Sunteți pe pagina 1din 152

Spin Dynamics in Novel Materials Systems

Dissertation
Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the
Graduate School of The Ohio State University
By
Howard Yu
Graduate Program in Physics
The Ohio State University
2015
Dissertation Committee:
Professor Ezekiel Johnston-Halperin, Advisor
Professor Jay Gupta
Professor Nandini Trivedi
Professor Louis DiMauro

Copyright by
Howard Yu
2015

Abstract
Spintronics and organic electronics are fields that have made considerable advances
in recent years, both in fundamental research and in applications. Organic materials have a
number of attractive properties that enable them to complement applications traditionally
fulfilled by inorganic materials, while spintronics seeks to take advantage of the spin
degree of freedom to produce new applications. My research is aimed at combining these
two fields to develop organic materials for spintronics use.
My thesis is divided into three primary projects centered around an organic-based
semiconducting ferrimagnet, vanadium tetracyanoethylene. First, we investigated the
transport characteristics of a hybrid organic-inorganic heterostructure. Semiconductors
form the basis of the electronics industry, and there has been considerable effort put
forward to develop organic semiconductors for applications like organic light-emitting
diodes and organic thin film transistors. Working with hybrid organic-inorganic
semiconductor device structures allows us to potentially take advantage of the
infrastructure that has already been developed for silicon and other inorganic
semiconductors. This could potentially pave the way for a new class of active hybrid
devices with multifunctional behavior.
Second, we investigated the magnetic resonance characteristics of V[TCNE]x, in
multiple measurement schemes and exploring the effect of temperature, frequency, and
chemical tuning. Recently, the spintronics community has shifted focus from static
ii

electrical spin injection to various dynamic processes, such as spin pumping and thermal
effects. Spin pumping in particular is an intriguing way to generate pure spin currents via
magnetic resonance that has attracted a high degree of interest, with the FMR linewidth
being an important metric for spin injection. Furthermore, we can potentially use these
measurements to probe the magnetic properties as we change the physical properties of the
materials by chemically tuning the organic ligand. We are therefore interested in exploring
the resonance properties of this materials system to lay the groundwork for future spin
pumping applications.
Third, we have made preliminary measurements of spin pumping in hybrid and allorganic bilayer structures. As mentioned above, FMR-driven spin pumping is method for
generating pure spin currents with no associated charge motion. This can be detected in a
number of ways, one of which is monitoring the FMR characteristics of two ferromagnets
in close contact, where spins injected from one magnet into the other changes the linewidth.
In conjunction with the magnetic resonance measurements, we have started to investigate
the FMR properties of these bilayer systems.

iii

Acknowledgments
First and foremost, I would like to thank my advisor Professor Ezekiel JohnstonHalperin, who has provided the resources, guidance, and support over my graduate career
that I needed to succeed.
I would also like to thank the many colleagues who have helped me along the way.
From my group, past and present, Yi-Hsin Chiu, Kurtis Wickey, Yu-Sheng Ou, Justin
Young, Mike Chilcote, Matt Sheffield, Ian Froning, Shawna Hollen, Yong Pu, Lei Fang,
Dongkyun Ko, Ke Li, Sarah Parks, and Patrick Truitt have all contributed to my research
efforts and my growth as a scientist. From our collaborators within the department, Megan
Harberts, Yu Lu, K. Deniz Bozdag, Chia-Yi Chen, Professor Art Epstein, Adam Hauser,
Professor Fengyuan Yang, Rohan Adur, and Professor P. Chris Hammel have helped make
our collaborations fruitful. Denis Pelekhov, Camelia Selcu, and Bob Wells all taught me
much about various pieces of equipment and measurements.
Furthermore, I would like to thank the support staff of the physics department. In
particular, Kris Dunlap, Amanda Zuurdeeg, and Jaimie Mollison have helped me navigate
the sometimes impenetrable bureaucracy of Ohio State, and John Gosser, Josh Gueth, and
Jonathan Shover from the machine shop made the endless stream of parts I needed. Without
their help, I would not have made it this far.
My wonderful friends have supported me throughout my time in graduate school.
Andrew Berger, Maggie Mills, Matt Steinpreis, Stephanie Arend, David Pignotti, Megan
iv

and Dan Harberts, Jai Salzwedel, Amy Holthusen, Greg Malen, Calen Henderson, Andrea
Chu, and many others have helped me in ways large and small. My life would be much
poorer without you all.
Finally, my parents and family have given me their support, encouragement, and
love even as I moved across the country. I would not be the person I am today were it not
for them, and I am eternally grateful.

Vita
June 2004Lynbrook High School, San Jose, CA
May 2008B.S. Physics, Harvey Mudd College, Claremont, CA
September 2008 to presentGraduate Teaching and Research Associate, Department of Physics,
The Ohio State University
Publications
Y. Lu, H. Yu, M. Harberts, A.J. Epstein, E. Johnston-Halperin. Vanadium [ethyl
tricyanoethylene carboxylate]x~2 : a new molecule-based magnet. Chemistry of Materials,
2015 (in review)
I. Froning, M. Harberts, Lu. Y, H. Yu, A.J. Epstein, E. Johnston-Halperin. Thin-film
encapsulation of the air-sensitive organic-based ferrimagnet vanadium
tetracyanoethylene. Applied Physics Letters, 2015. DOI: 10.1063/1.4916241
M. Harberts, Y. Lu, H. Yu, A.J. Epstein, E. Johnston-Halperin. Chemical vapor
deposition of an organic magnet, vanadium tetracyanoethylene, V[TCNE]x~2. Journal of
Visualized Experiments, 2015 (in press).
Y. Lu, M. Harberts, C.Y. Kao, H. Yu, E. Johnston-Halperin, A.J. Epstein. Thin film
deposition of an organic magnet based on vanadium methyl tricyanoethylenecarboxylate.
Advanced Materials, 2014. DOI: 10.1002/adma.201403834
H. Yu, M. Harberts, R. Adur, Y. Lu, P.C. Hammel, E. Johnston-Halperin, A.J. Epstein.
Ultra-narrow ferromagnetic resonance in organic-based thin films grown via lowtemperature chemical vapor deposition. Applied Physics Letters, 2014. DOI:
10.1063/1.4887924
H. Yu, M. Harberts, L. Fang, K. Deniz Bozdag, C. Y. Chen, A. J. Epstein, E. JohnstonHalperin. Electrical transport in a hybrid organic/inorganic heterostructure. Proceedings
of SPIE Spintronics IV, Aug 21, 2011. DOI: 10.1117/12.894641 *Non peer reviewed
Fields of Study
Major Field: Physics

vi

Table of Contents
Abstract .............................................................................................................................. ii
Acknowledgments ............................................................................................................. iv
Vita .................................................................................................................................... vi
List of Figures ................................................................................................................... ix
Chapters
1 Introduction ................................................................................................................... 1
1.1 Magnetism ........................................................................................................ 3
1.2 Ferromagnetic resonance ................................................................................ 10
1.3 Ferromagnetic resonance-driven spin pumping .............................................. 14
1.4 Organic materials and magnets ....................................................................... 16
1.5 Spin dynamics in novel materials systems - focus of my research .................. 18
References ............................................................................................................ 20
2 Background and theory ............................................................................................... 25
2.1 Introduction .................................................................................................... 25
2.2 Vanadium tetracyanoethylene and related materials ....................................... 27
2.3 Chemical vapor deposition process ................................................................. 29
2.4 Magnetic properties of V[TCNE]x~2 ................................................................ 31
2.5 V[TCNE]x~2 spintronics to date ....................................................................... 37
2.6 Conclusion ...................................................................................................... 40
References ............................................................................................................ 42
3 Transport in a hybrid organic-inorganic heterostructures ...................................... 42
3.1 Introduction .....................................................................................................42
3.2 Transport in p-i-n junctions ............................................................................. 44
3.2.1 Ideal p-n junction behavior ............................................................... 44
3.2.2 Transport in a p-i-n junction ............................................................. 48
3.2.3 Band structure simulation ................................................................ 49
3.3 Sample fabrication .......................................................................................... 51
3.4 Electrical transport in h7brid organic-inorganic diode .................................... 54
3.5 Conclusion ...................................................................................................... 59
References ............................................................................................................ 60
4 Ferromagnetic resonance in thin film organic-based magnets ................................. 61
4.1 Introduction .................................................................................................... 61
4.2 Theory of magnetic resonance ........................................................................ 63
4.3 Sample preparation ......................................................................................... 68
vii

4.4 Cavity ferromagnetic resonance ...................................................................... 69


4.4.1 Principles of operation ..................................................................... 69
4.4.2 Ultra-narrow ferromagnetic resonance in V[TCNE]x~2 .................... 71
4.4.3 Angular dependence of ferromagnetic resonance ............................. 73
4.4.4 Ferromagnetic resonance in other organic-based magnets ............... 78
4.4.4.1 FMR of V[MeTCEC]x ....................................................... 78
4.4.4.2 FMR of V[ETCEC]x .......................................................... 82
4.5 Stripline ferromagnetic resonance ................................................................... 85
4.5.1 Principles of operation ..................................................................... 85
4.5.2 Air-free magnetic resonance can ...................................................... 89
4.5.3 Frequency dependence ..................................................................... 91
4.5.4 Temperature dependence ................................................................. 94
4.6 Conclusion ...................................................................................................... 97
References ............................................................................................................ 98
5 Spin pumping in thin film organic magnetic bilayers .............................................. 101
5.1 Introduction .................................................................................................. 101
5.2 FMR detection of dynamic exchange coupling in magnetic bilayers ............ 103
5.3 Spin pumping in hybrid organic-inorganic magnetic bilayers ....................... 109
5.4 Spin pumping in all-organic magnetic heterostructures ................................ 115
5.5 Conclusion .................................................................................................... 119
References .......................................................................................................... 121
6 Conclusion .................................................................................................................. 123
References ..................................................................................................................... 129

viii

List of Figures
1.1 Types of magnetic ordering .......................................................................................... 5
1.2 Magnon excitation of a 1-dimensional magnet ............................................................. 8
1.3 Spin configuration of a sperimagnet ........................................................................... 10
1.4 Magnetization relaxation diagram .............................................................................. 12
1.5 Spin transfer torque ..................................................................................................... 15
2.1 Flexible organic electronics ........................................................................................ 26
2.2 Physical and magnetic structure of V[TCNE]x~2 ......................................................... 28
2.3 Chemical vapor deposition chamber ........................................................................... 30
2.4 V[TCNE]x~2 film deposited on Teflon tape ................................................................. 31
2.5 Magnetic properties of V[TCNE]x~2 ............................................................................ 33
2.6 Hysteresis loops at high and low temperature ............................................................. 35
2.7 Magnetic mechanisms in V[TCNE]x~2 ........................................................................ 36
2.8 Spin valves incorporating V[TCNE]x~2 layers .......................................................... 38
2.9 Circular polarization of light from V[TCNE]x~2 spin LED ....................................... 40
3.1 Doping profile of a p-n junction .................................................................................. 46
3.2 Ideal IV of a p-n junction ............................................................................................ 48
3.3 Layer structure of inorganic GaAs-AlGaAs diode ...................................................... 49
3.4 Simulated band structure of hybrid heterostructures ................................................... 51
3.5 Picture of air-free can .................................................................................................. 54
3.6 IV characteristics of hybrid organic-inorganic diode .................................................. 56
3.7 IV characteristics of V[TCNE]x~2 layer ...................................................................... 58
4.1 Zeeman effect on energy levels ................................................................................... 64
4.2 Magnetization relaxation diagram .............................................................................. 65
4.3 Schematic of FMR cavity ........................................................................................... 70
4.4 FMR spectrum of V[TCNE]x~2 ................................................................................... 73
4.5 Angular dependence of V[TCNE]x~2 FMR spectrum ................................................. 75
4.6 FMR characteristics of V[TCNE]x~2 with in-plane rotation ....................................... 76
4.7 Comparison of angular dependence of resonance fields ............................................. 77
4.8 Magnetization temperature dependence of V[MeTCEC]x .......................................... 80
4.9 Magnetization field dependence of V[MeTCEC]x ...................................................... 80
4.10 FMR spectrum of V[MeTCEC]x ............................................................................... 81
4.11 Magnetic properties of V[ETCEC]x .......................................................................... 83
4.12 Magnetic resonance properties of V[ETCEC]x ......................................................... 85
4.13 Simulation of field generated by stripline ................................................................. 87
4.14 Schematic of stripline measurement setup ................................................................ 89
ix

4.15 CAD model of magnetic resonance can .................................................................... 90


4.16 FMR spectra of V[TCNE]x~2 at various frequencies ................................................. 91
4.17 Resonance field and linewidth as functions of frequency ......................................... 94
4.18 Temperature dependence of FMR spectrum ............................................................. 96
5.1 Spin pumping schematic ........................................................................................... 105
5.2 FMR characteristics of inorganic bilayer .................................................................. 107
5.3 FMR linewidth change as peaks cross ....................................................................... 108
5.4 FMR peaks of YIG and V[TCNE]x~2 layers in isolation ............................................ 110
5.5 Angular dependence of FMR spectra of layers in isolation ...................................... 111
5.6 FMR characteristics as peaks cross in hybrid bilayer ................................................ 113
5.7 Linewidth of YIG and V[TCNE]x~2 peaks as function of angle ................................. 114
5.8 FMR of all-organic bilayer ........................................................................................ 116
5.9 Peak separation in all-organic bilayer as function of field angle ............................... 117
5.10 Two-Lorentzian simulation ..................................................................................... 118
6.1 IV characteristics of hybrid diode and bare V[TCNE]x~2 ........................................... 125
6.2 FMR of V[TCNE]x~2 layer ........................................................................................ 126
6.3 Spin pumping in YIG-V[TCNE]x~2 bilayer ............................................................... 127

Chapter 1 Introduction

In recent years, magnetism has been a driving force in both academic research and
practical applications [1, 2]. The discovery of the giant magnetoresistance (GMR) effect in
1988 kicked off a revolution in the field of spintronics that produced multi-billion dollar
industries, as well as the 2007 Nobel Prize in Physics [3]. At its core, spintronics involves
using the spin degree of freedom of electrons, analogous to electronics utilizing the charge
degree of freedom of electrons. Traditional electronics face fundamental limits as features
grow smaller, such as resistive heating leading to thermal runaway effects [4] and leakage
current [5], and spintronics potentially offers a way to circumvent these issues. Consumer
products like computer hard drives and magnetic random access memory (MRAM) are
enabled by our understanding of spintronics, and this field promises further novel
applications with potential advantages like nonvolatile memory, low heat transfer, and low
power consumption [6, 7].
Thus far, the most successful spintronics applications have involved all-metal
structures, which is unsurprising given that most ferromagnetic materials are metals. Nonmetallic magnets often have low Curie temperatures or are not air stable; however, they
also offer materials properties that cannot be found in metals. For example, in the field of
microwave electronics, the ferrimagnetic insulator yttrium iron garnet (YIG) is best in class
because its electrically-insulating nature and extremely narrow ferromagnetic resonance
1

(FMR) linewidth combine to give it very low microwave loss [8]. Furthermore, within the
field of spintronics interest in spin dynamics has grown significantly over the past few
years. The key operating principle of magnetic resonance imaging (MRI) machines widely
used in medical fields is magnetic resonance, and there is a strong interest in magnetic
materials for planar microwave devices, such as circulators, isolators, phase shifters, and
patch antennas, with much of the focus on soft ferrite materials [9]. Magnetization
dynamics have been exploited to demonstrate interesting physics such as chaos [10],
soliton production [11], and room-temperature Bose-Einstein condensation [12], and
appear to be an attractive way to generate pure spin currents with no net charge motion
[13-15]. Experiments in inorganic systems have also indicated that FMR linewidth is an
important metric for this type of spin injection [16].
A concurrent trend to the development of magnetoelectronics and spintronics has
been the exploration of organic materials. Existing applications for organic materials
largely focus on reproducing inorganic functionality, such as high conductivity [17] and
light emission or absorption [18, 19], combining those features with the advantages
inherent to organic materials. Organics have long shown low spin-orbit coupling [20],
which potentially means long spin lifetimes, can tolerate a wide variety of substrates, and
are compatible with low temperature and high pressure deposition processes [21-23].
Organic materials systems open new applications that inorganic materials are unsuited for
due to growth requirements like high temperature and lattice-matched substrates to obtain
optimal properties. This has manifested in organic light-emitting diodes [18, 24], thin film
transistors [25], and photovoltaic cells [19]. Our goal is to explore organic-based magnetic

systems, taking advantage of the unique properties of organics to similarly expand the field
of spintronics.
1.1 MagnetismEquation Chapter (Next) Section 1
Magnetism is a fundamentally quantum mechanical phenomenon, as a classical
system in thermal equilibrium cannot have a net magnetic moment, per the Bohr-van
Leeuwen theorem. The magnetic moment of a free atom can originate from three sources:
the inherent spin angular momentum of electrons and other fundamental particles, the
orbital angular momentum of electrons around the nucleus, and the change in the orbital
moment induced by an external magnetic field. In the absence of interactions between
electrons, a material can exhibit two types of magnetic response: diamagnetism, where the
applied magnetic field induces a magnetic moment opposite the direction of the applied
field; or paramagnetism, where the induced magnetic moment aligns with the direction of
the applied field. Formally, the application of a magnetic field adds the following terms to
the Hamiltonian of an atomic electron
H

ie
e2
( A A )
A2
2
2mc
2mc

(1.1)

If the magnetic field is uniform and in the z direction, this simplifies to

ieB
e2 B 2 2
(x y )
(x y2 )
x 8mc 2
2mc y

(1.2)

The first term is proportional to the orbital angular momentum (taking the nucleus as the
origin). First-order perturbation theory shows that the second term contributes a spherically
symmetric contribution

e2 B 2
E
r2
2
12mc

(1.3)

Which gives rise to a diamagnetic magnetic moment


e2 r 2
E

B
B
6mc 2

(1.4)

This agrees with the classical Langevin result for diamagnetism.


Paramagnetism arises from the Zeeman splitting of spin states, the populations of
which are determined by thermal statistics. For an atom with angular momentum J, which
has 2J+1 equally spaced energy levels, the magnetization is given by
M NgJ B BJ ( x), x gJ B B k BT

(1.5)

Where BJ is the Brillouin function for J [26].


BJ ( x)

2J 1
2J 1 1
1
x
coth
coth
2J
2J
2J
2J

(1.6)

At low fields this function is approximately linear, and at high fields the magnetization
saturates as all the moments align; however, for spin-1/2 paramagnets this can take tens of
Tesla.

Figure 1.1 Different types of magnetic ordering in the absence (left) and presence (right) of a magnetic field.
Black arrows represent individual spins, while green dots represent the exchange interaction between spins.
In a paramagnet, the spins do not interact, and in the absence of a magnetic field are randomly aligned. In a
ferromagnet, the exchange interaction tends to align neighboring spins, and so there is a net bulk
magnetization. In antiferromagnets and ferrimagnets, the exchange interaction makes it favorable for
neighboring spins to anti-align. In an antiferromagnet, the spins cancel out and so there is no net
magnetization; in a ferrimagnet, the spins to not cancel, and so there is a net magnetization.

For other types of magnetic effects in bulk materials, an additional interaction


between electrons is necessary. This is the exchange interaction (though it is not an
interaction in the classical sense), and the different types of magnetic ordering it can
generate are shown in Fig. 1-1. The exchange interaction is a direct result of the Pauli
exclusion principle and the requirement that electrons be indistinguishable. These two
principles means that the wavefunction of a two electron system can take two forms:
5

sym (1, 2) anti (1, 2)

(1.7)

anti (1, 2) sym (1, 2)

(1.8)

Or

Where (1,2) is the wavefunction that only involves the spatial coordinates and (1,2) is
the wavefunction that involves spin. Since the wavefunction must be antisymmetric under
exchange of the electrons, one of the spatial wavefunction or the spin wavefunction must
be antisymmetric, but not both. The solutions represent the singlet and triplet spin states
and take the form
sing A[ a (1) b (2) a (2) b (1)][ (1) (2) (2) (1)]
trip

(1) (2)

B[ a (1) b (2) a (2) b (1)] (1) (2) (2) (1)

(1) (2)

(1.9)

Where and are now the spatial and spin wavefunctions of single electrons. If we define
the system as two electrons orbiting two nuclei (for example, a hydrogen molecule) and
allow interactions between the electrons, the Hamiltonian is given by
e2 e2 e2 e2
H12

rab r12 r1b r2 a

(1.10)

Where a, b denote the nuclei and 1, 2 denote the electrons. Then, the energies of the two
states defined in equation (1.9) are

Esing A2 ( K12 J12 )


Etrip B 2 ( K12 J12 )

(1.11)

K12 is the average Coulomb interaction energy, while J12 is the exchange integral, and is
given by
6

J12 a * (1) b* (2) H12 a (2) b (1) d 1d 2

(1.12)

For further detail, see [27].


While the exchange interaction energy is electrostatic in nature, we can treat it as if
it were an interaction between atoms or electrons. For two atoms, i and j, the exchange
Hamiltonian is
H 2 J ij S i S j

(1.13)

Where Jij is the exchange integral, which gives the strength of the interaction, and S is the
total spin of an atom (assuming that Jij is the same for all electrons and the exchange energy
between electrons on the same atom is constant). In the case where it is energetically
favorable for the spins to align parallel to their neighboring spins, Jij is positive, and the
material is ferromagnetic. Ferromagnets have a spontaneous net bulk magnetization, i.e.
they have a net magnetic moment even in the absence of an applied magnetic field.
At zero temperature, all the spins in a ferromagnet will be pointing in the same
direction, as dictated by the exchange interaction. Excitations to the ferromagnetic system
are comprised of magnons, or spin waves. Consider a 1-dimensional chain of spins. The
lowest energy excitation that can be applied is to flip a single spin; however, simply
flipping a spin somewhere in the chain is a large energy perturbation, since in a ferromagnet
the exchange interaction strongly favors the spins being aligned. To alleviate this, rather
than flipping a single spin completely a spin wave cants each spin in the chain by a small
amount, so that over the entire chain a single spins worth of angular momentum has been
subtracted from the total while keeping the overall energy cost as low as possible by
keeping the spins in alignment as much as possible. Magnons are an important tool for
7

understanding the behavior of ferromagnetic materials, and form the basis of many
macroscopic properties. For example, the temperature dependence of the magnetization
can be described using the thermal population of magnons in the material. This yields a
function with the form
3
M (T )
1 T 2
M (0)

(1.14)

Which agrees well with the mean field theory prediction and experimental observations
[27].

Figure 1.2 Magnon excitation of a 1-dimensional chain in side view and top view.

