Sunteți pe pagina 1din 95

AN ABSTRACT OF THE THESIS OF

Nathaniel J. Coussens for the degree of Master of Science in Chemical Engineering


presented on August 19, 2011.

Title: Synthesis and Characterization of Zero Valent Iron Nanoparticles.

Abstract approved: ________________________________________________


Alexandre Yokochi

Synthetic methods for the preparation of magnetic nanocomposites with improved high

frequency (into the GHz range) performance over conventional magnetic materials were

examined. The most successful work involved synthesis of nanocrystalline zero valent

iron particles (nc-Fe) through a reverse micelle method. Here, aqueous Fe(II) solution

micelles formed using cetyl trimethylammonium bromide (CTAB) in heptane were

reacted with aqueous sodium borohydride micelles also using CTAB and heptane. This

work was carried out both in a batch reactor and in a T-mixer setup. Particle post-

synthesis processing involved vacuum drying followed by high temperature (500°C)

annealing of the resultant nanoparticles. This was done to improve the crystallinity of

the nanoparticles. Given the highly reactive nature of zero valent Fe, all materials made

consisted of mixtures of nc-Fe and Fe oxides (FeO, Fe2O3, Fe3O4). Additionally,

precipitation of NaBr was observed in some samples, which proved difficult to remove

without fully oxidizing the desired nc-Fe particles. Since the magnetic properties of the

nanoparticles are dependent on the size of coherent crystalline domains in the material,

which will be smaller than the size of agglomerated particles in suspension, size

characterization was carried out by X-ray diffraction analysis using the Scherrer formula.
Particle size was controlled by varying the concentration of Fe(II) in the solution used to

prepare the reverse micelles, and the concentration of Fe(II) in aqueous solution was

varied from 0.5M to 0.05M. XRD results confirmed average particle sizes in the desired

sub 25nm range, and indicated size dependence based on the amount of iron used in

solution in the case of the T-mixer setup. In the batch reaction process, magnetic

filtering of the materials resulted in samples with sizes consistently in the 10-20 nm

range. The nanocomposite was then prepared by mixing the resulting powder with

either epoxy or KBr matrices, with the use of KBr as a matrix proving to be the simpler

approach to making small toroid shaped materials due to the ease of higher pressure

casting.

Both epoxy and KBr composites yielded similar magnetic results, with typical quality

factors of about 100. By comparison, a commercial nanocrystalline nickel ferrite (nc-

NiFe2O4) standard had a quality factor of about 10. Likewise, materials based on fully

oxidized nc-Fe (nc-Fe3O4) produced magnetic materials with low quality factors. This

leads to the conclusion that the nc-Fe materials, even with heavy oxidation, represent a

potential avenue for improvement over air core inductors.


©Copyright by Nathaniel J. Coussens

August 19, 2011

All Rights Reserved


Synthesis and Characterization of Zero Valent Iron Nanoparticles

by
Nathaniel J. Coussens

A THESIS

submitted to

Oregon State University

in partial fulfillment of
the requirements for the
degree of

Master of Science

Presented August 19, 2011


Commencement June 2012
Master of Science thesis of Nathaniel J. Coussens presented on August 19, 2011.

APPROVED:

________________________________________________________________________

Major Professor, representing Chemical Engineering

________________________________________________________________________

Head of the School of Chemical, Biological, and Environmental Engineering

________________________________________________________________________

Dean of the Graduate School

I understand that my thesis will become part of the permanent collection of Oregon State
University libraries. My signature below authorizes release of my thesis to any reader upon
request.

________________________________________________________________________

Nathaniel J. Coussens, Author


TABLE OF CONTENTS

Page

1. Introduction .................................................................................................................... 1

1.1 Inductor Materials Previously Attempted in Research ............................................. 2

1.2 Other Material Property Considerations .................................................................. 3

1.3 Thesis Hypothesis ...................................................................................................... 3

2. Background ..................................................................................................................... 3

2.1 Background Material on Magnetism......................................................................... 3

2.1.1 A Brief Description of Superparamagnetism ...................................................... 3

2.1.2 Hysteresis Curves ................................................................................................ 4

2.1.3 Relating Superparamagnetism to Intrinsic Coercivity ........................................ 6

2.1.4 Determining Critical Particle Size for Nanoparticles .......................................... 7

2.1.5 Complex Permeability......................................................................................... 8

2.2 Zero Valent Iron Synthesis Techniques ................................................................... 10

2.2.1 Methods Using Iron Pentacarbonyl .................................................................. 10

2.2.2 Simple Reduction Methods .............................................................................. 11

2.2.3 Reduction in Reverse Micelles.......................................................................... 12

2.2.4 Emulsion Operating Conditions ........................................................................ 13

2.2.5 Other Methods ................................................................................................. 14


TABLE OF CONTENTS (Continued)

Page

2.3 Iron Nanoparticle Passivation ................................................................................. 14

2.4 Different Reactor Vessels and Equipment .............................................................. 15

2.4.1 Microreactors ................................................................................................... 15

2.4.2 Microwaves....................................................................................................... 16

2.4.3 Batch Reactor ................................................................................................... 16

2.4.4 Sonication ......................................................................................................... 16

2.4.5 Milling ............................................................................................................... 17

2.5 General Factors that Affect Particle Size and Shape ............................................... 17

2.6 Making and Characterizing Magnetic Nanocomposites ......................................... 17

3. Experimental Methods ................................................................................................. 19

3.1 Representative Synthetic Methodology ................................................................. 19

3.1.1 Reverse Micelle Synthesis Process using a Batch Reactor ............................... 19

3.1.2 Consideration of the Emulsion Phase Regime .................................................. 22

3.1.3 Adjustments to Reverse Micelle Synthesis when using a T-mixer ................... 23

3.1.4 Procedure for Making Iron Oxide Material ...................................................... 24

3.2 Creation of Magnetic Composite Material ............................................................. 24

3.3 Analysis of XRD Results ........................................................................................... 25


TABLE OF CONTENTS (Continued)

Page

4. Experimental Results .................................................................................................... 30

4.1 Initial Results ........................................................................................................... 30

4.2 TEM Imaging ............................................................................................................ 31

4.3 XRD Results.............................................................................................................. 34

4.4 Magnetic Characterization ...................................................................................... 35

4.4.1 Magnetic Curing Effects.................................................................................... 36

4.4.2 Epoxy Samples .................................................................................................. 37

