structure after applying

various VG for > 5 hours, for tGdOx = 4 nm and 30 nm.

Figures 5.5a-b show cyclic voltammetry (CV) data for a Pt(3nm)/GdOx(20nm)/Au(3nm) structure

under different atmospheric conditions (Fig. 5.5a) and at different sweep rates (Fig. 5.5b). We

assume that the top Au electrode acts as the working electrode while the bottom Pt electrode acts

as the counter/reference electrode166. When cyclic voltammetry is performed under different

atmospheres (Fig. 5.5a), significant anodic and cathodic currents are observed only when H2O is

present. Both currents are governed by the voltage sweep rate (v) (Fig 5.5b) which indicates a

process that is kinetically limited by mass transport rather than charge-transfer166. From both

information, we can conclude that water hydrolysis (VG > 0) and recombination (VG < 0) are indeed

Page 117: Dynamic Modulation of Material Properties by Solid State

111

taking place111–113. Despite the lack of a unique reference electrode, the classical exponential

behavior observed in oxidation is consistent with oxygen evolution reaction (OER) while the peak

observed in reduction is typical of oxygen reduction reaction (ORR)107,113.

Figure 5.5. Dependence of overall reaction rate on charge transfer and mass transport. a, Cyclic Voltammetry (CV) plot under different atmospheric conditions performed at v =

10mV/s.b, CV plot at different sweep rates performed in ambient.

Page 118: Dynamic Modulation of Material Properties by Solid State

112

5.3: Modulation of Magnetic Anisotropy

through Proton Injection

We now show that hydrogen insertion at the Co/GdOx interface allows the anisotropy to be

toggled from out-of-plane to in-plane without requiring redox reactions in the ferromagnetic Co

(Fig. 5.6a). Surface anisotropy is known to be sensitive to adsorbed H, as shown previously for

ultrathin ferromagnetic films in ultra-high vacuum upon exposure to molecular or atomic

hydrogen24–28. Here, we show the same behavior can be gated in solid-state devices. Starting from

a virgin state with PMA (Fig. 5.6b), the magnetization rotates in-plane when VG = +3V is applied

for 800s (Fig. 5.6c), corresponding to accumulation of hydrogen at the Co/GdOx interface. When

VG is set to 0V (grounded), PMA is spontaneously recovered (Fig. 5.6d) as the accumulated

hydrogen forms H+ and diffuses away from the bottom electrode. In-plane magnetization

reorientation is confirmed by Figs. 5.6e-g, which show that the PHE signal is absent when the film

has PMA and is present when the AHE signal vanishes (Figs. 5.6e-g) under positive bias. Figures

5.6h-j show cycling results for a device with tGdOx=4 nm, in which switching is much faster. VG

was cycled >2000 times between +3V and 0V at 0.5 Hz, and out-of-plane hysteresis loops were

acquired at 25 ms intervals using a polar magneto-optical Kerr effect (MOKE) polarimeter (see

Methods). Figure 5.6h shows the ratio of the remnant (Mr) to saturation magnetization (Ms) as a

function of time, for cycles 1-10 and 2060-2070, tracking the in-plane/out-of-plane transitions

(Fig. 5.6i). The square out-of-plane loop in the virgin state (Fig. 5.6i) is indistinguishable from

the loop after toggling the magnetization in plane, both after the first cycle (Fig. 5.6i) and after

cycle 2070 (Fig. 5.6j). We find, however, that the response time degrades slightly with repeated

cycling (Fig. 5.6h), which may be associated with increased leakage currents in the oxide. The

Page 119: Dynamic Modulation of Material Properties by Solid State

113

switching times at the rising and falling VG edge are 100ms and 400ms respectively (Fig. 5.7). This

switching speed is faster than any room temperature results in the literature for magneto-ionic

switching60,61. The asymmetry in switching can be mitigated by applying a negative VG to

accelerate H+ removal from the interface, but this can also lead to Co oxidation, which leads to a

progressive irreversible degradation of PMA due to irreversibility of oxygen insertion into the

magnetic layer62.

Figure 5.6. Magneto-ionic switching based on hydrogen accumulation at Co/GdOx

interface: a, Schematic of magneto-ionic switching scheme. b-d, AHE hysteresis loops in virgin

state (b), after VG = +3V is applied for 800s (c), and after VG is set to 0V for 800s (d). e-g, PHE

hysteresis loops corresponding to b-d respectively. h, Magnetization remnance ratio Mr/Ms

versus time as VG is cycled between +3V and 0V at 0.5 Hz for 2070 cycles, extracted from

hysteresis loops measured by the polar magneto-optical Kerr effect (MOKE). Results are shown

for the first and last ten cycles. i, Out-of-plane hysteresis loops corresponding to the virgin state

and the first switching cycle. j, Out-of-plane hysteresis loops corresponding to cycle 2070. The

hysteresis loop of the final relaxed state is identical to that in the virgin state.

Page 120: Dynamic Modulation of Material Properties by Solid State

114

Figure 5.7 Magneto-ionic switching speed of a Pt/Co/GdOx structure. a-b, Polar magneto-

optical Kerr effect magnetization remanence ratio Mr/Ms at rising (a) and falling edge (b) of VG.

To quantify the change in magnetic anisotropy energy induced by hydrogen insertion, we also

probed the magnitude of the out-of-plane magnetization in the virgin state as a function of an in-

plane field. The strength of the in-plane field required to tilt the magnetization in-plane would tell

us how large the starting perpendicular magnetic anisotropy is. Since the magnetization rotates

from out-of-plane to in-plane under VG > 0, the voltage-induced change in anisotropy would be at

least the magnitude of this starting perpendicular magnetic anisotropy. Figure 5.8a shows the out-

of-plane hysteresis loop probed through the anomalous Hall effect resistance (RAHE) for a Pt(3nm)/

Co(0.9nm)/ GdOx (30nm) structure while figure 5.8b shows RAHE as a function of in-plane field,

Hx applied along the current injection line for the virgin state. The red curve shows a fit to the

single-domain Stoner-Wohlfarth model, RAHE = Rsat cos(arcsin(Hx/HK)) in order to obtain the in-

plane saturation field, HK. Rsat is the AHE resistance when the magnetization is saturated in the

out-of-plane direction. The fitted HK is ~8.2kOe, and assuming Ms of 1400emu/cc for Co, the

uniaxial magnetic anisotropy, Ku = (HkMs)/2 is 5.7x106 erg/cc. Normalizing by the thickness of

the Co layer, tCo (0.9nm), we obtain interfacial magnetic anisotropy, Ks = tCo (HkMs)/2 of

0.52erg/cm2 for the device in virgin state. This value represents the lower bound for the change in

Page 121: Dynamic Modulation of Material Properties by Solid State

115

Ks when the magnetization rotates from out-of-plane to in-plane under VG > 0. With an electric

field of 1MV/cm (VG = +3V), this corresponds to a magnetoelectric efficiency of > 5200fJ/ Vm.

Figure 5.8. Quantification of magnetic anisotropy energy. a, Out-of-plane hysteresis loop

probed through the AHE resistance (RAHE) for a Pt(3nm)/ Co(0.9nm)/ GdOx (30nm) structure. b,

RAHE as a function of in-plane magnetic field. The red line shows the fit for a single domain

Stoner-Wohlfarth model.

Similar to the reduction of CoO described above, the accumulation of hydrogen which causes

this magnetization rotation in-plane only occurs in the presence of humidity. Figure 5.9a-b show

the rate of change in out-of-plane and in-plane magnetization probed using the anomalous Hall

effect (AHE) and planar Hall effect (PHE) as VG = +3V is applied to a Pt/Co/GdOx Hall cross

device under different atmospheric conditions (tGdOx = 30nm). The data clearly indicates that the

magnetization rotates from out-of-plane to in-plane only in the presence of moisture, again

confirming that H2O in the environment acts as the source of hydrogen which is accumulated at

the Co/GdOx interface167. Figures 5.9c-e show the corresponding AHE hysteresis loops in the

virgin state (Fig. 5.9c) and after VG = +3V is applied for 800s in vacuum (Fig. 5.9d) and wet N2

(Fig. 5.9e). Figures 5.9f-h show the corresponding PHE hysteresis loops.

Page 122: Dynamic Modulation of Material Properties by Solid State

116

Figure 5.9. Effect of positive VG in different atmospheric conditions starting from a Co

virgin state. a, Anomalous Hall effect resistance (RAHE) vs time at VG = +3V. b, Planar Hall

resistance (RPHE) vs time at VG = +3V. c-e, AHE hysteresis loops of the device in virgin state (c),

and after VG = +3V is applied for 800s in vacuum (d)and in wet N2 (e) . f-h, PHE hysteresis

loops of the device corresponding to the condition in c-e respectively.

Page 123: Dynamic Modulation of Material Properties by Solid State

117

5.4: Magnetic Response under Short

Circuit and Open Circuit

The electrochemical reactions at VG =+3V in the two half cells (top and bottom electrodes),

given by

cannot proceed without electron flow through the external circuit. Thus, one can realize two

operating modes by either grounding the device (VG = 0V) or leaving the device at open circuit

when VG is removed. Figure 5.10a shows that after VG = +3V is applied, the magnetization

transitions from out-of-plane (Fig. 5.10b) to in-plane (Fig. 5.10c), and when VG is set to open

circuit, the in-plane state is retained (Figs. 5.10c,d). This implies that the accumulated hydrogen

remains at the interface. When the device is subsequently set to closed circuit, PMA is

spontaneously recovered (Fig. 5.10e). Hence, the magnetization state can be switched in a

nonvolatile fashion between out-of-plane and in-plane states, or toggled with a unipolar voltage,

depending on whether the VG=0 condition is at open or closed circuit.

Anode (Au) : 2𝐻2𝑂 → 4𝐻+ + 𝑂2 + 4𝑒− Equation 5.3

Cathode (Pt) : 4𝐻+ + 4𝑒− → 4𝐻 Equation 5.4

Page 124: Dynamic Modulation of Material Properties by Solid State

118

Figure 5.10. Magnetic response under short circuit and open circuit. a, Evolution of Mr/Ms

vs time and the corresponding VG. Dashed line indicates open circuit while solid line at VG = 0V

indicates short circuit. VG = +3V is applied between time,t = 30s to t = 100s, at which point the

probe is lifted from the top electrode. At t = 5300s, the probe is landed again and VG set to 0V

(ground). b-e, Polar MOKE hysteresis loops corresponding to t = 25 s (b), t = 110 s (c), and t =

4200 s (d), and t = 5600s (e).

Such ability to modify the magnetic state over two very distinct timescales can be particularly

attractive for neuromorphic computing where one can modify the memory168,169 (non-volatile)

during computation and once the computation is complete, all memory elements are reinitialized

to the same state during a refresh stage (volatile).

These results show that removal of H from the bottom electrode requires the reactions in Eqs.

5.3 and 5.4 to occur in reverse. To be removed, H must first split into H+ and e- at the bottom

electrode, so that the H+ can be transported back up through the GdOx to the top electrode, where

it recombines with atmospheric oxygen to form H2O. If there is no electronic conduction path, the

reaction cannot proceed and as a result the in-plane magnetized state is retained at open circuit.

Hence, a finite leakage current through the GdOx reduces the stability of the hydrogen-loaded state

as it offers an alternate path for electron transfer from the bottom electrode to the top electrode.

Page 125: Dynamic Modulation of Material Properties by Solid State

119

From this, we show that H cannot simply diffuse from the bottom electrode; its insertion and

removal are governed by the anodic and cathodic electrochemical reactions in Eqs. 5.3 and 5.4.

Page 126: Dynamic Modulation of Material Properties by Solid State

120

5.5: Electrical Gating of Magnetic

Anisotropy at a Heavy-

metal/ferromagnet Interface

Since Pd is well known for its hydrogen loading capacity157, and atmospheric hydrogen loading

in Pd/Co/Pd has previously been shown to modulate PMA26, we next exploit this feature by

inserting a Pd layer between Co and GdOx. Figure 5.11a shows the layer schematic of a

Ta(4nm)/Pd(3nm)/Co(0.6nm)/Pd(4.5nm)/GdOx(10nm)/Au(3nm) heterostructure. Because the Co

layer is protected by Pd, negative VG does not result in oxidation of Co (Fig. 5.11b)), which allows

for applying negative gate bias to accelerate H removal and recovery of PMA. In this device,

positive and negative VG can pump hydrogen into and out of the Pd layer reversibly, switching the

anisotropy from out-of-plane to in-plane and back26 (Figs 5.11c-e). Figure 5.11f shows the

switching cycles of Mr/Ms as VG is cycled between +4V and -1V at 1Hz. Robust switching was

achieved, with a switching time of ~150 ms at both the rising and falling edges (Figs 5.11g and h).

Page 127: Dynamic Modulation of Material Properties by Solid State

121

Figure 5.11. Voltage gating of metal-metal interface by exploiting hydrogen loading in Pd.

a, Schematic of device operation in a Pd/Co/Pd/GdOx cell. b, Out-of-plane magnetization

remanence ratio Mr/Ms versus time at gate voltage VG = -3V. Inset shows polar magneto-optical

Kerr effect (MOKE) hysteresis loops at t = 0s and t = 4500s. c-e, Polar MOKE hysteresis loops

corresponding to virgin state (c), after VG = +4V (d), and after VG is set to -1V (e). f, Mr/Ms as VG

is cycled between +4 V and -1 V at 1 Hz. Each data point corresponds to a 25 ms MOKE

hysteresis loop g, Mr/Ms at rising edge of VG. h, Mr/Ms at falling edge of VG.

In order to directly evidence insertion of H into the heavy metal layer, we performed x-ray

absorption spectroscopy (XAS) at the 23-ID-1 beamline at National Synchrotron Light Source II.

Figure 5.12a shows an XAS spectrum of a virgin

Ta(4nm)/Pd(3nm)/Co(~0.6nm)/Pd(4.5nm)/GdOx(30nm)/Au(3nm) sample in the range between

525eV and 575eV, where one can observe the Pd M3 edge170 at ~532eV and O K-edge at

~538eV. Figure 5.12b shows comparison of the Pd M3 edge between a virgin and voltage-

Page 128: Dynamic Modulation of Material Properties by Solid State

122

modified device. For the voltage-modified device, VG = +3V is applied for >5min ex-situ before

the XAS spectrum was measured. From the data, one can clearly observe a peak shift of ~+0.7eV

for the Pd M3 edge upon voltage application. This shift arises from two sources. First, previous

XPS studies have shown that the binding energy of core electrons shifts slightly when Pd

becomes PdHx. This shift is ~+0.17eV for the 3d electrons and its magnitude is very small

because hydrogen in Pd is well screened in the lattice171. Secondly, there is a change in energy of

the final unoccupied 4d states of Pd by ~ +0.7eV when it becomes a hydride172,173. This change

accounts for most of the peak shift in the XAS spectrum. The total peak shift of the Pd M3 edge

should hence be ~+0.87eV, which is very close to the +0.7eV peak shift we observe in our data.