Another important consideration is magnetic anisotropy, or the existence of an


easy axis that the magnetization prefers to align with. Anisotropy can be generated by
several sources, including crystal structure, the shape of the material, magnetostrictive
effects from stress or tension, and exchange effects from neighboring magnetic materials.
For example, crystalline anisotropy is primarily a result of the spin-orbit interaction
between the orbital motion of electrons and the electric field of the crystal lattice [27].
8

Shape anisotropy is generated by the demagnetizing field of the ferromagnet, the stray field
generated by the magnetization (so named because it generally works to reduce the total
magnetic moment). The demagnetizing field is also the source of magnetic domains in
larger ferromagnets. Anisotropy effects change the energy required to excite magnons in
different directions, creating an easy axis that is energetically preferred. These effects can
be treated as internal magnetic fields to the magnet, as opposed to the external fields that
we apply.
The other types of magnetic ordering shown in Fig. 1.1 are less relevant for our
studies, but we will briefly describe them here. If the spins prefer to align antiparallel to
their neighbors, Jij is negative, and the material is antiferromagnetic. In materials where
the overall order is antiferromagnetic but the moments do not balance (for example, in
materials comprised of two different elements), the material retains a spontaneous net
magnetization and is ferrimagnetic. Though the microscopic structure of a ferrimagnet
differs significantly from a ferromagnet, the macroscopic properties of the two, such as the
hysteresis loop and magnetic resonance, are similar. Finally, a sperimagnet is a ferrimagnet
with random magnetic anisotropy, i.e. a system with two sub-networks of spins where one
or both are frozen in random directions [28]. The spin structure for a sperimagnet is shown
in Fig. 1.2. Above a characteristic threshold temperature (the Curie temperature for ferroand ferrimagnets, the Neel temperature for antiferromagnets) the thermal energy destroys
the spontaneous magnetic order. This condition can be expressed as
k BTC J ij Si S j

Where k BTC is the thermal energy at a critical temperature TC.


9

(1.15)

Figure 1.3 Spin configuration of a sperimagnet. Red and blue show the different sub-networks of spins [28].

1.2 Magnetic resonance


One of the ways to induce dynamic motion of a net magnetization is through the
application of an external field comprised of a static component and a dynamic component.
Resonance can be induced in both paramagnetic and ferro-/ferrimagnetic materials, called
electron paramagnetic resonance (EPR) and ferromagnetic resonance (FMR) respectively,
as well as in nuclear spins, in which case it is referred to as nuclear magnetic resonance
(NMR). From a quantum perspective, in the absence of a magnetic field the two spin states
of an electron are degenerate; once a field is applied, the energy of the spin +1/2 and -1/2
states is split via the Zeeman effect, creating an energy gap as described in equation (1.6).
The electron can transition between these two states through the radiation or absorption of
a photon with energy h, leading to the resonance condition
h g e B H 0

(1.16)

While this theoretically allows for resonance at any given combination of static
field/radiation, in general most EPR measurements are carried out with applied radiation
in the microwave range, with frequencies between 9 and 10 GHz corresponding to fields
10

around 3500 G. This is due to a combination of factors, including the wide availability of
microwave sources in the X band region (due to use in radars), the wavelength allowing
for a fairly large cavity size, and a field that is low enough to be easily generated. The
linewidth of an EPR line is usually quite sharp, on the order of 1 G.
At a basic level, ferromagnetic resonance follows the same principles laid out
above, but this picture is complicated by the presence of strong exchange forces. In this
case, we can describe magnetic resonance from a semi-classical perspective by treating the
magnetization as a single vector and considering the various torques on it. Formally, the
resonance condition is the same as equation (1.16), but now g can differ from 2 due to spinorbit interactions and Heff is the magnitude of the effective magnetic field, combining the
external static field with internal demagnetizing and anisotropy fields (as described above):
h g B H eff

(1.17)

The equation governing the time evolution of the magnetization is the Landau-LifshitzGilbert equation
dM

dM
(M H eff ) (M
)
dt
M
dt

(1.18)

where M is the magnetization vector; M is the magnitude of the magnetization, or the


saturation magnetization; Heff is the total effective magnetic field experienced by the
magnetization; is the gyromagnetic ratio; and is a phenomonological damping
parameter, also called the Gilbert damping parameter. Figure 1.4 shows the nature of each
term in equation (1.15), with the magnetization vector in red and the static field in black.
The first term in the equation describes the torque the effective field applies on the
magnetization, shown in green, causing M to precess about the effective field. The second
11

term describes a damping torque that restores the magnetization to alignment with the
effective field, shown in blue. Overall, the magnetization will relax to equilibrium along a
spiral path, shown as the blue dashed line. The magnetization is in equilibrium when
MHeff, or equivalently dM/dt = 0.

Figure 1.4 Magnetization dynamics as governed by the Landau-Lifshitz-Gilbert equation. At equilibrium, the
magnetization (red) is aligned with the static field (black). Upon excitation, the magnetization is subject to
two different torques (blue and green arrows). This leads to the magnetization relaxing to equilibrium by
following the path along the blue dashed line.

For resonance, the oscillating field applies an instantaneous torque to the


magnetization that drags it out of equilibrium. In accordance with the Landau-LifshitzGilbert equation above, the magnetization then relaxes back to equilibrium by following a
spiral path. The frequency at which the magnetization precesses is given by equation
12

(1.17). When the resonance condition is satisfied, the oscillating field applies another
torque just as the magnetization completes one precession, keeping the magnetization in
motion. This can be measured by applying a fixed microwave field and sweeping the DC
field (or equivalently, using a fixed DC field and sweeping the microwave frequency) and
monitoring the absorbed microwave power, which is expected to follow a Lorentzian:
L( x)

2
( x x0 ) 2
2

(1.16)
4

Where x0 is the resonance field and is the full width at half maximum linewidth.
The linewidth of a FMR spectrum can be narrowed or broadened by additional effects,
such as two-magnon scattering, sample mosaicity, and inhomogenous broadening [29, 30].
The way these two parameters change as functions of applied frequency, field direction,
and temperature can yield information about the magnetic structure and anisotropies in the
material. Broadly speaking, defects and impurities create inhomogeneous broadening,
while intrinsic effects create homogeneous broadening. Intrinsic relaxation is
phenomenologically described by the Gilbert damping parameter in eq. 1.18. If Gilbert
damping is the dominant mechanism, the linewidth should depend linearly on the
frequency. Mosaic effects describe the spread of physical parameters across the sample,
seen in variation in internal fields, thickness, crystal orientation, or other anisotropies. This
will lead to individual regions having slightly different resonance fields, and therefore the
overall signal is the superposition of the local FMR peaks. For frequency dependent
measurements, this contribution goes to zero. Two-magnon scattering describes a process
by which the uniform magnon mode excited in FMR ( k 0 ) scatters into degenerate states
13

of magnons with wave vectors k 0 . Finally, inhomogenous broadening describes


frequency and angle independent broadening which cannot be written in other forms
(Gilbert damping or two-magnon scattering terms).

1.3 Ferromagnetic resonance-driven spin pumping


A key concept to the field of spintronics is pure spin current. Electrical current is
the physical motion of charge, driven by a potential difference, resulting in the net
movement of some number of electrons from one point to another. Spin current is a related
concept whereupon a net spin moves from one point to another. A pure spin current is a
spin current that has no associated charge current. In the case of a conduction electron pure
spin current, for every electron that moves from point A to point B, another electron moves
from point B to point A. In the simplest case, all of the AB electrons are spin up and all
the BA electrons are spin down (or, more likely, the AB electrons are
disproportionately spin up). Another possibility is a spin wave, or magnon, spin current.
Spin currents carry angular momentum, and can apply torques to a magnetization,
a phenomenon called spin-transfer torque [31]. The itinerant spin-polarized electrons
interact with the magnetization upon reflection or transmission through the ferromagnet
and are reoriented to align with the magnetization, as shown in Fig. 1.5. By conservation
of angular momentum, this exerts an equal and opposite torque on the magnetization itself,
which is shown as a red arrow. This has been observed in a number of geometries, including
magnetic multilayers [32], nanopillars [33], nanowires [34], and magnetic tunnel junctions
[35], and with sufficiently high current densities can induce switching of the magnetization
and domain wall motion [36, 37].
14

Figure 1.5 Spin transfer torque process. An incident electron with some spin (purple arrow) interacts with the
bulk magnetization of a ferromagnet (blue arrow). The electron is reoriented to align with the magnetization,
a process which also applies a torque to the bulk magnetization (red arrow) [31].

Recently, FMR-driven spin pumping has emerged as an attractive method of


generating a pure spin current. This mechanism is the inverse of the spin current induced
magnetization switching described above spin pumping is a process by which a moving
magnetization vector emits a pure spin current [31, 38, 39]. As described in section 1.2,
when a ferromagnet is in resonance its magnetization is constantly in motion, precessing
about an external static magnetic field, and therefore continuously emitting a spin current.
This spin current can be injected into an adjacent normal metal, and if the metal exhibits
strong spin-flip scattering, this leads to enhanced damping in the ferromagnet.
Furthermore, if the adjacent material is another ferromagnet, this can generate an additional
torque on the magnetization absorbing the spin current, changing the dynamics of both
magnetic layers. These processes establish interesting interplay between charge, itinerant
15

spin, and bulk magnetization that can be exploited for various applications. For example,
it has been demonstrated that an electrical signal can be converted into a pure spin signal
via the spin Hall effect, transported through an insulating magnet as a spin wave, then
converted back to a voltage via the inverse spin Hall effect [40]. Spin pumping makes
ferromagnets into spin batteries, sources of pure spin current that can be used for
spintronics devices with no charge motion [41], with the potential the exploration of new
and exciting physics.

1.4 Organic materials and magnets


Organic materials are carbon-based small molecules or polymers, and the field of
organic electronics involves the study, design, and fabrication of organic materials with
desirable properties, primarily electric conductivity. Organic materials have made a large
impact in recent years, with devices like organic LEDs (OLED)[18], field effect and thin
film transistors (OFET, OTFT)[25, 42], and photovoltaic cells (OPV)[19] crossing over
from fundamental research to the realm of consumer electronics. In particular, OLEDs have
become a major component of display technology in mobile phones and other electronic
devices. In general, organics are lighter, less power hungry, less expensive, and more
compatible with flexible substrates than inorganic conductors, like metals; however, even
organic conductors have high resistances, and therefore are inefficient carriers of electric
current, when compared with the best inorganic materials, like copper. Therefore, finding
the appropriate niches for organic devices involves maximizing the utility of their favorable
qualities and minimizing their disadvantageous qualities. The goal is not to supplant
inorganic materials; rather, the goal is to expand the universe of possible applications. This
16

can be seen in the push for devices built on flexible substrates - this is an application for
which inorganic materials cannot be used, so organic materials are enabling a new class of
electronic devices.
In a similar vein, organic-based magnets could make a major impact in the fields
of spintronics and magneto-electronics. These materials bring the advantages of organics
in general, combined with magnetic functionality that is rare in organics. Since organic
materials mostly consist of low Z elements, they have relatively low spin orbit coupling,
potentially leading to long spin lifetimes [20]. Organic deposition methods involve
deposition conditions that are very benign compared to most inorganic materials, with
substrate temperatures close to room temperature and pressures close to ambient conditions
[43]. Organic-based magnets are also amenable to organic chemistry techniques that can
modify the material in interesting ways [44-46]. Finally, organic-based magnets retain their
most desirable properties almost independent of the substrate on which they are grown
[22]. In comparison, for optimal properties most inorganic magnets must be grown as
single crystals to minimize disorder, which in turn means depositing on specific latticematched substrates [47]. In inorganic magnets, structural disorder leads to the formation of
magnetic grains within the material as a whole.
Despite these advantages, magnetism rarely occurs in organic materials, and
existing organic-based magnets can be difficult to work with. Organic-based magnets are
sensitive to air exposure due to the relative strengths of the metal-organic and metal-oxygen
bonds, and little work has been done on integrating such materials into heterostructures.
Up to this point, most organic spintronics studies have focused on using an organic material
as a passive spin transport layer. Our goal is to expand the field, using an organic-based
17

magnet as an active material for spintronics and opening up new measurements and
applications for this class of materials.

1.5 Spin dynamics in novel materials systems - focus of my research


Spintronics and organic electronics are fields that have made considerable advances
in recent years, both in fundamental research and in applications. Organic materials have a
number of attractive properties that enable them to complement applications traditionally
fulfilled by inorganic materials, while spintronics seeks to take advantage of the spin
degree of freedom to produce new applications. My research is aimed at combining these
two fields to develop organic materials for spintronics use.
My thesis is divided into three primary projects centered around an organic-based
semiconducting ferrimagnet, vanadium tetracyanoethylene. First, we investigated the
transport characteristics of a hybrid organic-inorganic heterostructure. Semiconductors
form the basis of the electronics industry, and there has been considerable effort put
forward to develop organic semiconductors for applications like organic light-emitting
diodes and organic thin film transistors. Working with hybrid organic-inorganic
semiconductor device structures allows us to potentially take advantage of the
infrastructure that has already been developed for silicon and other inorganic
semiconductors. This could potentially pave the way for a new class of active hybrid
devices with multifunctional behavior.
Second, we investigated the magnetic resonance characteristics of V[TCNE]x, in
multiple measurement schemes and exploring the effect of temperature, frequency, and
chemical tuning. Recently, the spintronics community has shifted focus from static
18

electrical spin injection to various dynamic processes, such as spin pumping and thermal
effects. Spin pumping in particular is an intriguing way to generate pure spin currents via
magnetic resonance that has attracted a high degree of interest, with the FMR linewidth
being an important metric for spin injection. Furthermore, we can potentially use these
measurements to probe the magnetic properties as we change the physical properties of the
materials by chemically tuning the organic ligand. We are therefore interested in exploring
the resonance properties of this materials system to lay the groundwork for future spin
pumping applications.
Third, we have made preliminary measurements of spin pumping in hybrid and allorganic bilayer structures. As mentioned above, FMR-driven spin pumping is method for
generating pure spin currents with no associated charge motion. This can be detected in a
number of ways, one of which is monitoring the FMR characteristics of two ferromagnets
in close contact, where spins injected from one magnet into the other changes the linewidth.
In conjunction with the magnetic resonance measurements, we have started to investigate
the FMR properties of these bilayer systems.
These projects will be described in detail in chapters 3, 4, and 5 respectively. The
results of my research have established that V[TCNE]x~2 has a number of attractive
magnetic properties, especially with regards to magnetic resonance, that point toward
several promising avenues for fundamental studies and for potential applications.

19

References
1.

Wolf, S.A., et al., Spintronics: A spin-based electronics vision for the future.
Science, 2001. 294(5546): p. 1488-1495.

2.

Zutic, I., J. Fabian, and S. Das Sarma, Spintronics: Fundamentals and applications.
Reviews of Modern Physics, 2004. 76(2): p. 323-410.

3.

The Nobel Prize in Physics 2007.

4.

Nanver, L.K., et al., A back-wafer contacted silicon-on-glass integrated bipolar


process - Part I: The conflict electrical versus thermal isolation. Ieee Transactions
on Electron Devices, 2004. 51(1): p. 42-50.

5.

ITRS, International Technology Roadmap for Semiconductors. 2013.

6.

Wolf, S.A., et al., The Promise of Nanomagnetics and Spintronics for Future Logic
and Universal Memory. Proceedings of the Ieee, 2010. 98(12): p. 2155-2168.

7.

Wolf, S.A., A.Y. Chtchelkanova, and D.M. Treger, Spintronics - A retrospective


and perspective. Ibm Journal of Research and Development, 2006. 50(1): p. 101110.

8.

Chang, H.C., et al., Nanometer-Thick Yttrium Iron Garnet Films With Extremely
Low Damping. Ieee Magnetics Letters, 2014. 5.

9.

Pardavi-Horvath, M., Microwave applications of soft ferrites. Journal of


Magnetism and Magnetic Materials, 2000. 215: p. 171-183.

10.

Wigen, P.E., ed. Nonlinear Phenomena and Chaos in Magnetic Materials. 1994,
World Scientific.

11.

Mohseni, S.M., et al., Spin Torque-Generated Magnetic Droplet Solitons. Science,


2013. 339(6125): p. 1295-1298.

20

12.

Demokritov, S.O., et al., Bose-Einstein condensation of quasi-equilibrium magnons


at room temperature under pumping. Nature, 2006. 443(7110): p. 430-433.

13.

Ando, K., et al., Angular dependence of inverse spin-Hall effect induced by spin
pumping investigated in a Ni(81)Fe(19)/Pt thin film. Physical Review B, 2008.
78(1).

14.

Inoue, H.Y., et al., Detection of pure inverse spin-Hall effect induced by spin
pumping at various excitation. Journal of Applied Physics, 2007. 102(8).

15.

Saitoh, E., et al., Conversion of spin current into charge current at room
temperature: Inverse spin-Hall effect. Applied Physics Letters, 2006. 88(18).

16.

Wang, H.L., et al., Scaling of Spin Hall Angle in 3d, 4d, and 5d Metals from
Y3Fe5O12/Metal Spin Pumping. Physical Review Letters, 2014. 112(19).

17.

Shirakawa, H., et al., SYNTHESIS OF ELECTRICALLY CONDUCTING


ORGANIC POLYMERS - HALOGEN DERIVATIVES OF POLYACETYLENE,
(CH)X. Journal of the Chemical Society-Chemical Communications, 1977(16): p.
578-580.

18.

Tang, C.W. and S.A. Vanslyke, ORGANIC ELECTROLUMINESCENT DIODES.


Applied Physics Letters, 1987. 51(12): p. 913-915.

19.

Brabec, C.J., Organic photovoltaics: technology and market. Solar Energy


Materials and Solar Cells, 2004. 83(2-3): p. 273-292.

20.

Yu, Z.G., Spin-Orbit Coupling, Spin Relaxation, and Spin Diffusion in Organic
Solids. Physical Review Letters, 2011. 106(10).

21.

Manriquez, J.M., et al., A ROOM-TEMPERATURE MOLECULAR ORGANIC


BASED MAGNET. Science, 1991. 252(5011): p. 1415-1417.

22.

Pokhodnya, K.I., A.J. Epstein, and J.S. Miller, Thin-film V TCNE (x) magnets.
Advanced Materials, 2000. 12(6): p. 410-+.

21

23.

de Caro, D., et al., CVD-grown thin films of molecule-based magnets. Chemistry of


Materials, 2000. 12(3): p. 587-+.

24.

Baldo, M.A., et al., Highly efficient phosphorescent emission from organic


electroluminescent devices. Nature, 1998. 395(6698): p. 151-154.

25.

Forrest, S.R., The path to ubiquitous and low-cost organic electronic appliances on
plastic. Nature, 2004. 428(6986): p. 911-918.

26.

Kittel, C., Introduction to Solid State Physics. 8 ed. 2005.

27.

Morrish, A.H., The Physical Principles of Magnetism. 1980: IEEE Press.

28.

Cimpoesu, F., et al., Disorder, exchange and magnetic anisotropy in the roomtemperature molecular magnet V TCNE (x) - A theoretical study. Computational
Materials Science, 2014. 91: p. 320-328.

29.

Zakeri, K., et al., Spin dynamics in ferromagnets: Gilbert damping and two-magnon
scattering. Physical Review B, 2007. 76(10).

30.

Heinrich, B., et al., Dynamic exchange coupling in magnetic bilayers. Physical


Review Letters, 2003. 90(18).

31.

Brataas, A., A.D. Kent, and H. Ohno, Current-induced torques in magnetic


materials. Nature Materials, 2012. 11(5): p. 372-381.

32.

Myers, E.B., et al., Current-induced switching of domains in magnetic multilayer


devices. Science, 1999. 285(5429): p. 867-870.

33.

Mangin, S., et al., Current-induced magnetization reversal in nanopillars with


perpendicular anisotropy. Nature Materials, 2006. 5(3): p. 210-215.

34.

Fukami, S., et al., Current-induced domain wall motion in perpendicularly


magnetized CoFeB nanowire. Applied Physics Letters, 2011. 98(8).

22

35.

Kubota, H., et al., Quantitative measurement of voltage dependence of spin-transfer


torque in MgO-based magnetic tunnel junctions. Nature Physics, 2008. 4(1): p. 3741.

36.

Yamanouchi, M., et al., Current-induced domain-wall switching in a ferromagnetic


semiconductor structure. Nature, 2004. 428(6982): p. 539-542.

37.

Liu, L., et al., Spin-Torque Switching with the Giant Spin Hall Effect of Tantalum.
Science, 2012. 336(6081): p. 555-558.

38.

Tserkovnyak, Y., A. Brataas, and G.E.W. Bauer, Spin pumping and magnetization
dynamics in metallic multilayers. Physical Review B, 2002. 66(22).

39.

Tserkovnyak, Y., A. Brataas, and G.E.W. Bauer, Enhanced Gilbert damping in thin
ferromagnetic films. Physical Review Letters, 2002. 88(11).

40.

Kajiwara, Y., et al., Transmission of electrical signals by spin-wave


interconversion in a magnetic insulator. Nature, 2010. 464(7286): p. 262-U141.

41.

Brataas, A., et al., Spin battery operated by ferromagnetic resonance. Physical


Review B, 2002. 66(6).

42.

Crone, B., et al., Large-scale complementary integrated circuits based on organic


transistors. Nature, 2000. 403(6769): p. 521-523.

43.

Yu, H., et al., Ultra-narrow ferromagnetic resonance in organic-based thin films


grown via low temperature chemical vapor deposition. Applied Physics Letters,
2014. 105(1).

44.

Pokhodnya, K.I., B. Lefler, and J.S. Miller, A methyl tricyanoethylenecarboxylatebased room-temperature magnet. Advanced Materials, 2007. 19(20): p. 3281-+.

45.

Kao, C.Y., et al., Thin films of organic-based magnetic materials of vanadium and
cobalt tetracyanoethylene by molecular layer deposition. Journal of Materials
Chemistry C, 2014. 2(30): p. 6171-6176.

23

46.

Lu, Y., et al., Thin-Film Deposition of an Organic Magnet Based on Vanadium


Methyl Tricyanoethylenecarboxylate. Advanced Materials, 2014. 26(45): p. 76327636.

47.

Sun, Y.Y., Y.Y. Song, and M.Z. Wu, Growth and ferromagnetic resonance of
yttrium iron garnet thin films on metals. Applied Physics Letters, 2012. 101(8).