4.4.3 KBr Composite Samples .................................................................................... 39

4.4.4 KBr Composite Samples Re-tested (5 Weeks Later) ......................................... 42

4.4.5 Assessment of Instrument Repeatability ......................................................... 45

4.4.6 Oxidized KBr Composite Data ........................................................................... 49

5. Conclusions and Future Work ....................................................................................... 52

Appendix A: Quality Factors for Epoxy Composites ......................................................... 55

Appendix B: Magnetic Composite Test Results, KBr ......................................................... 65

Appendix C: Quality Factor Graphs, KBr ........................................................................... 67

Appendix D: XRD Peaks ..................................................................................................... 70

Appendix E: XRD Results ................................................................................................... 71


TABLE OF CONTENTS (Continued)

Page

Works Cited ....................................................................................................................... 77


LIST OF TABLES
Table Page

Table 1: Particle sizes as determined from the aggregates of each image. ..................... 33

Table 2: Experimental particle size results for each data point ....................................... 35

Table 3: Quality factors for differing quantities of Fe ...................................................... 40

Table 4: KBr measurement real and imaginary part averages and standard deviation ... 46

Table 5: 1x Fe, real and imaginary part averages and standard deviation ....................... 48
1

1. Introduction

As technology progresses, devices are becoming smaller, and, as a result, higher

frequency electronic circuits are required.[1] One problematic area that needs to be

investigated is the development of high frequency inductors that can operate into the

gigahertz range. To accomplish this, several different materials are being investigated.

One such material is metallic nanoparticles. More specifically, iron has promising

properties.

Elemental iron has very promising magnetic properties, as has been shown from its

experimental room temperature (20°C) saturation magnetization of 218 A-m2/kg.[2] This

is better than both Co and Ni, with saturation magnetizations of 161 and 54 A-m2/kg,

respectively.[2] Further, zero valent iron nanoparticles can achieve unique properties at

sizes less than about 10-20 nm in size.[3] One such property is superparamagnetism.

One device that would benefit greatly from the use of zero valent iron is an

ultracompact micropower module,[4] where the inductor used could be improved by

using an iron nanoparticle-polymer composite type material. The synthesis and

characterization of this type of material is the focus of the thesis.


2

1.1 Inductor Materials Previously Attempted in Research

Figure 1 shows different inductor configurations that have been used in research,

including air, permalloy, and ferrite thin films.[5] Here, a cross-section of the inductor

coils used for experiments is provided. In the top image, the wire coil used in making an

inductor is only encased with air. Below this image, the inductor is partially covered by

permalloy. Below this image, the wires are encased with permalloy. Finally, in the

bottom image, ferrite is introduced as a material to encompass the wires.

[5]
Figure 1: Different inductor configurations as proposed by Galle et al. Copyright © 2007, IEEE.

All of the materials used in research as shown in Figure 1 encounter issues that make

them difficult to use for high frequency inductor applications. Ferrite cores suffer from

high energy losses via magnetic hysteresis.[6] Meanwhile, the permalloy films are very

conductive, meaning that they need to be applied as a very thin film (less than 200 nm)

to avoid high energy losses due to eddy currents.[7] An inductor made using zero valent
3

iron nanoparticles is a possible alternative material, which to date has not been

investigated for this specific type of application.

1.2 Other Material Property Considerations

Energy losses from eddy currents also need to be considered, since they will compound

the losses from ferromagnetic resonance. Also, metal nanoparticles exhibit higher

permeability than non-conducting materials such as air. Here, the permeability is

needed to facilitate appropriate inductance improvement over an air core type

inductor.[8] Establishing a matrix using insulator material is required to provide the

necessary high material resistivity for an inductor. Finally, magnetic hysteresis needs to

be kept to a minimum to reduce losses from energy dissipation. This is fulfilled by

synthesizing nanoparticles that are superparamagnetic.

1.3 Thesis Hypothesis

Therefore, the hypothesis of the thesis is that creating a nanocomposite material with

metal nanoparticles in an insulating matrix will result in a magnetic material with

magnetic property improvements over air for high frequency inductor applications.

2. Background

2.1 Background Material on Magnetism

2.1.1 A Brief Description of Superparamagnetism

Superparamagnetism is a property possessed by magnetically soft materials.

Briefly, soft magnetic materials reach saturation upon exposure to a small applied

magnetic field whereas hard magnetic materials require a much more substantial
4

applied field to reach saturation.[9] Specifically, superparamagnetism, has to do with

how the time to magnetize something compares against the time it takes for the

magnetic state to randomly change orientation (referred to as the relaxation time). For

a superparamagnetic material, the relaxation time is much smaller than the time it takes

for the magnetization of the material to be reversed.

2.1.2 Hysteresis Curves

The effects of a reverse applied field via a hysteresis curve can illustrate the differences

between hard and soft magnetic materials. The hysteresis curves shown in Figure 2

demonstrate how magnetization is affected by an applied magnetic field for the

different types of material. Here, the arrows indicate the direction in which

magnetization will change when going from a negative to positive applied magnetic field

or vice versa.

Figure 2: Hysteresis curves showing the effects of an applied magnetic field on magnetization for both soft
and hard magnetic materials.
5

The energy that goes into the magnetization of a material is the summation of potential

energy and work required to change the material from one magnetic state to another.

Since for a hysteresis cycle, the difference in potential energy will be zero, the energy

lost between cycles will be the work required to change magnetic states. This work is

given as:[10]

M2

W  HdM
M1
(1)

Here, M is the magnetization, and H is the applied magnetic field. This integral can be

expressed visually as the area under the hysteresis curve, as is shown by the red regions

in Figure 3. The area encompassed between the line intersecting the saturation

magnetization point and the hysteresis curve (shown as the green region in Figure 3)

represents the potential energy that is stored at the point of saturation.

Figure 3: Energy loss (red region) and energy storage (green region) at the point of magnetic saturation
6

As is shown in Figure 3, the key difference between material types is that very little

energy is lost between cycles when the magnetization is reversed in a soft magnetic

material, whereas a large quantity of energy is necessary to reverse magnetization in a

hard magnetic material.