Figure 5.12 X-ray absorption spectroscopy of Pd M3 edge. a, X-ray absorption spectrum of a

Pd (3nm)/Co (0.6nm)/Pd (4.5nm)/GdOx (30nm)/Au(3nm) sample in the range between 525eV

and 575eV. b, X-ray absorption spectra of a virgin and voltage-modified (VG = +3V for >5min)

sample, clearly indicating an energy shift in the Pd M3 peak for the voltage-modified sample.

We have also confirmed this with a thick Mg as the hydrogen loading layer174 in a

Ti(3nm)/Mg(40nm)/Pd(5nm)/GdOx (30nm)/Au(3nm) structure. Fig 5.13 shows X-ray absorption

spectra the device in the virgin state and after a positive bias is applied to the top electrode in

ambient (VG =+3V, for 5 minutes). Comparison of the XAS data of the biased device and the

literature data174 shows that the Mg layer becomes hydrided upon application of a positive gate

Page 129: Dynamic Modulation of Material Properties by Solid State

123

bias, VG = +3V. We should note that before such a thick Mg (40nm) can be loaded with

hydrogen, the entire Pd layer will have to be penetrated by hydrogen. Such a substantial amount

of hydrogen loaded into the system is far more than what would be required to explain the

magnetic property changes observed in the thin film structures in our experiments.

Figure 5.13 X-ray absorption spectra of Mg K-edge. Mg K-edge of a Ti(3nm)/Mg (40nm)/Pd

(5nm)/GdOx (30nm)/Au(3nm) structure in virgin state (black) and after VG = +3V is applied for

5min (red). Mg K-edge of the biased device shows remarkable similarity to that of MgHx174.

With this, we show for the first time that the magnetic anisotropy at a metal/metal interface can

be modulated substantially by an electric field using electrochemical gating of a metal to its

hydride phase in an appropriately designed solid state heterostructure. This effect could never be

achieved by any other known mechanism since electric field vanishes in a metal.

Page 130: Dynamic Modulation of Material Properties by Solid State

124

5.6 Comparison between Au and Pt Top

Electrodes For the Pt(4nm)/Co(0.9nm)/GdOx(4nm) structure described in chapter 5.3 (figure 5.6), we also

compared the difference in magneto-ionic switching speed between a 3nm Au and 3nm Pt top

electrode. Both electrodes are deposited on the same Pt/Co/GdOx film. The MOKE results for

magnetic switching in both devices at VG = +3V and 0V cycled at 0.5Hz are shown in figure

5.14. Similar to the Au devices, the Pt devices demonstrate uniform 90° switching with excellent

cyclability. Because the proton for magnetic gating is sourced from H2O in ambient, Pt devices

would be expected to toggle much faster than Au devices due to its higher activity for water

splitting. However, this is only true for the switching process from in-plane to out-of-plane at VG

= 0V. During switching from out-of-plane to in-plane state, the speed of the Pt device is

surprising slower. The reason for this is still unclear at this point.

Page 131: Dynamic Modulation of Material Properties by Solid State

125

Figure 5.14. MOKE switching with VG cycled between +3V and 0V at 0.5Hz. (a) Comparison

of 10 switching cycles between Au and Pt top electrodes. (b) Switching transient at falling edge

(left) and rising edge (right) of VG.

Page 132: Dynamic Modulation of Material Properties by Solid State

126

Chapter 6: Voltage

Gating of Magnetic

Damping and Spin-

Orbit Torques using

Proton

Page 133: Dynamic Modulation of Material Properties by Solid State

127

In heavy metal/ ferromagnet/oxide heterostructures, a rich set of properties such as spin-orbit

torques175, Dzyaloshinskii-Moriya interaction (DMI)159, and magnetic damping176 can affect a

wide range of behaviors in a ferromagnet. For instance, spin-orbit torques determine how

efficient the injected spin current can drive magnetization precession or switch the

magnetization. DMI affects how domain walls are configured and how fast domain wall can be

driven by a current177. Magnetic damping on the other hand represents an energy loss mechanism

which determines dynamic properties like spin wave propagation and relaxation timescale for

magnetic precession. An effective mean to gate these fundamental properties would allow a wide

range of device behaviors to be accessible to control which would significantly boost the

functionalities and performance of current magnetic devices.

A promising approach to voltage gating is through voltage-induced ionic modulation of magnetic

interfaces, dubbed the magneto-ionic effect58–65. In this approach, a gate voltage drives ionic

migration and electrochemical reactions which in turn changes the properties of the magnetic

layer. Two ions that have been studied are oxygen ions and protons. In the former case, it has

been shown that voltage-induced Co oxidation in a Pt/Co/GdOx can remove its magnetization,

upon which a voltage of the opposite polarity can reverse the effect. For the latter case, voltage-

induced hydrogen accumulation near the metallic Co layer in a Pt/Co/GdOx structure changes its

interfacial magnetic anisotropy which results in the magnetization rotation from out-of-plane

state to in-plane (chapter 5). Thus far, the main interest has been to modulate magnetic

anisotropy in order to reduce barrier for switching. However, one should expect ions to change

other interfacial magnetic properties too.

In this chapter, we demonstrate voltage-induced ionic gating of magnetic damping and spin-orbit

torques in addition to magnetic anisotropy, probed by spin-torque ferromagnetic resonance39,40.

Page 134: Dynamic Modulation of Material Properties by Solid State

128

Spin torque ferromagnetic resonance is a very useful technique where a current-induced torque

drives magnetic resonance which manifests itself electrically through a change in resistance. This

can then be measured as a DC mixing voltage, 𝑉𝑚𝑖𝑥 which can be expressed as39,136

𝑉𝑚𝑖𝑥 = 𝑆𝑊2

(𝐻− 𝐻𝐹𝑀𝑅)2+𝑊2+ 𝐴

𝑊(𝐻− 𝐻𝐹𝑀𝑅)

(𝐻− 𝐻𝐹𝑀𝑅)2+𝑊2 Equation 6.1

Here, 𝑆 and 𝐴 are the amplitudes of the symmetric and antisymmetric Lorentzian components of

the resonance peak respectively, 𝑊 and 𝐻𝐹𝑀𝑅 are the width and position of both these

Lorentzian components.

Page 135: Dynamic Modulation of Material Properties by Solid State

129

6.1 Experimental Methods

Experiments focus on Ta(4nm)/Pt(3nm)/Co(tCo)/GdOx (36nm)/Au(3nm) structures which were

fabricated on thermally oxidized Si (100) substrates using magnetron sputtering at room

temperature and 3mTorr of Ar pressure. The metal layers were grown by DC sputtering. The

GdOx layer was deposited either using reactive sputtering with PO2 of 0.07mTorr or RF

sputtering with PO2 of 0.7mTorr O2. The base pressure during sputtering is 5x10-7 Torr. Figure

6.1(a) shows the device and measurement schematic for the gated ST-FMR device while figure

6.1(b) shows an optical micrograph from a top-down view. The waveguide has dimension of

10µm x 5µm, 20µm x 10µm, 10µm x 10µm, and 20µm x 5µm. The in-plane field, H is applied

45º to the current line. 15dbm of power at frequencies, f ranging from 5.5GHz to 12GHz is

supplied by a signal generator to the waveguide while 𝑉𝑚𝑖𝑥 is measured by a lock-in amplifier.

The modulation amplitude and frequency from the lock-in amplifier (to the signal generator) are

1V and 2kHz respectively. Figure 6.1(c) shows a set of FMR spectra measured at different f for

a tCo =4.8nm device. These spectra are fitted with two Lorentzian (symmetric and

antisymmetric) components according to equation 6.1. HDemag is then obtained from a fit of 𝐻𝐹𝑀𝑅

to the Kittel formula given by equation 6.2.

2𝜋𝑓 = 𝛾 √𝐻𝐹𝑀𝑅(𝐻𝐹𝑀𝑅 + 𝐻𝐷𝑒𝑚𝑎𝑔) Equation 6.2

A sample fit is shown in figure 6.1(d). Similarly, the widths, W of the FMR spectra at different

frequency can also be fitted to equation 6.3 to obtain the damping constant, 𝛼. The fit is shown in

figure 6.1(e).

𝑊 = 2𝜋𝑓

𝛾𝛼 Equation 6.3

Page 136: Dynamic Modulation of Material Properties by Solid State

130

For in situ probing of the magnetic properties during voltage gating, 𝑉𝑚𝑖𝑥 is continuously

measured at 8GHz(15dbm) of injected power while a constant gate bias VG is applied. The FMR

spectrum for each acquisition cycle is then fitted with the two Lorentzian components (equation

6.1) so that the change in magnetic parameters can be plotted as a function of time. For fittings

which require multiple frequencies such as the damping constant, the spectra for different

frequencies were acquired during specific times by setting the gate bias to a holding voltage.

Figure 6.1. Voltage gating of magnetic properties probed using spin-torque ferromagnetic

resonance (ST-FMR). (a) Device schematic of a 10µm x5µm gated device. 𝑉𝑚𝑖𝑥 is measured

using a lock-in amplifier at modulation frequency of 2kHz. (b) Optical micrograph of the gated

device for ST-FMR measurement. The magnetic field is applied at 45° to the current line. The

contact pad at the top left corner is connected to the top gate. (c) ST-FMR spectra of a

Pt(3nm)/Co(4.8nm)/GdOx (36nm)/Au(3nm) device at different frequencies. All the spectra are

fitted to equation 6.1 to obtain S,A, 𝑊, and 𝐻𝐹𝑀𝑅 (d) Fit of 𝐻𝐹𝑀𝑅 to the Kittel equation

(equation 6.2) to obtain HDemag. (e) Fit of W to equation 6.3 to obtain α.

Page 137: Dynamic Modulation of Material Properties by Solid State

131

6.2 Spin Torque Ferromagnetic

Resonance to Probe Voltage Gating of

Spin Orbit Torque and Magnetic Damping

Figure 6.2(a) shows the ST-FMR spectra at 8GHz for different Co thicknesses. The spectra show

three trends with increasing Co thickness (1) 𝑊 decreases, (2) HDemag shifts to smaller field

values, and (3) there is an increasing contribution from an antisymmetric Lorentzian component.

To quantify these trends, the Kittel formula (Equation 6.2) is first fitted for different frequencies

at each Co thickness to obtain 𝐻𝐷𝑒𝑚𝑎𝑔. 𝐻𝐷𝑒𝑚𝑎𝑔 as a function of Co thickness is shown in figure

6.2(b). At small thicknesses, 𝐻𝐷𝑒𝑚𝑎𝑔 is small at ~3kOe due to the large interfacial perpendicular

anisotropy at the Pt/Co and Co/GdOx interfaces. At larger thicknesses (>6nm), the demagnetizing

field approaches ~14kOe which is equivalent to 4πMCo. This indicates shape anisotropy is much

larger than the interfacial anisotropy and dominates at such thicknesses. Figure 6.2(c) shows the

trend in magnetic damping constants as a function of Co thickness. The data shows decreasing

damping constant with increasing Co thickness, consistent with previous literature. To get the

spin orbit torque quantities, the values obtained from the Lorentzian fits are substituted in

equation 6.4 to obtain 𝜉𝐹𝑀𝑅.

𝜉𝐹𝑀𝑅 = 𝑆

𝐴

𝑒𝜇0𝑀𝑆𝑡𝐶𝑜𝑡𝑃𝑡

ℏ√1 +

4𝜋𝑀𝑆

𝐻𝐹𝑀𝑅 Equation 6.4

Here 𝜉𝐹𝑀𝑅 is the spin hall angle in the absence of a field-like torque from the spin current, 𝑀𝑆 is

the saturation magnetization, 𝑡𝐶𝑜 and 𝑡𝑃𝑡 are the Co and Pt thicknesses respectively, and 𝑒 and ℏ

are the electron charge and Planck’s constant. By plotting the 1/𝜉𝐹𝑀𝑅 versus 1/tCo (figure

Page 138: Dynamic Modulation of Material Properties by Solid State

132

6.2(d)),we can then get damping-like (𝜉𝐷𝐿) and field like (𝜉𝐹𝐿) torque coefficients from equation

6.5.

1

𝜉𝐹𝑀𝑅=

1

𝜉𝐷𝐿 (1 +

𝑒)

𝜉𝐹𝐿

𝜇0𝑀𝑆𝑡𝐶𝑜𝑡𝑃𝑡 Equation 6.5

At Co thickness of less than ~2.5nm, 1/𝜉𝐹𝑀𝑅 is negative because the field-like torque is

dominated by the torque from the spin-current whereas at >3nm, 1

𝜉𝐹𝑀𝑅 is positive because it is

dominated by the Oersted field. 𝜉𝐷𝐿 obtained from the intercept of the linear fit is ~0.03,

consistent with literature178. A plot of 𝑆

𝐴 versus 𝑡𝐶𝑜 is also shown in figure 6.1(e) for reference.

In this case, 𝑆

𝐴 shows the ratio of the total damping-like torque to the total field-like torque

regardless of whether the torques are due to spin current or the Oersted field.

Figure 6.2. Thickness dependence of HDemag, α, and 1/𝝃𝑭𝑴𝑹. (a) ST-FMR spectra of

Pt(3nm)/Co(tCo)/GdOx (36nm)/Au(3nm) device at 8GHz. (b) HDemag as a function of tCo. (c) α as a

function of tCo. (d) 1/𝜉𝐹𝑀𝑅 as a function of 1/tCo (e)S/A ratio as a function of tCo.

Page 139: Dynamic Modulation of Material Properties by Solid State

133

Figure 6.3(a) shows the ST-FMR spectra of a Pt/Co/GdOx/Au device with 1.2nm thick Co at

different frequencies. At this thickness, the spectra are dominated by a symmetric Lorentzian

component, indicating a large damping-like torque relative to the field-like torque. Figure 6.3(b)

shows the ST-FMR spectra after VG = +4V has been applied for 8200s while figure 6.3(c) shows

the ST-FMR spectra after VG is set to 0V for >400000s. Quick inspection shows a clear change

in resonance peak positions, widths and shapes. Figure 6.3(d)-(f) show the fits of the extracted

Lorentzian parameters to equation 6.2, 6.3, and 6.4 to obtain to obtain HDemag, α, and 𝜉𝐹𝑀𝑅

respectively. At VG of +4V a large increase in demagnetizing field from 3.5kOe to 8kOe is

observed due to large increase in in-plane magnetic anisotropy. Similarly, the magnitude of 𝜉𝐹𝑀𝑅

also increases from -0.040 to -0.084. On the other hand, the magnetic damping constant, α

decreases from 0.090 to 0.037. These changes are largely reversible when VG is set to 0V, where

HDemag, α, and 𝜉𝐹𝑀𝑅 revert back to values of 3.7kOe, 0.088, and -0.042 respectively. Figure 6.2

(b) –(d) shows the corresponding time series data of three magnetic parameters extracted from

the ST-FMR spectra under the gate biases. The black data points correspond to values extracted

by directly inserting the resonance peak field at 8GHz into equations 6.2 and 6.4 while the red

data points correspond to values obtained from fits to 4 different frequencies (5.4GHz, 6GHz,

7GHz, and 8GHz). During the measurement of the spectra for the 4 frequencies, a holding

voltage of +3V was used. Regarding the mechanism which leads to the modulation of the

magnetic anisotropy, damping, and spin torques, we propose that there are two electrochemical

processes which are taking place. The first process is voltage-induced hydrogen generation and

removal. In the previous chapter, we proposed that hydrogen is primarily accumulated at the

Co/GdOx interface. However, since magnetic damping in Co/Pt heterostructures primarily

originate from the Co/Pt interface, it is very likely that some hydrogen also penetrates the Co and

Page 140: Dynamic Modulation of Material Properties by Solid State

134

disrupts this interface. The second process is voltage-induced oxidation and reduction of the

magnetic Co layer. In this case, the device starts out with some oxidized dead layer and a

positive bias reduces it to magnetic Co. When the VG is next set to 0V, the reduced Co gradually

reoxidizes again. The redox of Co modulates the magnetic parameters primarily by changing the

effective thickness of the Co layer.