24

Chapter 2 Background and theory

2.1 Introduction
Organic materials are carbon-rich compounds with molecular and physical
structures that can be tailored to the desired functionality, whether that is charge mobility
[48] or luminescence [18, 24] or other properties. There has been widespread interest in
organic materials for electronics applications due to their ability to be deposited on a
variety of inexpensive substrates, including flexible materials, and the relative ease of
processing through chemical tools [21, 23, 44]. Several organic systems have already made
considerable inroads in consumer electronics by combining properties found in inorganic
systems with the advantages of organic materials. These include organic light-emitting
diodes (OLEDs) [18], organic photovoltaic cells (OPVs) [19], and organic field effect and
thin film transistors (FETs, TFTs) [25, 42] - OLEDs alone have a large and growing
presence in cell phone screens and televisions, and are entering the lighting industry. Figure
2.1 shows two such applications, a flexible cell phone display utilizing OLEDs [49] and a
flexible photovoltaic cell [50]. It is important to note that in most cases these organic
materials are not directly replacing inorganic materials in high performance electronics;
rather, the variability and low cost deposition of organic materials opens the door to
applications that would otherwise not exist.

25

In this light, organic magnetic materials could make a similar impact in the arenas
of spintronics and high frequency electronics where inorganic materials have been
dominant, but share similar difficulties in growing well-ordered, crystalline materials on
different substrates that optimize magnetic performance. Broadly, there are two barriers to
the development of this industry: (1) the demonstration of the desired spintronic or high
frequency behavior and related effects; and (2) bringing the cost of fabricating such devices
to a point of economic competitiveness. We are primarily concerned with the first
challenge, demonstrating desired functionality. Organic materials have been used as a
passive layer in various spintronics studies - demonstrated device geometries include the
spin transport layer of vertical and lateral spin valves, magnetic tunnel junctions, and the
spin pumping structures. However, we are interested in using an organic material not only
as a passive spin transport layer, but as an active spin injecting layer. For these purposes,
we are studying a family of organic-based magnets, vanadium tetracyanoethylene and
related materials.

Figure 2.1 (a) Flexible cell phone display powered by organic light-emitting diodes [49]. (b) Flexible
photovoltaic cell [50].
26

2.2 Vanadium tetracyanoethylene and related materials


Vanadium tetracyanoethylene (V[TCNE]x~2) is an organic-based magnet that can
be prepared in a solvent-based solution [21] or as a thin film [22]. There are several ways
to generate the thin film, including chemical vapor deposition (CVD), physical vapor
deposition (PVD), and molecular layer deposition (MLD) [51, 52]. In both the solution and
the thin film forms, the Curie temperature has been above room temperature, with thin film
samples reaching Curie temperatures of over 600 K [43]. Transport measurements of thin
films have found that it is a semiconductor with a band gap around 0.5 eV [21]. Physically,
V[TCNE]x~2 is a disordered network of vanadium ions and tetracyanoethylene (TCNE)
molecules, as shown in Fig. 2.2(a). While there is no long range crystalline structure, there
is strong local ordering that arises from a vanadium ion coordinating with up to six TCNE
molecules [53]. This structure means that there is considerable flexibility in constructing
related materials - the magnetic ion can be changed, which has been demonstrated with
cobalt, manganese, and iron [54, 55]; or the organic ligand can be changed, as demonstrated
with various functionalizations of the TCNE molecule, like methyl and ethyl
tricyanoethylene carboxylate (MeTCEC and ETCEC respectively) [46].
Figure 2.2(b) shows the electronic structure of V[TCNE]x~2. The magnetic order in
V[TCNE]x~2 arises from direct antiferromagnetic exchange coupling between the unpaired
spins in the t2g orbital of V2+ (S = 3/2) and the unpaired spins in the * orbitals of the TCNE
anions- (S = 1/2). Polarized neutron diffraction studies confirm the TCNE- has one unpaired
electron that occupies the * antibonding molecular orbital, and determine the spin on the
27

TCNE- is distributed over the entire radical anion. There is a splitting of the * band into
two subbands due to Coulomb repulsion (UC ~ 2 eV), with the occupied * and unoccupied
*+UC subbands exhibiting antiparallel spin polarization. The t2g levels lie within the
Coulomb gap and define the valence band, while the *+UC levels define the conduction
band with a 0.5 eV bandgap. This electronic structure has ferromagnetically aligned
conduction and valence bands, and is consistent with both theoretical calculations [56-58]
and experimental studies [59-61].

Figure 2.2 (a) Physical structure of V[TCNE]x~2 films. Vanadium ions and their spins are in red, while TCNE
molecules and their spins are in blue. There is no long range crystalline order, but there is strong local order
arising from vanadium ions coordinating with the TCNE molecules. (b) Electronic structure of V[TCNE]x,
with vanadium atomic energy levels in red and TCNE molecular energy levels in blue. Magnetism arises
from antiferromagnetic coupling between the spins on the vanadium and the neighboring TCNE molecules
with a superexchange pathway to neighboring vanadium ions. Electrons can be thermally activated to the
conduction band from the vanadium levels with a band gap of ~0.5 eV.

28

2.3 Chemical vapor deposition process


The V[TCNE]x~2 films reported here are grown by a low temperature (60 C) CVD
process in a custom built reactor shown in Fig. 2.3. The films are formed through the
reaction of precursors TCNE and V(CO)6, which are carried in from opposite sides of the
reactor by argon gas. Commercially available TCNE is purified through sublimation and
V(CO)6 is synthesized from tetraethylammonium vanadium hexacarbonyl and phosphoric
acid. The heating coil creates a temperature gradient determined by the density of the coil
windings. The precursors combine in the reaction zone and deposit on the substrates. Initial
reports of V[TCNE]x~2 grown via CVD yielded the first thin film organic magnetic material
with a TC above room temperature (TC~400 K). It was observed that the CVD grown films
show sharper magnetization switching compared to previous solvent-based powders,
which is attributed to the absence of the noncoordinating CH2Cl2 solvent required for
powder synthesis. Previous studies have also shown that the sublimation rate and ratio of
the V[TCNE]x~2 precursors affect the coercivity and sharpness of the switching for M(H)
measurements. We have extended these studies by further exploring the impact of the
relevant conditions including: the temperature of the precursors, the ratio of the precursors
and the impact of a clean solvent-free substrate.

29

Figure 2.3 Chemical vapor deposition chamber for V[TCNE]x. Precursors are flowed in from either side by
argon carrier gas, react in the green reaction zone, and deposit on the substrates.

For ideal films, (1) the TCNE precursor temperature is kept at 50 C while the
V(CO)6 is kept at 10 C; (2) the ratio of TCNE to V(CO)6 is set at 10:1 by weight; and (3)
the film is deposited on a clean, solvent-free substrate. Depositions that deviated from these
conditions generated films with magnetic properties in line with previous reports. Over 70
depositions we consistently observe that these growth conditions generate more
magnetically homogenous samples, exhibiting behaviors such as sharper switching in
hysteresis loops and higher extrapolated TC, with extremely sharp ferromagnetic resonance
features. Due to the air sensitive nature of the films, the CVD process is performed inside
a series of argon glove boxes (O2, H2O < 0.5 ppm). The resulting V[TCNE]x~2 film is black
and opaque, and is extremely air sensitive if left in air, the film will disappear within a
30

few hours, likely due to oxidation. Figure 2.4 shows a V[TCNE]x~2 film deposited on a
piece of Teflon tape, with a magnet held to it to show that it is still magnetic on the flexible
substrate. For further details of the CVD process, see [62].

Figure 2.4 V[TCNE]x~2 deposited on a piece of teflon tape and still exhibiting a magnetic response.

2.4 Magnetic properties of V[TCNE]x


One of the more surprising aspects of the magnetic ordering in V[TCNE]x~2 is that
it is quite robust, even in the original powder synthesis, despite a significant degree of
structural disorder. Specifically, extended x-ray absorption fine-structure (EXAFS) studies
show that the vanadium atoms are locally coordinated with 6 nitrogen atoms in a distorted
octahedral environment with a vanadium-nitrogen spacing of 2.084(5) [53]. In contrast,
x-ray diffraction and transmission electron microscopy (TEM) show no signatures of longrange structural order [63]. While unusual, the presence of strong long-range magnetic
order without corresponding long-range structural order makes V[TCNE]x~2 particularly
31

amenable to deposition on a variety of both crystalline and amorphous substrates. The


antiferromagnetic exchange interaction between the t2g electrons on the V2+ and the *
electrons on the [TCNE]- molecules results in a net ferrimagnetic ordering with the
[TCNE]- acting as a superexchange pathway for adjacent V2+ ions (Fig. 2.2(b) green
arrow).
SQUID

(superconducting

quantum

interference

device)

magnetometry

measurements of V[TCNE]x~2 thin films have revealed several interesting properties. First,
M vs H measurements over a wide range of temperatures do not exhibit any significant
change in coercivity - at all temperatures, the hysteresis loop has a coercive field around
20 Oe. This number does not change with different field orientations (in-plane versus outof-plane), which is not very surprising given the lack of crystalline structure in the material.
The magnetization has an initial sharp transition followed by a more gradual,
paramagnetic-like asymptotic behavior until it saturates around 1 T, as shown in Fig. 2.5(a).
Various developments in thin film deposition techniques have increased the Curie
temperature extracted from M vs T measurements, but qualitatively the M vs T curve is
quite similar across all the deposition methods. The magnetization is nonzero at room
temperature and peaks at an intermediate temperature, usually in the 100-150 K region,
before dropping off as the temperature decreases.
The Curie temperature of V[TCNE]x~2 cannot be measured directly, as the film
decomposes at temperatures higher than ~330 K. However, TC can be extracted from the
temperature dependence of the magnetization through a fit to the Bloch Law:

32

M s (T ) M s (0)(1 T 2 )

(2.1)

where the saturation magnetization is given by Ms(T) and is a fitting parameter. Figure
2.5(b) shows the reported TC at various stages of development, starting with the powderbased preparation and continuing with several different thin film techniques. Our most
recent published result shows a TC of 606 K, which is 100 K higher than the previous record
high [43]. This number should be taken as an estimate - the Bloch Law monotonically
decreases, and so we can only fit the data at temperatures higher than the peak (though this
is consistent with the methods used in previous reports). Furthermore, the magnetization
does not exhibit much variation over this temperature range, making the fit less accurate.
However, we do note that the TC we extract is lower than what we would get from a linear
extrapolation to zero magnetization (632 K).

Figure 2.5 (a) Magnetization versus temperature for a V[TCNE]x~2 thin film with 100 Oe applied field. Solid
line shows a Bloch Law fit with extracted TC of 606 K. Inset: Magnetization versus field at 300 K. (b)
Evolution of extracted Curie temperature over time. Open circle is the initial powder synthesis[21], open
squares are thin film reports[22, 64, 65], and the solid star is the highest Curie temperature we have measured
in our optimized films[43].
33

The peak in the magnetization has been linked to a magnetic phase transition in the
material. Initial reports indicated that the material behaves as a spin glass below the
magnetization peak. In the spin glass state, the exchange interaction between the spins is
frustrated and the spins are frozen in place. This is evidenced by the peak in the
magnetization as a function of temperature as well as a splitting between the field-cooled
and zero-field-cooled magnetization as the temperature decreases. However, other
behaviors do not match up with canonical spin glass properties. A true spin glass has no
overall magnetic ordering, and upon removal of an external field the magnetization will
decay quickly to the remanent value and then decay slowly with no characteristic time scale
to zero (hence the analogy to glass). We should be able to observe this in the frequency
dependence of an AC susceptibility measurement and in the M vs H hysteresis loop, but
these do not show the expected spin glass behaviors in V[TCNE]x~2 [66]. It has been
proposed that the low temperature behavior of V[TCNE]x~2 is best described by a relatively
exotic magnetic phase called sperimagnetism [28].

34

Figure 2.6 Magnetization versus field at 300 K (left) and 5 K (right). Contrary to expectations, the coercivity
appears higher and the switching appears sharper at high temperature.

Figure 2.6 shows the M vs H characteristics of V[TCNE]x~2 at 300 K and 5 K.


Generally we expect magnetic properties to improve as temperature decreases, as thermal
energy normally competes with the exchange interaction and destroys magnetic ordering.
However, in V[TCNE]x, we observe that magnetic properties such as the coercivity and the
sharpness of the switching appear to be better at high temperature. We attribute this to the
existence of two exchange pathways in V[TCNE]x, illustrated in Fig. 2.7 the first is local
antiferromagnetic exchange, which creates overall ferromagnetic ordering with the TCNE
molecules acting as a superexchange pathway for adjacent vanadium ions, while the second
is a long range carrier-mediated mechanism. The inset to Fig. 2.7 shows the local exchange,
with an island of well-ordered V[TCNE]x~2 having a net magnetization illustrated by the
black arrow. The long range magnetic interaction is maintained by electrons hopping
between these islands, with the extent of electron motion represented as the blue clouds.
35

The electrical resistivity of V[TCNE]x~2 increases exponentially as temperature decreases,


and so at low temperature, there is little to no electron motion throughout the film.
Therefore the islands of magnetization are isolated and the overall film behaves as a
magnetic material with a large number of domains, resulting in the more rounded hysteresis
loop. At high temperature, the conduction electrons are mobile enough to keep the islands
aligned, and so the film behaves more like a single domain magnet, with sharper switching.

Figure 2.7 Diagram of magnetism mechanisms in V[TCNE]x. Local structural order and exchange
interactions between spins (red and blue arrows) create islands with a net magnetization, represented by the
black arrows. Blue clouds represent the range of electron motion. At low temperature (left), electrons are not
mobile, which is also observed in the electrical resistivity sharply increasing, meaning each of the islands of
magnetization are independent. At high temperature (right), electrons are very mobile, which allows the
smaller islands to interact and increasing the strength of the magnetic order.

36

2.5 V[TCNE]x~2 spintronics to date


V[TCNE]x~2 has been used in a number of spintronics studies as the active spin
injection layer, in spin valves and spin LEDs. Spin valves are the first spintronics device,
forming the basis for most types of magnetic memory. The device depends on an effect
called giant magnetoresistance where the resistance of a ferromagnetic heterostructure
consisting of two distinct ferromagnetic layers depends on the relative orientations of the
magnetizations. This can be understood by thinking of the carriers in the material as two
populations, those with spin aligned with the magnetization (majority carriers) and those
with spin anti-aligned (minority carriers), each with their own density of states. When the
magnetizations are aligned, the majority carriers in the source electrode are also the
majority carriers in the drain electrode, and therefore have a large number of states
available to them for transport through the layer. Thus, the resistance is low. When the
magnetizations are anti-aligned, the majority carriers in the source are the minority carriers
in the drain, and therefore have a low number of states available to them. Conversely, the
minority carriers in the source are the majority in the drain and have many available states,
but there are a low number of carriers available to fill those states. Therefore, the resistance
is high.
Spin valves can be built in two different geometries. In a vertical structure, the
layers are deposited on top of each other, and so the independent control of the
magnetizations can be controlled by using two different magnetic materials, two different
thicknesses of the same ferromagnet, or by pinning a ferromagnet via an adjacent
antiferromagnetic layer. In a lateral structure, magnetic electrodes are deposited next to one
another on a conductive channel. In this case, the magnetic electrodes are distinguished by
37

their shape and size. The signature of the spin valve is in its magnetoresistance - by
sweeping the field, switching can be observed from low resistance to high resistance and
back again.

Figure 2.8 Magnetoresistance curves from hybrid (left) and all-organic (right) spin valves[59, 67]. Hybrid
spin valve sample structure is V[TCNE]x~2/rubrene/LAO/LSMO. All-organic spin valve sample structure is
V[TCNE]x~2/rubrene/V[TCNE]x~2.

V[TCNE]x~2 has been successfully incorporated into vertical spin valves. This was
first achieved using V[TCNE]x~2 as one of the electrodes and LSMO as the other in a hybrid
organic-inorganic heterostructure, as shown in Fig. 2.8 [59]. This result was extended to
an all-organic spin valve, using two different deposition methods for the V[TCNE]x~2
electrodes (chemical vapor deposition and molecular layer deposition) [67]. These results
were the first demonstration of using an organic magnet as the active magnetic layer in a
spintronics device. However, the spin valve is not a definitive demonstration of spin
injection from the magnetic layer - the signature spin valve signal can be reproduced by a
38

number of artifacts, and has even been observed in systems with only one magnetic layer
[68]. In inorganic systems, the lateral spin valve geometry can also be used for a
measurement of the Hanle effect by changing the direction of the magnetic field, which
leaves no doubt. Unfortunately, the Hanle effect has not yet been observed in organic
materials, and due to the different transport mechanisms in organics it is unclear if it can
be observed. Therefore, different methods must be used to demonstrate spin injection from
the V[TCNE]x~2 layer.
An alternate technique to detect spin injection is to use a spin light emitting diode
(spin LED). In this measurement geometry, a ferromagnet is deposited on top of a
semiconducting LED structure and a current is applied to the heterostructure. When the
current passes through the ferromagnet, it becomes spin polarized depending on the
orientation of the magnetization. The electrons then enter the LED, where they recombine
and emit light. Based on the spin orientation of the recombining electron, the emitted
photon will either be right or left circularly polarized. This is because the recombination
process must obey strict optical selection rules that determine the polarity of the emitted
photon. Since the current has been spin polarized by the ferromagnetic layer, the light
emitted by the diode will have a net circular polarization.
Figure 2.9 shows the polarization response of a hybrid spin LED consisting of a
V[TCNE]x~2 layer deposited on a GaAs p-i-n LED. (We will further discuss the electrical
transport characteristics of the hybrid structure in chapter 3.) As described above, current
is passed through the V[TCNE]x~2 layer and into the LED, and the circular polarization of
the light is detected. In this case, the light emitted from the LED can be resolved as
originating from the heavy holes or the light holes in the semiconductor by wavelength.
39

The magnetization of the V[TCNE]x~2 layer as measured by SQUID magnetometry is


shown in green. The circular polarization of the electroluminescence tracks the
magnetization, demonstrating that spins are being injected from the organic layer. This is
the first and only direct demonstration of electrical spin injection from an organic magnet
to date.

Figure 2.9 Spin LED circular polarization response. Red and blue triangles show the response from heavy
holes and light holes respectively, while the green line shows the magnetization of the V[TCNE]x~2 layer.
The heavy hole polarization and the magnetization match well, providing direct evidence that spins from the
organic layer are being injected into the inorganic LED structure.

2.6 Conclusion
V[TCNE]x~2 is a unique organic-based material that combines robust magnetic
properties, including a Curie temperature well above room temperature, with
semiconducting transport and benign deposition conditions. The lack of crystalline
40

structure enables deposition on a wide variety of surfaces, including flexible substrates, but
also means there is little to no magnetic anisotropy. Electrical spin injection from a
V[TCNE]x~2 layer has been demonstrated in multiple geometries, including spin valves and
spin LEDs. These results point to V[TCNE]x~2 as an extremely attractive material for
spintronics, and we look to extend these results for spin dynamics.

41

References
18.

Tang, C.W. and S.A. Vanslyke, ORGANIC ELECTROLUMINESCENT DIODES.


Applied Physics Letters, 1987. 51(12): p. 913-915.

19.

Brabec, C.J., Organic photovoltaics: technology and market. Solar Energy


Materials and Solar Cells, 2004. 83(2-3): p. 273-292.

21.

Manriquez, J.M., et al., A ROOM-TEMPERATURE MOLECULAR ORGANIC


BASED MAGNET. Science, 1991. 252(5011): p. 1415-1417.

22.

Pokhodnya, K.I., A.J. Epstein, and J.S. Miller, Thin-film V TCNE (x) magnets.
Advanced Materials, 2000. 12(6): p. 410-+.

23.

de Caro, D., et al., CVD-grown thin films of molecule-based magnets. Chemistry


of Materials, 2000. 12(3): p. 587-+.

24.

Baldo, M.A., et al., Highly efficient phosphorescent emission from organic


electroluminescent devices. Nature, 1998. 395(6698): p. 151-154.

25.

Forrest, S.R., The path to ubiquitous and low-cost organic electronic appliances
on plastic. Nature, 2004. 428(6986): p. 911-918.

28.

Cimpoesu, F., et al., Disorder, exchange and magnetic anisotropy in the roomtemperature molecular magnet V TCNE (x) - A theoretical study. Computational
Materials Science, 2014. 91: p. 320-328.

42.

Crone, B., et al., Large-scale complementary integrated circuits based on organic


transistors. Nature, 2000. 403(6769): p. 521-523.

43.

Yu, H., et al., Ultra-narrow ferromagnetic resonance in organic-based thin films


grown via low temperature chemical vapor deposition. Applied Physics Letters,
2014. 105(1).

44.

Pokhodnya, K.I., B. Lefler, and J.S. Miller, A methyl


tricyanoethylenecarboxylate-based room-temperature magnet. Advanced
Materials, 2007. 19(20): p. 3281-+.
42

46.

Lu, Y., et al., Thin-Film Deposition of an Organic Magnet Based on Vanadium


Methyl Tricyanoethylenecarboxylate. Advanced Materials, 2014. 26(45): p. 76327636.

48.

Cao, Y., P. Smith, and A.J. Heeger, COUNTERION INDUCED


PROCESSIBILITY OF CONDUCTING POLYANILINE AND OF CONDUCTING
POLYBLENDS OF POLYANILINE IN BULK POLYMERS. Synthetic Metals,
1992. 48(1): p. 91-97.

49.

LG: Curved smartphones like G Flex 40% of market by 2018. 2013.

50.

Whitwam, R., Harvard uses distributed computing and quantum mechanics to


dream up new solar cells. 2013.

51.

Kao, C.Y., et al., Investigation of thin films of organic-based magnets grown by


physical vapor deposition. Applied Physics Letters, 2014. 105(14).

52.

Kao, C.-Y., et al., Molecular Layer Deposition of an Organic-Based Magnetic


Semiconducting Laminate. Acs Applied Materials & Interfaces, 2012. 4(1): p.
137-141.

53.

Haskel, D., et al., Local structural order in the disordered vanadium


tetracyanoethylene room-temperature molecule-based magnet. Physical Review
B, 2004. 70(5).

54.

Bozdag, K.D., et al., Optical control of magnetization in a room-temperature


magnet: V-Cr Prussian blue analog. Physical Review B, 2010. 82(9).

55.

Caruso, A.N., et al., Direct evidence of electron spin polarization from an


organic-based magnet: Fe-II(TCNE)(NCMe)(2) (FeCl4)-Cl-III. Physical Review
B, 2009. 79(19).

56.

Matsuura, H., K. Miyake, and H. Fukuyama, Theory of Room Temperature


Ferromagnet V(TCNE)(x) (1.5 < x < 2): Role of Hidden Flat Bands. Journal of
the Physical Society of Japan, 2010. 79(3).

43

57.

Tchougreeff, A.L. and R. Dronskowski, A computational study of the crystal and


electronic structure of the room temperature organometallic ferromagnet
V(TCNE)(2). Journal of Computational Chemistry, 2008. 29(13): p. 2220-2233.

58.

De Fusco, G.C., et al., Density functional study of the magnetic coupling in


V(TCNE)(2). Physical Review B, 2009. 79(8).