2.1.3 Relating Superparamagnetism to Intrinsic Coercivity

Intrinsic coercivity is a very important property with respect to permanent magnet

applications. The intrinsic coercivity can be defined as the applied field necessary to

take a material from its saturation point back to zero.[9] However, for the current

application, the coercivity should in fact be as close to zero as possible. At an intrinsic

coercivity of zero, the material is said to be superparamagnetic.

Also, for low coercivities, a material is said to exhibit “soft” magnetic behavior. When

relating this to particle size, there is a critical diameter above which the particle will no

longer display superparamagnetic properties, which will be referred to as D’.[9] Below

and near D’ the particles will therefore exhibit soft magnetic behavior.

For higher coercivities, the material exhibits “hard” magnetic behavior. At particle sizes

above the critical diameter, the coercivity rises until it reaches a maximum value.[9] The

particle size where this maximum coercivity value is reached, which will be referred to

as D’’, is where particles are most likely to exhibit hard magnetic behavior.

The equation that relates intrinsic coercivity to particle diameter is: [9]

h
H ci  g 
D3/ 2 (2)
7

Here, Hci is the intrinsic coercivity, g and h are empirically determined constants that are

dependent on the material used, and D is the particle diameter.

2.1.4 Determining Critical Particle Size for Nanoparticles

The critical particle size can be determined by the Néel-Arrhenius equation, which is

applicable only for superparamagnetic particles in the nano-regime. This equation can

be used to determine the time it takes for the particle to change between its two stable

magnetic states.[11]

1  KV 
 exp  
f  B 
k T
(3)

Here:  represents the time typically taken for a flip in the particle’s magnetic state (the

relaxation time), f is the attempt frequency, K = particle magnetic anisotropy constant, V

= particle volume, kB = the Boltzmann constant, and T = temperature. Using Equation 2,

and assuming a spherical particle:

3
4  D' 
V     , where D’ is the critical particle diameter.
3  2

  4  D' 3  
 K    
 3  2   
  exp   
1
f  k BT 
 
 
 4  D'  
3

k BT ln f   K     
 3  2  
 6k T ln f  
1/ 3

D'   B 
 K  (4)
8

Using equation 3, the critical particle size can be found. A rough calculation is given

below:

First, it should be noted that the magnetic anisotropy constant for nanosized iron is

roughly one order of magnitude higher than that for bulk iron.[12] Here, the value for

nanosized iron (~2nm diameter) determined by Lacroix et al. will be used:

K  5.2 105 J / m3 .[12] For the attempt frequency, a typical value is about f=1010s-1. The

relaxation time will be taken to be the inverse of the desired frequency (200 MHz), or:

1
  5 10 9 s . Then, at room temperature (assumed to be 20°C):
200 10 6 s 1

    
 6 1.38  10  23 J / K 293K  ln 5  10 9 s  1010 s 1
D'  
  1/ 3

 (5.2  10 5 J / m 3 ) 
 
D'  3.9  10 9 m  3.9 nm

2.1.5 Complex Permeability

There are cases where the magnetization cannot change immediately in response to the

applied magnetic field, leading to time dependence. This occurs when there is an

abrupt change in magnetization. Such a change could be from going from one set value

to another, or from an alternating magnetic field.[13] The case that will be considered is

an applied alternating external magnetic field at high frequencies. The expressions used

to reflect the time dependence of the magnetization and the applied magnetic field are

provided below:

In terms of the magnetization[13]:

B  B0 e i (t  ) (5)
9

Here, B is the magnetization, B0 is the initial magnetization, σ is the phase difference,

and t is the time.

In terms of the applied magnetic field[13]:

H  H 0 e it (6)

Here H and H0 represent the applied magnetic field and initial applied magnetic field,

respectively.

In terms of permeability[13]:

B
 (7)
H

Substituting in equations 5 and 6 into equation 7 this becomes:[13]


B e i (t  ) B
  0 it  0 e i
H 0e H0

From the relationship e i  cos   i sin  [13]


:

e i  cos   i sin  , so that:[13]

B0
 cos   i sin  
H0
B0 B
 cos   i 0 sin 
H0 H0

Here there will be a real and imaginary component so that:[13]

B0
'  cos  is the real component of permeability, and:[13]
H0

B0
' '  sin  is the imaginary component of permeability.
H0

The overall expression then becomes:[13]

   'i ' ' (8)


10

The significance of the real component of permeability in equation 8 is that this is the

component that coincides with the applied magnetic field.[13] The imaginary component

of permeability, however, has a 90° phase shift from the applied magnetic field.[13] As

the imaginary component of permeability approaches zero, the real component of

permeability approaches the overall permeability, and energy loss approaches zero. As

the imaginary component increases, more energy is required to sustain the applied

magnetic field, which is the cause of energy loss.[13] As a result, in an ideal situation

where energy loss is minimal, the imaginary component of permeability will be very

small. On the other hand, since the real component of permeability is in phase with the

applied magnetic field it represents the energy that is stored. The quality factor can be

expressed as:

'
Q (9)
''

A higher quality factor value reflects less energy loss and better energy storage, which is

what is desired for inductors operating at high frequencies.

2.2 Zero Valent Iron Synthesis Techniques

Two different synthesis routes are used in producing zero valent iron samples: using iron

pentacarbonyl and reacting reducing agents with cationic iron in solution. The

reduction method can be space confined by using micelles, as will be discussed in 2.2.3.

2.2.1 Methods Using Iron Pentacarbonyl

Via thermal degradation, iron pentacarbonyl, after a series of complicated reactions,

forms zero valent iron. The biggest obstacle to using this process is the very pyrophoric
11

nature of iron pentacarbonyl. To use a pyrophoric chemical, a high-quality glove box

would be required (which was not readily available). Another problem with iron

pentacarbonyl is devising a process to achieve a monodisperse particle distribution. Pei

et al. attempted achieving better particle uniformity by using the principle of Ostwald

ripening. Here, the iron nanoparticle-containing solvent was heated at a constant

temperature over a prolonged period of time. This allowed time for the smaller

particles to dissolve, while the larger nanoparticles grew.[14] Eventually, all the smaller

particles would be dissolved, leaving only the larger ones and a narrower particle size

distribution.

2.2.2 Simple Reduction Methods

In a “simple reduction”-type process, a strong reducing agent, such as sodium

borohydride, is used in an aqueous reaction process to form iron nanoparticles. Other

alternative reducing agents which are weaker include: hydrazine and ascorbic acid[15, 16].