Figure 6.3. Voltage gating of HDemag, α, and 𝝃𝑭𝑴𝑹. (a)-(c) ST-FMR spectra of a

Pt(3nm)/Co(1.2nm)/GdOx (36nm)/Au(3nm) device in virgin state (a), after VG = +4V for 8200s

(b) and after setting VG back to +0V for 410000s. (d)-(f) Fit to the Kittel equation 6.1(d),

equation 6.2(e) and equation 6.3(f) to obtain HDemag, α, and 𝜉𝐹𝑀𝑅 respectively.

Page 141: Dynamic Modulation of Material Properties by Solid State

135

Figure 6.4. Time series data for tCo = 1.2nm device. (a) HDemag ,(b) α, and (c) 𝜉𝐹𝑀𝑅 at VG =

+4V and 0V for device shown in figure 6.3

To verify this conclusion, we also performed voltage gating on a Pt/Co/GdOx/Au structure with

𝑡𝐶𝑜=4.2nm. In this case, under a positive bias, there is minimal change in the demagnetizing field

(magnetic anisotropy) and damping constant, while there is 30% reduction in the torque ratio. If

we look at figure 6.1, while the increase in the effective thickness of the Co layer due to

reduction of any oxidized layer only lead to small changes in damping constant at 4nm, it should

lead to relatively large change in demagnetizing field. The fact that this is not observed strongly

suggests that the large changes observed in the 1.2nm Co device is primarily due to hydrogen-

Page 142: Dynamic Modulation of Material Properties by Solid State

136

induced modulation of the magnetic interfaces. In the 4.2nm Co device, the Co thickness is too

large for hydrogen to penetrate it, as a result there is minimal change in the magnetic anisotropy

and damping. As for the spin orbit torques, it has been shown that the strength of the field-like

torque depends crucially on the ferromagnet/oxide interface; as a result hydrogen accumulation

at the Co/GdOx can alter the ratio of the different torques even if the Pt/Co interface is not

affected.

Figure 6.5. Time series data for tCo = 4.2nm device. (a) HDemag ,(b) α, and (c) 𝜉𝐹𝑀𝑅 at VG =

+4V.

Page 143: Dynamic Modulation of Material Properties by Solid State

137

In conclusion, we have shown that magnetic properties which manifest themselves at interfaces

can be significantly modulated by ions driven using a gate voltage. These properties include

magnetic anisotropy, magnetic damping, and spin-orbit torques. Dynamic modulation of these

properties represent an important advancement in the field of voltage control of magnetism, and

can enable important new functionalities and device design.

Page 144: Dynamic Modulation of Material Properties by Solid State

138

Chapter 7: Voltage-

induced Magneto-

Ionic Effect in

Pt/Co/MOx

Heterostructure (M=

Gd, Y, Zr, and Ta)

Page 145: Dynamic Modulation of Material Properties by Solid State

139

In this chapter, we demonstrate the generality of voltage-induced magneto-ionic effect in

Pt/Co/MOx structure (M = Gd,Y,Zr, Ta) probed magnetically by magneto-optical Kerr effect. For

both rates, ROxide of voltage-induced Co oxidation and reduction in the Pt/Co/MOx structure,

there is a trend which follows the order: RGdOx > RYOx > RZrOx > RTaOx. RGdOx is very similar to

RYOx while RZrOx is ~ 2 orders of magnitude smaller. For the TaOx device, no redox of Co is

observed under an applied bias throughout the experiment for 24hours. The difference in

magneto-ionic rates for the four oxides is attributed to the difference in their proton

conductivities. This shows that significant improvement in speed of magneto-ionic switching can

be achieved through careful selection of material as the solid state electrolyte.

Electrical gating of magnetism has shown tremendous versatility for dynamic control of

various magnetic properties. Pioneering work has demonstrated reversible control of magnetic

anisotropy using an electric field at a metal/metal oxide interface12,51,52. However, such electronic

effect is small due to Coulombic screening in a metal. An alternative approach called magneto-

ionic switching60–63,66 exploits mobile ionic species in a solid state heterostructure to modulate

magnetic anisotropy with an efficiency of up to 5000fJ/Vm60,61 compared to ~10fJ/Vm in electric

field gating12. However, its speed is slower because it relies on ionic motion instead of electron

flow. To improve this speed, it is necessary to find an optimal metal oxide dielectric which is

simultaneously an excellent electrolyte to allow for fast ionic transport of the active species.

Thus far, studies on magneto-ionic switching has mostly focused on thin film GdOx60–62as

the electrolyte in a heavy metal/ferromagnet/dielectric (electrolyte) structure. The reliance on

only one material makes it difficult to identify trends or desirable characteristics of the

electrolyte. In this chapter, we introduce three new materials, namely YOx, ZrOx, and TaOx and

compare their rates of voltage-induced magneto-ionic effect to that of GdOx in a Pt/Co/MOx (M

Page 146: Dynamic Modulation of Material Properties by Solid State

140

= Gd, Y, Zr, and Ta) structures. We chose these materials for three reasons. 1) They are binary

oxides which allow for simple comparison to GdOx144, 2) they cover a wide spectrum in terms of

the similarity of their crystal structures179–183 to GdOx , and 3) they are high-k dielectric179 with

good breakdown characteristics.

Page 147: Dynamic Modulation of Material Properties by Solid State

141

7.1 Experimental Methods

Experiments focus on Ta(4nm)/Pt(3nm)/Co(0.9nm)/MOx (10nm) (M = Gd,Y,Zr, and Ta)

structures which are fabricated by DC magnetron sputtering on thermally oxidized Si substrates.

All layers are sputtered at argon pressure,PAr of 3mT and base pressure of 5 x 10-7 torr. For the

Co layer, the initial state is either metallic or oxidized. The oxidized state is sputtered with an

additional oxygen pressure PO2 of 0.07mT. For the MOx layer, reactive sputtering was performed

at PO2 of 0.06mT, followed by PO2 of 0.7mT during the last 2 minutes of sputtering. This last step

acts like an additional plasma oxidation step. To allow for electrical contact, a portion of the

Ta/Pt layers (bottom electrode) is uncovered by the MOx layer while round Au (3nm) electrodes

of 200µm diameter (top electrode) is sputtered on top of the MOx layer using shadow mask

lithography (Figure 7.1). The bottom Pt is grounded while the gate voltage, VG is applied to the

top Au by contacting the edge of the electrode with a CuBe probe. All magnetic measurements

are done using magneto-optical Kerr effect (MOKE) polarimeter with a 1mW laser at

wavelength of 655nm. The laser is focused at the middle of the Au electrode.

Figure 7.1. Measurement schematic for in situ probe of Co magnetization using magneto-

optical Kerr polarimetry.

Page 148: Dynamic Modulation of Material Properties by Solid State

142

7.2 Rate of Voltage-Induced Magnetic

Modulation at Positive Bias

Figure 7.2(a) shows the change in ratio of the remnant (Mr) to saturation magnetization

(Ms) obtained from hysteresis loops as a gate voltage, VG = +3V is applied to Pt/CoO/MOx

structures. In its virgin state, Mr /Ms ~0 because the Co layer is oxidized. As VG = +3V is applied,

CoO layer is reduced to Co leading to an increase and eventual plateau in Mr /Ms of the GdOx ,

YOx and ZrOx structures, which is consistent with literature60,61. As VG is applied for longer,

interestingly, the Mr /Ms values decreases. For the TaOx structure, no discernable reduction of the

Co layer is observed over a period of 24 hours. The data shows a trend in the rates of Co

reduction, Roxide which proceeds in the following order: RGdOx > RYOx > RZrOx > RTaOx .

To identify the magnetization states during application of VG = +3V, we plotted the

hysteresis loops at different times of the experiment in figure 7.2(a). Figure 7.2(b)-(e) show the

hysteresis loops for GdOx, YOx , ZrOx and TaOx devices respectively. At t =0s, in all cases, there

is no magnetic signal. As VG = +3V is applied, we observe a similar trend in the GdOx, YOx, and

ZrOx devices, where upon complete reduction of CoO to Co, there is large perpendicular

magnetic anisotropy (PMA) in the layer. As VG = +3V is applied for longer, the hysteresis loops

evolve into a hard-axis loop, which corresponds to the rotation of the magnetization in-plane.

The observed reduction in Mr /Ms in figure 7.2(a) originates from this rotation and not from

vanished magnetization. This rotation is due to hydrogen accumulation at the Co/MOx interface

(chapter 4) as a positive VG is applied beyond reduction of CoO to Co. For the TaOx device,

figure 7.2(e) confirms that the Co layer remains oxidized throughout the experiment.

Page 149: Dynamic Modulation of Material Properties by Solid State

143

Figure 7.2. Voltage-induced magneto-ionics at positive bias. (a) Rates of change in Mr / Ms at

VG = +3V for different Pt/Co/MOx /Au structures, where M = Gd, Y, Zr, and Ta. (b)-(d) MOKE

hysteresis loops at different times in (a), corresponding to M = (b) Gd, (c)Y, (d) Zr, and (e) Ta.

Figure 7.3(a) shows the change in Mr /Ms for the GdOx, YOx, and ZrOx devices at VG = 0V after

VG = +3V has rotated the magnetization in plane. The data shows relaxation of the magnetization

back to an out-of-plane state, with the rate of relaxation for the four oxides following the same

Page 150: Dynamic Modulation of Material Properties by Solid State

144

trend as the Co reduction rate. Figure 7.3(b) shows a representative set of hysteresis loops for the

GdOx, YOx and ZrOx devices before and after the relaxation, clearly indicating a rotation of the

magnetization from in-plane to out-of-plane. This rotation is due to removal of hydrogen from

the bottom Co/MOx interface as the hydrogen forms protons and is transported to the top Au

electrode where it recombines with O2 to form water through the oxygen reduction reaction.

Figure 7.3. Voltage induced magneto-ionics at 0V after positive bias gating. (a) Rate of

change in Mr / Ms for the Pt/Co/MOx structures at VG = 0V after VG = +3V has been applied. (b)

Hysteresis loops before and after VG is set to 0V.

Page 151: Dynamic Modulation of Material Properties by Solid State

145

7.3 Rate of Voltage-Induced Magnetic

Modulation at Negative Bias

Figure 7.4(a) shows the change in Mr /Ms obtained from hysteresis loops as a negative

bias, VG = -2V is applied to the Pt/Co/MOx structures. All four devices exhibit large PMA in their

initial state. As VG = -2V is applied, we observe a gradual decrease in Mr /Ms of the GdOx, YOx

and ZrOx devices to ~0. Unlike the situation at VG = +3V where the reduction in Mr /Ms

corresponds to rotation of the magnetization in-plane, in this case Mr /Ms decays due to the

oxidation of Co to CoO60,61. This is confirmed by hysteresis loops in figure 7.4(b) to (d)

corresponding to GdOx, YOx and ZrOx devices respectively where instead of a hard-axis loop,

we observe no magnetic signal even at a maximum field of ~500Oe after a prolonged negative

VG. For the TaOx device, we do not observe any oxidation of Co after 24hours. The rates of Co

oxidation for the different oxides again follow the same trend as CoO reduction in figure 7.2(a).

Page 152: Dynamic Modulation of Material Properties by Solid State

146

Figure 7.4. Voltage induced magneto-ionics at negative bias. (a) Rates of change in Mr / Ms at

VG = -2V for different Pt/Co/MOx /Au structures. (b)-(e) MOKE hysteresis loops for

Pt/Co/MOx/Au structure at different times in (a), where M = (b) Gd, (c)Y, (d) Zr, and (e) Ta.

The metal oxide layer plays a very important role in the overall magneto-ionic process. For

voltage-induced Co oxidation, the difference in rates can be attributed to the basicity of the

different oxides. GdOx and YOx have very high basicity compared to ZrOx and TaOx, and form

hydroxides143,144 when they are in contact with H2O. During voltage-induced Co oxidation, it is

the H2O stored in the form of hydroxides that oxidizes the Co layer, as shown in chapter 4. As a

Page 153: Dynamic Modulation of Material Properties by Solid State

147

result, the Co oxidation rate is much higher in GdOx and YOx compared to ZrOx and TaOx.

Besides acting as a store for H2O, the metal oxide layer also acts as the electrolyte for proton

transport. In fact, the trend in the rates of all voltage-induced magneto-ionic effects tracks the

proton conductivities of the different oxides studied81,82,137. This is consistent since proton

transport has to take place for all electrochemical reactions to happen. GdOx and YOx have high

proton conductivities due to large concentration of carriers and high carrier mobility. On the

other hand, ZrOx and TaOx have low proton conductivities; as a result their rates of voltage-

induced magneto-ionic effect are very slow.

In conclusion, we have demonstrated that the choice of metal oxide adjacent to the

ferromagnet affect substantially the rate of voltage- induced magneto-ionic effect. Through

judicious choice of materials, we can potentially reduce the timescale of voltage gating by a few

orders of magnitude, making it viable for practical low power memory applications.

Page 154: Dynamic Modulation of Material Properties by Solid State

148

Chapter 8: Room

Temperature

Reversible Solid Oxide

Fuel Cell

Page 155: Dynamic Modulation of Material Properties by Solid State

149

The scientific community has approached miniaturization of power sources for

electronics by focusing mostly on solid state Li-ion microbatteries184–188. Another power storage

and generation system: the solid oxide fuel cells, produce power through recombination of

hydrogen fuel with oxygen to form water. However, the operation of SOFCs have largely been

limited to very high temperature (>700C) due to sluggish kinetics of charge transfer and mass

transport. In addition, SOFCs were mainly studied in large bulky form factors with gaseous

hydrogen as fuel; these are specifically customized for large scale power generation system, not

microelectronics. In fact, most of the gaseous hydrogen fuel is generated by a separate methane

reforming cell, and the fuel is stored in a gas tank. In this chapter, we demonstrate a miniaturized

room temperature reversible solid oxide fuel cell for small-scale systems such as

microelectronics. The cell can be “charged” by splitting water in ambient111,112. It can then

produce power when needed through the water recombination reaction113. Hydrogen is stored in

a thin GdOx film where good scalability in terms of area and thickness of the GdOx storage layer

is observed. This is a significant because we demonstrate for the first time the miniaturization of

a solid oxide fuel cell for room temperature operation. This opens up a new range of applications

such as microelectronics where solid oxide cells can be employed. The use of a rechargeable

hydrogen cell also puts forth a new alternative to lithium-ion battery.