59.

Yoo, J.W., et al., Spin injection/detection using an organic-based magnetic


semiconductor. Nature Materials, 2010. 9(8): p. 638-642.

60.

Tengstedt, C., et al., X-ray magnetic circular dichroism and resonant


photomission of V(TCNE)(x) hybrid magnets. Physical Review Letters, 2006.
96(5).

61.

Kortright, J.B., et al., Bonding, backbonding, and spin-polarized molecular


orbitals: Basis for magnetism and semiconducting transport in V TCNE (x similar
to 2). Physical Review Letters, 2008. 100(25).

62.

Harberts, M., et al., Chemical vapor deposition of an organic magnet, vanadium


tetracyanoethylene, V[TCNE]x~2. Journal of Visualized Experiments, 2015.

63.

Miller, J.S., Oliver Kahn Lecture: Composition and structure of the V TCNE (x)
(TCNE = tetracyanoethylene) room-temperature, organic-based magnet - A
personal perspective. Polyhedron, 2009. 28(9-10): p. 1596-1605.

64.

Yoo, J.-W., et al., Photoinduced magnetism and random magnetic anisotropy in


organic-based magnetic semiconductor V(TCNE)(x) films, for x similar to 2.
Physical Review Letters, 2007. 99(15).

65.

Plachy, R., et al., Ferrimagnetic resonance in films of vanadium


tetracyanoethanide (x), grown by chemical vapor deposition. Physical Review B,
2004. 70(6).

66.

Lundgren, L., P. Svedlindh, and O. Beckman, MEASUREMENT OF COMPLEX


SUSCEPTIBILITY ON A METALLIC SPIN-GLASS WITH BROAD RELAXATION
SPECTRUM. Journal of Magnetism and Magnetic Materials, 1981. 25(1): p. 3338.
44

67.

Li, B., et al., Magnetoresistance in an All-Organic-Based Spin Valve. Advanced


Materials, 2011. 23(30): p. 3382-+.

68.

Gould, C., et al., Tunneling anisotropic magnetoresistance: A spin-valve-like


tunnel magnetoresistance using a single magnetic layer. Physical Review Letters,
2004. 93(11).

45

Chapter 3 Transport in a hybrid organic-inorganic heterostructures

3.1 Introduction
Semiconducting materials form the basis of modern electronics, and the design and
manufacture of semiconductor devices is a several hundred billion dollar industry. The
basic building block of semiconductor electronics is the p-n junction, from which devices
such as diodes, transistors, solar cells, light-emitting diodes, and integrated circuits can be
built. The junction consists of a single crystal semiconductor with two doped regions, one
p-type and one n-type, referring to doping with electron acceptors and donors respectively.
In general, a doped semiconductor is relatively conducting, but by placing these two
differently doped regions in contact, a region is created in the neighborhood of the interface
where the electron and hole dopants recombine and charge is depleted. Depending on the
direction of the electrical bias on the junction, this depletion region can be extended or
reduced, thus allowing current to flow in one bias direction but not the other. The basic
physical principles that govern the behavior of p-n junctions are well understood, and
therefore they are useful tools for investigating other materials, in particular the organic
materials that we are interested in.
Furthermore, the field of semiconductor spintronics promises the extension of spinbased electronics beyond memory and magnetic sensing into active electronic components
with implications for next-generation computing and quantum information[69]. Organic42

based magnets with room temperature magnetic ordering and semiconducting functionality
promise to further broaden this impact by providing a route to all-organic spintronic
devices and hybrid organic/inorganic structures. These devices can potentially exploit the
multifunctionality and ease of production in organic systems as well as the well-established
spintronic functionality of inorganic materials [21, 70-72]. In this vein, electrical spin
injection in a hybrid organic/inorganic spin-resolved light-emitting diode (spin-LED)
structure has been demonstrated, revealing the ability to probe the properties of organic
and molecular systems by utilizing inorganic systems [73].
The long term goal for organic materials is to find applications that take advantage
of their unique properties. It is unlikely for any material to supplant silicon as the
semiconductor of choice for electronics; rather, the goal is to find new applications and
device structures for new materials. From this perspective, studying how organics interact
with well-known semiconductor devices could eventually lead to hybrid devices that can
make use of existing infrastructure. Here, we deposit V[TCNE]x~2 on a GaAs p-i-n junction
and explore the impact on the electrical transport properties in these hybrid devices in order
to better understand the spin injection process. We compare temperature-dependent
transport in three samples, the full V[TCNE]x~2 diode, a bare GaAs p-i-n diode and a
V[TCNE]x~2 layer on glass with metallic contacts to isolate the response of the individual
components from emergent behavior at the organic/inorganic interface.

43

3.2 Transport in p-i-n junctionsEquation Chapter (Next) Section 1


3.2.1 Ideal p-n junction behaviorEquation Chapter (Next) Section 1
The simplest p-n junction consists of a single crystal semiconductor with two
regions with different doping types, one p-type and one n-type. The doping profile of a
junction can be varied to engineer the band structure of the junction and therefore obtain
different properties suitable for different applications. For this section, we will focus on
the simplest case of an abrupt junction, where the two regions both have uniform impurity
concentrations separated by a discontinuous change. This can be represented
mathematically as
N , x 0
N d ( x) d

0, x 0
0, x 0
N a ( x)

N a , x 0

(3.1)

where Ni(x) is the density of donors or acceptors, Ni is the constant donor concentration,
and the discontinuity at x = 0, as shown in Fig. 3.1(a). Assuming the crystal is in
equilibrium (no external bias is applied, the system is in thermal equilibrium, etc.), the total
potential drop across the junction is:

N N
e Eg k BT ln d a
N c Pv

(3.2)

where Eg is the band gap, Nc is the concentration of electrons in the conduction band, and
Pv is the concentration of holes in the valence band.
The electrochemical potential across the junction is shown in Fig. 3.1. Far away
from the interface, the energy levels are unchanged; at the interface, the chemical potential
bends (equivalently the energy levels bend and the chemical potential is constant), as
44

shown in Fig. 3.1(b). Additionally, a depletion region with sharply reduced carrier densities
forms around the interface. This can be intuitively understood in terms of charge diffusion
- in the initial condition, a high concentration of electrons (holes) exists in the n-side (pside) of the junction. Correspondingly, there is a low concentration of electrons (holes) in
the p-side (n-side). This cannot be maintained, and so the electrons begin to diffuse across
the interface into the p-side, with holes diffusing in the opposite direction. This charge
transfer builds up an electric field until the diffusive drive is exactly canceled out. In doped
semiconductors the carriers are mobile, and so the carrier density is very low where the
electric field exists. The depletion region typically extends from 10 to 1000 nm depending
on materials parameters like the dielectric permittivity, the impurity densities, and contact
potential, as well as the bias voltage.

45

Figure 3.1 (a) Doping profile of an ideal p-n junction with interface at x = 0 and uniform donor/acceptor
densities. (b) Energy diagram of this doping profile.

With this basic picture in place, we will now consider the behavior of the junction
under an external bias voltage V (we will take V to be positive when it raises the potential
of the p-side relative to the n-side). When there is no applied bias, the depletion region
extends several nm around the border. The reduced carrier density means the depleted layer
46

has a high electrical resistance, especially in comparison to the conductive doped


semiconducting layers, and so the system can be treated as a high resistance layer
sandwiched between two low resistance layers in series. Therefore, the majority of the
voltage drop across the junction will occur across the depleted region, and application of a
positive voltage will shrink the depletion region while a negative voltage will extend the
region. The total current is the sum of the hole and electron currents
j e J hgen J egen e eV

k BT

(3.3)

This equation showcases the highly rectifying behavior of a p-n junction, with
practically no current flowing in reverse bias and exponential current flowing in forward
bias, shown in Fig. 3.2 (a full derivation of this behavior can be found in [74]). Real diodes
can be characterized by how well they match this ideal scenario via the Shockley diode
equation
I I s eV / nVT 1

(3.4)

where I is the measured current, IS is the reverse bias saturation current, VT is the thermal
voltage ( k BT e ), and n is the ideality factor, which is equal to 1 for an ideal diode. This
equation only takes drift, diffusion, and thermal recombination-generation into account,
and therefore does not describe phenomena like reverse breakdown (where a large enough
reverse bias will make the diode conduct).

47

Figure 3.2 Ideal current-voltage response of a p-n junction. Current increases exponentially in forward bias,
while reverse bias generates no current. Blue dashed line shows the turn on voltage, where V = nVT.

3.2.2 Transport in a p-i-n junction


A p-i-n diode adds an intrinsic layer (i.e. with no impurities added) between the pand n-sides of the junction. The intrinsic region changes the characteristics of the junction
so that the rectification is not as strong, but makes the structure more suitable for a range
of applications, such as LEDs, photodetectors, RF switches, attenuators, and phase shifters
[75]. Typically, the depletion region is much larger than in a p-n junction and exists almost
entirely in the intrinsic region. For our purposes, p-i-n photodiodes are attractive because
of this extended depletion region, which creates a larger area where recombination and
photon emission can take place.
48

3.2.3 Band structure simulation


The inorganic system we studied is a GaAs-AlGaAs p-i-n diode structure. The
wafer is grown by III-V molecular beam epitaxy (MBE), and was obtained from the
Semiconductor Epitaxy and Analysis Laboratory (SEAL) at The Ohio State University.
The III-V material is grown on a p-doped (p = 11018 cm-3) GaAs (100) substrate with
layer structure as follows: 300 nm p-GaAs (11017 cm-3) buffer layer/ 200 nm pAl0.1Ga0.9As (p = 11016 cm-3) p-contact/ 25 nm i-Al0.1Ga0.9As barrier/ 10 nm i-GaAs QW/
25 nm i-Al0.1Ga0.9As barrier/ 100 nm n-Al0.1Ga0.9As (n = 11016 cm-3 ) drift layer/ 15 nm
nn+Al0.1Ga0.9As grading layer going from n = 11016 cm-3 to 51018 cm-3/ 15 nm n+Al0.1Ga0.9As (n+ = 51018 cm-3) Schottky contact. A schematic is shown in Fig. 3.2. This
layer structure is optimized for spin-LED measurements [73, 76] as mentioned above,
the p-i-n layer structure is useful for photoemission, with the intrinsic layer forming a
quantum well for optical detection (see Fig. 3.3).

Figure 3.3 Layer structure of inorganic LED used for these measurements. Thicknesses are not to scale.

49

BandEng is a program that simulates band structures by solving the Poisson and
Schrodinger equations [77]. We can use this to calculate the band structure of the inorganic
diode at zero applied bias, which is shown in Fig. 3.4, with the p-GaAs buffer layer and
substrate omitted. The n-AlGaAs layer is on the left and the p-AlGaAs layer is on the right.
The V[TCNE]x~2 layer is not simulated since we do not know all of the materials properties
necessary for the calculation. In the full device, the V[TCNE]x~2 layer would be deposited
over the n-AlGaAs layer, so the exact nature of the interface located on the left side of the
band structure is also unknown. We show simulations with the Fermi energy (solid red
line) pinned at the valence (1.5487 eV) and conduction (0 eV) bands, as well as in the
middle of the AlGaAs band gap (0.75 eV), revealing that the overall band structure is only
affected by interface states near the valence band of the top AlGaAs layer (the n+ doping
at the surface makes this situation unlikely). The GaAs quantum well is clearly visible near
the center.

50

Figure 3.4 BandEng simulation of the band structure of the inorganic diode with the Fermi energy (red) pinned at the
middle of the band gap (black), at the valence band (blue), and at the conduction band (green). The V[TCNE]x~2 layer
would be at the left side of the diagram.

3.3 Sample fabrication


In order to perform electrical measurements with the hybrid diodes, we prepare
samples using a combination of standard inorganic processing techniques and specialized
processes necessitated by the air-sensitive nature of V[TCNE]x~2. We begin by cleaving 5
mm x 5 mm pieces of a molecular beam epitaxy (MBE) -grown III-V wafer (layer structure
described in previous section). After cleaving, 7 nm of Al and 23 nm of Au are thermally
deposited on the back of the p-GaAs substrate to form an Ohmic back contact. The wafer
51

is pattered with photolithography to define a 200 m x 1.5 mm mesa in the LED structure
using a standard Shipley 1813 photoresist, developing for 2 minutes in MF-319 developer,
and baking at 115 C for 10 minutes. The wafer is then wet etched using a solution of 100
mL deionized water, 5 mL hydrogen peroxide (H2O2), and 1 mL hydrofluoric acid (HF)
for 60 seconds, giving an etch depth of around 300 nm. The remaining photoresist is
stripped with a solvent rinse. In order to isolate electrical contact to the top of the mesa,
the wafer is again coated with insulating Shipley 1813 photoresist. Using photolithography,
a 100 m x 1 mm window is opened over the mesa, followed by developing for 2 minutes
in MF-319 and baking at 115 for 20 minutes to remove any solvents that may react with
the V[TCNE]x~2.
Due to the air-sensitive nature of V[TCNE]x~2 processing involving the organic
layer takes place in a series of interconnected argon glove boxes (O2, H2O < 0.5 ppm). A
layer of V[TCNE]x~2 roughly 400 nm in thickness is grown on the entire sample surface
via a CVD process. The film is formed by a gas phase reaction of TCNE and V(CO)6 at a
rate of approximately 5.5 nm/min [22]. The sample is then transferred in situ to an
integrated vacuum chamber for top contact deposition. A 7 nm layer of Al is thermally
deposited over the entire window using a shadow mask, followed by a 23 nm thick stripe
of Au that partially covers the mesa. The thicker gold layer ensures good electrical contact
to the aluminum and to the mesa. The device is then wired to a custom-designed sample
mount (Fig. 3.5). The top contact is connected with gold wire to a pressed indium contact,
and the back contact is connected to the sample mount using double-sided copper tape. An
additional strip of Kapton tape is used to secure the sample to the mount.

52

For electrical measurements, the sample is loaded in a custom-designed air-free


sample mount, shown as a Solidworks model in Fig. 3.5(a). Because V[TCNE]x~2 degrades
in air, the sample mount is built to prevent any exposure to air while the sample is being
transported between the glovebox and the measurement apparatus. The entire assembly
consists of a copper mount, where the sample is connected as described above, and an
aluminum can that fits around the mount and seals with a screw cap. The body of the can
is fitted with a pair of windows to allow for optical access to the sample. The windows
were purchased from Oxford Instruments and measure 10 mm in diameter. The windows
are sealed to the can by Eccobond and a spring assembly to account for differential
contraction at low temperatures. The spring assembly is depicted as the dark rim of the
window in Fig. 3.5(a). The red surface indicates where the epoxy is applied. A picture of
the various parts of the sample mount is shown in Fig. 3.5(b). The sample is wired and
loaded while in the argon glove box. The can is sealed with a strip of indium wire pressed
between the body of the can and the sample mount as shown by the green surface in Fig.
3.5(a). This assembly allows for completely air-free handling of the samples from the
deposition of the V[TCNE]x~2 layer through full electrical and optical characterization.
Figure 3.5(c) shows a picture of a hybrid diode mounted in the full assembly. The copper
sample mount and the strips of Kapton tape securing the sample can be seen through the
window. The mesa, the aluminum and gold contact, the indium dot and gold wiring on the
sample are also visible.

53

Figure 3.5 (a) CAD model of air-free can for low temperature electrical and optical measurements. Green
surface is sealed with indium, while red surfaces are sealed with epoxy. Blue surfaces are windows to allow
optical access. (b) Photo of disassembled can and copper sample mount. (c) Photo of sample mounted in can.

3.4 Electrical transport in hybrid organic-inorganic diode


Figure 3.6(a) shows the temperature dependence of the hybrid V[TCNE]x~2 diode.
The inset shows a schematic of the full device, indicating the V[TCNE]x~2 deposited over
a photoresist window onto the inorganic p-i-n diode with top Al/Au contact. Currentvoltage (IV) curves are measured from room temperature down to 60 K, and we observe
that the turn-on voltage shifts higher as the temperature decreases. Since the V[TCNE]x~2
layers resistivity increases sharply with temperature, going to temperatures lower than 60
K pushes the turn-on voltage past 20 V and runs the risk of burning out the device. The IV
curve also deviates from purely exponential behavior. Figure 3.6(b) shows the same data
on a semi-log plot, which more clearly highlights the deviation from ideal diode behavior
54

as temperature is decreased. At 300 K the IV is close to linear on the semi-log scale,


indicating exponential behavior, while at lower temperatures the IVs bend away from linear
behavior.
Figure 3.6(c) compares the IV characteristics of a bare diode, lacking a V[TCNE]x~2
layer, and a hybrid V[TCNE]x~2 diode at a temperature of 60 K. We observe the turn-on
voltage of the bare diode at +1.5 V comparable to the band gap of the AlGaAs layer (1.5487
eV). The addition of the V[TCNE]x~2 shifts the turn-on voltage to +14 V. We can quantify
the deviation from ideal behavior of the diode by fitting the IV curve to the Shockley diode
equation (eq. 3.4). The ideality factor is 1 for a perfect diode, and typical diodes have an
ideality less than 2. Fitting the 60 K bare diode data yields an ideality of 32.8; this value is
high compared to what we would expect from a bare p-i-n diode, and is probably due to
the poor quality of the contact to the n-AlGaAs. (For these characterization measurements
we chose an Al/Au top contact to maintain process consistency and to match the work
function of the V[TCNE]x~2 layer to the best of our knowledge, rather than using a
traditional Ohmic stack.) Adding the V[TCNE]x~2 layer gives n = 324.4, decreasing the
ideality by an order of magnitude. This decrease in ideality is also reflected in the semi-log
plot of the same data, shown in the inset to Figure 3.6(c). The bare diode shows linear
behavior; in contrast, the V[TCNE]x~2 diode shows a strong deviation from linearity. We
observe that the p-i-n diode exhibits little or no temperature dependence, and therefore we
can conclude the temperature dependence of the hybrid diode is primarily due to the
V[TCNE]x~2 layer. The shift in turn-on voltage and deviation from ideal diode behavior at
low temperatures indicate the addition of a considerable series resistance, which could be

55

a contribution from bulk transport through the V[TCNE]x~2 layer or an effect of the
interface between the organic V[TCNE]x~2 layer and the inorganic portion of the device.

Figure 3.6 (a) Inset shows a device schematic, a V[TCNE]x~2 layer deposited on a n-i-p diode with Al/Au contacts.
Current-voltage scans of the hybrid diode structure are shown at 300 K, 250 K, 200 K, 150 K, 100 K, 75 K, and 60 K.
Decreasing temperature pushes the diode turn-on to higher voltages, with the effect becoming more pronounced at lower
temperature. (b) The same data as Figure 3.4(a) on a semi-log plot. Decreasing the temperature causes the IV to deviate
from ideal diode behavior, which manifests as a deviation from linearity. (c) Comparison between a bare diode (black
open triangles) and a hybrid V[TCNE]x~2 diode (black filled squares) at 60 K. Inset shows the same data on a semi-log
plot.

56

To determine whether the IV characteristics of the V[TCNE]x~2 diode are due to


interface effects or bulk effects, additional measurements were taken on a device consisting
solely of a V[TCNE]x~2 layer with two metallic contacts, shown schematically in the inset
to Figure 3.7(a). Current-voltage curves are shown on a linear scale in Fig. 3.7(a) and on a
semi-log scale in Fig. 3.7(b). Without the diode structure, the IV does not rectify and is
symmetric about zero bias. At high temperature, the IV is linear and the V[TCNE]x~2 layer
has relatively low resistance. However, as temperature decreases, the IV becomes
significantly nonlinear and the low bias resistance grows quickly, which is consistent with
thermal activation in a narrow-gap semiconductor. The semi-log plot in Fig. 3.7(b) exhibits
similar IV characteristics to that of the V[TCNE]x~2 diode at low temperature.

57

Figure 3.7 (a) Inset shows a device schematic, a V[TCNE]x~2 layer between two Al/Au contacts on a glass substrate.
The temperature dependence of the IV curve is shown, with data taken at 200 K, 175 K, 150 K, 125 K, 100 K, 75 K, and
60 K. At temperatures higher than 200 K the IVs are similar on this scale, and are not shown. Without the diode, no
rectifying behavior is observed, though strong nonlinearity is seen. (b) The same data as Fig. 3.5(a) on a semi-log plot.
As temperature decreases, the IV bends away from linear behavior in a similar fashion to the V[TCNE]x~2 diode.

The combination of the temperature dependence of the hybrid diode and the bare
diode suggest that the V[TCNE]x~2 layer contributes a series resistance to the electrical
transport characteristics of the diode, which shifts the turn-on voltage and decreases the
ideality. Removing the diode and measuring the transport behavior of a bulk V[TCNE]x~2
58

layer reveals similar temperature-dependent behavior as the hybrid device without the
rectifying effect of the diode. Taken together, this suggests that bulk behavior of the
V[TCNE]x~2 layer is dominating the transport response of the hybrid diode rather than
interface effects between the organic and inorganic layers.

3.5 Conclusion
In conclusion, we observe that bulk V[TCNE]x~2 properties dominate the
characteristics of a hybrid organic/inorganic device. By better understanding the properties
of this hybrid device we identify constraints for V[TCNE]x~2.based spintronic devices. This
class of materials shows significant promise for room-temperature spintronic devices, as it
is apparent that the organic/inorganic interface is not a fundamental barrier to either spin
transport (as demonstrated by the spin-LED [73]) or charge-based functionality in hybrid
devices. Further, this approach validates the use of extensively-studied inorganic systems
as a probe to study organic and molecular systems, significantly accelerating the
development curve for these materials.

59

References
21.

Manriquez, J.M., et al., A ROOM-TEMPERATURE MOLECULAR ORGANIC


BASED MAGNET. Science, 1991. 252(5011): p. 1415-1417.

22.

Pokhodnya, K.I., A.J. Epstein, and J.S. Miller, Thin-film V TCNE (x) magnets.
Advanced Materials, 2000. 12(6): p. 410-+.

69.

Divincenzo, D.P., QUANTUM COMPUTATION. Science, 1995. 270(5234): p.


255-261.

70.

Prigodin, V.N., et al., Spin-driven resistance in organic-based magnetic


semiconductor V TCNE (x). Advanced Materials, 2002. 14(17): p. 1230-+.

71.

Yoo, J.-W., et al., Spin injection/detection using an organic-based magnetic


semiconductor. Nature Materials, 2010. 9.

72.

Yoo, J.-W., et al., Giant magnetoresistance in ferromagnet/organic


semiconductor/ferromagnet heterojunctions. Physical Review B, 2009. 80(20).

73.

Fang, L., et al., Electrical Spin Injection from an Organic-Based Ferrimagnet in a


Hybrid Organic-Inorganic Heterostructure. Physical Review Letters, 2011.
106(15).