For this method, a surfactant is typically used to bind to the particle surface and keep

particles from sticking to each other. Sodium borohydride based reduction reactions

can typically be done at room temperature. However, for weaker reducing agents, heat,

and in some cases, pressurized processes, are sometimes required. An example of this

is preparation of nanoparticles using the “ferrite route” as done by Kuroda et al. [16]

Here, heating in an autoclave is required to reform ferrite nanoparticles into iron

nanoparticles.[16]

Another process that has been attempted for simple reduction is the use of a seed

material to help with the nucleation phase, and, ideally, facilitate better control of
12

particle size. This process involved the use of Palladium as the seed material, and

polyacrylic acid as the dispersing agent.[17]

2.2.3 Reduction in Reverse Micelles

One simple method to control particle size is to kinetically restrict the space in which the

particle can grow. This can be achieved using the reverse micelle synthesis process.

Here, the surfactant must reach the critical micelle concentration (CMC) for micelles to

form. For the reverse micelle synthesis process, two solutions are used. Within each

solution, the solvent is organic, and the micelles are aqueous. A solution where the

micelles contain cationic iron is mixed with a solution where the micelles contain the

reducing agent. Upon mixing, zero valent iron precipitates within the micelles, where

growth is limited by the size of the micelle. This is shown in the diagram in Figure 4

below.

[18]
Figure 2: The general reverse micelle reaction process. Here, the microemulsion can be nanoscale with
sufficient surfactant concentrations. Reprinted with permission from Elvesier.
13

Several researchers have used this method. One example is the research work of Wang

et al., which used a water-in-ethanol micelle system. Here, the surfactant used was

sodium lauryl sulfate.[19] The ethanol-water emulsion initially contained FeCl2, which

was reacted with the reducing agent, KBH4.[19] Then, a nickel-containing solution and

additional reducing agents were added after black precipitate was produced to promote

further reaction.[19]

One typical method used in more recent literature to form micelles is as follows: octane

is used as the organic solvent, CTAB is used as the surfactant, n-butanol is used as the

co-surfactant, and water is used for the solvent in the space within the

micelles.[20,21,22,23] Here, the Fe-containing solution is typically mixed with a

borohydride-containing reducing agent to form zero valent iron.[20,21,22,23]

2.2.4 Emulsion Operating Conditions

Figure 5: Tertiary phase diagram for CTAB+butanol, water, and octane system as proposed by Ayyub et
[24]
al. Here, L1 is the oil-in-water emulsion phase and L2 is the water-in-oil emulsion phase. Reprinted with
permission by the Royal Society of Chemistry.

When using the reverse micelle method, the correct phase region needs to be chosen to

ensure that the solution is truly an emulsion. The above tertiary phase diagram shows
14

the different states for different chemical compositions. Here, L2 represents the reverse

micelle emulsion area of interest. This diagram was taken into consideration for all

chosen system chemical compositions in the present thesis work, which are all in the

water-in-oil emulsion phase. For more discussion on how the general operating

conditions in the phase diagram relates to this work, refer to the experimental section.

2.2.5 Other Methods

It should briefly be noted that other compounds can be used for iron nanoparticle

formation. One example involves the use of Fe[N(SiMe3)2]2, which when subjected to a

temperature of 150°C and fed with hydrogen to a pressure of 3 bar, will produce iron

nanoparticles.[25]

2.3 Iron Nanoparticle Passivation

Since zero valent iron is reactive with air, real world applications with zero valent

nanoparticles would be problematic without some type of surface layer passivation.

There are a few different approaches to accomplishing this. One such method is the use

of an outer layer of metal which is less reactive with air. Gold is one candidate for

achieving this. Gold has been introduced after the formation of zero valent iron as a

dilute auric acid solution. This is then mixed with sodium borohydride to make a zero

valent gold passivation layer.[21,22,26,27] One advantage of gold is that it allows for

passivation with limited effect on the magnetic properties of the zero valent iron

nanocomposite material.[26] The gold layer is typically added using the reverse micelle

process. A more reactive metal that has been used in research is nickel. Here, a reverse
15

micelle process with sodium borohydride as the reducing agent is used in a fashion

similar to a gold coating layer.[28,29] However, even with a surface coating, oxidation

may still slowly persist over time.[27]

Polymer coatings have also been used to facilitate passivation of the surface.[16,30,31,32]

Finally, in some of the literature, very little is done to passivate the outer surface. Here,

it would be assumed that enough of a passivating layer of oxide is formed on the outer

surface to prohibit further oxidation.[29] Many of the procedures in these synthesis

papers require surfactant, which may remain on the surface after drying to form some

degree of protection from oxidation. Finally, one last possibility to passivate the surface

is to use a different material that enjoys magnetic properties similar to those of iron.

Permalloy has higher permeability than zero valent iron alone.[33] Permalloy is

essentially Ni and Fe combined in alloy form, and can come in many different ratios of

Ni:Fe. However, the highest permeability permalloy will have higher Ni compositions.[33]

2.4 Different Reactor Vessels and Equipment

In addition to the chemical processes, there are different physical processes which can

be used in the synthesis of zero valent iron nanoparticles.

2.4.1 Microreactors

One example of this is the use of microreactors. In one such case, nanoparticles of silver

and gold were produced using a series of microreactors.[34] Here, the particle size was

controlled by adjusting the concentration of borohydride reducing agent. Flow rate

adjustment was also used for particle size distribution control.[34] Ideally, the main

advantage of adopting a microreactor process is narrower particle size distribution, as


16

all the solutes in solution should have a similar history when the process is adequately

mixed.

2.4.2 Microwaves

Microwaves have the advantage that they can very evenly distribute heat in a solution.

Processes requiring heat, such as reduction processes which need heat for the reducing

agent to work properly (as is the case with hydrazine reduction), may benefit from the

use of microwaves. This uniform distribution of heat can increase the speed with which

a solution can reach a desired temperature. This can help with particle size distribution

since nucleation would be occurring mostly at a set temperature leading to better

efficiency. In one case, synthesis speed was increased by a factor of ten.[35]

2.4.3 Batch Reactor

A batch reactor setup could be used. This is the reactor setup described for all reverse

micelle reaction processes. While easier to set up than a micromixer-type process, it

can have the disadvantage of broader particle size distribution. Reverse micelles help to

eliminate this disadvantage by making the particle growth regions space-constrained, so

that only a limited size of nanoparticles can be grown in solution.