Page 156: Dynamic Modulation of Material Properties by Solid State

150

8.1 Experimental Methods

Experiments focus on Ta(4nm)/Pt(10nm)/GdOx (tGdOx nm) /Au cross bar structures (figure

8.1) which are fabricated on thermally oxidized Si substrates. The Ta, Pt and Au layers are

sputtered using DC magnetron sputtering at argon pressure,PAr between 3mTorr and 3.5mTorr.

The GdOx layer is RF sputtered at PAr of 3mTorr and oxygen partial pressure, PO2 of 0.7mTorr.

The base pressure in all cases is ~5 x 10-7 torr. CuBe probes are used for electrical contact. For

electrical characterization, a Keithley 6430 source meter unit is used for charging and to measure

the discharge and power density curves. All measurements are performed in a CPX-VF probe

station using mechanically compliant CuBe probes. Experiments under different gas environments

were performed by backfilling the chamber with either O2 gas (99.999% purity) or N2 gas

(99.999% purity). Humidity was introduced into the N2 gas flow by bubbling through water.

Ambient condition at 25°C corresponds to 12mT of H2O partial pressure respectively. All

experiments were performed at room temperature.

Figure 8.1. Device schematic of cross bar solid oxide cell.

As mentioned in chapter 4, the Pt/Co/GdOx/Au cell stores hydrogen when VG > 0 is applied to

the top Au gate. Figure 8.2 summarizes the charging, storage, and discharging processes for a

Pt/GdOx/Au cell. In this chapter, we are primarily interested in the capacity and power output of

Page 157: Dynamic Modulation of Material Properties by Solid State

151

the cell for energy storage applications. During charging, H2O is split into O2 and H+ through the

oxygen evolution reaction. The H+ which is produced is driven to the bottom Pt electrode,

reduced to neutral hydrogen, and stored in GdOx layer. During discharge, the stored hydrogen

provides the electrons which drives the load in the external circuit.

Figure 8.2. Reaction schematic of reversible solid oxide fuel cell.

Page 158: Dynamic Modulation of Material Properties by Solid State

152

8.2 Proton Conductivity of GdOx

To characterize the lower bound for proton conductivity of the GdOx layer, we measured

its dc conductivity. Figure 8.3(a) and (b) show the total current flow at VG = +2.5V for a

hydrated GdOx in ambient and in vacuum respectively while figure 8.3(c) and (d) show the same

measurements for a non-hydrated GdOx. We observe a large difference in dc current flow

between the hydrated GdOx cell in ambient and in vacuum indicating that a majority of current

flow is due to ionic current alone. We also observe that the current is significantly larger in a

hydrated GdOx compared to a non-hydrated GdOx in ambient, indicating a much larger

conductivity in the hydrated GdOx cell. Specifically, it is the ionic conductivity which is much

larger in the hydrated vs the non-hydrated GdOx, as will be shown later.

Figure 8.3. Charging current of solid oxide cell at VG = +2.5V at different temperatures (a)

Hydrated GdOx in ambient. (b) Hydrated GdOx in vacuum. (c) Non-hydrated GdOx in ambient.

(d) Non-hydrated GdOx in vacuum.

Page 159: Dynamic Modulation of Material Properties by Solid State

153

By subtracting the current that flows in a hydrated GdOx sample in vacuum from the

current flow in ambient, we obtain the current flow due to reactions and proton transport alone.

This is because in vacuum, the water splitting reactions depicted in equation 2.6 and 2.7 cannot

take place, whereas in ambient, they do. Figure 8.4 shows the conductivity calculated from this

current flow as a function of temperature. At room temperature, the total conductivity is between

10-12 and 10-11 S/cm. This value fits well with the extrapolated proton conductivity of doped-

GdOx in literature82,92 of between 10-11and 10-10 S/cm. Note however that the ionic conductivity

calculated from dc current include resistance contributions from reactions at both electrodes; ie

the oxygen evolution and hydrogen evolution reactions (equation 2.6 and 2.7). Hence, the

conductivity value given here only represents a lower bound to actual proton conductivity of the

oxide. By linear fitting the conductivity values, we obtained an activation energy of 0.23eV.

While this activation energy is not a true representation of the proton mobility in the oxides

(since it includes several other contributions), it is still interesting to note that the value is very

low compared to the activation energy for proton transport in other oxides81,82,92,137.

Figure 8.4. Conductivity of solid oxide cell.

Page 160: Dynamic Modulation of Material Properties by Solid State

154

8.3 Cell Performance and Scalability

Figure 8.5 summarizes the performance of the cell. At VG = +2.5V, the hydrated GdOx

cell can be charged to 100% capacity of 1.4µAh/cm2 in ~60min while the non-hydrated cell does

not show any stored charge up to 80min of charging time. The amount of charge stored in the

hydrated cell is more than 3 orders of magnitude larger than what is calculated from the expected

capacitance of the GdOx layer. The absence of any stored charge in the non-hydrated cell shows

that the small current flow in figure 8.2 is primarily due to low proton conductivity of the non-

hydrated oxide. This is consistent with the findings in chapter 4,5, and 6 which show voltage-

induced magnetic change in a Pt/Co/GdOx heterostructures are only observed in hydrated GdOx

devices. Figure 8.5(b) shows the discharge curve of the GdO at ~4µA/cm2 for the different

charging times. We can observe a plateau at ~1V which corresponds well to the thermodynamic

potential for the water recombination at room temperature. Figure 8.5(c) shows power density

curves for the Pt/GdOx(40nm)/Au cells at different charging times where a maximum power

density of ~50µW/cm2 was achieved at current density of 80µA/cm2. Figure 8.5(d) shows the

cyclability of a 40nm thick cell where charging was done at VG = +3V for 10min. A constant

capacity of ~0.5µAh/cm2 was obtained with no significant decrease in capacity for 50cycles.

Figure 8.5(e) shows the corresponding discharge curves for 5 out of the 50 cycles, with

reproducible behavior throughout the 50 cycles. Figure 8.5(f) shows the efficiency of the

charging process at VG = +2.5V. The cell exhibits efficiency of ~10% at room temperature. The

efficiency values are calculated as the ratio of the stored energy to the input energy. The stored

energy is obtained by integrating the area under the discharge curves, while the input energy is

simply the product of the charging voltage, charging current, and charging time. In the GdOx

Page 161: Dynamic Modulation of Material Properties by Solid State

155

cell, possible losses include large overpotentials from oxygen reduction reaction113, ohmic losses

and electronic leakage across the oxide.

Figure 8.5. Performance of solid oxide cell. (a) Capacity as function of charging time. (b)

Discharge curves. (c) Power density curves (d) Cyclabililty. (e) Discharge curves at different

cycles in (d). (f) Cell efficiency.

In figure 8.6 (a)-(c), we show that the capacity and power density of the cell scale with

the thickness of the GdOx. This thickness scalability confirms that the dominant charge storage

in the device is electrochemical in nature and not capacitive. It also shows that the charge in the

form of hydrogen is stored across the thickness of the GdOx instead of at the interfaces. Note

however that maximum cell voltage for a 80nm thick cell is only ~ 0.78V compared to ~1V in

thinner cells. This is likely due to larger Ohmic overpotential for proton transport during

discharge of hydrogen stored further away from the Pt interface. In figure 8.6(d)-(e), we also

demonstrate that the capacity scales with cell area. The largest capacity of 1.8x10-3µAh is

Page 162: Dynamic Modulation of Material Properties by Solid State

156

achieved with an area of 4mm2. In these GdOx thin film however, the exact charge storage

mechanism is still not well understood; one possibility is hydrogen is stored in the crystal

structure of GdOx as interstitials.

Figure 8.6. Scalability of solid oxide cell. (a) Capacity and peak power density as function of

GdOx thickness. (b) Discharge curves of cell with different GdOx thicknesses. (c) Power density

curves of cell with different GdOx thicknesses. (d) Capacity as function of cell area for a 40nm

thick GdOx. (e) Discharge curves of cell for different cell areas

In chapter 4, we showed that voltage-induced reduction of CoO in a Pt/CoO/GdOx/Au

heterostructure at VG > 0 (applied to Au) is due to hydrogen. This was done by probing the

magnetic signal of the CoO layer using Hall magnetometry; when charged in vacuum, the

magnetic signal remain absent due to . On the other hand, when the cell is charged in ambient

(50% humidity), there is large magnetic signal indicating the reduction of CoO to Co by

hydrogen generated at the bottom Pt electrode. Here, we again show that the stored charge is

hydrogen by comparing the capacity of the cell charged in ambient vs in vacuum. While the

Page 163: Dynamic Modulation of Material Properties by Solid State

157

maximum capacity is 1.4µAh/cm2 when charged in ambient (figure 8.3), no measurable charge is

detected after charging in vacuum. This is expected since the calculated capacitance for the

GdOx cell dimensions (1.23x10-3cm2) is ~0.5nF assuming no hydrogen storage. When charged

at +2.5V in vacuum, the total stored charge is only ~1nC, which is too small to be measured.

Figure 8.7 also shows the power density curves for a 1mm2, 20nm thick GdOx cell charged at

+3V for 30s and discharged under different atmospheric conditions. The results are again

consistent with the hypothesis that H2O acts as the source of hydrogen for energy storage in the

cell.

Figure 8.7. Power density of GdOx cell in different atmospheric conditions. The charging

and discharging conditions are listed in order.

Page 164: Dynamic Modulation of Material Properties by Solid State

158

8.4 Gating of Magnetism using Built-in

Voltage

As discussed in chapter 5, in a Pt/Co/GdOx/Au heterostructure, a positive gate voltage,

VG applied to the top Au electrode drives the Co magnetization from out-of-plane to in-plane due

to the generated hydrogen. When VG is next set to 0V (short circuit), the magnetization reverts

back to an out-of-plane state due to removal of hydrogen from the Co interface. If the VG is

instead set to open circuit after +3V, the magnetization remains in the in-plane state as the

hydrogen remain “trapped” at the Co interface due to the absence of an external circuit through

which it can donate its electron.

In this section, we further show that the Co can be switched between a high and a low

perpendicular anisotropy state by toggling between short circuit (SC) and open circuit (OC). In

the experiment, we first charge the Pt(4nm)/Co(0.9nm)/GdOx(10nm)/Au(3nm) device at VG =

+3V for 600s before alternating between short and open circuit(figure 8.8). Figure 8.9 shows the

switching cycles of Mr/Ms and coercivity between the two conditions. From the data, we can

clearly observe an increase in Mr/Ms and coercivity at short circuit which corresponds to an

increased perpendicular anisotropy while the opposite is observed for open circuit. The

corresponding polar MOKE hysteresis loops in the first cycle are also shown. The magnitude of

change is largest during the first cycle, and slowly decreases over cycles. This is because during

the first cycle immediately after VG = +3V, the amount of hydrogen stored in GdOx is maximum.

With increasing short circuit duration, the amount of charge stored is gradually decreased due to

the removal of hydrogen through water recombination reaction. The toggling between high and

Page 165: Dynamic Modulation of Material Properties by Solid State

159

low perpendicular anisotropy states happen in the absence of an external power source; this

means it is the internal voltage in the GdOx which is driving this magnetic change. At short

circuit, the voltage difference between the top and bottom electrode is zero while at open circuit,

the voltage difference is given by the potential for the water recombination reaction. The detailed

mechanism which drives the magnetic modulation is however not well understood at this point.

Figure 8.8. Experimental sequence for toggling Co magnetization between short circuit and

open circuit

Figure 8.9. Modulation of magnetic anisotropy at open circuit (OC) and short circuit (SC)

conditions. (a) Switching cycles of coercivity at OC and SC. (b) Switching cycles of Mr/Ms at

OC and SC. (c) Polar MOKE hysteresis loops corresponding to the first cycle.

Page 166: Dynamic Modulation of Material Properties by Solid State

160

In conclusion, we demonstrated a room temperature miniaturized reversible solid oxide fuel cell

based on hydrogen storage in a thin film GdOx. Hydrogen is sourced from ambient through the

water splitting reaction during charging while power is produced during discharge through the

water recombination reaction. For a self-contained energy storage unit, we propose the

integration of a hydrogen storage unit on top of the GdOx which will enable implementation in

practical devices. We also showed that by inserting a thin magnetic layer beneath the GdOx, the

voltage of the cell can be used to modulate interfacial magnetic anisotropy by toggling between

short and open circuit. This work demonstrates the viability of oxide fuel cell miniaturization for

solid state device applications.

Page 167: Dynamic Modulation of Material Properties by Solid State

161

Chapter 9: Voltage

Gating of Optical

Properties

Page 168: Dynamic Modulation of Material Properties by Solid State

162

Besides gating of magnetic properties, we also observe reversible color change in a

Pt/GdOx/Au heterostructure induced by a gate voltage. Electrochromic films present an attractive

way to build energy efficient smart windows, coatings and mirrors189. These films exhibit a

change in their optical transmission and absorption under an applied voltage. Typically,

electrochromic films consists of three layers: an ion storage layer, an electrolyte and an active

layer such as WO3 or NiOx189–191. The three layers are then sandwiched by transparent

conductors. When voltage is applied across the device, ions are pumped into and out of the active

layer through the electrolyte, modulating the optical characteristics. Due to the trilayer design,

the complexity and cost of such films have limited their broader application, and for certain

applications, the optical switching speed of common electrochromic materials is still

prohibitively low.

In this chapter, we describe a thin-film reflective electrochromic device consisting of a

single ultra-thin layer of GdOx that acts simultaneously as electrolyte and active layer and operates

without a separate ion storage layer. We achieve gate-voltage-induced reversible modulations in

reflectivity of ~10% and switching time down to 20 ms at room temperature. Measurements of

the change in reflectivity as a function of inserted charge192 show a figure of merit of ~37.6 cm2/C

for a 20nm thick film, comparable to conventional electrochromic materials, making this simple

structure a viable alternative to more complex multilayer stacks. Systematic experiments

demonstrate a strong dependence of device behavior on the top electrode thickness, which

indicates that the rate-limiting step is the water splitting reaction at the top electrode112 (chapter

5). Because the GdOx layer serves both as electrolyte and active layer, and the atmosphere serves

as a hydrogen reservoir, this eliminates the need for an ion storage layer and greatly simplifies the

device architecture.

Page 169: Dynamic Modulation of Material Properties by Solid State

163

9.1 Experimental Methods

Our work focuses on Ta(4 nm)/Pt(10 nm)/GdOx(tox)/Au(tAu) sputter-deposited on a Si

substrate with 50nm of native oxide. Here tox and tAu represent the thickness of the corresponding

layers. Figure 9.1(a) shows a schematic of the device structure. The metal layers were deposited

by dc magnetron sputtering at room temperature with an Ar pressure of 3 mTorr. The GdOx layer

was RF sputtered from a stoichiometric Gd2O3 target at room temperature at a rate of ~ 0.2nm/min

in an oxygen partial pressure of 0.7mT. The Ta layer acts as an adhesion layer while the Pt layer

acts as the back electrode and serves as a reflecting mirror. The GdOx is the electrolyte and active

layer whose optical properties change depending on the applied voltage. On top of these layers,

Au electrodes with a diameter of ~200µm and with tAu=3 nm except where noted, were patterned

using shadow mask lithography to serve as the top gate. At this thickness, the Au layer is semi-

transparent and porous61. Experiments were performed by contacting the top electrode with a CuBe

microprobe and grounding the back electrode (figure 9.1). The optical reflectivity was monitored

in situ with a wide-field microscope and CCD camera, as well as with a with a focused laser spot

(655 nm wavelength) whose reflected intensity was recorded by a photodiode for time-resolved

measurements.