74.

Ashcroft, N.W. and N.D. Mermin, Solid State Physics. 1976: Thomson Learning,
Inc.

75.

Taniyasu, Y., M. Kasu, and T. Makimoto, An aluminium nitride light-emitting


diode with a wavelength of 210 nanometres. Nature, 2006. 441(7091): p. 325-328.

76.

Adelmann, C., et al., Spin injection and relaxation in ferromagnet-semiconductor


heterostructures. Physical Review B, 2005. 71(12).

77.

Grundmann, M. and U.K. Mishra. Available from:


http://my.ece.ucsb.edu/mgrundmann/bandeng.htm.

60

Chapter 4 Ferromagnetic resonance in thin film organic-based magnets

4.1 IntroductionEquation Chapter (Next) Section 1


Magnetic resonance is a powerful tool for studying materials properties, probing
the spins of unpaired electrons (electron paramagnetic resonance, EPR), atomic nuclei
(nuclear magnetic resonance, NMR), and magnetic materials (ferromagnetic resonance,
FMR). The most prominent application is in medical imaging, as magnetic resonance
imaging (MRI), but there is also strong interest in magnetic materials for planar microwave
devices, such as circulators, isolators, phase shifters, and patch antennas, with much of the
focus on soft ferrite materials [9]. FMR is also a sensitive probe of the effective fields and
relaxation processes that govern magnetization dynamics in the ferromagnet [29, 78-80].
Furthermore, the field of spintronics promises the control of the electrons spin degree of
freedom in addition to the charge, and is therefore predicated on generating, manipulating,
and measuring spin polarizations [1, 76, 81]. In this context, FMR has emerged as a
promising method for generating pure spin currents through the spin pumping process,
which has been observed in a magnetic bilayer [30] and in ferromagnet-normal metal
heterostructures, detected via the inverse spin Hall effect [13-15].
In terms of materials, yttrium iron garnet (Y3Fe5O12, YIG) has enabled the
development of high frequency electronic devices that take advantage of its electrically
61

insulating properties, extremely narrow ferromagnetic resonance (FMR) linewidth, and


consequent low microwave loss [82]. However, only films grown on specific latticematched substrates (such as gadolinium gallium garnet, GGG) yield optimal properties
[47]. These properties make YIG the material of choice for a range of microwave
applications, but also limit it to device structures that can accommodate these specific
substrates. Specifically, devices must either be compatible with high temperature growth
or integrated with a separately grown YIG layer. Given these constraints, there is space for
a narrow linewidth magnetic material that can be deposited on varied substrates, ideally
post fabrication, to complement YIG.
Organic and organic-based conductors and semiconductors have been shown to
tolerate a wide variety of substrates and are compatible with low temperature and ambient
pressure deposition. Recently, these materials have made inroads in consumer electronics
in the form of organic light-emitting diodes (OLEDs) [18], organic photovoltaics (OPVs)
[19], and organic field effect transistors (FETs) [25, 42]. In this light, an organic material
with magnetic resonance properties comparable to YIG could similarly enable future
organic-based microwave devices. A promising candidate for these purposes is
V[TCNE]x~2 (x~2,

TCNE:

tetracyanoethylene),

an

organic-based

ferrimagnetic

semiconductor with room temperature magnetic ordering and EG ~ 0.5 eV [21].


V[TCNE]x~2 can be grown via a low temperature (60 C) chemical vapor deposition (CVD)
process, enabling conformal deposition on a variety of surfaces such as flexible and threedimensional substrates [21-23]. However, previous studies of FMR in V[TCNE]x~2 have
yielded complex resonance spectra that evolve as a function of the external field orientation
and vary significantly from sample to sample [65].
62

This chapter describes the work we have done to characterize and improve the
magnetic resonance properties of V[TCNE]x~2 and similar thin film organic-based magnets.
First, we describe the theory of magnetic resonance and relevant spin relaxation
mechanisms. Second, we discuss the sample preparation process. Third, we discuss studies
with an EPR cavity. Finally, we discuss measurements taken with a microwave stripline.

4.2 Theory of magnetic resonance


Consider an electron with magnetic moment and spin quantum number s = 1/2,
with magnetic components ms = 1/2. By applying an external magnetic field H, the energy
levels of the ion will be split by the Zeeman effect, creating an energy gap E g e B H ,
where ge is the electron g-factor and B is the Bohr magneton, as illustrated in Fig. 4.1. An
electron can transition between the split levels by absorbing or emitting a photon with
energy h E , leading to the resonance condition:
g e B H

(4.1)

From a quantum mechanics standpoint, resonance can be driven by applying a combination


of electromagnetic radiation and a static magnetic field that satisfies the resonance
condition.

63

Figure 4.1 Electron energy level split by the Zeeman effect. For magnetic resonance, the energy of the applied
electromagnetic radiation must match the energy gap opened by the static magnetic field.

Ferromagnetic resonance is distinct from other types of magnetic resonance in that


there is a strong exchange interaction between the spins. As a result, we can treat the spins
in the ferromagnet collectively as the magnetization (we also assume that the applied
magnetic field is strong enough to remove the domain structure). The equation governing
the time evolution of the magnetization is the Landau-Lifshitz-Gilbert (LLG) equation
[83]:

dM

dM
M HT
M

dt
Ms
dt

(4.2)

Where M is the magnetization vector, is the electron gyromagnetic ratio, HT is the total
magnetic field experienced by the magnetization, is the dimensionless Gilbert damping
parameter, and Ms is the saturation magnetization, or the magnitude of M. The physical
interpretation of this equation can be seen in Fig. 4.2.

64

Figure 4.2 Magnetization dynamics as governed by the Landau-Lifshitz-Gilbert equation. At equilibrium, the
magnetization (red) is aligned with the static field (black). Upon excitation, the magnetization is subject to
two different torques (blue and green arrows). This leads to the magnetization relaxing to equilibrium by
following the path along the blue dashed line.

At equilibrium (left), the magnetization (red arrow) is aligned with the total
magnetic field (black arrow). After some process excites the magnetization and torques it
away from equilibrium (right), the magnetization will try to relax to its equilibrium
position. However, it does not do so by directly realigning with the field. The first term in
the LLG equation describes the torque the applied field exerts on the magnetization (green
arrow), which causes the magnetization to precess around the total field. The second term
describes damping of the precession (blue arrow), which pushes the magnetization back to
equilibrium. Therefore, overall the magnetization will follow the spiral path (blue dashed
line) as it relaxes.

65

Ferromagnetic resonance is a special case of this precession process. The


magnetization is torqued away from equilibrium by incoming electromagnetic radiation (in
general, an AC magnetic field) and begins to relax as described above. However, as the
magnetization completes a precession cycle, the applied radiation pulls it away from
equilibrium again, keeping the precession going. Therefore, the magnetization is in a semisteady state - as long as the static and oscillating fields are unchanged, the magnetization
will continue to precess. This can be measured by monitoring the microwave absorbance
of the sample. The lower level will be more highly populated than the higher one because
the statistical distribution of electrons is determined by the Maxwell-Boltzmann
distribution, assuming the system is in thermal equilibrium:

nhigher
nlower

Elower
E
exp higher
k BT

E
exp

k BT


exp

k BT

(4.3)

Therefore, there is net absorption of radiation. Unlike the simpler case of EPR or NMR,
the resonance field is contingent on a number of factors, and in general the magnetic free
energy depends on the Zeeman energy, the microwave energy, the demagnetizing field, the
crystalline anisotropy field, and the exchange energy. In the case of small deviations from
equilibrium, the resonance frequency res is given by

res

E
M sin

E E2

(4.4)

Where Eij is the second derivative of the free energy E with respect to the angles i, j. By
varying the angle of the applied field, the position of the resonance field can yield
information about the anisotropy fields in the sample. In practice, this yields predictions

66

like the expected angular-dependent resonance condition for a thin film ferromagnet
rotated from in-plane to out-of-plane (assuming H>>4Ms):

res

H 4 M

cos 2 H H 4 M s cos 2 H

(4.5)

Where res is the resonance frequency in MHz, H is the applied external field in Oe and H
is the angle of the applied external field with respect to the normal of the film plane.
In a thin film ferromagnet, the resonance condition can further be easily solved for
two cases: when the magnetization lies in the plane of the film and when it aligns with the
normal vector of the film. In the absence of significant crystalline anisotropy the dominant
field is the demagnetizing field, or the field generated by the magnetization of the
ferromagnet. This field is the energy cost of magnetic moments polarized normal to the
film plane having dipolar fields opposite to neighboring moments, equivalent to producing
magnetic charges at the film surfaces and so it manifests as a shape anisotropy field that
tries to keep the magnetization in the plane of the film.
The effective field has the form H eff H 4 M s , where H is the external field.
Using equation 4.2 with 0 , the resonance condition is simple to compute for external
field and magnetization out of plane, as H eff ( H 4 M s ) z and the magnetization with
a small circular precessing component is M M s z me it x ime it y where m M s .
The resulting resonance condition for out-of-plane magnetization is

H 4 M s

67

(4.6)

For in-plane magnetization it is important to consider the dipole field from the precessing
magnetization, which results in an effective field with a time-dependent component

H eff Hz 4 mx eit x where z is the direction of the field and x is the direction normal
to the film plane. Together with the magnetization M M s z mx eit x imy eit y this
results in the resonance condition for in-plane magnetization.
2


H H 4 M s

(4.7)

Furthermore, the linewidth of the absorption spectrum can yield information on the
relaxation times in the system. Several mechanisms can contribute to the FMR linewidth generally we are more interested in the materials intrinsic relaxation properties, but shape
effects, interfaces, magnetic impurities, and other defects can change the anisotropy fields
and therefore the linewidth. These mechanisms include Gilbert damping, mosaicity, twomagnon scattering, and inhomogenous broadening, and are further described in section 1.3.

4.3 Sample preparation


Due to the air sensitive nature of the films, samples are deposited inside a series of
argon glove boxes (O2, H2O < 0.5 ppm), and then transferred out of the glove box in a
sealed tube for measurements. Films of thicknesses between 250 and 350 nm are deposited
onto sapphire substrates for FMR measurements. Sapphire is used for FMR because of its
lower microwave absorption. All samples are sealed in evacuated ESR grade quartz tubes
to prevent air exposure during transport; for FMR measurements, samples are mounted on
a custom sample mount to maintain their orientation. The sample mount consists of a
68

cylindrical piece of ceramic with a slot cut out, as shown in Fig. x. The cylindrical shape
keeps the sample upright, which is important since small angular deviations in the sample
orientation (and therefore the orientation of the applied field) can introduce shifts in the
resonance field and linewidth.

4.4 Cavity ferromagnetic resonance


4.4.1 Principles of operation
Measurements in this section were taken with a Bruker ESR Spectrometer, also
known as an ESR cavity. Essentially, the spectrometer consists of a radiation source, a
sample space, and a detector; the source and detector are contained in a microwave bridge.
A diagram of the microwave bridge is shown in Fig. 4.3, with the red arrows showing the
direction the microwave radiation is propagating.

69

Figure 4.3 Schematic diagram of Bruker EPR cavity.

The microwave source generates the microwave radiation for the measurement (in
this spectrometer and generally in ESR spectrometers, this is X-band radiation, around 9.8
GHz). It is difficult to change the output power at the source, and so a variable attenuator
is used to modify the microwave power delivered to the sample. This is followed by a
circulator, which is a microwave device that directs the flow of radiation from port 1 to 2,
2 to 3, and 3 to 1 (as labeled in Fig. 4.3). This ensures that the microwaves from the source
(input to port 1) is directed to the sample space (port 2), rather than to the detection portion
of the spectrometer. After interacting with the sample in the sample space, the reflected
70

microwaves enter the circulator at port 2, and so are directed to the detection arm (port 3).
The sample space is a microwave cavity, a metal space that amplifies signals from the
sample.
The reflected microwaves are detected by Schottky barrier diode, which converts
the incoming radiation to electrical current. The diode operates in two general regimes, one
where the current is directly proportional to the microwave power (the square law regime)
and one where the current is proportional to the square root of the microwave power (the
linear regime). The diode is most sensitive in the linear regime, and in particular when the
output current is around 200 A, and so the microwave bridge contains a reference arm
that consists of a second attenuator and a phase shifter that matches the phase of the
reference microwaves with the reflected microwaves. Some of the radiation generated by
the source is split off to the reference arm, and the attenuator works to keep the detection
diode in the optimal sensitivity regime.
The microwave cavity is a metal box that resonates with the microwave radiation,
which means it stores the energy - when the incoming radiation matches the resonance
frequency of the cavity, no microwaves are reflected. The cavity is characterized by its
quality factor Q, which indicates how efficiently the cavity stores microwave energy.
Microwaves are coupled into the cavity through an iris, with a screw that effectively
changes the size of the aperture.

4.4.2 Ultra-narrow ferromagnetic resonance in V[TCNE]x


The magnetically homogenous nature of the films is also well captured by FMR
measurements. Figure 4.4 shows the FMR spectrum of a thin film of V[TCNE]x~2 (black)
71

as well as the precursors used in the V[TCNE]x~2 CVD process, V(CO)6 (red) and TCNE
powder (blue). Measurements are taken at a temperature of 300 K with 200 W of applied
microwave power at 9.63 GHz and Hext applied in-plane. The applied microwave frequency
is tuned between 9 and 10 GHz for optimal microwave cavity performance before starting
the measurement. The inset of Fig. 4.4 shows the precursor spectra at a 1000 expanded
scale. The peak in the V[TCNE]x~2 spectrum has an exceptionally narrow FWHM linewidth
of 1.4 G, and there is no feature in the two precursor spectra that resembles the sharpness
and position of the resonance peak seen in the V[TCNE]x~2 spectrum. Previous reports of
FMR in V[TCNE]x~2 thin films found spectra with similarly sharp features but with
multiple resonances that evolve with the field orientation with respect to the plane of the
film. The observed linewidth of the V[TCNE]x~2 film is comparable to that of YIG films of
similar thickness (1.1 Oe) and substantially better than all other known thin film
ferromagnets.

72

Figure 4.4 Ferromagnetic resonance spectrum of a V[TCNE]x~2 thin film taken at 300 K and an applied
microwave frequency of 9.388 GHz, with the static field in the sample plane (black). The linewidth of the
peak is 1.4 G, comparable to the narrowest line seen in inorganic materials. FMR spectra of the precursors
TCNE (blue) and V(CO)6 (red) show the peak only appears in V[TCNE]x~2 and is therefore a response of the
overall system. Inset: Zoomed in view of the TCNE (blue) and V(CO)6 (red) FMR spectra.

4.4.3 Angular dependence of ferromagnetic resonance


Fig. 4.5 shows the FMR spectrum of an optimized film at various angles of the
applied microwave and DC fields, rotating from in-plane (90) to out-of-plane (0) at 300
K with an applied microwave frequency of 9.39 GHz. The in-plane resonance field is
shifted from the spectrum shown in Fig. 4.4, which can be attributed to slightly different
73

cavity conditions and sample-to-sample variation. The only variation with angle is the
expected shift in resonance field predicted by the Kittel equation, assuming there is little
or no crystalline anisotropy and shape anisotropy is dominant, as is the case here. Figure
4.5(b) shows the extracted resonance field as a function of angle at 300 K. The expected
angular-dependent resonance condition for a thin film ferromagnet rotated from in-plane
to out-of-plane when H>>4Ms (as is the case here) is:

res

H 4 M

cos 2 H H 4 M s cos 2 H

(4.8)

where res is the resonance frequency in MHz, is the gyromagnetic ratio 2 x 2.8
MHz/Oe, Ms is the saturation magnetization in emu/cm3, H is the applied external field in
Oe and H is the angle of the applied external field with respect to the normal of the film
plane. The black line in Fig. 4.4(b) shows the expected angular dependence for 4Ms = 95
G (7.56 emu/cm3), which fits the angular data well and is consistent with SQUID data
within our ability to estimate the volume of the sample. This validates the assumption that
the magnetization angle exactly follows the external field angle since the shape anisotropy
field 4Ms is much smaller than the applied field. This angular variation is weak compared
to other inorganic ferromagnets as a result of the lack of crystalline anisotropy and the low
Ms (for comparison, thin film YIG has a saturation magnetization of 1750 G at room
temperature).

74

Figure 4.5 (a) Angular dependence of FMR spectrum rotating from 90 (field in-plane) to 0 (field out-ofplane) at 300 K. (b) Extracted resonance fields from angular dependence. Black line is the expected behavior
of a thin film ferromagnet with a saturation magnetization of 95 G.

We also measure the FMR response while rotating the magnetic field in the sample
plane. Previous SQUID magnetometry measurements have indicated that there is no
detectable magnetic anisotropy in V[TCNE]x, with no change seen in the coercivity or
saturation magnetization whether the magnetic field was applied in-plane or out-of-plane.
This can be explained by the structural order in the material being strong at the local level
but weak at the global level, leading to a lack of crystal structure and a corresponding lack
of crystalline anisotropy.
75

Figure 4.6 (a) FMR spectrum with field in-plane at 300 K. (b) Resonance field (black) and linewidth (red) as
a function of in-plane rotation. There is little systematic variation with angle, indicating the lack of anisotropy
in V[TCNE]x.

A single in-plane FMR spectrum of a V[TCNE]x~2 thin film is shown in Fig. 4.6(a).
The spectrum is single peaked and very clean, as expected from previous measurements.
The linewidth is slightly broader than the narrowest lines we have measured, which could
be due to sample-to-sample variation. Figure 4.6(b) shows the angular dependence of the
76

resonance field (black, left axis) and linewidth (red, right axis) of a V[TCNE]x~2 film, with
magnetic field rotated in-plane. Since V[TCNE]x~2 does not have crystalline anisotropy,
the absolute value angle is arbitrary, and so the angle used measures relative rotation. As
expected, the resonance field and linewidth do not change significantly, though both appear
to drift to higher values (since scans were taken in order from 0 to 180, this is also drift
with laboratory time). The jump in resonance field around 110 is due to the cavity
retuning, which changed the frequency of the applied microwave radiation. Figure 4.7
shows the resonance field as a function of angle when rotated purely in-plane (black)
compared with rotating in-plane to out-of-plane (blue).

Figure 4.7 Comparison of resonance field as a function of in-plane rotation (black) versus out-of-plane
rotation (blue).
77

4.4.4 Ferromagnetic resonance in other organic-based magnets


In addition to V[TCNE]x~2, we can fabricate a number of related materials that
follow the general pattern of M[Acceptor], where M is a metallic or magnetic ion.
Switching the magnetic ion can generate different types of magnetic ordering [45]. On the
other hand, substituting the organic ligand can also affect the magnetism more subtly,
chiefly by changing the physical structure of the material and thereby changing the
magnetic structure [46]. This can be achieved by taking advantage of various organic
chemistry techniques, and we have successfully generated solution-based and thin films of
two such materials, V[MeTCEC]x (methyl tricyanoethylene carboxylate) and V[ETCEC]x
(ethyl tricyanoethylene carboxylate), by replacing a cyano group on the TCNE molecule
with a methyl group and an ethyl group respectively. Studying these materials can shed
light on the relationship between physical structure and other emergent properties, such as
magnetism.

4.4.4.1 Ferromagnetic resonance of V[MeTCEC]x


Vanadium methyl tricyanoethylene carboxylate uses MeTCEC in place of TCNE
as the organic ligand. The MeTCEC molecule is shown in Fig. 4.8, and consists of a TCNE
molecule with one cyano group replaced by a methyl group. In V[TCNE]x, the nitrogen at
the end of each TCNE leg can coordinate with a vanadium ion, as described in section 2.3.
By replacing a cyano group with a methyl group, that leg can no longer coordinate with
the magnetic ion and provide an exchange pathway due to the lack of an appropriate
nitrogen atom. Furthermore, the bulkier molecule is not as free to rotate, limiting the range
of vanadium ions in reach. This has the effect of decreasing the overall strength of the
78

magnetism, lowering the Curie temperature and potentially tipping the scale toward the
sperimagnetic phase.
Figure 4.8 shows the temperature dependence of the magnetization of a
V[MeTCEC]x thin film. We observe a Curie temperature of approximately 317 K extracted
from the Bloch Law fit (eq 2.1), slightly higher than room temperature. Similar to
V[TCNE]x~2, the magnetization peaks at an intermediate temperature, in this case 190 K.
However, the peak is sharp and the drop in magnetization at low temperature is close to
linear, in contrast to V[TCNE]x~2, which is consistent with the disorder-induced random
magnetic anisotropy observed in other organic-based magnets. Figure 4.9 shows the field
dependence of the magnetization, at 5 K and 300 K. At 5 K the hysteresis loop is slightly
more rounded than at 300 K, but the coercive field is unchanged. The inset to Fig. 4.7b
shows the low field behavior, which shows that the coercivity is nonzero (in contrast, the
solution preparation of V[MeTCEC]x shows no coercivity down to 1 Oe), making
V[MeTCEC]x a soft magnet. Based on this, we speculate that the material transitions from
paramagnetism to ferrimagnetism around 300 K, followed by a second transition to
sperimagnetism at the peak in the magnetization, i.e. the freezing temperature.

79

Figure 4.8 Temperature dependence of the magnetization of a V[MeTCEC]x~2 thin film. Curie temperature is
approximately 317 K. Inset: MeTCEC molecular diagram.

Figure 4.9 Hysteresis loop of a V[MeTCEC]x~2 at 5 K (pink) and 300 K (blue). Inset: Zoomed in view around
zero field to show the coercive field.

80

Figure 4.10 shows the FMR response of a V[MeTCEC]x~2 thin film at 10 K. At


room temperature, there is no FMR peak, which is another piece of evidence that it is the
magnetic order in these materials that produces the magnetic resonance characteristics. The
linewidth is broader compared to V[TCNE]x, at 15 G. Ideally, we would track the FMR
properties with the temperature as the most potentially interesting region is around the
magnetization peak, but stabilizing at these intermediate temperatures is difficult with the
cavity. Nevertheless, we have found that the MeTCEC substitution still produces a clean
FMR spectrum.

Figure 4.10 In-plane FMR spectrum of a V[MeTCEC]x~2 thin film at 10 K with applied microwave frequency
of 9.383 GHz.