2.4.4 Sonication

Sonication can be used to break up chemical compounds by adding sufficient energy to

break up the chemical bonds. In the case of iron pentacarbonyl, sonication can be used

to remove the carbonyl groups, leaving only zero valent iron behind. When sonication is

conducted in a bath of coating solution, as in the work done by Nikitenko et al [31,36], the

iron can also be coated and passivated through this process.


17

2.4.5 Milling

Milling is a physical process by which iron can be broken down into nanoscale particles

by simply grinding bulk iron until the appropriate grain size is reached. One large

problem with this process is possible contamination introduced (such as machine oil).

Also, the particle size distribution is significantly broader for this process than for other

processes described here.[3]

2.5 General Factors that Affect Particle Size and Shape

There are several factors which affect the particle size. With the case of particle size,

one factor to consider is the reducing agent. The use of a stronger reducing agent will

lead to smaller nanoparticles.[37] This has to do with the rate of reaction. With weaker

reducing agents, the rate of reaction is slower. Also, the particle size is larger. Particle

size distribution, on the other hand, can be affected both ways. For a slow rate of

reaction, new particles are continuously formed, and particle size distribution will be

more broad. However, where further nucleation is limited, there is diffusion limited

growth. This will have a more narrow particle distribution.[37]

Solute concentration is another factor affecting particle size. By maintaining low solute

concentrations, a diffusion controlled particle growth process can be achieved, leading

to a more monodisperse distribution of nanoparticles.[37] Also, it is common to have a

layer of polymer coating on the particle surfaces as a means of providing stability to the

particle suspensions.[37] This polymer coating also further limits diffusion, and, as a

result, particle growth.[37]

2.6 Making and Characterizing Magnetic Nanocomposites


18

A nanocomposite material consists of a dispersion of nanoparticles in a matrix. Matrix

materials used for the preparation of nanocomposites have included rubber and

epoxy.[38,39] When molding the nanocomposite materials, a popular process is

compression molding. In one such instance in literature, the molding took place at

175°C and 10 MPa.[40] It should be noted that permeability is affected by a variety of

factors, including: the particle size, particle shape, extent to which the particles are

monodisperse, and the particle loading volume percent.[40] Here, particle loading can

directly be changed during the process of molding nanocomposites. Particle-particle

interactions significantly affect magnetic and electric properties only above a particle

loading volume fraction of about 0.2.[40] Below this volume fraction, a continuous

conductive network does not exist.[40] Also for this reason, the permittivity of the

composite is expected to increase linearly with particle loading volume percent as long

as the particle loading remains low.[40]

How the particles are affected by the molding process also needs to be considered.

Here, Jamal et al. found that, when taking XRD analysis of a composite material

containing nickel nanoparticles after processing, the nanomaterial maintained the same

chemical composition.[38] The absence of Ni oxidation is encouraging for iron

applications since nickel will also oxidize.


19

3. Experimental Methods

Hexadecyltrimethylammonium bromide (CTAB) for molecular biology, ≥99% was

purchased from Sigma-Aldrich, and used as received. Heptane, 99.0% minimum, was

purchased from EMD, and used as received. Butyl alcohol, ferrous sulfate 7-hydrate

(granular), and methyl alcohol (anhydrous) were purchased from Mallinckrodt

Chemicals, and used as received. Sodium borohydride, 98%, was purchased from Alfa

Aesar, and used as received.

3.1 Representative Synthetic Methodology

3.1.1 Reverse Micelle Synthesis Process using a Batch Reactor

A simple batch reaction was used for all synthesis experiments. As has been discussed

previously, the reverse micelle synthesis approach was used. The “standard”synthesis

method was chosen to closely follow the paper by Lin et al.[23] This method was then

adjusted when optimizing the process for particle size control. Procedurally, two

separate solutions were prepared. One had reverse micelles containing dissolved iron

ions. The other had reverse micelles containing dissolved reducing agent (sodium

borohydride). For both solutions, CTAB and butanol were cosurfactants. These two

solutions were then mixed together to form zero valent iron nanoparticles. The

“standard” method was as follows:

Solution A contained: 0.34 g (0.5M) iron (II) sulfate heptahydrate, 2.4 mL deionized

water, 23 mL heptane, 6.0 g CTAB, and 6.2 mL butanol.

Solution B contained: 0.09 g (1.0M) sodium borohydride, 2.4 mL deionized water, 23 mL

heptane, 6.0 g CTAB, and 6.2 mL butanol.


20

Additional lab equipment (for each experiment) included:

One 2-necked 125 mL flask (for solution A preparation and zero valent iron reaction)

One 100 mL bottle (for solution B preparation)

1 septum (on 2-necked flask)

2 magnetic stir bars

2 glass frits (for sparging)

1 magnetic stir plate

One 0.4T magnet (for particle separation from solution)

1 weighing scale

2 weighing dishes

3 plastic scoopulas

One 25 mL graduated cylinder

One 10 mL graduated cylinder

One 1 mL pipette

First, all 2-necked flasks, magnetic stir bars, and glass frits used were soaked in 20

volume percent nitric acid to ensure any iron from prior experiments was removed from

the surface of all equipment. Then, this was rinsed with deionized water. Then, the

components for solution A were measured out, in the proportions described above, and

mixed in the 2-necked 125 mL flask. Then, the components for solution B were

measured out, and mixed in the 100 mL bottle. Both of these solutions were degassed

using the 2 glass frits and (99.9% purity minimum) nitrogen gas for a minimum of 15
21

minutes to remove any oxygen from solution. Then, these two solutions were mixed

together, with vigorous stirring, using a magnetic stir bar.

The procedure then closely follows the one by Cho et al.[22] The solution was allowed

one hour of reaction time. Then, the stir bar was removed from the solution. Finally,

the solution was situated on the 0.4T magnet to separate out the nanoparticles. After

separation, the nanoparticles were rinsed with methanol under an inert nitrogen

atmosphere. The solution was then vacuum dried. Unlike the paper by Cho et al, the

surface oxide layer was assumed to be sufficient to passivate the particle surface, and a

gold coating was not used.