Page 170: Dynamic Modulation of Material Properties by Solid State

164

9.2 Voltage Gating of Optical Reflectivity

in Pt/GdOx/Au Heterostructure

Figure 9.1(b) shows optical micrographs of the electrode in its virgin state, after applying

+3V for 10 s, and after applying -2 V for 10 s, respectively, to the top gate. The reflected color

changes from yellow to brown under positive bias, and this color change can be reversed by

inverting the gate voltage polarity.

Figure 9.1. Voltage gated optics. (a) Schematic illustration of the Ta(4nm)/Pt

(10nm)/GdOx(10nm) /Au(3nm) optical device. (b) Optical micrographs of the top electrode in its

virgin state, at V = +3V and at V= -2V.

Page 171: Dynamic Modulation of Material Properties by Solid State

165

Figure 9.2(a) shows the normalized reflectivity measured using the laser probe during

several cycles of a square-wave voltage waveform applied to the gate. Here the voltage amplitude

was 3V and the frequency was 0.5 Hz. We find that in the initial cycle (not shown here) the

reflectivity change proceeds rather slowly and is accompanied by some irreversible darkening,

whereas in subsequent cycles, the device response is quite fast and largely reversible. As can be

seen in figure 6.2(a), the reflectivity can be reversibly switched by ~10% over many cycles. Figure

9.2(b) shows the reflectivity near the rising and falling edge of the voltage waveform. For the

rising edge, the reflectivity abruptly decreases by 5% within ~ 40ms, followed by a slower

continued decrease. At the falling edge, the response is faster; reflectivity increases by 8% within

20 ms. Robust reflectivity switching is observed over >200 cycles, with an overall decrease in

reflectivity indicating some degree of irreversibility. We then correlate the change in reflectivity

with the amount of inserted charge in order to determine the figure of merit for the device193–195.

We performed the experiment by applying a positive gate voltage, which we anticipate leads to

the injection of hydrogen into the system (chapter 5), and measuring the current and reflectivity

simultaneously. Figure 9.2(c) shows data for the device reflectivity, normalized to the virgin state,

versus integrated charge, for a typical device after initial cycling. A linear fit to the data yields a

slope of 37.6 cm2/C. We note that while this figure of merit was measured in reflectivity, similar

coloration efficiency is expected for transmission195.

Page 172: Dynamic Modulation of Material Properties by Solid State

166

Figure 9.2. Switching transient. (a) Switching cycles of reflectivity and the corresponding

voltage profile. (b) Reflectivity at rising edge and falling edge of the gate voltage. The voltage

was applied at relative time = 0ms. (c) Reflectivity as a function of inserted charge density. The

slope yields a figure of merit of 37.6 cm2/C.

We first examine the variation of optical properties on tAu, which provides evidence that

water splitting and proton transport are key to achieving the observed optical changes. As has

been shown in chapter 5, the rate of water splitting depends on the thickness and morphologies of

the anodes. In the GdOx optical device, the same trend is observed. In figure 9.3(b)-(e) we show

optical micrographs for four different electrodes with different tAu deposited on the same GdOx

film, after applying a gate bias of + 3V for 600 s. For tAu=3 nm, darkening of the film happens

rapidly and relatively uniformly across the electrode area, which we attribute to the porous Au

microstructure (see figure 9.3(a), inset). As tAu is increased, it is evident that the change in optical

Page 173: Dynamic Modulation of Material Properties by Solid State

167

properties begins at the electrode edge and proceeds inward. This corresponds to the triple phase

boundary, where electrons, protons and H2O are all present and electrochemical reactions114,115 are

expected to proceed most efficiently. Figure 9.3(f)-(i) show a sequence of images during voltage

application for tAu=4 nm, where it is evident that darkening progressively moves from the edge to

the center.

Page 174: Dynamic Modulation of Material Properties by Solid State

168

Figure 9.3. In situ probe of optical gating. (a)-(d) Optical micrographs of device with tAu=

3nm (a), 4nm(b), 6nm(c), and 10nm(d) after +3V for 600s. Scale bar is 500nm for inset of (a).

(e)-(h) Optical micrographs tAu= 4nm device at +3V after 0s (e), 250s (f), 450s (g) and 1200s

(h). The red dot indicates the laser spot.

Page 175: Dynamic Modulation of Material Properties by Solid State

169

9.3 Source of Irreversible Optical Change

in Pt/GdOx/Au Heterostructure

Finally, we consider the origin of the long-term irreversible decrease in reflectivity in the

device, which contributes to degradation during repeated cycling. The optical micrographs in

Figure 9.5(a)-(b) shows the virgin state and the state after applying a long-term bias of +3V for

10000s. The dramatic decrease in reflectivity (by 40% compared to the virgin state), is largely

irreversible. In order to understand the chemical change causing the irreversibility, we performed

X-ray photoelectron spectroscopy (XPS) on the sample in its virgin state and at the end of the

applied bias. Figures 9.5(c) shows the XPS spectra for Au4f where we observe a 0.5eV shift

towards a higher binding energy for the Au4f peaks, which likely indicates the oxidation of Au at

large anodic potential96. This oxidation in turn causes the irreversible browning of the device.

Figure 9.4: XPS analysis of irreversible optical changes. (a)-(b) Optical micrographs of the

top electrode in its virgin state(a) and after +3V for 10000s(b).(c) XPS spectrum of Au4f (c)

energy region for the device in virgin state and after bias application. Au4f has two peaks

corresponding to the doublet f 5/2 and f 7/2.

Page 176: Dynamic Modulation of Material Properties by Solid State

170

Figure 9.5 shows that the irreversibility in optical reflectivity is more pronounced when a larger

VG is applied. In the case of VG = +2V, although optical gating takes longer, the changes are largely

reversible. On the other hand, the optical changes at VG = +4V while large, are mostly irreversible.

In these Pt/GdOx/Au structures, the magnitude of reversible reflectivity change is ~0.1. One can

also notice the appearance of a bright region surrounding the electrode in figure 9.5(b). The details

will be provided in chapter 9.5.

Figure 9.5. Dependence of reversibility on VG. (a)-(b) Modulation of optical reflectivity at VG =

+2V (a) and +4V(b)

Figure 9.6 shows the change in optical reflectivity at different VG. The reflectivity is

calculated by taking the ratio of the current reflected intensity to its initial intensity. Note that VG’s

are applied to virgin devices in these experiments; as a result, the optical changes are slower.

Subsequent cycles yield much faster optical changes. At this point, it is not clear why VG = +1V

Page 177: Dynamic Modulation of Material Properties by Solid State

171

produces a change in optical reflectivity even though the thermodynamic water splitting potential

is 1.23V at room condition.

Figure 9.6. Rate of modulation of optical reflectivity at different VG’s.

Page 178: Dynamic Modulation of Material Properties by Solid State

172

9.4 Voltage Gating of GdOx

Heterostructures with different Top and

Bottom Electrodes

Figure 9.7 shows modulation of optical reflectivity in Pt(10nm)/GdOx(20nm)

heterostructures with top Pt(3nm) and Cu(3nm) top electrodes respectively while figure 9.8

shows optical modulation in Au(10nm)/GdOx(20nm)/Au(3nm) and

Pt(10nm)/YOx(10nm)/Au(3nm) heterostructures (different bottom electrode and oxides). The

presence of optical modulation in all devices show that all three layers might contribute to

changes in the optical properties.

Page 179: Dynamic Modulation of Material Properties by Solid State

173

Figure 9.7. Voltage gating of optical reflectivity in Pt/GdOx structures with Pt(a) and

Cu(top) electrodes

Page 180: Dynamic Modulation of Material Properties by Solid State

174

Figure 9.8. Voltage gating of optical reflectivity in Au/GdOx/Au (a) and Pt/YOx/Au (b)

heterostructures.

Page 181: Dynamic Modulation of Material Properties by Solid State

175

9.5 Optical Modulation outside Active

Region due to Hydrogen Diffusion

Figure 9.9 shows the appearance of a bright region around the circular Au electrode in a

Pt(10nm)/GdOx (20nm)/Au(3nm) heterostructure. This region grows at a rate of up to 100nm/s at

VG = +2.3V. In figure 9.10, we observe that upon reaching another device which was previously

gated, the curvature of the region changes (there are signs of necking). The propagation of the

bright region is however not affected by electric field applied between two adjacent devices; in

this case the bright region retains its circular profile (results not shown). This indicates a

diffusion instead of a drift process. We also inserted a thin CoO (2nm) layer between the Pt and

GdOx layers and observed reduction of CoO to metallic Co within the bright region. These

results show that neutral hydrogen is responsible for this optical change. The hydrogen is

injected into the film through the water splitting reaction at VG > 0.

Figure 9.9. Bright region surrounding active device.

Page 182: Dynamic Modulation of Material Properties by Solid State

176

Figure 9.10. Propagation of bright region.

Figure 9.11. Reduction of CoO within the bright region in a Pt/CoO/GdOx/Au

heterostructure.

In conclusion, we have demonstrated voltage gating of optical properties in different thin film

oxide heterostructures. We introduced a simple electrochromic device where a single layer of

oxide acts as both as an electrolyte and as the active layer. The ion storage layer is absent because

the atmosphere can act as a reservoir of hydrogen. We also discovered the presence of neutral

Page 183: Dynamic Modulation of Material Properties by Solid State

177

hydrogen in the region surrounding the active device when VG > 0 is applied. This leads to

increased reflectivity in the region.

Page 184: Dynamic Modulation of Material Properties by Solid State

178

Chapter 10: Electrical

Properties of GdOx

Page 185: Dynamic Modulation of Material Properties by Solid State

179

Because a majority of this thesis is focused on heterostructures with GdOx as the proton

conducting layer, we will also provide some characterizations on the electrical properties of

GdOx. We deposited 30nm of GdOx on Ta(4nm)/Pt(3nm) by reactive sputtering at different

partial pressure of oxygen (PO2). 200µm of circular Pt(3nm) electrodes were then patterned on

the continuous GdOx film with a short vacuum break between the fabrication of the GdOx and

the Pt top electrodes. The gate voltage (VG) is applied to the top circular electrodes while the

bottom electrode is grounded. The device schematic is shown in figure 10.1a. Figure 10.1b

shows the total current which flows at VG between ±1V for GdOx deposited at different oxygen

flow rates. VG is swept according to the sequence: 0V→1V→-1V→0V. The PO2 which

corresponds to the oxygen flow rates are shown in figure 10.1c. We observe that GdOx deposited

at higher PO2 shows a higher total resistance compared to lower PO2 samples, with the difference

in resistance between the 2.3sccm and 5sccm samples being more than 3 orders of magnitude. In

all three samples, we observe rectifying behavior in electrical conduction.

Figure 10.2 shows the voltage breakdown characteristics of the different GdOx layers

deposited at different PO2. The VG sequence follows 0V→11V→-11V→0V. Consistent with

earlier results, deposition at higher PO2 leads to samples with larger breakdown voltage. For the

samples in figure 10.2(e), the GdOx layer is first deposited at oxygen flow rate of 2.5sccm

followed by a plasma oxidation step for 2minutes. This step is performed by sputtering the

metallic Gd target at 5.5sccm which causes target poisoning. In the case of GdOx sputtered at

5sccm, no breakdown is observed up to more than ±11V.

Page 186: Dynamic Modulation of Material Properties by Solid State

180

Figure 10.1. Electrical properties of GdOx. (a) Schematic of a

Ta(4nm)/Pt(3nm)/GdOx(30nm)/Pt(3nm) device. (b) IV curves of GdOx deposited at different

PO2. (c) PO2 as a function of oxygen flow rate for reactive sputtering of GdOx. The shutter covers

the chimney around the sputter source. (d) Data in (b) in linear scale.

Page 187: Dynamic Modulation of Material Properties by Solid State

181

Figure 10.2. Voltage breakdown characteristics of GdOx reactively sputtered under different

oxygen flow.

To correlate the performance of magneto-ionic device to the electrical properties of

GdOx, we also fabricated Ta(4nm)/Pt(3nm)/Co(0.9nm)/GdOx(4nm or 10nm) structures with

either Au(5nm) or Ta(1.4nm)/Au(5nm) circular electrodes patterned on top of the GdOx. The

Page 188: Dynamic Modulation of Material Properties by Solid State

182

four commonly observed IV curves are shown in figure 10.4, where the VG sequence follows

0V→11V→-11V→0V. The electrical characteristic in figure 10.4(a) and 10.4(b) are very similar

where they both start out with a rectifying behavior; however there is a change to Ohmic

behavior (breakdown) at large voltage for the case in figure 10.4(b). The electrical behavior in

figure 10.4(c) and 10.4(d) are also very similar where they start out with an Ohmic behavior;

however in the case of 10.4(d), there is a change to rectifying behavior at large voltage. The most

likely explanation for these observations is the presence of parallel electronic conduction paths

in the GdOx. Ohmic behavior is caused by an electronically conductive filament formed through

the GdOx, while rectifying behavior can either mean the absence of any filaments or the presence

of a filament with Schottky junction at the interfaces.

Typically, devices which exhibit IV curve shown in figure 10.4(c) show minimal

modulation in magnetic properties with a gate voltage. In this case, the magnetization remains

out-of-plane throughout the voltage sweep sequence (figure 10.5(a)). The “best” devices have IV

curves shown in figure 10.4(a) or (b) where they exhibit large modulation in magnetic properties

with a small gate voltage. In this case, the magnetization rotates in-plane (figure 10.5(b)) at VG >

0V, which is reversible upon setting VG = 0V (figure 10.5(c)). At VG < 0V, the magnetization

vanishes as the Co layer oxidizes (figure 10.5(d)). Device in figure 10.4(d) also exhibits

modulation in magnetic properties; however a large gate voltage has to be applied during the first

cycle to “break” the conductive filament. In this case, the magnetization remains out-of-plane

with minimal change (figure 10.5(a)) from VG = 0V to 6V. At +7V, upon “breaking” the

filament, the magnetization rotates in-plane (figure 10.5(b)). The device now operates with IV

curve as shown in figure 10.4(a) or (b).

Page 189: Dynamic Modulation of Material Properties by Solid State

183

Note that while we mention that the “best” devices have rectifying behavior, the opposite is not

always true. In another word, devices with rectifying behavior does not necessarily show

magnetic modulation under a gate voltage. The rectifying behavior only implies low electronic

current but it does not tell us much about the ionic current that flows. The total current consists

of both electronic and ionic current superimposed. In order to have large ionic current from

proton transport, the GdOx has to be hydrated. Essentially, magneto-ionic devices with no

electronic current but maximum ionic (proton) current is desired.

To give a sense of the ionic current required to induce magneto-ionic switching in a

0.9nm thick Co layer, there is 5x10-12mol of Co atoms in a 200µm diameter circular device.