81

4.4.4.2 Ferromagnetic resonance of V[ETCEC]x


Another organic-based thin film magnet is V[ETCEC]x~2 (vanadium ethyl
tricyanoethylene carboxylate). The molecular structure of ETCEC is shown in Fig. 4.11(a),
and consists of a TCNE molecule with one cyano group replaced by an ethyl group. ETCEC
is similar to MeTCEC, but slightly bulkier because the ethyl group is larger than the methyl
group, and therefore should weaken the magnetic properties further. This is borne out by
the DC magnetometry, shown in Figs. 4.11(b) and 4.11(c). The magnetization again peaks
at an intermediate temperature, around 90 K (depending on the applied field), similar to
both V[TCNE]x~2 and V[MeTCEC]x. The Curie temperature is reduced to 161 K, and both
the field-cooled and zero-field-cooled magnetizations rise and fall sharply around the peak.
The hysteresis loop is again similar in shape to the other organic-based magnets we have
studied, with small coercive fields and relatively sharp switching. Compared to
V[MeTCEC]x, it appears that the M vs T behavior of V[ETCEC]x~2 rises more sharply upon
cooling past the Curie temperature. Furthermore, the rise to the magnetization peak in
V[MeTCEC]x~2 appears to have two distinct phases with different slopes, as opposed to the
single slope of V[ETCEC]x. We attribute this to the ETCEC ligand decreasing the strength
of the magnetism to the point where below the Curie temperature the sperimagnetic phase

82

Figure 4.11 (a) Chemical structure of an ethyl tricyanoethylene carboxylate (ETCEC) molecule. (b) Fieldcooled and zero-field-cooled M vs T of V[ETCEC]x~2 thin film at various applied fields. Inset: Bloch law fit
of saturation magnetization, yielding a Curie temperature of 161 K. (c) M vs H at 5 K (blue), 50 K (red), and
100 K (black). Inset: Zoom in at zero field to show coercivity.

Figure 4.12(a) shows the FMR response of a V[ETCEC]x~2 thin film at 5 K. As with
V[MeTCEC]x~2 thin films, the linewidth is broader compared to V[TCNE]x, at 6 G. Figure
83

4.12(b) shows the angular dependence of the FMR spectra. It is notable that the resonance
field only shifts a few Gauss at this temperature - in general, we would associate a peak
that does not shift with the direction of the magnetic field with EPR because that is the
response of non-interacting spins. However, low temperature SQUID measurements have
indicated that the magnetization in this material drops sharply below the magnetization
peak, which is also consistent with this angular behavior. Nevertheless, we have found that
the ETCEC substitution also still produces a clean FMR spectrum despite the weakened
magnetism as reflected by the lowered Curie temperature. Furthermore, the angular
dependence of the resonance field is broadly consistent with the expected behavior of a
thin film ferromagnet with a very low saturation magnetization, which agrees with DC
magnetometry.

84

Figure 4.12 (a) In-plane FMR spectrum of a V[ETCEC]x~2 thin film at 5 K with applied microwave frequency
of 9.388 GHz. (b) FMR spectra as a function of angle. Peak only shifts by a few Gauss, indicating a small
saturating magnetization, consistent with SQUID magnetometry.

4.5 Stripline ferromagnetic resonance


4.5.1 Principles of operation
Measurements in this section were taken using a microstrip geometry rather than a
microwave cavity. While the cavity offers greater measurement resolution, it can only
apply microwave excitations at a single frequency and has limited capability to measure at
85

temperatures below 300 K. We can therefore complement cavity measurements with


measurements utilizing a microstrip, as shown in Fig. 4.13(a). This has the advantage of
being able to tune the frequency of the applied microwave field, as well as allowing the
direction of the microwave field to be rotated with respect to the static field. In our case,
we can also use different measurement systems that can access higher magnetic fields and
more precisely control the temperature, which gives us additional information about the
magnetic resonance characteristics.
In general, microstrips consist of a conductor that carries the current that generates
the oscillating magnetic field, a dielectric, and a ground plane. This is shown in cross
section in Fig. 4.13(a), along with a top view that shows the geometry of our particular
microstrip. This geometry is designed so that the sample will sit as indicated and experience
a microwave field along the z-axis with the static magnetic field in the xy-plane (axes can
be seen in Fig. 4.13(b)). The microwave field generated by the stripline is shown in Fig.
4.13(b), as simulated by HFSS, a finite element solver for electromagnetic structures. The
size and color of the arrows indicate the strength of the field. The strength of the field dies
off rapidly above the conductor, with little to no field further than 100 m away.

86

Figure 4.13 (a) Top view and cross sectional view of microwave stripline. (b) HFSS simulation of magnetic
field generated by the stripline. Arrows indicate direction of the field, and size and color of arrows indicate
the field strength.

The constraints of our measurement setup dictate that we utilize a reflection


geometry, as shown in Fig. 4.14 - the output power is shown with the red arrows, and the
reflected power is shown with blue arrows. Microwaves are continuously generated by a
variable frequency microwave source and pass through a circulator to the microstrip (an
isolator is included to prevent any current from returning to the source and potentially
damaging it). The circulator is a piece of equipment that directs microwave radiation from
one port to the next - microwaves that enter port 1 exit via port 2, enter port 2 exit via port
3, and enter port 3 exit via port 1. The microwaves pass through a launcher to the
microstrip, then are reflected at the end of the microstrip and return to the circulator. The
87

circulator directs the returned microwaves to the detection equipment - we most often used
a microwave diode that converts microwave radiation into a DC voltage, which is then read
by a Keithley 2400 sourcemeter. Alternatively, we can combine the microwave generation
and detection with a vector network analyzer, which works as a pulse measurement rather
than a continuous wave. In this scheme, we are measuring the reflected power, and so we
expect a Lorentzian peak rather than the Lorentzian derivative we see in the cavity. The
disadvantage of the reflection geometry is that many things in the microwave circuit can
generate reflections, which will all be detected and contribute to noise in the measurement.

88

Figure 4.14 Schematic of measurement setup for stripline magnetic resonance. Microwaves are generated by
a source (red) and passed through an isolator to a circulator and then the stripline. A sample is mounted as
shown. When the microwaves reach the end of the stripline, they are reflected (blue) and return to the
circulator, where they are directed to the detector. For some measurements, this is a microwave diode that
converts microwave power to DC voltage, which can then be detected by a voltmeter. A static field must also
be generated, but is not shown in this diagram.

4.5.2 Air-free magnetic resonance can


As mentioned in chapter 2, V[TCNE]x~2 is highly air sensitive, necessitating the
development of specialized air-free cans to protect samples during transport and
measurement. The can designed for microwave measurements is shown in Fig. 4.15, with
89

the inset showing the can assembled. The cap and the body of the can are made of
nonmagnetic stainless steel to accommodate the microwave feedthrough, which is laser
welded in place. The cap also contains several DC electrical feedthroughs for electrical
measurements. The can seals with indium wire, which is discarded and reapplied every
time, and is marked to gauge the quality of the seal and to index the sample to the external
magnetic field. This can is compatible with a number of systems, including an Oxford
Spectromag, which allows us to access static magnetic fields up to 8 T and temperatures
ranging from 300 K to 1.2 K with a high degree of field and temperature resolution.

Figure 4.15 CAD model of air-free can for stripline FMR. Seal is formed with indium wire that is placed on
the rim of the can. Inset: air-free can fully assembled.

90

4.5.3 Frequency dependence


By varying the frequency of the applied microwave radiation, we can observe the
changing magnetic resonance response and extract a number of properties. The linewidth
of the FMR peak can tell us about the relaxation mechanisms in the material. The resonance
field allows us to extract a saturation magnetization and gyromagnetic ratio.

Figure 4.16 FMR spectrum at various frequencies from 3 to 13 GHz at 50 K, measured in an Oxford
Spectromag. Data is normalized to the initial point in the scan to compare across frequencies, as the stripline
itself has a frequency response.

Figure 4.16 shows various FMR scans taken from 3 to 13 GHz using a vector
network analyzer for detection. The y-axis shows the raw voltage generated by the diode.
Since this is DC rather than AC detection, we expect the peak to be a Lorentzian rather
than the Lorentzian derivative seen in the cavity FMR shown in section 4.4. We also see a
larger background, more drift with laboratory time, and a higher noise level than in the
cavity. This is all due to the inherent drawbacks of this particular measurement system, as
91

described above. The FMR peak is not always as prominent due to the frequency response
of the stripline and overall microwave circuit independent of the sample.
We fit the peak to a Lorentzian to extract the resonance field and the full width-half
maximum (FWHM) linewidth. Figure 4.17(a) shows the extracted resonance fields at
different applied microwave frequencies. The resonance condition for in-plane FMR is
2

2 K eff

H H 4 M s
Ms

(4.9)

Where /2 is the microwave frequency, is the gyromagnetic ratio, H is the magnetic


field, Ms is the saturation magnetization, and Keff is the effective anisotropy energy. The fit
is shown in red in Fig. 4.17(a). The extracted saturation magnetization of 49.1 G is broadly
consistent with SQUID magnetometry, as was the case with the angular dependence (Fig.
4.7). The electron gyromagnetic ratio from the fit is 26735.33 MHz/T, compared to the
literature value for a free electron of 28024.95 MHz/T [84]. The deviation from the free
electron gyromagnetic ratio could be an indication of electron interactions within the
material.
As described in section 4.1, there are several mechanisms that can contribute to the
linewidth, including Gilbert damping, mosaicity, two-magnon scattering, and
inhomogenous broadening [29]. Gilbert damping leads to the linewidth increasing linearly
with increasing frequency:
B Gilbert

92

2
3 cos

(4.10)

Where is the angle between the magnetization M and the external field B. Two-magnon
scattering will contribute a nonlinear term, appearing as deviation from linearity at low
frequency:

B 2 mag xi f (B xi ) arcsin

xi

2 (0 2) 2 0 2
U ( eq c )
2 (0 2)2 0 2

(4.11)

Where xi is the strength of two-magnon scattering along a crystal axis

xi , and

0 0 M eff , the effective magnetization that takes anisotropy into account. The function
f (B xi ) allows for the two-magnon contribution to vary with the in-plane direction of
B if scattering centers are not distributed uniformly in the sample, and U ( eq c ) is a step

function used to describe the switching off of two-magnon scattering at a critical out-ofplane angle. Mosaicity and inhomogenous broadening both do not contribute to the
frequency dependence of the linewidth.
The linewidth of the FMR peak as a function of frequency at 50 K is shown in Fig.
4.17(b). Due to the lack of crystalline structure, and therefore crystalline anisotropy, we
only measure in one orientation. The linewidth does not appear to vary linearly with
frequency, which makes it difficult to extract , the Gilbert damping parameter, and also
obscures any curvature that would indicate two-magnon scattering. We do not have all the
material parameters necessary to fit for , the strength of the two-magnon scattering
process. We can compare to the lowest damping observed in an inorganic system, YIG, at

2.3 104 , shown as the red line in Fig. 4.15(b). The linewidth in V[TCNE]x~2 is

93

broadly comparable to that in YIG, following the pattern seen in the cavity FMR
measurements.

Figure 4.17 (a) Resonance field as a function of applied microwave frequency. Data is plotted with the center
field as the x-axis to use the displayed equation for the curve fit, shown in red. Fit yields a saturation
magnetization of 49.1 G, broadly consistent with SQUID data and FMR angular dependence. (b) Linewidth
as a function of applied microwave frequency. Red line shows the expected behavior for Gilbert damping
with 2.3 10 .
4

4.5.4 Temperature dependence of FMR in V[TCNE]x


We use the measurement setup described above to explore the temperature
dependence of the magnetic resonance characteristics of V[TCNE]x. For this measurement,
we utilize an Oxford Spectromag to precisely control the temperature of the system - while
we do not have direct thermometry on the sample probe itself, there are a number of
temperature probes in the system, and the system can stabilize to within ~0.1 K from 5 K
to 300 K.
94

Figure 4.18(a) shows FMR spectra of a V[TCNE]x~2 thin film measured at 9 GHz
from 50 K to 300 K, with 50 K steps between points. We see relatively little change in the
magnetic resonance as the temperature changes. The resonance field shifts by a few Gauss,
but the linewidth remains sharp (except for an anomalous point at 250 K). Figure 4.18(b)
shows the extracted resonance field (black) and linewidth (red) as a function of
temperature. From M vs H measurements, the static field is large enough that we expect
the magnetization is close to saturated, and so we do not see any effects from the different
magnetic mechanisms. All magnetic anisotropy constants are expected to be temperaturedependent and connected to the temperature dependence of the magnetization, but this is
not well understood in thin magnetic films[85]. In V[TCNE]x~2, this is further complicated
by the lack of both crystalline and magnetic anisotropy, as described in chapter 2. It is
possible that as the decreasing temperature introduces a sperimagnetic phase,
inhomogenous broadening is strengthened, which is a potential area for further
investigation. However, we therefore cannot draw many conclusions about the system from
this data, other than we do not see anything unexpected.

95

Figure 4.18 (a) FMR spectra as a function of temperature from 50 K to 300 K. Center field shifts by a few
Gauss to higher field as temperature decreases. (b) Extracted center field (black) and linewidth (red) as a
function of temperature.

96

4.6 Conclusion

Since its discovery, magnetic resonance has been used to study a variety of
fundamental properties and has found a number of applications. We have measured the
magnetic resonance characteristics of V[TCNE]x~2 and related materials in a variety of
ways, using both cavity and stripline geometries and as a function of the applied field angle,
microwave frequency, and temperature. These measurements have shown that V[TCNE]x~2
has a FMR linewidth comparable to the narrowest line seen in inorganic systems, YIG.
Frequency dependent measurements indicate that the Gilbert damping parameter is
correspondingly low, though further study is necessary to fully characterize the relaxation
mechanisms in this material. Finally, FMR measurements of related organic-based magnets
involving different organic ligands, MeTCEC and ETCEC, see broader FMR lines, though
they are still narrow in comparison to most inorganic ferromagnets. This supports the
model of magnetic interactions described in chapter 2, as the larger functional groups
interfere with the carrier-mediated exchange pathway. Overall, these measurements point
to a potential role for these thin film organic-based magnets to complement YIG for high
frequency electronics and lay the foundation for further study of spin dynamics in these
systems.

97

References

1.

Wolf, S.A., et al., Spintronics: A spin-based electronics vision for the future.
Science, 2001. 294(5546): p. 1488-1495.

9.

Pardavi-Horvath, M., Microwave applications of soft ferrites. Journal of


Magnetism and Magnetic Materials, 2000. 215: p. 171-183.

13.

Ando, K., et al., Angular dependence of inverse spin-Hall effect induced by spin
pumping investigated in a Ni(81)Fe(19)/Pt thin film. Physical Review B, 2008.
78(1).

14.

Inoue, H.Y., et al., Detection of pure inverse spin-Hall effect induced by spin
pumping at various excitation. Journal of Applied Physics, 2007. 102(8).

15.

Saitoh, E., et al., Conversion of spin current into charge current at room
temperature: Inverse spin-Hall effect. Applied Physics Letters, 2006. 88(18).

18.

Tang, C.W. and S.A. Vanslyke, ORGANIC ELECTROLUMINESCENT DIODES.


Applied Physics Letters, 1987. 51(12): p. 913-915.

19.

Brabec, C.J., Organic photovoltaics: technology and market. Solar Energy


Materials and Solar Cells, 2004. 83(2-3): p. 273-292.

21.

Manriquez, J.M., et al., A ROOM-TEMPERATURE MOLECULAR ORGANIC


BASED MAGNET. Science, 1991. 252(5011): p. 1415-1417.

22.

Pokhodnya, K.I., A.J. Epstein, and J.S. Miller, Thin-film V TCNE (x) magnets.
Advanced Materials, 2000. 12(6): p. 410-+.

23.

de Caro, D., et al., CVD-grown thin films of molecule-based magnets. Chemistry of


Materials, 2000. 12(3): p. 587-+.

25.

Forrest, S.R., The path to ubiquitous and low-cost organic electronic appliances on
plastic. Nature, 2004. 428(6986): p. 911-918.
98

29.

Zakeri, K., et al., Spin dynamics in ferromagnets: Gilbert damping and two-magnon
scattering. Physical Review B, 2007. 76(10).

30.

Heinrich, B., et al., Dynamic exchange coupling in magnetic bilayers. Physical


Review Letters, 2003. 90(18).

42.

Crone, B., et al., Large-scale complementary integrated circuits based on organic


transistors. Nature, 2000. 403(6769): p. 521-523.

45.

Kao, C.Y., et al., Thin films of organic-based magnetic materials of vanadium and
cobalt tetracyanoethylene by molecular layer deposition. Journal of Materials
Chemistry C, 2014. 2(30): p. 6171-6176.

46.

Lu, Y., et al., Thin-Film Deposition of an Organic Magnet Based on Vanadium


Methyl Tricyanoethylenecarboxylate. Advanced Materials, 2014. 26(45): p. 76327636.

47.

Sun, Y.Y., Y.Y. Song, and M.Z. Wu, Growth and ferromagnetic resonance of
yttrium iron garnet thin films on metals. Applied Physics Letters, 2012. 101(8).

65.

Plachy, R., et al., Ferrimagnetic resonance in films of vanadium


tetracyanoethanide (x), grown by chemical vapor deposition. Physical Review B,
2004. 70(6).

76.

Adelmann, C., et al., Spin injection and relaxation in ferromagnet-semiconductor


heterostructures. Physical Review B, 2005. 71(12).

78.

Urban, R., G. Woltersdorf, and B. Heinrich, Gilbert damping in single and


multilayer ultrathin films: Role of interfaces in nonlocal spin dynamics. Physical
Review Letters, 2001. 87(21).

79.

Arias, R. and D.L. Mills, Extrinsic contributions to the ferromagnetic resonance


response of ultrathin films. Physical Review B, 1999. 60(10): p. 7395-7409.

80.

Heinrich, B. and J.F. Cochran, ULTRATHIN METALLIC MAGNETIC-FILMS MAGNETIC ANISOTROPIES AND EXCHANGE INTERACTIONS. Advances in
Physics, 1993. 42(5): p. 523-639.
99

81.

Fert, A., Nobel Lecture: Origin, development, and future of spintronics. Reviews
of Modern Physics, 2008. 80(4): p. 1517-1530.

82.

Sun, Y., et al., Growth and ferromagnetic resonance properties of nanometer-thick


yttrium iron garnet films. Applied Physics Letters, 2012. 101(15).

83.

Gilbert, T.L., A phenomenological theory of damping in ferromagnetic materials.


Ieee Transactions on Magnetics, 2004. 40(6): p. 3443-3449.

84.

NIST. Available from: http://physics.nist.gov/cgi-bin/cuu/Value?gammaebar.

85.

Farle, M., Ferromagnetic resonance of ultrathin metallic layers. Reports on


Progress in Physics, 1998. 61(7): p. 755-826.

100

Chapter 5 Spin pumping in thin film organic magnetic bilayers

Equation Chapter (Next) Section 1


5.1 Introduction

Spin angular momentum is an intrinsic property of elementary and composite


particles that gives rise to magnetism and provides an additional degree of freedom to
manipulate. Recent work in spintronics has been focused on the active use of the spin of
electrons to perform operations. Electrical spin injection is the most successful method of
generating spin currents, with a major impact in hard drives in the consumer electronics
industry. This involves passing an electrical current through a ferromagnet, which polarizes
the carriers, and has been demonstrated in geometries like spin valves [86], magnetic tunnel
junctions [87], and spin LEDs [88]. However, direct electrical spin injection necessarily
entangles spin and charge motion and generates heat, and the spin polarizations that can be
generated are limited by the density of states in the material system. Therefore, there have
been considerable efforts to find alternative ways to generate, manipulate, and detect spin
and spin currents [31]. These include thermal spintronics, using thermal gradients to
generate and transport spin [89]; magnonics, engineering materials such that only certain
spin wave modes are allowed [90]; valleytronics, manipulating the band structure of
semiconducting materials [91]; and topological insulators, materials where certain surface
states exhibit spin-momentum locking [92].
101

Another method that has emerged is ferromagnetic resonance (FMR) spin pumping.
When a magnet is undergoing FMR, the magnetization is canted away from its equilibrium
position, and so the projection of the magnetization along that direction is reduced. Due to
conservation of angular momentum, this must be dispersed as spin motion, which can be
detected in a number of ways. By attaching another material to the magnet in question and
forming a bilayer, a spin imbalance is created at the interface between the two layers, and
a pure spin current is injected from the magnet. While a spin current is simply net spin
motion, a pure spin current is a special case where the spin motion has no corresponding
charge current. In this case, while charge does move across the interface, it is not driven
by any applied voltage, and for every charge that moves from material A to B another
charge moves from B to A. If the second material in the bilayer is nonmagnetic, the spin
current can be detected via the inverse spin Hall effect, which converts the spin current into
a DC voltage [15]. This has been observed in materials like platinum and tantalum [13,
37]. Another method is to use a second magnetic layer and excite FMR in both layers [30,
93]. By rotating the direction of the applied field, the two FMR peaks will evolve
differently with angle, and at some point they will cross. Around this point, spins from one
magnet will be pumped into the other, which will change the linewidth of the FMR peak.
Through these spin pumping studies, it has been observed that the FMR linewidth
is an important metric for spin injection. In that light, we have measured extremely narrow
linewidths in V[TCNE]x, on par with the narrowest lines measured in inorganic systems. It
is therefore a natural fit to evaluate V[TCNE]x~2 as a source for spin pumping, which has
not been reported in any organic-based magnet. This chapter describes the preliminary
work we have done to measure spin pumping in V[TCNE]x. First, we describe past
102

measurements in inorganic systems and outline the underlying theory. Second, we fabricate
both hybrid organic-inorganic samples incorporating V[TCNE]x~2 and YIG and measure
the magnetic resonance characteristics. Third, we measure all-organic structures using two
V[TCNE]x~2 layers with differing thicknesses. While we have not yet obtained a complete
characterization, we believe the data in the hybrid system show clear signatures of spin
pumping.

5.2 FMR detection of dynamic exchange coupling in magnetic bilayers

In static spintronics devices, one of the primary magnetic interactions is the


Ruderman-Kittel-Kasuya-Yoshida (RKKY) interlayer exchange between magnetic layers
in heterostructures [94]. RKKY refers to a coupling mechanism of localized inner d or f
shell spins in a metal interacting through the conduction electrons, and has been integral to
the understanding of giant magnetoresistance (GMR). RKKY theory accurately predicts
how the coupling between magnetic layers separated by nonmagnetic spacers was observed
to oscillate between ferromagnetic and antiferromagnetic depending on the thickness of the
spacer [95, 96]. This is a static interaction, which can be suppressed by a thick spacer layer
or a tunnel barrier. In contrast, in the regime where there is no observed static interaction
between magnetic layers there is still coupling between magnetizations, which can be
explained through the emission and absorption of spin currents. This dynamic coupling has
attracted a high degree of interest as it is qualitatively quite different from static coupling,
and potentially has a large effect on relaxation and switching behavior in magnetic
multilayer structures.