After the vacuum drying process was finished, the nanoparticles were moved into a

glass tube, which contained an inert nitrogen gas environment when sealed at both

ends. This was then annealed in a tube furnace for 10 hours (with an additional 50

minutes of ramp-up time) at 500°C to improve the crystallinity of the sample. After

annealing, the sample was made into a nanocomposite for magnetic testing.

Using this method, several different adjustments were attempted. First, an outer

coating of nickel was used to passivate the nanoparticles. Here, nickel was used. For

this experiment, the same setup was used except now a third solution was added after

the first two were mixed together:

Solution C contained: 0.1 g nickel (II) sulfate heptahydrate, 2.4 mL deionized water, 23

mL heptane, 6.0 g CTAB, and 6.2 mL butanol.

Another process attempted involved creating a permalloy material (20% Fe / 80% Ni);

this material would oxidize less quickly than zero valent Fe. Here, the procedure was
22

similar to above except that the iron and nickel ions were in the same micelle solution.

The method is given below:

Solution A contained: 0.07 g iron (II) sulfate heptahydrate, 0.26 g nickel (II) sulfate

heptahydrate, 2.4 mL deionized water, 23 mL heptane, 6.0 g CTAB, and 6.2 mL butanol.

Solution B contained: 0.09 g sodium borohydride, 2.4 mL deionized water, 23 mL

heptane, 6.0 g CTAB, and 6.2 mL butanol.

Post-annealing, both experiments involving nickel for passivation showed XRD peaks not

characteristic of the expected material. As a result, the procedure using only iron

nanoparticles was adopted as the standard procedure. Variations of the “standard

method” were then used to change particle size. Here, by increasing the surfactant

amount while maintaining a similar CTAB:butanol co-surfactant ratio, the micelle size

was changed, which permitted differences in overall particle size. Also, if the particle

size is significantly smaller than micelle size due to a lack of Fe ions in solution, adjusting

the iron concentration may have an effect on particle size.

3.1.2 Consideration of the Emulsion Phase Regime

The reverse micelle method described must operate within the emulsion regime, as

previously described in the background section. There are two significant differences

that should be considered when using Figure 5 for the present experimental work. First,

the CTAB:butanol ratio is slightly different from the 1:0.73 ratio represented in the

figure (it is actually 1:0.83).[22] Second, heptane has been used in the place of octane.

In retrospect, the 1:0.73 ratio could have been used to get a better means of

comparison.[22] However, the 1:0.83 ratio is the ratio used by Lin et al.[23] This ratio was
23

used for the “standard method,” with some TEM images providing a means of

estimation for expected particle size. However, some differences in particle size can be

anticipated since gold coating is not adopted in the present work. Also, in this work,

heptane was chosen instead of octane, which should have a fairly inconsequential effect

on the system given that both have similar properties as organic solvents.

Since the present conditions do not exactly correlate with the conditions used to make

Figure 3, some margin of error was allowed, with none of the chosen conditions lying

near the phase boundary regions.

3.1.3 Adjustments to Reverse Micelle Synthesis when using a T-mixer

The reverse micelle synthesis process previously described was modified to use a T-

mixer rather than a batch reactor setup as detailed above. Here, solutions were

prepared as above, with the exception of the sodium borohydride solution. Since the

reaction time was substantially decreased, the concentration of sodium borohydride

was increased to compensate. The concentration was approximately 2.5 times that of

the original process detailed above, or: 0.225g sodium borohydride added instead of

0.09g. The solutions were mixed together via a T-mixer apparatus, and then introduced

to a tube across which the reaction took place. The syringe pumps forcing the solutions

through the T-mixer were set for a residence time of 5 minutes. However, the pressure

produced by the hydrogen offgas from the sodium borohydride solution forced the

solution through the tubing substantially faster, at a residence time of 3.5 minutes.

After exiting the tube, the solution was then quenched in 55 mL of deoxygenated

methanol. In this way, the reaction should only be taking place in the tubing.
24

After creating the nanoparticles, a procedure had to be established to properly mold the

material into a nanocomposite suitable for testing.

3.1.4 Procedure for Making Iron Oxide Material

An oxide sample was prepared by taking a solution of 1.4 grams FeSO4.7H2O dissolved in

96 mL deionized water and adding to it: 0.75 grams sodium borohydride dissolved in 20

mL deionized water. This was done under vigorous stirring. After reaction, the

precipitate was magnetically separated from solution, rinsed with deionized water, and

put in the oven for 10 hours at 500°C while under air to reach full oxidation.

3.2 Creation of Magnetic Composite Material

Two different composite materials were made. The first composite material used epoxy

as the matrix material. For molding this epoxy, a silicone putty molding template was

used. Silicone putty was chosen as a material since it will not stick to epoxy. West

System Epoxy two-part system comprising of mixtures 105 (resin) and 205 (hardener)

was used for the epoxy. The iron nanoparticles were dispersed and mixed in a 1:4 ratio

of hardener to clear resin in various different weight percents. The nanocomposite

material was then vacuum cured. Part of the epoxy mix is volatile, so that longer than

one minute of curing will cause part of the mixture to spill out of the mold. The main

purpose of the vacuum curing process was to remove any significant void spaces. The

composite material was then given a minimum of 4 hours of drying time.

The second matrix material used was KBr. To make a KBr composite, first a small

fraction of KBr was ground into a fine powder using a mortar and pestle. Then, this

powder was oven-dried to remove any moisture it may have absorbed from the
25

surroundings. With the KBr powder prepared, the nanoparticles to be used in the

composite were weighed, with ten weight percent KBr powder being added to the

powder immediately after. This mixture was then ground using mortar and pestle until

a fine powder was formed. This powder was placed in a KBr press, and after

compression using the press a composite pellet was formed. A hole was then created in

the middle of the pellet from manually scraping at the pellet with a drill bit. Holes were

drilled manually since the pellets were too fragile to survive the mechanical hand drill.

After the hole in the middle was made, the resulting toroid-shaped composite sample

would be used for magnetic testing.

3.3 Analysis of XRD Results

The background to the XRD pattern first must be removed if an XRD peak is to be fitted

for size. To do this, a blank XRD pattern was used. This pattern was then multiplied by a

constant and subtracted from the original XRD pattern. A second constant was also

subtracted to reach an intensity of zero for all 2-theta angles without a peak. To get

these two constants, the solver function was used in excel to get the least sum of

squares error for a portion of the pattern known to be zero. In equation form:

I '  I  I 0 * C1  C2 (4)

Here, I ' is the zeroed XRD intensity, I is the original XRD intensity, I 0 is the blank XRD

pattern, and C1 and C 2 are constants.