Assuming 1H is needed for every Co atom, ~500nC of charge will need to flow. If the timescale

of switching is ~100ms, there will be 5µA of ionic current.

Figure 10.3. Schematic of a magneto-ionic device.

Page 190: Dynamic Modulation of Material Properties by Solid State

184

Figure 10.4. Typical IV curves for magneto-ionic device. (a) Rectifying behavior. (b)

Rectifying behavior with breakdown. (c) Ohmic behavior. (d) Ohmic behavior with abrupt

change to rectifying behavior at large voltage.

Page 191: Dynamic Modulation of Material Properties by Solid State

185

Figure 10.5. Polar MOKE hysteresis loops at varying VG. (a)-(d) Hysteresis loop in initial

state (a), at VG > 0V (b), at VG back to 0V (c), and at VG < 0V (d).

Page 192: Dynamic Modulation of Material Properties by Solid State

186

Chapter 11: Summary

and Outlook

Page 193: Dynamic Modulation of Material Properties by Solid State

187

11.1 Summary

In summary, we have significantly advanced the mechanistic understanding of voltage-

induced ferromagnet oxidation in a ferromagnet/oxide heterostructures. We debunked the

conventional belief that Co oxidation in a Pt/Co/GdOx heterostructure is dominated by oxygen-

ion migration from the GdOx layer into the Co. Instead, we showed that Co oxidation is primarily

caused by water stored in the GdOx matrix under a gate voltage. The conclusion was supported

by magnetic data showing voltage-induced Co oxidation only in Pt/Co/GdOx device which has

been hydrated under high partial pressure of water and chemical spectroscopic data showing Co

oxidation.

We also demonstrated for the first time voltage gating of magnetism through solid state

proton gating. This significant breakthrough allowed for magneto-ionic gating of magnetism in

solid state devices without oxidation of the ferromagnetic layer. We achieve 90° magnetization

switching by proton injection for >2000 cycles at timescale down to 100ms and observed new

magnetic responses which could be employed for neuromorphic applications. We also modulated

magnetic damping and spin torques in a planar waveguide by solid state proton pumping which

allows us to tune a range of dynamical properties in a magnetic device. Finally, by insertion of

hydrogen in a heavy metal layer like Pd, we also showed for the first time gating of magnetic

anisotropy at a metal-metal interface. Because a wide range of magnetic interactions such as

Dzyaloshinskii-Moriya interaction and RKKY coupling manifest themselves at such interfaces,

this work paves the way to a new field of voltage-controlled spin orbitronics.

Besides voltage-induced modulation of magnetic properties, we also demonstrate that solid state

proton gating can be similarly employed to tune the optical properties of a thin film

Page 194: Dynamic Modulation of Material Properties by Solid State

188

metal/oxide/metal device. We showed gating of the optical reflectivity by ~10% at timescale

down to 20ms. The observation of optical gating is general across a wide range of thin film

structures with different electrodes and oxides, indicating the versatility of proton gating for

optical modulation.

Finally, we demonstrate a miniaturized room temperature reversible solid oxide fuel cell

for small scale system which is based on hydrogen storage. The cell can be “charged” by

splitting water in ambient. It can then produce power when needed through the water

recombination reaction. Hydrogen is stored in a thin GdOx film where good scalability in terms

of area and thickness of the GdOx storage layer is observed. This is a significant development

because we demonstrate for the first time miniaturization of a solid oxide fuel cell for room

temperature operation. This opens up a new range of applications such as microelectronics where

solid oxide cells can be employed. The use of a rechargeable hydrogen cell also puts forth a new

alternative to lithium-ion battery.

Page 195: Dynamic Modulation of Material Properties by Solid State

189

11.2 Outlook

In this thesis, we bridge the concept of proton-based solid oxide cell which has largely been

confined to high temperature energy storage applications to dynamic modulation of material

properties in nanoscale devices. Below, we provide a brief outlook into potential developments

and applications in the future.

11.2.1 Integration of Hydrogen Storage in

Proton Magneto-Ionic Device

So far, we have primarily relied on proton sourced from humidity in ambient (through

water splitting) to modulate magnetic properties or for energy storage. This represents a

significant barrier towards practical implementation in electronics because operation which

varies according to season and places is simply too unreliable for use. Besides, the water splitting

reaction is a complicated process which involves multiple charge transfer and mass transport

steps. As a result, the reaction is very energy inefficient and can be rate limiting. For practical

devices, a hydrogen storage layer on top of the electrolyte will be needed to enable fast sourcing

of proton and to keep the device entirely self-contained. Some of these of hydrogen storage layer

includes oxides like WO3196 or alloys like Ni-Mg158.

Page 196: Dynamic Modulation of Material Properties by Solid State

190

11.2.2 Proton Magneto-Ionics for Memory

and Logic Devices

Because solid state proton gating can be used to modulate a wide range of magnetic

properties, it can be used to generate exotic spin textures such as skyrmions197 or control their

motions. One of the most promising architectures for magnetic memory device is the magnetic

racetrack, where magnetic bits can be dynamically created and moved for reading and writing

(figure 10.2)198. In a magnetic racetrack memory, a gate voltage can be used to generate

skyrmions as magnetic bits, which can then be moved by a spin current. By placing a series of

gate pads, a gate voltage can then used to synchronize the motion of these skyrmions or tune the

velocity of the skyrmionic bits (figure 11.2b)199.

Figure 11.1. Proton magneto-

ionic device with an integrated

hydrogen storage

Page 197: Dynamic Modulation of Material Properties by Solid State

191

Figure 11.2. Magnetic racetrack memory. (a) Schematic of a skyrmionic magnetic racetrack

memory where skymion is used to represent data bits. (b) Periodic gate pads to synchronize

current driven skyrmionic motion. Adapted from reference199,200.

An example application for proton magneto-ionics in logic operation is a logic gate as shown in

figure 11.3201. In this device, a gate voltage, VG can be used to control whether a cell is switched

by a current (IIN) or not. The cell output is then the Hall voltage readout (Rxy) from the cell.

Through a combination of a few cells, one can get a complementary AND or an OR gate.

Page 198: Dynamic Modulation of Material Properties by Solid State

192

Figure 11.3. Logic gate based on voltage controlled spin-orbit torques. (a) Schematic of the

logic gate. (b) Rxy output as a function of VG and IIN. (c) Table for the logic operations. Adapted

from reference201.

Page 199: Dynamic Modulation of Material Properties by Solid State

193

11.2.3 Proton Magneto-Ionics to Quantify

Proton Conductivity in Thin Film Oxides

While there is a rich literature of proton conductivity in bulk oxides at high temperature,

relatively little is understood about proton conductivity of oxides at room temperature and in

ultra-thin dimensions. By fabricating Pt/Co/MOx heterostructures using different metal oxides as

the electrolyte, proton magneto-ionics can be utilized to provide quantitative information about

proton conductivity and even proton mobility in these oxides.

In these experiments, MOKE or Hall magnetometry can be used for time resolved probe

of the Co magnetization while we keep as constant both the electrodes, and the ambient

humidity. The thickness of the oxide can be varied to study the effect of electric field on proton

conductivity. The effect of microstructure and phase on proton conductivity of the different

oxides can also be studied by varying the fabrication parameters of the oxide.

Page 200: Dynamic Modulation of Material Properties by Solid State

194

References

Page 201: Dynamic Modulation of Material Properties by Solid State

195

1. Hwang, H. Y. et al. Emergent phenomena at oxide interfaces. Nat. Mater. 11, 103–113

(2012).

2. Chappert, C., Fert, A. & Van Dau, F. N. The emergence of spin electronics in data

storage. Nat. Mater. 6, 813–823 (2007).

3. Reiss, G., Schmalhorst, J., Thomas, A., Hütten, A. & Yuasa, S. Magnetic tunnel junctions.

Springer Tracts Mod. Phys. 227, 291–333 (2007).

4. O’Handley, R. C. Modern Magnetic Materials: Principles and Applications. John Wiley

Sons New York Chichester Weinheim Brisbane Singapore Toronto (2000).

doi:10.1109/MEI.2005.1490004

5. Beach, G. S. D. Hysteresis in Ferromagnetic Materials. Lecture

6. Bayreuther, G. Magnetic Anisotropy. Lecture

7. Emori, S. Magnetic Domain Walls Driven by Interfacial Phenomena. PhD Thesis (2013).

8. Ross, C. A. Patterned Magnetic Recording Media. Annu. Rev. Mater. Res. 203–35 (2001).

9. Iwasaki, S. ichi & Ouchi, K. Co-Cr recording films with perpendicular magnetic

anisotropy. IEEE Trans. Magn. 14, 849–851 (1978).

10. Iwasaki, S. Perpendicular Magnetic Recording. IEEE Trans. Magn. 16, 71–76 (1980).

11. Tudu, B. & Tiwari, A. Recent Developments in Perpendicular Magnetic Anisotropy Thin

Films for Data Storage Applications. Vacuum 146, 329–341 (2017).

12. Maruyama, T. et al. Large voltage-induced magnetic anisotropy change in a few atomic

Page 202: Dynamic Modulation of Material Properties by Solid State

196

layers of iron. Nat. Nanotechnol. 4, 158–161 (2009).

13. Lin, C. J. et al. Magnetic and structural properties of Co/Pt multilayers. J. Magn. Magn.

Mater. 194–206, (1991).

14. Carcia, P. F., Meinhaldt, A. D. & Suna, A. Perpendicular magnetic anisotropy in Pd/Co

thin film layered structures. Appl. Phys. Lett. 47, 178–180 (1985).

15. G. H. O. Daalderop, P. J. K. and M. F. H. S. Magnetic anisotropy of a free-standing Co

monolayer and of multilayers which contain Co monolayers. Phys. Rev. B 50, 9989

(1994).

16. Johnson, M. T., Bloement, P. J. H., Broeder, F. J. A. den & Vries, J. . de. Magnetic

Anisotropy in Metallic Multilayers. Rep. Prog. Phys. 59, 1409–1458 (1996).

17. Nakajima, N. et al. Perpendicular Magnetic Anisotropy Caused by Interfacial

Hybridization via Enhanced Orbital Moment in Co/Pt Multilayers: Magnetic Circular X-

Ray Dichroism Study. Phys. Rev. Lett. 81, 5229–5232 (1998).

18. Heinrich, B. & Cochran, J. F. Ultrathin metallic magnetic films: magnetic anisotropies and

exchange interactions. Adv. Phys. 42, 523–639 (1993).

19. D. Weller, Y. Wu, J. Stohr, and M. G. Samant, B. D. Hermsmeier, C. C. Orbital magnetic

moments of Co in multilayers with perpendicular magnetic anisotropy. Phys. Rev. B 49,

888–896 (1994).

20. Yang, H. X. et al. First-principles investigation of the very large perpendicular magnetic

anisotropy at Fe|MgO and Co|MgO interfaces. Phys. Rev. B 84, 054401 (2011).

21. Dieny, B. & Chshiev, M. Perpendicular magnetic anisotropy at transition metal/oxide

Page 203: Dynamic Modulation of Material Properties by Solid State

197

interfaces and applications. Rev. Mod. Phys. 89, (2017).

22. Manchon, A. et al. Analysis of oxygen induced anisotropy crossover in Pt / Co / MOx

trilayers. J. Appl. Phys. 104, 043914 (2008).

23. Manchon, A. et al. X-ray analysis of the magnetic influence of oxygen in Pt/Co/AlOx

trilayers. J. Appl. Phys. 103, (2008).

24. Mankey, G. J., Kief, M. T., Huang, F. & Willis, R. F. Hydrogen chemisorption on

ferromagnetic thin film surfaces. J. Vac. Sci. Technol. A Vacuum, Surfaces, Film. 11,

2034–2039 (2002).

25. Valvidares, S. M., Dorantes-Dávila, J., Isern, H., Ferrer, S. & Pastor, G. M. Interface-

driven manipulation of the magnetic anisotropy of ultrathin Co films on Pt(111): Substrate

deposition of hydrogen and model calculations. Phys. Rev. B 81, 1–11 (2010).

26. Munbodh, K., Perez, F. A., Keenan, C. & Lederman, D. Effects of hydrogen / deuterium

absorption on the magnetic properties of Co / Pd multilayers. Phys. Rev. B 83, 094432

(2011).

27. Sander, D. et al. Reversible H-induced switching of the magnetic easy axis in Ni/Cu(001)

thin films. Phys. Rev. Lett. 93, 247203 (2004).

28. Santos, B. et al. Hydrogen-induced reversible spin-reorientation transition and magnetic

stripe domain phase in bilayer Co on Ru(0001). Phys. Rev. B 85, 3–8 (2012).

29. Hsu, P., Rózsa, L., Finco, A., Schmidt, L. & Palotás, K. Inducing skyrmions in ultrathin

Fe films by hydrogen exposure. arXiv 1711.06784v1

30. Dijken, S. Van, Vollmer, R., Poelsema, B. & Kirschner, J. The influence of CO and H

Page 204: Dynamic Modulation of Material Properties by Solid State

198

adsorption on the spin reorientation transition in Ni / Cu ( 0 0 1 ). J. Magn. Magn. Mater.

210, 316–328 (2000).

31. Gallego, S., Sanchez, N., Martin, S., Muñoz, M. C. & Szunyogh, L. Reversible

enhancement of the magnetism of ultrathin Co films by H adsorption. Phys. Rev. B -

Condens. Matter Mater. Phys. 82, 1–6 (2010).

32. Barman, A. & Haldar, A. Time-Domain study of magnetization dynamics in magnetic thin

films and micro- and nanostructures. Solid State Physics - Advances in Research and

Applications 65, (Elsevier Inc., 2014).

33. Hoffmann, A. Spin Hall Effects in Metals. Ieee Trans. Magn. 49, 5172–5193 (2013).

34. Liu, L. et al. Spin-Torque Switching with the Giant Spin Hall Effect of Tantalum. Science

(80-. ). 336, 555–559 (2012).

35. Liu, L., Lee, O. J., Gudmundsen, T. J., Ralph, D. C. & Buhrman, R. A. Current-Induced

Switching of Perpendicularly Magnetized Magnetic Layers Using Spin Torque from the

Spin Hall Effect. Phys. Rev. Lett. 109, 096602 (2012).

36. Emori, S., Bauer, U., Ahn, S. M., Martinez, E. & Beach, G. S. D. Current-driven

dynamics of chiral ferromagnetic domain walls. Nat. Mater. 12, 611–616 (2013).

37. Sankey, J. C. et al. Spin-transfer-driven ferromagnetic resonance of individual

nanomagnets. Phys. Rev. Lett. 96, 50–53 (2006).

38. Tulapurkar, A. A. et al. Spin-torque diode effect in magnetic tunnel junctions. Nature 438,

339–342 (2005).

39. Liu, L., Moriyama, T., Ralph, D. C. & Buhrman, R. A. Spin-torque ferromagnetic

Page 205: Dynamic Modulation of Material Properties by Solid State

199

resonance induced by the spin Hall effect. Phys. Rev. Lett. 106, 1–4 (2011).

40. Fuchs, G. D. et al. Spin-torque ferromagnetic resonance measurements of damping in

nanomagnets. Appl. Phys. Lett. 91, 20–23 (2007).