103

One method of studying the dynamic processes is with magnetic resonance. FMR
is a technique in which the magnetization of a ferromagnet can be resonantly excited by
the application of a combination of microwave radiation and static magnetic fields. This is
a quasi-static state, as the magnetization is kept in a precessing state around the static field,
with its motion described by the Landau-Lifshitz-Gilbert equation:
dM
dM
M H eff M
dt
dt

(5.1)

This is generally measured by looking at the absorbed microwave power of the system,
yielding information about the magnetic anisotropies, exchange interactions, and
relaxation processes in the material. FMR is discussed in more detail in chapter 4.
By putting two magnetic layers in contact, new and interesting effects arise. In
inorganic systems, a nonmagnetic spacer is placed between the magnetic layers to suppress
the RKKY exchange. Spin can be carried by a driven charge current between the layers,
which in turn exerts a torque on the magnetizations, and at sufficiently high current
densities this process can drive magnetization dynamics [31]. However, even in the
absence of a driving voltage, spins are injected into the normal material by a ferromagnet
with a magnetization in motion [39]. Assuming the spin-flip scattering in the normal
material is small, this spin current will be absorbed at the interface of the other magnetic
layer, exerting a torque on the receiving magnets magnetization. Each magnet is therefore
acting as a spin sink for the spin current generated by its counterpart. Under special
conditions, the two magnetizations are resonantly coupled by spin currents and their motion
is synchronized, like coupled pendulums.

104

Figure 5.1 Cartoon of spin pumping process. On the left, layer F1 is in resonance and a spin current is pumped
into the spacer, while layer F2 is stationary. On the right, both films are in resonance, creating equal and
opposite spin currents. The shaded regions around the interfaces represent where the spin current is absorbed
[30].

The theory behind this measurement rests upon a few assumptions. A precessing
magnetization m i M i / M i pumps a spin current I sipump m i into the normal material (i, j
are indices that denote the magnetic layers). The magnetic bilayers are weakly excited close
to the parallel alignment, so the spin current I sipump m j for arbitrary i, j 1, 2 . Spin
momentum perpendicular to the magnetization direction cannot penetrate the magnetic
layer beyond the spin coherence length, sc. The magnetization of each layer is constant
across its thickness, which is larger than sc, and therefore the transverse spin current is
completely absorbed. The spin current generated by the precessing magnetization is [39]
I sipump

dm i

g mi
4
dt

(5.2)

Where g is the spin mixing conductance, a dimensionless tensor that describes the
efficiency of spin transfer across an interface (in this case, between the ferromagnet and
the normal material) [97]. Assuming spin-flip scattering in the normal material is small,
the spin current entering the spacer exits the normal metal by equal outgoing spin currents
105

to the left and right [98]. On the time scale of FMR processes, this happens close to
instantaneously, and so the torque on each interface is
1 I spump
I spump
2 2
2
1

(5.3)

When one of the magnetizations is stationary, as shown on the left in Fig. 5.1, the time
evolution of the other one is governed by the LLG equation with an enhanced Gilbert
damping parameter. The modified damping parameter is i i(0) i , where i(0) is the
intrinsic value and i g 8i , where i is the total magnetic moment of the
precessing magnet. The magnetization scales linearly with the volume of the magnetic
layer and g scales with interface area, and so the enhanced damping parameter is
inversely proportional to film thickness.
If both magnetizations are in motion, as shown on the right of Fig. 5.1, the LLG
can be expanded to take the additional spin torque into account:

dm j

dmi
dmi
dmi
mi H ieff i(0)mi
i mi
mj

dt
dt
dt
dt

(5.4)

Where j 1 if i 2 and vice versa. Assuming magnetization deviations from the


equilibrium position are small and that the precession is circular, it can be shown that the
average magnetization deviation is damped with the intrinsic Gilbert parameter i(0) , while
the difference u u1 u 2 relaxes with an enhanced Gilbert parameter (0) 1 2
[30]. To measure the spin torques, the ability to independently excite FMR in the two
magnetic layers is required, so that they can be measured in isolation where the intrinsic
Gilbert damping parameter dominates.
106

Figure 5.2 Resonance field as a function of the angle of the static magnetic field with respect to the Fe
[100] crystal axis of the F1 (orange) and F2 (green) layers. Upper left inset: Cartoon of how the microwave
field h drives the magnetization precession (red). Upper right inset: The measured absorption peaks of F1
and F2 at 60 [30].

Figure 5.2 shows the FMR characterization of one such bilayer sample, consisting
of two Fe layers of different thicknesses separated by a thin Au spacer, as the field is rotated
in the plane of the sample. The F1 layer is 16 monolayers thick, while the F2 layer is 40
monolayers thick. The top right inset shows the measured absorption peaks at 60 ,
where the FMR peaks are far apart and can be excited independently. The main Fig. shows
the resonance field of F1 and F2 in orange and green respectively as the direction of the
field is rotated. At most angles the peaks are well separated, but in the shaded region the

107

resonance fields cross over, and the dynamic exchange coupling leads to collective
behavior of the magnetic layers.

Figure 5.3 FMR characteristics of a magnetic bilayer. Left and right frames show the FMR signals at
resonance field differences of -78 Oe and +161 Oe respectively. Center shows the linewidths of F1 (orange)
and F2 (green) as functions of the resonance field difference, along with theory (blue solid line) [30].

The resonance characteristics in the shaded region are further explored in Fig. 5.3.
The center frame shows the linewidth of F1 and F2 extracted from a two Lorentzian fit as
a function of the difference between the resonance fields of F1 and F2 (so at zero, the
magnets are in resonance at the same field). The left and right frames show the FMR spectra
at H 2 H1 of -78 Oe and +168 Oe, where the two peaks are clearly visible. When the
resonance fields are identical, the precessing components of F1 and F2 are parallel to each
other, and therefore the total spin current across the interfaces vanish, resulting in no excess
damping. This is consistent with the data shown in Fig. 5.3, where the linewidths are at a
108

minimum where H 2 H1 0 . The solid lines show theoretical predictions generated by


solving equation 5.4, and it matches the data well. Therefore, both theory and experiment
show evidence of this dynamic spin pumping process, and it has been observed in a number
of systems.

5.3 Spin pumping in hybrid organic-inorganic bilayers

While spin pumping has been measured in inorganic systems, it has not been
reported in organic systems. However, given the promising FMR characteristics of
V[TCNE]x, incorporating it into a magnetic bilayer and looking for spin pumping is a
promising direction.
For this measurement, the magnetic bilayer consists of an yttrium iron garnet (YIG,
Y3Fe5O12) layer and a V[TCNE]x~2 layer. The YIG is 20 nm thick and is grown by
sputtering on a (111)-oriented gadolinium gallium garnet (GGG, Gd3Ga5O12) substrate.
The GGG substrate allows for lattice-matched growth and yields the highest quality YIG
films [16]. As described in the previous section, thinner films generate a larger change in
the Gilbert damping, and we therefore deposit a thinner layer of V[TCNE]x, on the order
of 50 nm directly on the YIG (we are unable to measure the thickness of the deposition
directly, and so this is an estimate based on previous depositions). We did not include the
spacer layer as the nature of the magnetic interactions between the organic and inorganic
layers is unknown, and adding an ex-situ spacer layer could interfere with spin transport
between the magnetic layers.
Measurements were taken in a Bruker ESR spectrometer (principles of operation
can be found in section 4.4.1), rotating the field from in-plane to out-of-plane. V[TCNE]x~2
109

has no magnetic anisotropy in the plane of the sample due to the lack of crystalline
structure, and this field geometry is needed to shift the resonance fields of both the
V[TCNE]x~2 and the YIG FMR peaks. All measurements in this chapter are taken at room
temperature. Figure 5.4 shows the FMR peak with the field in-plane (90) of (a) the YIG
and (b) the V[TCNE]x~2 layers. The YIG response is very well fit by the expected
Lorentzian behavior. The V[TCNE]x~2 FMR spectrum has additional satellite features and
a broad shoulder, unlike V[TCNE]x~2 deposited on sapphire, though the FMR feature in
V[TCNE]x~2 remains sharp with a peak-to-peak linewidth of 1.5 G. Previous magnetometry
measurements have indicated that properties such as the saturation magnetization, Curie
temperature, and coercivity are largely substrate-independent; however, the FMR
characteristics appear to be more sensitive to deposition conditions, and we are working on
optimizing the deposition process for future magnetic resonance experiments.

Figure 5.4 (a) FMR spectrum of YIG layer with field in-plane. (b) FMR spectrum of V[TCNE]x~2 layer with
field in-plane.

110

Figure 5.5 shows the FMR spectra at various angles for (a) the YIG layer and (b)
the V[TCNE]x~2 layer, in order to find where the peaks cross. The resonance field in both
layers behave as expected for a thin film ferromagnet, with the difference between the two
explained by the larger saturation magnetization and magnetic anisotropies in the YIG.
Again, the FMR characteristics of the V[TCNE]x~2 layer in the heterostructures appear to
be degraded compared to single layers as the field is rotated, the shape of the FMR peak
evolves and various additional peaks appear and fade away. This coarse angular
dependence suggests that the peak crossover should occur around 40.

Figure 5.5 (a) FMR spectrum of YIG layer at field angles from 90 to 30. (b) FMR spectrum of V[TCNE]x~2
layer at field angles from 90 to 0.

111

Figure 5.6 shows the resonance characteristics of the bilayer in the crossover
region. We focus on tracking the YIG peak as we are interested in the precise behavior of
the linewidth and resonance field, and so the V[TCNE]x~2 peak is not fit well enough to
extract those parameters with a high degree of confidence. Individual FMR spectra are
shown with field angles of 41, 44, and 48 in panels (a), (b), and (c) respectively. In (a) and
(c), the peaks are clearly separated, with the V[TCNE]x~2 peak at higher field in (a) and the
YIG peak at higher field in (c). When this is the case, we can fit the YIG peak and obtain
a resonance field and linewidth, and the parameters extracted from the broad angular
dependent measurements in Fig. 5.5(a) are shown in Fig. 5.6(d). However, for a range of
angles the two peaks overlap and it is difficult to disentangle the YIG response from the
V[TCNE]x~2 response, as shown in Fig. 5.6(b).

112

Figure 5.6 (a-c) FMR spectrum at 300 K of the magnetic bilayer at field angles of 48, 44, and 41 respectively.
All three have the same y scale. (a) and (c) show the two resonance peaks clearly separated, while (b) shows
the two peaks crossing over. (d) Extracted resonance field (black) and linewidth (red) of YIG layer. (e)
Linewidth extracted from the bilayer as a function of angle. Black data is the same as in (d), which shows
the behavior of the YIG peak in isolation. Red data shows the linewidth extracted from the YIG peak in the
region where it crosses the V[TCNE]x~2 peak. The linewidth deviates strongly from the monolayer behavior,
indicating the presence of spin pumping.

Figure 5.6(e) shows the extracted linewidth of the YIG peak in the crossover region,
with the broader behavior from the YIG peak in isolation in black and the detailed behavior
of the bilayer in red. When the peaks are separated, the linewidth of the YIG varies
113

smoothly with angle, similar to the expected behavior of a ferromagnet in isolation.


However, once the peaks begin to cross (between 43 and 47), the linewidth diverges
significantly from the isolated behavior. Ideally we would be able to track the YIG through
the entire crossing region, but we still observe a 6.2% (7.7%) deviation in the linewidth on
the low-to-high angle (high-to-low angle) side. This is suggestive of spin pumping between
the V[TCNE]x~2 and the YIG. The exact behavior of the linewidth does not agree with the
predicted behavior shown in Fig. 5.3, but this can be explained by the different rotational
geometry. However, while we cannot use a curve fit on the V[TCNE]x~2 peak, we can
measure the peak-to-peak linewidth, which does not show the same divergence as the YIG
peak, as shown in Fig. 5.7. It is possible that the V[TCNE]x~2 is affecting the YIG more
than vice versa, as the nature of the interface and the interactions there are unknown, but
for now it is unclear what the nature of the signal is.

Figure 5.7 Full-width half-maximum linewidth of the YIG peak (red) and peak-to-peak linewidth of the
V[TCNE]x~2 peak (blue) as a function of angle. While the YIG linewidth diverges around the peak crossover,
no similar behavior is observed in the V[TCNE]x~2.
114

5.4 Spin pumping in all-organic magnetic heterostructures

In addition to the hybrid magnetic bilayer, we also fabricated all-organic


heterostructures. The fundamental FMR work we have done with V[TCNE]x~2 and the
related materials V[MeTCEC]x~2 and V[ETCEC]x~2 gives us a catalogue of options for
these measurements. The spin pumping measurements described above requires the
resonance fields of the two magnetic layers involved evolve differently with field direction
and that they cross over at some angle. The first condition can be achieved by the choice
of material or by the thickness of the layer, while the second is not directly under our
control and a suitable system can be found through trial and error. The most successful
system we tried consists of two V[TCNE]x~2 layers of different thicknesses with a thin
aluminum spacer layer. Aluminum is chosen here to match the work function of
V[TCNE]x~2 to the best of our knowledge. Again, measurements are taken with a Bruker
ESR spectrometer at room temperature.

115

Figure 5.8 (a) FMR spectrum of the all-organic bilayer with external field at 90. (b) FMR spectrum of the
all-organic bilayer with external field at 0. (c) Angular dependence of the FMR spectrum.

Figure 5.8(a) and (b) show the FMR spectrum of the all-organic bilayer with the
external field applied at 90 and 0 respectively. It is clear, especially in (b), that there are
two FMR peaks, one that is broader and noisier than the other. In (a) the broader feature is
at higher field, while in (b) it is at lower field, and therefore they must cross at some point.
However, it is not clear which feature corresponds with which layer while in inorganic
ferromagnets thinner layers tend to have broader FMR peaks, we have not observed this in
V[TCNE]x. Furthermore, the thin layer is deposited after the thick layer, and while the
deposition temperature is low compared to most inorganic deposition processes, a sample
temperature of 30 C is hot enough to degrade the magnetic properties. It is plausible that
116

this degradation could broaden the FMR response. Neither of the peaks is well fit by a
Lorentzian, which also points to a potential area for improvement in the deposition. Figure
5.8(c) shows the full angular dependence of the FMR spectrum from in-plane to out-ofplane. Unlike the hybrid bilayer, the peaks here do not separate enough to consider them
in isolation.

Figure 5.9 (a) Peak position of the high field (blue) and low field (red) peaks in the FMR spectrum of the allorganic bilayer as a function of angle. (b) Peak separation (high field low field) as a function of angle. The
separation appears to hit a minimum value for a range of angles around the crossover angle. (c) Individual
FMR scans of the bilayer at 50, 55, and 60. Red and blue circles show the peak position extracted for (a).

The FMR response of the all-organic bilayer is summarized in Fig. 5.9. Since the
peaks do not neatly conform to a Lorentzian, we instead extract measurable parameters
low
from the scans, primarily peak position of the feature at low field H res
(red) and at high

117

high
field H res
(blue), shown in Fig. 5.9(a). Figure 5.9(c) shows individual FMR spectra at 50,

55, and 60, with red and blue circles corresponding to the peak positions extracted in Fig.
5.9(a). Since the peaks cross, the low field feature will be attached to the response from
one of the layers at low angle and the other at high angle, separated by the crossover angle.
high
low
The peak separation, defined as H res
, is shown as a function of angle in Fig. 5.9(b).
H res

The separation starts high at low angle, then gradually decreases as the peaks move closer
together, until it flattens while the peaks are crossing. This may be an indication that the
magnetizations are coupled and resonating together, as described in section 5.2 and shown
in Fig. 5.2, but the effect here is subtle and not as clear as the change in linewidth seen in
Fig. 5.6.

Figure 5.10 (a) Simulated FMR spectrum of two Lorentzians with linewidths of 2 and 8 with differing peak
separations. (b) High (blue) and low (red) field peak positions with peak crossover at 0. (c) Peak separation
as a peaks cross over. Qualitatively, this captures the minimum seen in Fig. 5.8, but appears more abrupt.
118

To determine whether the signal seen could be spin pumping, we do a simple


simulation of two non-interacting Lorentzians - if two non-interacting peaks can generate
the same behavior we see, then it is unlikely to be spin pumping. This is accomplished by
adding two Lorentzians with linewidths of 2 and 8, approximately what we measured in
the all-organic bilayer. Figure 5.10(a) shows the two Lorentzian spectrum as the peaks
cross over. This response does not resemble the experimental data, where the two peaks
can be resolved even when they are close to overlapping. Figure 5.10(b) shows the high
field (blue) and low field (red) peak positions and Fig. 5.10(c) shows the peak separation,
extracted by the same procedure as with the experimental data. Figure 5.10(c) does appear
to capture the peak separation reaching a minimum value and staying there directly around
the crossover point, but there is a sharp drop to the minimum value that does not agree with
the data. Conversely, if the linewidth of the two Lorentzians is reduced to 1 and 2, the
spectra more closely resemble those in Fig. 5.9(c), but the peak separation becomes a vshape and does not exhibit the same floor behavior. From this, it is difficult to conclude
whether spin pumping is being detected while most of the qualitative features of the data
reproduced by the simulations, no single simulation captures all of the features. While it is
inconclusive for now, this is also a work in progress, and we have laid the ground work for
further studies.

5.5 Conclusion

Ferromagnetic resonance-driven spin pumping is an exciting development in the


field of spintronics with the potential to efficiently generate pure spin currents. Spin
pumping has been observed in a number of inorganic magnetic systems, detected via the
119

inverse spin Hall effect or enhanced Gilbert damping in magnetic resonance measurements.
We have extended these studies to both hybrid organic-inorganic bilayers and all-organic
heterostructures with the hope of taking advantage of the extremely narrow FMR linewidth
in V[TCNE]x. Measurements of a hybrid YIG-V[TCNE]x~2 bilayer show indications of spin
pumping, as the linewidth of the YIG peak deviates strongly from its isolated behavior
when it crosses the V[TCNE]x~2 peak, though no similar deviation is seen in the linewidth
of the V[TCNE]x~2 peak. Further investigation of the hybrid system is needed to determine
the nature of this signal.

120

References

13.

Ando, K., et al., Angular dependence of inverse spin-Hall effect induced by spin
pumping investigated in a Ni(81)Fe(19)/Pt thin film. Physical Review B, 2008.
78(1).

15.

Saitoh, E., et al., Conversion of spin current into charge current at room
temperature: Inverse spin-Hall effect. Applied Physics Letters, 2006. 88(18).

16.

Wang, H.L., et al., Scaling of Spin Hall Angle in 3d, 4d, and 5d Metals from
Y3Fe5O12/Metal Spin Pumping. Physical Review Letters, 2014. 112(19).

30.

Heinrich, B., et al., Dynamic exchange coupling in magnetic bilayers. Physical


Review Letters, 2003. 90(18).

31.

Brataas, A., A.D. Kent, and H. Ohno, Current-induced torques in magnetic


materials. Nature Materials, 2012. 11(5): p. 372-381.

37.

Liu, L., et al., Spin-Torque Switching with the Giant Spin Hall Effect of Tantalum.
Science, 2012. 336(6081): p. 555-558.

39.

Tserkovnyak, Y., A. Brataas, and G.E.W. Bauer, Enhanced Gilbert damping in thin
ferromagnetic films. Physical Review Letters, 2002. 88(11).

86.

Xiong, Z.H., et al., Giant magnetoresistance in organic spin-valves. Nature, 2004.


427(6977): p. 821-824.

87.

Yuasa, S., et al., Giant room-temperature magnetoresistance in single-crystal


Fe/MgO/Fe magnetic tunnel junctions. Nature Materials, 2004. 3(12): p. 868-871.

88.

Fiederling, R., et al., Injection and detection of a spin-polarized current in a lightemitting diode. Nature, 1999. 402(6763): p. 787-790.

89.

Jaworski, C.M., et al., Giant spin Seebeck effect in a non-magnetic material.


Nature, 2012. 487(7406): p. 210-213.

121

90.

Kruglyak, V.V., S.O. Demokritov, and D. Grundler, Magnonics. Journal of Physics


D-Applied Physics, 2010. 43(26).

91.

Nebel, C.E., VALLEYTRONICS Electrons dance in diamond. Nature Materials,


2013. 12(8): p. 690-691.

92.

Hsieh, D., et al., A tunable topological insulator in the spin helical Dirac transport
regime. Nature, 2009. 460(7259): p. 1101-U59.

93.

Heinrich, B., et al., Spin Pumping at the Magnetic Insulator (YIG)/Normal Metal
(Au) Interfaces. Physical Review Letters, 2011. 107(6).

94.

Slonczewski, J.C., MECHANISM OF INTERLAYER EXCHANGE IN MAGNETIC


MULTILAYERS. Journal of Magnetism and Magnetic Materials, 1993. 126(1-3): p.
374-379.

95.

Ruderman, M.A. and C. Kittel, INDIRECT EXCHANGE COUPLING OF


NUCLEAR MAGNETIC MOMENTS BY CONDUCTION ELECTRONS. Physical
Review, 1954. 96(1): p. 99-102.

96.

Kasuya, T., A THEORY OF METALLIC FERROMAGNETISM AND


ANTIFERROMAGNETISM ON ZENERS MODEL. Progress of Theoretical
Physics, 1956. 16(1): p. 45-57.

97.

Brataas, A., Y.V. Nazarov, and G.E.W. Bauer, Finite-element theory of transport
in ferromagnet-normal metal systems. Physical Review Letters, 2000. 84(11): p.
2481-2484.

98.

Bauer, G.E.W., et al., Universal angular magnetoresistance and spin torque in


ferromagnetic/normal metal hybrids. Physical Review B, 2003. 67(9).

122

Chapter 6 Conclusion

Spintronics promises to extend spin functionality to the existing charge-based


electronics paradigm in a number of ways, with existing applications in magnetic memory
and high frequency magnetoelectronics. With traditional semiconducting electronics
facing fundamental limits as features grow smaller, the manipulation of spin offers a path
to greater computing power and to new applications. This has been demonstrated in a
number of device geometries, with the largest impact made by giant magnetoresistance
spin valves and tunneling magnetoresistance tunnel junctions in magnetic hard drives.
Recently, interest has also been growing in the area of magnetization dynamics, which can
be used to generate pure spin currents via a process called spin pumping. Most of the
research done in the arena of spintronics has been with metallic and insulating inorganic
ferromagnets, which are well-understood materials systems. However, another growing
field is that of organic materials. These carbon-based systems can reproduce several
favorable properties found in inorganic materials, such as electrical conductivity and light
emission and absorption, and combine them with properties unique to organics, like
physical flexibility. Organic materials can thus be used for complementary applications to
inorganics, opening up areas that would otherwise be closed off. The goal of this thesis has
been to extend the use of organic materials to spintronics, utilizing the organic-based
magnet V[TCNE]x, and specifically examining the dynamic magnetization properties. To
this end, we pursued three major projects: electrical transport in a hybrid organic-inorganic
123

heterostructures, ferromagnetic resonance in V[TCNE]x, and FMR-driven spin pumping in


V[TCNE]x.
First, we measured electrical transport in a hybrid organic-inorganic system
consisting of a V[TCNE]x~2 layer deposited on a GaAs-AlGaAs diode. As shown in Fig.
6.1(a), the IV characteristics of the hybrid diode showed strong temperature dependence,
in contrast to the bare diode, which suggests that the V[TCNE]x~2 layer contributes a series
resistance to the diode, which shifts the turn-on voltage and decreases the ideality. Figure
6.1(b) shows the behavior of a bulk V[TCNE]x~2 layer, and reveals similar temperaturedependent behavior as the hybrid device without the rectifying effect of the diode. Taken
together, this indicates that bulk behavior of the V[TCNE]x~2 layer is dominating the
transport response of the hybrid diode. By better understanding the properties of this hybrid
device we identify constraints for V[TCNE]x~2.based spintronic devices, and this result
validates the use of extensively-studied inorganic systems as a probe to study organic and
molecular systems, significantly accelerating the development curve for organic materials.