The result of blanking the pattern can be seen in Figure 6 and Figure 7 below.
26

original xrd
5000
4500
4000
3500
3000
2500
2000
1500
1000
500
0
25 30 35 40 45 50 55 60 65 70
2-theta

Figure 6: XRD result of “standard solution” reverse micelle synthesis process, post-annealing.

zeroed xrd
2000
1800
1600
1400
1200
1000
800
600
400
200
0
25 35 45 55 65
2-theta

Figure 7: XRD result of zeroed “standard solution” processed via the reverse micelle synthesis process,
post-annealing.
27

In Figure 7, a fair amount of noise exists. This makes it more difficult for the XRD sizing

software to get a good fit of a given XRD peak. To remedy this, a 5-point moving boxcar

average was used. The result can be found in the XRD result in Figure 6 below:

Moving boxcar average, 5-pt


2000

1500

1000

500

0
25 30 35 40 45 50 55 60 65 70
2-theta

Figure 8: XRD result of zeroed “standard solution” after 5-point moving boxcar average.

To determine the particle size from this image, a program called Winfit was used. Winfit

is a program that was created by Stefan Krumm of the Institute for Geology in Erlangen,

Germany. It was last updated in 1995, and can be accessed from the following URL:

ftp://eps.unm.edu/pub/xrd/index.htm.

Using this program, the XRD peak at 43.3 could be fit. This peak was chosen due to the

peak at 44.5 for iron being skewed to the left when taking measurements. The cause of

this was the powder surface on which the XRD radiation diffracted not being perfectly

flat.

Then, using the XRD pattern of a standard material, an approximation of particle size

could be determined using Winfit. Here, Winfit uses Fourier analysis based on the
28

integral breadth of the peak. All XRD sizing results were obtained using this process.

An example of a peak deconvolution using Winfit is given in Figure 9 for a 0.75x CTAB,

0.5x Fe sample:

Fe3O4 Fe

ϴ1=42.76 ϴ2=43.96

Figure 9: Peak deconvolution for 0.75x CTAB, 0.5x Fe sample.

From here, Fourier analysis is performed by Winfit on a single peak (at 43.3) to obtain

the sizing data. Here, the settings selected were 1024 data points and 5 angstrom

intervals. Specifically, the Scherrer formula can be used to approximate the average

particle size. This can be done by following the example procedure below[41]:
29

The Scherrer formula equation:

0.9
D
1/2 cos 

Here: 0.9 is the assumed shape factor, 1 / 2 is the half width of the XRD pattern taken at

its base, and theta is the angle which can be approximated by taking the average of the


two angles ϴ1 and ϴ2 at the base. Also,   1.54 A is the value used for the copper

radiation. As an example, for the Fe peak in Figure 9:

0.9
D
1/2 cos 

1   2 42.76  43.96
   43.36
2 2

43.96  42.76
1 / 2   2  1   0.6
2

 

0.91.54 A 
D   
 182.0 A  18.2nm
 π radians 
(0.6)  cos(43.36)
 180 degrees 
30

4. Experimental Results

4.1 Initial Results

The first important result obtained was confirmation that zero valent iron was being

made. To accomplish this, the typical cation and anion-containing solutions detailed in

the experimental section were used. However, no micelle-type process was adopted.

Rather, these solutions were prepared with water only. Then, a second sample was

prepared in a similar fashion, except that a very small amount of CTAB was added

relative to the iron in solution (1 mole%). This small addition of iron significantly

broadens the xrd reading of the sample, as can be seen in Figure 10. Further, higher

mole% samples were tried (5-25 mole%), with each showing no distinct pattern for Fe.

1 mole%
CTAB:Fe
Intensity (a.u.)

No CTAB

35 40 45 50 55 60 65 70 75 80 85
2θ (degrees)

Figure 10: Comparison of sample without CTAB vs. one with 1 mole% CTAB:Fe

To get distinct iron peaks when analyzing data, it became apparent that annealing

would be required. The XRD pattern in Figure 11 was a post-annealing result. Here, no
31

distinct pattern would have been recognizable without annealing, as the sample was

prepared at 25 mole% CTAB:Fe.

Figure 11: 25mole% CTAB:Fe sample, post-annealing

4.2 TEM Imaging

TEM results were obtained for several samples. However, with some contaminants in

the actual process (from CTAB that wasn’t removed), obtaining a perfect picture of

particle size couldn’t be achieved with TEM. The image in Figure 12, and its

corresponding particle sizes, is shown to provide at least some insight into particle size

distribution. XRD results, which were depended on to reach the final conclusions for the

present thesis work, do not provide such size distribution results.


32

Figure3:xxx:
Figure ZeroZero valent
valent ironiron nanoparticles
nanoparticles as seen
as seen through
through TEM.
TEM. ThisThis
waswas
for for
the the “standard
“standard solution,”
solution,” as as
highlightedininthe
highlighted theexperimental
experimentalsection
sectio

Particle count
80
70
60
50
number

40
30
20
10
0
5 10 15 20 25 30 35 40
particle size (nm)

Figure 4: Particle size distribution as determined from the TEM image of the “standard solution.”

As can be seen from the particle size distribution data in Figure 13, the bulk of

nanoparticles tended to skew to the size of 15 nm. However, given that this was for

roughly one hundred nanoparticles, and that there were millions actually present, this
33

result may not be accurate for the entirety of the sample. Further, since sizing had to be

done manually, some human error may have occurred in the counting process.

More relevant to the present situation, however, is an approximation of average particle

size, with relative polydisperisity determined from standard deviation data. Using the

standard experimental procedure detailed in the experimental methods section, the

following data in Table 1 was obtained from manually sizing TEM images with the help

of an imaging program called imagej:

Table 1: Particle sizes as determined from the aggregates of each image.

Image# Sample Particle Size (avg) Standard Deviation


M801 1x-1 11.39 7.31
M799 1x-2 11.50 3.24
M787 1.25x-3 13.36 5.15
M788 1.25x-3 15.40 6.36
M790 1.75x-1 10.27 4.79
M792 1.75x-1 11.44 4.15
M793 1.75x-2 10.15 4.23
M795 1.75x-2 10.79 3.64

As can be seen from the data above, particle size does not vary by a significant amount.