41. Dietl, T. A ten-year perspective on dilute magnetic semiconductors and oxides. Nat.

Mater. 9, 965–974 (2010).

42. Ohno, H. Making Nonmagnetic Semiconductors Ferromagnetic. Science (80-. ). 281, 951–

957 (1998).

43. Matsukura, F., Tokura, Y. & Ohno, H. Control of magnetism by electric fields. Nat.

Nanotechnol. 10, 209–220 (2015).

44. Ohno, H. et al. Electric- field control of ferromagnetism. Nature 408, 944–946 (2000).

45. Chiba, D., Yamanouchi, M., Matsukura, F. & Ohno, H. Electrical Manipulation of

Magnetization Reversal in a Ferromagnetic Semiconductor. Science (80-. ). 301, 943–946

(2003).

46. Eerenstein, W., Mathur, N. D. & Scott, J. F. Multiferroic and magnetoelectric materials.

Nature 442, 759–765 (2006).

47. Kim, S., Lee, J., Shin, S. & Wook, H. Voltage control of a magnetization easy axis in

piezoelectric / ferromagnetic hybrid films. J. Magn. Magn. Mater. 267, 127–132 (2003).

48. Zheng, H. et al. Multiferroic BaTiO3-CoFe2O4 Nanostructures. Science (80-. ). 303, 661–

664 (2004).

49. Zavaliche, F., Zheng, H., Yang, S. Y. & Zhan, Q. Electric Field-Induced Magnetization

Switching in Epitaxial Columnar Nanostructures. Nano Lett. 4, 1793–1796 (2005).

Page 206: Dynamic Modulation of Material Properties by Solid State

200

50. Ma, J., Hu, J., Li, Z. & Nan, C. Recent Progress in Multiferroic Magnetoelectric

Composites : from Bulk to Thin Films. Adv. Mater. 1062–1087 (2011).

51. Weisheit, M. et al. Electric field-induced modification of magnetism in thin-film

ferromagnets. Science (80-. ). 315, 349–351 (2007).

52. Shiota, Y., Maruyama, T., Nozaki, T. & Shinjo, T. Voltage-assisted Magnetization

Switching in Ultrathin Fe80Co20 Alloy Layers. arXiv 0905.0185, (2009).

53. Duan, C. et al. Surface Magnetoelectric Effect in Ferromagnetic Metal Films. Phys. Rev.

Lett. 101, 137201 (2008).

54. Tsujikawa, M. & Oda, T. Finite electric field effects in the large perpendicular magnetic

anisotropy surface Pt/Fe/Pt(001): A first-principles study. Phys. Rev. Lett. 102, 247203

(2009).

55. Nakamura, K., Shimabukuro, R., Fujiwara, Y., Akiyama, T. & Ito, T. Giant Modification

of the Magnetocrystalline Anisotropy in Transition-Metal Monolayers by an External

Electric Field. Phys. Rev. Lett. 102, 187201 (2009).

56. Duan, C. G. et al. Surface magnetoelectric effect in ferromagnetic metal films. Phys. Rev.

Lett. 101, 137201 (2008).

57. Shiota, Y. et al. Voltage-assisted magnetization switching in ultrathin Fe 80Co20 alloy

layers. Appl. Phys. Express 2, (2009).

58. Bauer, U., Emori, S. & Beach, G. S. D. Voltage-controlled domain wall traps in

ferromagnetic nanowires. Nat. Nanotechnol. 8, 411–416 (2013).

59. Emori, S., Bauer, U., Woo, S. & Beach, G. S. D. Large voltage-induced modification of

Page 207: Dynamic Modulation of Material Properties by Solid State

201

spin-orbit torques in Pt/Co/GdOx. Appl. Phys. Lett. 105, (2014).

60. Bi, C. et al. Reversible control of Co magnetism by voltage-induced oxidation. Phys. Rev.

Lett. 113, 267202 (2014).

61. Bauer, U. et al. Magneto-ionic control of interfacial magnetism. Nat. Mater. 14, 174–181

(2015).

62. Gilbert, D. A. et al. Structural and magnetic depth profiles of magneto-ionic

heterostructures beyond the interface limit. Nat. Commun. 7, 12264 (2016).

63. Gilbert, D. A. et al. Controllable positive exchange bias via redox-driven oxygen

migration. Nat. Commun. 7, 11050 (2016).

64. Sakamaki, M. & Amemiya, K. Observation of an electric field-induced interface redox

reaction and magnetic modification in GdOx/Co thin film by means of depth-resolved X-

ray absorption spectroscopy. Phys. Chem. Chem. Phys. 20, 20004–20009 (2018).

65. Mishra, R. et al. Electric-field control of spin accumulation direction for spin-orbit

torques. Nat. Commun. 10, 248 (2019).

66. Li, H. B. et al. Electric-field control of ferromagnetism through oxygen ion gating. Nat.

Commun. 8, 2156 (2017).

67. Lu, N. et al. Electric-field control of tri-state phase transformation with a selective dual-

ion switch. Nature 546, 124–128 (2017).

68. Zhou, X. et al. Role of Oxygen Ion Migration in the Electrical Control of Magnetism in

Pt/Co/Ni/HfO2 Films. J. Phys. Chem. C 120, 1633–1639 (2016).

69. Li, Q. et al. Electrical control of exchange bias via oxygen migration across CoO-ZnO

Page 208: Dynamic Modulation of Material Properties by Solid State

202

nanocomposite barrier. Appl. Phys. Lett. 109, 1–5 (2016).

70. Hirai, T. et al. Control of magnetic anisotropy in Pt/Co system using ionic liquid gating.

Appl. Phys. Express 9, (2016).

71. Dasgupta, S. et al. Intercalation-driven reversible control of magnetism in bulk

ferromagnets. Adv. Mater. 26, 4639–4644 (2014).

72. Dasgupta, S. et al. Toward On-and-Off Magnetism: Reversible Electrochemistry to

Control Magnetic Phase Transitions in Spinel Ferrites. Adv. Funct. Mater. 26, 7507–7515

(2016).

73. Zhang, Q. et al. Lithium-Ion Battery Cycling for Magnetism Control. Nano Lett. 16, 583–

587 (2016).

74. Di, N. Electric and chemical manipulation of magnetic anisotropy of ultrathin

ferromagnetic films. PhD Thesis. (2015).

75. Jani, M. G., Halliburton, L. E. & Halperin, A. Observation of simple lithium-associated

electron trap in crystalline SiO2. Phys. Rev. Lett. 56, 1392–1395 (1986).

76. Minh, N. Recent advances in solid oxide fuel cell technology. Am. Ceram. Soc. Bull. 82,

305–313 (2003).

77. Haile, S. M. Fuel cell materials and components. Acta Mater. 51, 5981–6000 (2003).

78. Laguna-Bercero, M. A. Recent advances in high temperature electrolysis using solid oxide

fuel cells: A review. J. Power Sources 203, 4–16 (2012).

79. Malavasi, L., Fisher, C. A. J. & Islam, M. S. Oxide-ion and proton conducting electrolyte

materials for clean energy applications: Structural and mechanistic features. Chem. Soc.

Page 209: Dynamic Modulation of Material Properties by Solid State

203

Rev. 39, 4370–4387 (2010).

80. Bi, L., Boulfrad, S. & Traversa, E. Steam electrolysis by solid oxide electrolysis cells

(SOECs) with proton-conducting oxides. Chem. Soc. Rev. 43, 8255–8270 (2014).

81. Kreuer, K.-D. Proton Conductivity: Materials and Applications. Chem. Mater. 8, 610–641

(1996).

82. Kreuer, K. D. Proton-Conducting Oxides. Annu. Rev. Mater. Res. 33, 333–359 (2003).

83. Skinner, S. J. & Kilner, J. A. Oxygen ion conductors. Mater. Today 6, 30–37 (2003).

84. Sun, C., Hui, R. & Roller, J. Cathode materials for solid oxide fuel cells: A review. J.

Solid State Electrochem. 14, 1125–1144 (2010).

85. Cowin, P. I., Petit, C. T. G., Lan, R., Irvine, J. T. S. & Tao, S. Recent progress in the

development of anode materials for solid oxide fuel cells. Adv. Energy Mater. 1, 314–332

(2011).

86. Dusastre, V., Tao, S. & Irvine, J. T. . A redox-stable efficient anode for solid-oxide fuel

cells. Mater. Sustain. Energy 259–262 (2012).

87. Kreuer, K.-D., Paddison, S. J., Spohr, E. & Schuster, M. Transport in Proton Conductors

for Fuel-Cell Applications: Simulations, Elementary Reactions, and Phenomenology.

Chem. Rev. 104, 4637–4678 (2004).

88. Kreuer, K. Proton Conductivity : Materials and Applications. Chem.Mater 8, 610–641

(1996).

89. Medvedev, D. et al. BaCeO3: Materials development, properties and application. Prog.

Mater. Sci. 60, 72–129 (2014).

Page 210: Dynamic Modulation of Material Properties by Solid State

204

90. Pergolesi, D. et al. High proton conduction in grain-boundary-free yttrium-doped barium

zirconate films grown by pulsed laser deposition. Nat. Mater. 9, 846–852 (2010).

91. Fabbri, E., Pergolesi, D. & Traversa, E. Materials challenges toward proton-conducting

oxide fuel cells: A critical review. Chem. Soc. Rev. 39, 4355–4369 (2010).

92. Larring, Y. & Norby, T. Protons in rare earth oxides. Solid State Ionics 77, 147–151

(1995).

93. Truls, N. Solid-state protonic conductors: principles, properties, progress and prospects.

Solid State Ionics 125, 1–11 (1999).

94. Haugsrud, R. & Norby, T. Proton conduction in rare-earth ortho-niobates and ortho-

tantalates. Nat. Mater. 5, 193–196 (2006).

95. Tanner, C. W. & Virkar, A. V. InstabiIiiy of BaCeO3 in H20-Containing Atmospheres. J.

Electrochem. Soc. 143, 1386–1389 (1996).

96. Fabbri, E. et al. Developments and perspectives of oxide-based catalysts for the oxygen

evolution reaction. Catal. Sci. Technol. 4, 3800–3821 (2014).

97. Gregori, G., Shirpour, M. & Maier, J. Proton conduction in dense and porous

nanocrystalline ceria thin films. Adv. Funct. Mater. 23, 5861–5867 (2013).

98. Scherrer, B. et al. On proton conductivity in porous and dense yttria stabilized zirconia at

low temperature. Adv. Funct. Mater. 23, 1957–1964 (2013).

99. Shirpour, M., Gregori, G., Merkle, R. & Maier, J. On the proton conductivity in pure and

gadolinium doped nanocrystalline cerium oxide. Phys. Chem. Chem. Phys. 13, 937–940

(2011).

Page 211: Dynamic Modulation of Material Properties by Solid State

205

100. Yang, J. J., Strukov, D. B. & Stewart, D. R. Memristive devices for computing. Nat.

Nanotechnol. 8, 13–24 (2013).

101. Waser, R. Solubility of Hydrogen Defects in Doped and Undoped BaTiO3. J. Am. Ceram.

Soc. 71, 58–63 (1988).

102. Hong, W. T. et al. Toward the rational design of non-precious transition metal oxides for

oxygen electrocatalysis. Energy Environ. Sci. 8, 1404–1427 (2015).

103. Suen, N. T. et al. Electrocatalysis for the oxygen evolution reaction: Recent development

and future perspectives. Chem. Soc. Rev. 46, 337–365 (2017).

104. Man, I. C. et al. Universality in Oxygen Evolution Electrocatalysis on Oxide Surfaces.

ChemCatChem 3, 1159–1165 (2011).

105. Kim, J. S., Kim, B., Kim, H. & Kang, K. Recent Progress on Multimetal Oxide Catalysts

for the Oxygen Evolution Reaction. Adv. Energy Mater. 8, 1–26 (2018).

106. Zeng, M. & Li, Y. Recent advances in heterogeneous electrocatalysts for the hydrogen

evolution reaction. J. Mater. Chem. A 3, 14942–14962 (2015).

107. Stamenkovic, V. R., Strmcnik, D., Lopes, P. P. & Markovic, N. M. Energy and Fuels from

Electrochemical Interfaces. Nat. Mater. 16, 57–69 (2017).

108. Sheng, W., Gasteiger, H. A. & Shao-Horn, Y. Hydrogen Oxidation and Evolution

Reaction Kinetics on Platinum: Acid vs Alkaline Electrolytes. J. Electrochem. Soc. 157,

B1529 (2010).

109. Strmcnik, D. et al. Improving the hydrogen oxidation reaction rate by promotion of

hydroxyl adsorption. Nat. Chem. 5, 300–306 (2013).

Page 212: Dynamic Modulation of Material Properties by Solid State

206

110. Nielsen, J. & Hjelm, J. Impedance of SOFC electrodes: A review and a comprehensive

case study on the impedance of LSM:YSZ cathodes. Electrochim. Acta 115, 31–45

(2014).

111. Rossmeisl, J., Qu, Z., Zhu, H., Kroes, G. & Norskov, J. K. Electrolysis of Water on Oxide

Surfaces. J. Electrochem. Chem. 607, 83–89 (2007).

112. Rossmeisl, J., Logadottir, A. & Nørskov, J. K. Electrolysis of water on ( oxidized ) metal

surfaces. Chem. Phys. 319, 178–184 (2005).

113. Nørskov, J. K. et al. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell

Cathode. 17886–17892 (2004).

114. Adler, S. B. Factors Governing Oxygen Reduction in Solid Oxide Fuel Cell Cathodes.

Chem. Rev. 104, 4791–4844 (2004).

115. Peng, R., Wu, T., Liu, W., Liu, X. & Meng, G. Cathode processes and materials for solid

oxide fuel cells with proton conductors as electrolytes. J. Mater. Chem. 20, 6218–6225

(2010).

116. Tsai, T. & Barn, S. A. Effect of LSM-YSZ cathode on thin electrolyte solid oxide fuel cell

performance. Solid State Ionics 93, 207–217 (1997).

117. Primdahl, S. Oxidation of Hydrogen on Ni/Yttria-Stabilized Zirconia Cermet Anodes. J.

Electrochem. Soc. 144, 3409 (2006).

118. Barsoukov, E. & Macdonald, J. R. Theory , Experiment , and. America 177, (2005).

119. Lai, W. & Haile, S. M. Impedance spectroscopy as a tool for chemical and

electrochemical analysis of mixed conductors: A case study of ceria. J. Am. Ceram. Soc.

Page 213: Dynamic Modulation of Material Properties by Solid State

207

88, 2979–2997 (2005).

120. Jaminik, J. Generalised Equivalent Circuits for Mass and Charge Transport.Pdf. Phys.

Chem. Chem. Phys. 3, 1668–1678 (2001).

121. Harrington, D. A. & Van Den Driessche, P. Mechanism and equivalent circuits in

electrochemical impedance spectroscopy. Electrochim. Acta 56, 8005–8013 (2011).

122. Tuller, H. L. Ionic conduction in nanocrystalline materials. Solid State Ionics 131, 143–

157 (2000).