124

Figure 6.1 (a) IV characteristics of the hybrid organic-inorganic diode at various temperatures from 300 K to
60 K. Inset: Layer structure of hybrid diode. (b) IV characteristics of bare V[TCNE]x~2 layer from 200 K to
60 K. Inset: Sample structure for electrical transport measurement of V[TCNE]x~2 layer.

Second, we measured the magnetic resonance characteristics of V[TCNE]x~2 and


related materials in a variety of ways, using both cavity and stripline geometries and as a
function of the applied field angle, microwave frequency, and temperature. In particular,
the FMR linewidth in V[TCNE]x~2 is comparable to the narrowest line seen in inorganic
systems, YIG, as shown in Fig. 6.2(a). Control measurements of the V[TCNE]x~2
precursors are also shown in Fig. 6.2(a) and its inset, and do not have any features
comparable to the FMR peak. This combined with the angular dependence of the resonance
field shown in Fig. 6.2(b) confirms that we are seeing FMR rather than an artifact or EPR
peak. Frequency dependent measurements indicate that the Gilbert damping parameter is
low (in agreement with the narrow linewidth), but further study is necessary to fully
characterize the relaxation mechanisms in this material. Finally, FMR measurements of
related organic-based magnets V[MeTCEC]x~2 and V[ETCEC]x~2 show broader FMR lines,
though they are still relatively sharp compared to most inorganic ferromagnets. This
supports the model of magnetic interactions described in chapter 2, as the larger functional
groups interfere with the carrier-mediated exchange pathway. Overall, these measurements

125

point to a potential role for these thin film organic-based magnets to complement YIG for
high frequency electronics and lay the foundation for further study of spin dynamics in
these systems.

Figure 6.2 (a) FMR spectrum of V[TCNE]x~2 thin film at 300 K with external field in-plane in black. FMR
spectra of the precursors for the CVD process, TCNE (blue) and V(CO)6 (red) are also shown. Inset: Zoomed
in plot of FMR spectra of CVD precursors (same colors). (b) Resonance field of the V[TCNE]x~2 FMR peak
as a function of applied field angle at 300 K. Solid line shows fit to expected thin film ferromagnet behavior
with a saturation magnetization of 95 G, consistent with SQUID measurements.

Third, we are pursuing ferromagnetic resonance-driven spin pumping from a


V[TCNE]x~2 layer, a dynamic process with the potential to efficiently generate pure spin
currents. Spin pumping has been observed in a number of inorganic magnetic systems, and
126

we have extended these studies to both hybrid organic-inorganic bilayers and all-organic
heterostructures in hopes of taking advantage of the extremely narrow FMR linewidth in
V[TCNE]x. By measuring FMR as a function of field direction in a magnetic bilayer, we
can observe changes in the linewidth of the FMR peaks that indicate spin pumping. In the
all-organic system, we have not seen conclusive evidence of the presence or absence of
spin pumping. Measurements of a hybrid YIG-V[TCNE]x~2 bilayer show indications of
spin pumping, as the linewidth of the YIG peak deviates strongly from its isolated behavior
when it crosses the V[TCNE]x~2 peak, as shown in Fig. 6.3. However, this is not observed
in the linewidth of the V[TCNE]x~2 peak, and so it is ambiguous whether this is spin
pumping or not.

Figure 6.3 Full-width half-maximum linewidth of the YIG peak (red) and peak-to-peak linewidth of the
V[TCNE]x~2 peak (blue) as a function of angle. While the YIG linewidth diverges around the peak crossover,
no similar behavior is observed in the V[TCNE]x~2.

These results taken together significantly push the state of organic spintronics
forward. We successfully incorporated V[TCNE]x~2 into a hybrid organic-inorganic
127

heterostructure, demonstrating that organic materials are compatible with the existing
inorganic semiconductor industry. We explored the magnetic resonance properties of
V[TCNE]x, finding a ferromagnetic resonance linewidth comparable to the narrowest peak
in inorganic materials, and of related chemically modified materials, V[MeTCEC]x~2 and
V[ETCEC]x, yielding information about the nature of magnetism in these materials.
Finally, we began investigating FMR-driven spin pumping in V[TCNE]x, in both allorganic and hybrid organic-inorganic geometries. In addition to the projects detailed here,
a major technical breakthrough has addressed the air sensitivity of V[TCNE]x~2 thin films,
one of the primary issues with the material. Therefore, the work presented here can serve
as a launching pad for a variety of directions, from high frequency applications for
consumer electronics to fundamental studies of spin currents and spin dynamics.

128

References

1.

Wolf, S.A., et al., Spintronics: A spin-based electronics vision for the future.
Science, 2001. 294(5546): p. 1488-1495.

2.

Zutic, I., J. Fabian, and S. Das Sarma, Spintronics: Fundamentals and applications.
Reviews of Modern Physics, 2004. 76(2): p. 323-410.

3.

The Nobel Prize in Physics 2007.

4.

Nanver, L.K., et al., A back-wafer contacted silicon-on-glass integrated bipolar


process - Part I: The conflict electrical versus thermal isolation. Ieee Transactions
on Electron Devices, 2004. 51(1): p. 42-50.

5.

ITRS, International Technology Roadmap for Semiconductors. 2013.

6.

Wolf, S.A., et al., The Promise of Nanomagnetics and Spintronics for Future Logic
and Universal Memory. Proceedings of the Ieee, 2010. 98(12): p. 2155-2168.

7.

Wolf, S.A., A.Y. Chtchelkanova, and D.M. Treger, Spintronics - A retrospective


and perspective. Ibm Journal of Research and Development, 2006. 50(1): p. 101110.

8.

Chang, H.C., et al., Nanometer-Thick Yttrium Iron Garnet Films With Extremely
Low Damping. Ieee Magnetics Letters, 2014. 5.

9.

Pardavi-Horvath, M., Microwave applications of soft ferrites. Journal of


Magnetism and Magnetic Materials, 2000. 215: p. 171-183.

10.

Wigen, P.E., ed. Nonlinear Phenomena and Chaos in Magnetic Materials. 1994,
World Scientific.

11.

Mohseni, S.M., et al., Spin Torque-Generated Magnetic Droplet Solitons. Science,


2013. 339(6125): p. 1295-1298.
129

12.

Demokritov, S.O., et al., Bose-Einstein condensation of quasi-equilibrium magnons


at room temperature under pumping. Nature, 2006. 443(7110): p. 430-433.

13.

Ando, K., et al., Angular dependence of inverse spin-Hall effect induced by spin
pumping investigated in a Ni(81)Fe(19)/Pt thin film. Physical Review B, 2008.
78(1).

14.

Inoue, H.Y., et al., Detection of pure inverse spin-Hall effect induced by spin
pumping at various excitation. Journal of Applied Physics, 2007. 102(8).

15.

Saitoh, E., et al., Conversion of spin current into charge current at room
temperature: Inverse spin-Hall effect. Applied Physics Letters, 2006. 88(18).

16.

Wang, H.L., et al., Scaling of Spin Hall Angle in 3d, 4d, and 5d Metals from
Y3Fe5O12/Metal Spin Pumping. Physical Review Letters, 2014. 112(19).

17.

Shirakawa, H., et al., SYNTHESIS OF ELECTRICALLY CONDUCTING


ORGANIC POLYMERS - HALOGEN DERIVATIVES OF POLYACETYLENE,
(CH)X. Journal of the Chemical Society-Chemical Communications, 1977(16): p.
578-580.

18.

Tang, C.W. and S.A. Vanslyke, ORGANIC ELECTROLUMINESCENT DIODES.


Applied Physics Letters, 1987. 51(12): p. 913-915.

19.

Brabec, C.J., Organic photovoltaics: technology and market. Solar Energy


Materials and Solar Cells, 2004. 83(2-3): p. 273-292.

20.

Yu, Z.G., Spin-Orbit Coupling, Spin Relaxation, and Spin Diffusion in Organic
Solids. Physical Review Letters, 2011. 106(10).

21.

Manriquez, J.M., et al., A ROOM-TEMPERATURE MOLECULAR ORGANIC


BASED MAGNET. Science, 1991. 252(5011): p. 1415-1417.

22.

Pokhodnya, K.I., A.J. Epstein, and J.S. Miller, Thin-film V TCNE (x) magnets.
Advanced Materials, 2000. 12(6): p. 410-+.

23.

de Caro, D., et al., CVD-grown thin films of molecule-based magnets. Chemistry of


Materials, 2000. 12(3): p. 587-+.
130

24.

Baldo, M.A., et al., Highly efficient phosphorescent emission from organic


electroluminescent devices. Nature, 1998. 395(6698): p. 151-154.

25.

Forrest, S.R., The path to ubiquitous and low-cost organic electronic appliances on
plastic. Nature, 2004. 428(6986): p. 911-918.

26.

Kittel, C., Introduction to Solid State Physics. 8 ed. 2005.

27.

Morrish, A.H., The Physical Principles of Magnetism. 1980: IEEE Press.

28.

Cimpoesu, F., et al., Disorder, exchange and magnetic anisotropy in the roomtemperature molecular magnet V TCNE (x) - A theoretical study. Computational
Materials Science, 2014. 91: p. 320-328.

29.

Zakeri, K., et al., Spin dynamics in ferromagnets: Gilbert damping and two-magnon
scattering. Physical Review B, 2007. 76(10).

30.

Heinrich, B., et al., Dynamic exchange coupling in magnetic bilayers. Physical


Review Letters, 2003. 90(18).

31.

Brataas, A., A.D. Kent, and H. Ohno, Current-induced torques in magnetic


materials. Nature Materials, 2012. 11(5): p. 372-381.

32.

Myers, E.B., et al., Current-induced switching of domains in magnetic multilayer


devices. Science, 1999. 285(5429): p. 867-870.

33.

Mangin, S., et al., Current-induced magnetization reversal in nanopillars with


perpendicular anisotropy. Nature Materials, 2006. 5(3): p. 210-215.

34.

Fukami, S., et al., Current-induced domain wall motion in perpendicularly


magnetized CoFeB nanowire. Applied Physics Letters, 2011. 98(8).

35.

Kubota, H., et al., Quantitative measurement of voltage dependence of spin-transfer


torque in MgO-based magnetic tunnel junctions. Nature Physics, 2008. 4(1): p. 3741.

36.

Yamanouchi, M., et al., Current-induced domain-wall switching in a ferromagnetic


semiconductor structure. Nature, 2004. 428(6982): p. 539-542.
131

37.

Liu, L., et al., Spin-Torque Switching with the Giant Spin Hall Effect of Tantalum.
Science, 2012. 336(6081): p. 555-558.

38.

Tserkovnyak, Y., A. Brataas, and G.E.W. Bauer, Spin pumping and magnetization
dynamics in metallic multilayers. Physical Review B, 2002. 66(22).

39.

Tserkovnyak, Y., A. Brataas, and G.E.W. Bauer, Enhanced Gilbert damping in thin
ferromagnetic films. Physical Review Letters, 2002. 88(11).

40.

Kajiwara, Y., et al., Transmission of electrical signals by spin-wave


interconversion in a magnetic insulator. Nature, 2010. 464(7286): p. 262-U141.

41.

Brataas, A., et al., Spin battery operated by ferromagnetic resonance. Physical


Review B, 2002. 66(6).

42.

Crone, B., et al., Large-scale complementary integrated circuits based on organic


transistors. Nature, 2000. 403(6769): p. 521-523.

43.

Yu, H., et al., Ultra-narrow ferromagnetic resonance in organic-based thin films


grown via low temperature chemical vapor deposition. Applied Physics Letters,
2014. 105(1).

44.

Pokhodnya, K.I., B. Lefler, and J.S. Miller, A methyl tricyanoethylenecarboxylatebased room-temperature magnet. Advanced Materials, 2007. 19(20): p. 3281-+.

45.

Kao, C.Y., et al., Thin films of organic-based magnetic materials of vanadium and
cobalt tetracyanoethylene by molecular layer deposition. Journal of Materials
Chemistry C, 2014. 2(30): p. 6171-6176.

46.

Lu, Y., et al., Thin-Film Deposition of an Organic Magnet Based on Vanadium


Methyl Tricyanoethylenecarboxylate. Advanced Materials, 2014. 26(45): p. 76327636.

47.

Sun, Y.Y., Y.Y. Song, and M.Z. Wu, Growth and ferromagnetic resonance of
yttrium iron garnet thin films on metals. Applied Physics Letters, 2012. 101(8).

132

48.

Cao, Y., P. Smith, and A.J. Heeger, COUNTERION INDUCED PROCESSIBILITY


OF CONDUCTING POLYANILINE AND OF CONDUCTING POLYBLENDS OF
POLYANILINE IN BULK POLYMERS. Synthetic Metals, 1992. 48(1): p. 91-97.

49.

LG: Curved smartphones like G Flex 40% of market by 2018. 2013.

50.

Whitwam, R., Harvard uses distributed computing and quantum mechanics to


dream up new solar cells. 2013.

51.

Kao, C.Y., et al., Investigation of thin films of organic-based magnets grown by


physical vapor deposition. Applied Physics Letters, 2014. 105(14).

52.

Kao, C.-Y., et al., Molecular Layer Deposition of an Organic-Based Magnetic


Semiconducting Laminate. Acs Applied Materials & Interfaces, 2012. 4(1): p. 137141.

53.

Haskel, D., et al., Local structural order in the disordered vanadium


tetracyanoethylene room-temperature molecule-based magnet. Physical Review B,
2004. 70(5).

54.

Bozdag, K.D., et al., Optical control of magnetization in a room-temperature


magnet: V-Cr Prussian blue analog. Physical Review B, 2010. 82(9).

55.

Caruso, A.N., et al., Direct evidence of electron spin polarization from an organicbased magnet: Fe-II(TCNE)(NCMe)(2) (FeCl4)-Cl-III. Physical Review B, 2009.
79(19).

56.

Matsuura, H., K. Miyake, and H. Fukuyama, Theory of Room Temperature


Ferromagnet V(TCNE)(x) (1.5 < x < 2): Role of Hidden Flat Bands. Journal of the
Physical Society of Japan, 2010. 79(3).

57.

Tchougreeff, A.L. and R. Dronskowski, A computational study of the crystal and


electronic structure of the room temperature organometallic ferromagnet
V(TCNE)(2). Journal of Computational Chemistry, 2008. 29(13): p. 2220-2233.

58.

De Fusco, G.C., et al., Density functional study of the magnetic coupling in


V(TCNE)(2). Physical Review B, 2009. 79(8).

133

59.

Yoo, J.W., et al., Spin injection/detection using an organic-based magnetic


semiconductor. Nature Materials, 2010. 9(8): p. 638-642.

60.

Tengstedt, C., et al., X-ray magnetic circular dichroism and resonant photomission
of V(TCNE)(x) hybrid magnets. Physical Review Letters, 2006. 96(5).

61.

Kortright, J.B., et al., Bonding, backbonding, and spin-polarized molecular


orbitals: Basis for magnetism and semiconducting transport in V TCNE (x similar
to 2). Physical Review Letters, 2008. 100(25).

62.

Harberts, M., et al., Chemical vapor deposition of an organic magnet, vanadium


tetracyanoethylene, V[TCNE]x~2. Journal of Visualized Experiments, 2015.

63.

Miller, J.S., Oliver Kahn Lecture: Composition and structure of the V TCNE (x)
(TCNE = tetracyanoethylene) room-temperature, organic-based magnet - A
personal perspective. Polyhedron, 2009. 28(9-10): p. 1596-1605.

64.

Yoo, J.-W., et al., Photoinduced magnetism and random magnetic anisotropy in


organic-based magnetic semiconductor V(TCNE)(x) films, for x similar to 2.
Physical Review Letters, 2007. 99(15).

65.

Plachy, R., et al., Ferrimagnetic resonance in films of vanadium


tetracyanoethanide (x), grown by chemical vapor deposition. Physical Review B,
2004. 70(6).

66.

Lundgren, L., P. Svedlindh, and O. Beckman, MEASUREMENT OF COMPLEX


SUSCEPTIBILITY ON A METALLIC SPIN-GLASS WITH BROAD RELAXATION
SPECTRUM. Journal of Magnetism and Magnetic Materials, 1981. 25(1): p. 33-38.

67.

Li, B., et al., Magnetoresistance in an All-Organic-Based Spin Valve. Advanced


Materials, 2011. 23(30): p. 3382-+.

68.

Gould, C., et al., Tunneling anisotropic magnetoresistance: A spin-valve-like tunnel


magnetoresistance using a single magnetic layer. Physical Review Letters, 2004.
93(11).

69.

Divincenzo, D.P., QUANTUM COMPUTATION. Science, 1995. 270(5234): p.


255-261.

134

70.

Prigodin, V.N., et al., Spin-driven resistance in organic-based magnetic


semiconductor V TCNE (x). Advanced Materials, 2002. 14(17): p. 1230-+.

71.

Yoo, J.-W., et al., Spin injection/detection using an organic-based magnetic


semiconductor. Nature Materials, 2010. 9.

72.

Yoo, J.-W., et al., Giant magnetoresistance in ferromagnet/organic


semiconductor/ferromagnet heterojunctions. Physical Review B, 2009. 80(20).

73.

Fang, L., et al., Electrical Spin Injection from an Organic-Based Ferrimagnet in a


Hybrid Organic-Inorganic Heterostructure. Physical Review Letters, 2011.
106(15).

74.

Ashcroft, N.W. and N.D. Mermin, Solid State Physics. 1976: Thomson Learning,
Inc.

75.

Taniyasu, Y., M. Kasu, and T. Makimoto, An aluminium nitride light-emitting


diode with a wavelength of 210 nanometres. Nature, 2006. 441(7091): p. 325-328.

76.

Adelmann, C., et al., Spin injection and relaxation in ferromagnet-semiconductor


heterostructures. Physical Review B, 2005. 71(12).

77.

Grundmann,
M.
and
U.K.
Mishra.
http://my.ece.ucsb.edu/mgrundmann/bandeng.htm.

78.

Urban, R., G. Woltersdorf, and B. Heinrich, Gilbert damping in single and


multilayer ultrathin films: Role of interfaces in nonlocal spin dynamics. Physical
Review Letters, 2001. 87(21).

79.

Arias, R. and D.L. Mills, Extrinsic contributions to the ferromagnetic resonance


response of ultrathin films. Physical Review B, 1999. 60(10): p. 7395-7409.

80.

Heinrich, B. and J.F. Cochran, ULTRATHIN METALLIC MAGNETIC-FILMS MAGNETIC ANISOTROPIES AND EXCHANGE INTERACTIONS. Advances in
Physics, 1993. 42(5): p. 523-639.

81.

Fert, A., Nobel Lecture: Origin, development, and future of spintronics. Reviews
of Modern Physics, 2008. 80(4): p. 1517-1530.
135

Available

from:

82.

Sun, Y., et al., Growth and ferromagnetic resonance properties of nanometer-thick


yttrium iron garnet films. Applied Physics Letters, 2012. 101(15).

83.

Gilbert, T.L., A phenomenological theory of damping in ferromagnetic materials.


Ieee Transactions on Magnetics, 2004. 40(6): p. 3443-3449.

84.

NIST. Available from: http://physics.nist.gov/cgi-bin/cuu/Value?gammaebar.

85.

Farle, M., Ferromagnetic resonance of ultrathin metallic layers. Reports on


Progress in Physics, 1998. 61(7): p. 755-826.

86.

Xiong, Z.H., et al., Giant magnetoresistance in organic spin-valves. Nature, 2004.


427(6977): p. 821-824.

87.

Yuasa, S., et al., Giant room-temperature magnetoresistance in single-crystal


Fe/MgO/Fe magnetic tunnel junctions. Nature Materials, 2004. 3(12): p. 868-871.

88.

Fiederling, R., et al., Injection and detection of a spin-polarized current in a lightemitting diode. Nature, 1999. 402(6763): p. 787-790.

89.

Jaworski, C.M., et al., Giant spin Seebeck effect in a non-magnetic material.


Nature, 2012. 487(7406): p. 210-213.

90.

Kruglyak, V.V., S.O. Demokritov, and D. Grundler, Magnonics. Journal of Physics


D-Applied Physics, 2010. 43(26).

91.

Nebel, C.E., VALLEYTRONICS Electrons dance in diamond. Nature Materials,


2013. 12(8): p. 690-691.

92.

Hsieh, D., et al., A tunable topological insulator in the spin helical Dirac transport
regime. Nature, 2009. 460(7259): p. 1101-U59.

93.

Heinrich, B., et al., Spin Pumping at the Magnetic Insulator (YIG)/Normal Metal
(Au) Interfaces. Physical Review Letters, 2011. 107(6).

94.

Slonczewski, J.C., MECHANISM OF INTERLAYER EXCHANGE IN MAGNETIC


MULTILAYERS. Journal of Magnetism and Magnetic Materials, 1993. 126(1-3): p.
374-379.
136

95.

Ruderman, M.A. and C. Kittel, INDIRECT EXCHANGE COUPLING OF


NUCLEAR MAGNETIC MOMENTS BY CONDUCTION ELECTRONS. Physical
Review, 1954. 96(1): p. 99-102.

96.

Kasuya, T., A THEORY OF METALLIC FERROMAGNETISM AND


ANTIFERROMAGNETISM ON ZENERS MODEL. Progress of Theoretical
Physics, 1956. 16(1): p. 45-57.

97.

Brataas, A., Y.V. Nazarov, and G.E.W. Bauer, Finite-element theory of transport
in ferromagnet-normal metal systems. Physical Review Letters, 2000. 84(11): p.
2481-2484.

98.

Bauer, G.E.W., et al., Universal angular magnetoresistance and spin torque in


ferromagnetic/normal metal hybrids. Physical Review B, 2003. 67(9).

137

S-ar putea să vă placă și