123. Tuller, H. L. Ionic conduction in nanocrystalline materials. Solid State Ionics 131, 143–

157 (2000).

124. Greene, J. E. Review Article: Tracing the recorded history of thin-film sputter deposition:

From the 1800s to 2017. J. Vac. Sci. Technol. A Vacuum, Surfaces, Film. 35, 05C204

(2017).

125. Kaiser, N. Review of the fundamentals of thin-film growth. Appl. Opt. 41, 3053 (2007).

126. Thornton, J. A. Influence of apparatus geometry and deposition conditions on the structure

and topography of thick sputtered coatings. J. Vac. Sci. Technol. 11, 666–670 (2002).

127. Hentzell, H. T. G., Smith, D. A. & Grovernor, C. The development of grain structure

during growth of metallic thin films. Acta Metall. 32, 773–781 (1984).

128. Kerkache, L., Layadi, A. & Mosser, A. Effect of oxygen partial pressure on the structural

and optical properties of dc sputtered ITO thin films. J. Alloys Compd. 485, 46–50 (2009).

129. Choi, C. G. et al. Effects of oxygen partial pressure on the microstructure and electrical

properties of indium tin oxide film prepared by d.c. magnetron sputtering. Thin Solid

Page 214: Dynamic Modulation of Material Properties by Solid State

208

Films 258, 274–278 (1995).

130. Price, E. P. Characterization of transport processes in magnetic tunnel junctions. (2001).

131. Hamrle, J. Magneto-optical Kerr effect ( MOKE ) Photon-photon spectroscopies (

absorption ) I : (2014).

132. Qiu, Z. Q. & Bader, S. D. Surface magneto-optic Kerr effect (SMOKE). J. Magn. Magn.

Mater. 200, 664–678 (1999).

133. Bauer, U. Voltage Programmable Materials. PhD Thesis. (2015).

134. Garello, K. et al. Symmetry and magnitude of spin–orbit torques in ferromagnetic

heterostructures. Nat. Nanotechnol. 8, 587–593 (2013).

135. Nagaosa, N., Sinova, J., Onoda, S., MacDonald, A. H. & Ong, N. P. Anomalous Hall

effect. Rev. Mod. Phys. 82, 1539–1592 (2010).

136. An, H., Kageyama, Y., Kanno, Y., Enishi, N. & Ando, K. Spin-torque generator

engineered by natural oxidation of Cu. Nat. Commun. 7, 1–8 (2016).

137. Norby, T., Widerøe, M., Glöckner, R. & Larring, Y. Hydrogen in oxides. Dalt. Trans. 0,

3012–3018 (2004).

138. Tsuruoka, T. et al. Humidity effects on the redox reactions and ionic transport in a

Cu/Ta2O5/Pt atomic switch structure process. Jpn. J. Appl. Phys. 55, 06091–09095

(2016).

139. Lübben, M., Wiefels, S., Waser, R. & Valov, I. Processes and Effects of Oxygen and

Moisture in Resistively Switching TaOxand HfOx. Adv. Electron. Mater. 4, 1–11 (2018).

Page 215: Dynamic Modulation of Material Properties by Solid State

209

140. Tsuruoka, T. et al. Effects of moisture on the switching characteristics of oxide-based,

gapless-type atomic switches. Adv. Funct. Mater. 22, 70–77 (2012).

141. Messerschmitt, F., Kubicek, M. & Rupp, J. L. M. How Does Moisture Affect the Physical

Property of Memristance for Anionic-Electronic Resistive Switching Memories? Adv.

Funct. Mater. 25, 5117–5125 (2015).

142. Sediva, E., Bowman, W. J., Gonzalez-Rosillo, J. C. & Rupp, J. L. M. Investigation of the

Eightwise Switching Mechanism and Its Suppression in SrTiO3 Modulated by Humidity

and Interchanged Top and Bottom Platinum and LaNiO3 Electrode Contacts. Adv.

Electron. Mater. 1800566, 1–11 (2018).

143. Moeller, T. & Kremers, H. E. The basicity characteristics of scandium, yttrium, and the

rare earth elements. Chem. Rev. 37, 97–159 (1945).

144. Adachi, G. & Imanaka, N. The Binary Rare Earth Oxides. Chem. Rev. 98, 1479–1514

(1998).

145. Fanciulli, M. & Scarel, G. Rare earth oxide thin films: Growth, Characterization and

Applications. Springer (2007). doi:10.1007/11499893_20

146. Cohen, S. et al. The interaction of H2O with the surface of polycrystalline gadolinium at

the temperature range of 300-570 K. Surf. Sci. 617, 29–35 (2013).

147. Jeon, S. & Hwang, H. Effect of hygroscopic nature on the electrical characteristics of

lanthanide oxides (Pr2O3, Sm2O3, Gd2O3, and Dy2O3). J. Appl. Phys. 93, 6393–6395

(2003).

148. Külah, E. et al. Surface chemistry of rare-earth oxide surfaces at ambient conditions:

Page 216: Dynamic Modulation of Material Properties by Solid State

210

Reactions with water and hydrocarbons. Sci. Rep. 7, 1–10 (2017).

149. Villars, P. Gd2O3 C-phase Crystal Structure. Inorganic Solid Phases, SpringerMaterials

(online database). (Springer-Verlag Berlin Heidelberg & NIMS, Japan, 2016).

150. Villars, P. Gd(OH)3 Crystal Structure. Inorganic Solid Phases, SpringerMaterials (online

database). (Springer-Verlag Berlin Heidelberg & NIMS, Japan, 2016).

151. Aruna, I., Mehta, B. R., Malhotra, L. K. & Shivaprasad, S. M. Stability and hydrogenation

of ‘bare’ gadolinium nanoparticles. Adv. Funct. Mater. 15, 131–137 (2005).

152. Brundle, C. R. & Wandelt, K. The Interactions of oxygen with Gadolinium: UPS ans XPS

Studies. Surf. Sci. 157, 162–182 (1985).

153. Lasia, A. in Handbook of Fuel Cells- Fundamentals, Technology and Applications 417–

434 (2003). doi:10.1002/9780470974001.f204033

154. Tymoczko, J., Calle-Vallejo, F., Schuhmann, W. & Bandarenka, A. S. Making the

hydrogen evolution reaction in polymer electrolyte membrane electrolysers even faster.

Nat. Commun. 7, 10990 (2016).

155. Tan, A. J. et al. Magneto-ionic control of magnetism using a solid-state proton pump. Nat.

Mater. 18, 35–42 (2019).

156. Zhu, X. et al. In Situ Nanoscale Electric Field Control of Magnetism by Nanoionics. Adv.

Mater. 28, 7658–7665 (2016).

157. Adams, B. D. & Chen, A. The role of palladium in a hydrogen economy. Mater. Today

14, 282–289 (2011).

158. Schlapbach, L. & Züttel, A. Hydrogen-storage materials for mobile applications. Nature

Page 217: Dynamic Modulation of Material Properties by Solid State

211

414, 353–358 (2002).

159. Yang, H., Thiaville, A., Rohart, S., Fert, A. & Chshiev, M. Anatomy of Dzyaloshinskii-

Moriya Interaction at Co / Pt Interfaces. Phys. Rev. Lett. 115, 267210 (2015).

160. S.S.P.Parkin, R.Bhadra & Roche, K. P. Oscillatory magnetic exchange coupling through

thin copper layers. Phys. Rev. Lett. 66, 2152–2155 (1991).

161. Campbell, C. T. Catalyst–support interactions: Electronic perturbations. Nat. Chem. 4,

597–598 (2012).

162. Rodriguez, J. A., Liu, P., Hrbek, J., Evans, J. & Pérez, M. Water gas shift reaction on Cu

and Au nanoparticles supported on CeO 2(111) and ZnO(0001): Intrinsic activity and

importance of support interactions. Angew. Chemie - Int. Ed. 46, 1329–1332 (2007).

163. Valov, I. & Lu, W. D. Nanoscale electrochemistry using dielectric thin films as solid

electrolytes. Nanoscale 8, 13828–13837 (2016).

164. Tsuruoka, T. et al. Effects of moisture on the switching characteristics of oxide-based,

gapless-type atomic switches. Adv. Funct. Mater. 22, 70–77 (2012).

165. Yin, Q. et al. Cathode bubbles induced by moisture electrolysis in TiO2-x-based resistive

switching cells. J. Phys. D. Appl. Phys. 49, 09LT01 (2016).

166. Bard, A. J., Faulkner, L. R., Swain, E. & Robey, C. Electrochemical Methods:

Fundamentals and Applications. (John Wiley and Sons).

167. Yoon, H. et al. Reversible phase modulation and hydrogen storage in multivalent VO2

epitaxial thin films. Nat. Mater. 15, 1113–1119 (2016).

168. Burr, G. W. et al. Neuromorphic computing using non-volatile memory. Adv. Phys. X 2,

Page 218: Dynamic Modulation of Material Properties by Solid State

212

89–124 (2017).

169. Pershin, Y. V. & Di Ventra, M. Experimental demonstration of associative memory with

memristive neural networks. Neural Networks 23, 881–886 (2010).

170. Thompson, A. C. & D.Vaughn. X-RAY DATA BOOKLET. (2001).

171. Bennett, P. A. & Fuggle, J. C. Electronic structure and surface kinetics of palladium

hydride studied with x-ray photoelectron spectroscopy and electron-energy-loss

spectroscopy. Phys. Rev. B 26, 6030–6039 (1982).

172. Davoli, I. et al. Palladium L3 absorption edge of PdH0.6 films: Evidence for hydrogen

induced unoccupied states. Solid State Commun. 71, 383–390 (1989).

173. Tew, M. W., Miller, J. T. & van Bokhoven, J. A. Particle Size Effect of Hydride

Formation and Surface Hydrogen Adsorption of Nanosized Palladium Catalysts: L 3 Edge

vs K Edge X-ray Absorption Spectroscopy. J. Phys. Chem. C 113, 15140–15147 (2009).

174. Richardson, T. J. et al. X-ray absorption spectroscopy of transition metal-magnesium

hydride thin films. J. Alloys Compd. 356–357, 204–207 (2003).

175. Gambardella, P. & Miron, I. M. Current-induced spin–orbit torques. Philos. Trans. R. Soc.

A 369, 3175–3197 (2011).

176. Hellman, F. et al. Interface-Induced Phenomena in Magnetism. Rev. Mod. Phys. 89,

025006 (2017).

177. Emori, S., Bauer, U., Ahn, S., Martinez, E. & Beach, G. S. D. Current-driven dynamics of

chiral ferromagnetic domain walls. Nat. Mater. 12, 611–616 (2013).

178. Pai, C. F., Ou, Y., Vilela-Leão, L. H., Ralph, D. C. & Buhrman, R. A. Dependence of the

Page 219: Dynamic Modulation of Material Properties by Solid State

213

efficiency of spin Hall torque on the transparency of Pt/ferromagnetic layer interfaces.

Phys. Rev. B - Condens. Matter Mater. Phys. 92, (2015).

179. Robertson, J. High dielectric constant oxides. Eur. Phys. J. Appl. Phys. 28, 265–291

(2004).

180. Paton, M. G. & Maslen, E. N. A refinement of the crystal structure of yttria. Acta

Crystallogr. 19, 307–310 (1965).

181. French, R., Glass, S., Ohuchi, F., Xu, Y. & Ching, W. Experimental and theoretical

determination of the electronic structure and optical properties of three phases of ZrO2.

Phys. Rev. B 49, 5133–5142 (1994).

182. Scott, H. G. Phase relationships in the zirconia-yttria system. J. Mater. Sci. 10, 1527–1535

(1975).

183. Stephenson, N. C. & Roth, R. S. Structural systematics in the binary system Ta2O5–WO3.

V. The structure of the low-temperature form of tantalum oxide L-Ta2O5. Acta

Crystallogr. 27, 1037–1044 (1971).

184. Wang, Y. et al. Lithium and lithium ion batteries for applications in microelectronic

devices: A review. J. Power Sources 286, 330–345 (2015).

185. Pikul, J. H., Gang Zhang, H., Cho, J., Braun, P. V & King, W. P. High-power lithium ion

microbatteries from interdigitated three-dimensional bicontinuous nanoporous electrodes.

Nat. Commun. 4, 1732 (2013).

186. Lai, W. et al. Ultrahigh-energy-density microbatteries enabled by new electrode

architecture and micropackaging design. Adv. Mater. 22, 139–144 (2010).

Page 220: Dynamic Modulation of Material Properties by Solid State

214

187. Bates, J. B. B. et al. Thin-film rechargeable lithium batteries. J. Power Sources 54, 58–62

(1995).

188. Dudney, N. J. Solid-state thin-film rechargeable batteries. Mater. Sci. Eng. B 116, 245–

249 (2005).

189. Gillaspie, D. T., Tenent, R. C. & Dillon, A. C. Metal-oxide films for electrochromic

applications: present technology and future directions. J. Mater. Chem. 20, 9585 (2010).

190. Niklasson, G. a. & Granqvist, C. G. Electrochromics for smart windows: thin films of

tungsten oxide and nickel oxide, and devices based on these. J. Mater. Chem. 17, 127

(2007).

191. Granqvist, C. G. Recent progress in thermochromics and electrochromics: A brief survey.

Thin Solid Films 1, 1–7 (2015).

192. Triana, C. a, Granqvist, C. G. & Niklasson, G. a. Electrochromic Properties of Li + -

Intercalated Amorphous Tungsten (aWO3- x ) and Titanium (aTiO2- x ) Oxide Thin

Films. J. Phys. Conf. Ser. 559, 012004 (2014).

193. Cai, G. F. et al. Ultra fast electrochromic switching of nanostructured NiO films

electrodeposited from choline chloride-based ionic liquid. Electrochim. Acta 87, 341–347

(2013).

194. Cai, G. et al. Hierarchical structure Ti-doped WO3 film with improved electrochromism

in visible-infrared region. RSC Adv. 3, 6896–6905 (2013).

195. Lu, Y., Liu, L., Mandler, D. & Lee, P. S. High switching speed and coloration efficiency

of titanium-doped vanadium oxide thin film electrochromic devices. J. Mater. Chem. C 1,

Page 221: Dynamic Modulation of Material Properties by Solid State

215

7380 (2013).

196. Granqvist, C. G. Electrochromic tungsten oxide: Review of progress 1993-1998. Sol.

Energy Mater. Sol. Cells 60, 201–262 (2000).

197. Woo, S. et al. Observation of room-temperature magnetic skyrmions and their current-

driven dynamics in ultrathin metallic ferromagnets. Nat. Mater. 15, 501–6 (2016).

198. Parkin, S. S. P., Hayashi, M. & Thomas, L. Magnetic Racetrack Memory. 320, 190–195

(2008).

199. Zhao, W. et al. Complementary Skyrmion Racetrack Memory With Voltage

Manipulation. IEEE Electron Device Lett. 37, 924–927 (2016).

200. Zhang, S., Baker, A. A., Komineas, S. & Hesjedal, T. Topological computation based on

direct magnetic logic communication. Sci. Rep. 5, 1–11 (2015).

201. Baek, S. H. C. et al. Complementary logic operation based on electric-field controlled

spin-orbit torques. Nat. Electron. 1, 398–403 (2018).