This was thought to be due to the magnetic stir bars, which possibly removed a portion

of particles larger than a certain size, skewing the data. Further, it is possible that there

were smaller particles within aggregated clusters that were not accounted for.

Attempts were made at dispersing the particles with sonication. Even sonicating the

particles for an hour did not disperse particle aggregates. To remedy the possible

problems caused by the stir bars, the T-mixer method was used later. To address

problems caused by the tendency of particles to aggregate, XRD-based particle sizing


34

was used. The XRD results in the next section are all for samples synthesized using the

T-mixer method.

4.3 XRD Results

For the reverse micelle process itself, peaks associated with NaBr and possibly Fe3O4 and

Fe2O3 in addition to zero valent Fe were observed. The peak of interest for all particle

sizing is taken to be at 43.3. For a complete comparison of peaks, refer to Appendix D.

For the actual XRD patterns, refer to Appendix E. While the presence of these other

compounds is less than optimal, it does provide a higher level of surface passivation,

making it possible to work with the iron sample in air.

In Figure 14, particle size information can be found as a function of the fraction of the

“standard amount” of Fe per the standard procedure discussed in the prior

experimental procedure section. The particle size trendline decreases as the amount of

Fe used for an otherwise “standard solution” reverse micelle experiment decreases.

The error was taken to be one standard deviation (1σ). The trendline (a second order

polynomial) is within the margin of error for all data points with the exception of 0.75x

Fe. This could be explained by the inaccuracy associated with fitting the XRD peaks

when making size measurements. All data used in the chart can be found in the table

following it (Table 2).


35

XRD Sizing Data With Trendline and Error Bars


1000
900

Particle Size (angstroms)


800
700
600
500
400
300
200
100
0
0 0.2 0.4 0.6 0.8 1
Fraction Fe

Figure 5: Sizing as determined by Fourier analysis using Integral Breadth for varying Fe concentrations,
with experimental triplicates taken at each data point.

Table 2: Experimental particle size results for each data point

Fraction Size Average Standard Deviation


1 919
1 730
779 123
1 687
0.75 825
0.75 807
787 52
0.75 728
0.5 230
0.5 405
439 228
0.5 683
0.25 229
0.25 389
438 237
0.25 695
0.1 390
0.1 382
334 91
0.1 229

4.4 Magnetic Characterization


36

All magnetic measurement readings were determined with the help of Han Song in the

department of electrical engineering at Oregon State. Initial tests were conducted with

epoxy as the matrix medium in which the iron was contained. Later, this material was

changed to KBr.

4.4.1 Magnetic Curing Effects

It was experimentally found that magnetic curing has little to no significant effect on the

sample. This is highlighted in Figure 15 below, which details the effects of curing on a

commercial NiFe2O4 magnetic nanomaterial from MTI Corporation, 30nm particle size.

Such curing involved taking an already vacuum cured epoxy sample and subjecting it to

a 0.4T magnetic field while drying. As can be seen from Figure 15, the effect is relatively

small, to such an extent that the differences could be attributed to being within the

margin of error for accuracy in producing and testing a given sample.


37

Real Part Permeability, various NiFe2O4 samples


1.40E+00

1.35E+00

1.30E+00

1.25E+00
Real Part Permeability

1.20E+00

1.15E+00

1.10E+00

1.05E+00

1.00E+00
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)

Real, 10wgt% no field real, 10 wgt% with field real, 25 wgt%, no field
real, 25wgt%, with field real, 50wgt%a no field real, 50wgt%a, with field
real, 50wgt%b with field real, 50wgt%b without field

Figure 6: Permeability Measurements on NiFe2O4 showing effects of magnetic curing.

4.4.2 Epoxy Samples

A variety of quality factor plots are provided in Appendix A showing the magnetic

characteristics of various epoxy composite samples. Here, the quality factor is taken as
38

the real part of permeability divided by the imaginary part of permeability. For all these

samples, there is a large degree of noise until about 200MHz. Further, these samples

have inconsistent quality factor values. The triplicate result at high frequencies (600

MHz - 1 GHz) at 1.5x surfactant has two at 300, and one at 400. The curves for each

quality factor graph do not tend to follow any type of general trend. The only exception

to this seems to be with the large toroid samples, which all have minimal noise and

follow the same trend, where they increase rapidly and then gradually decrease in

quality factor for increasing frequency. One outcome that did go as expected is how

NiFe2O4 compares against Fe samples, with most NiFe2O4 having a quality factor of less

than fifty at high frequencies (600 MHz-1GHz range) for both samples reported in

Appendix A, while all Fe samples have quality factors of typically higher than 100 at high

frequencies. Figure 16 is illustrative of another problem encountered when using epoxy

as a matrix material for a composite:

Figure 7: 60 weight percent zero valent iron composite permeability measurement


39

Even with the highest loading found to be possible with reasonable sample uniformity

(~60 weight percent Fe), the permeability readings for the real part of the material

indicate that it is very near a value of one. For this reason, higher weight percent iron

loading was investigated by using KBr as the matrix medium, along with pressure (via a

pellet press) to better compact the composite material.

4.4.3 KBr Composite Samples

All initial tests are summarized in Appendix C. Here, the magnetic readings all had real

parts of permeability near one. However, the imaginary component of permeability was

very small, leading to quality factors of roughly ~100. With the real part of permeability

close to one, the material was only weakly magnetic at higher frequencies. This was the

reason that the switch was made to KBr as a matrix medium. With the use of epoxy,

there was no clear way of increasing the weight fraction of iron in the composite

without a pressurizing device of some kind. Further, when pressure was applied,

difficulty was encountered determining a method to prevent the epoxy from sticking to

the pressurizing device. The use of KBr allowed for the use of a KBr press for

pressurization purposes. This effectively allowed for an increase of the weight load of

iron in a given sample.

To magnetically characterize KBr samples, toroids were formed as described in the

experimental section. Here, the toroids containing 90 weight% Fe samples mixed with

10 weight% KBr were analyzed for permeability over frequencies up to 109 Hz. Table 3

highlights the measurements taken. Here, a “standard solution” was prepared, as

detailed in the experimental section, with differing amounts of iron. The table also
40

includes re-test values from five weeks later, which will be discussed in the next section.

All values given are the average for high frequencies (600-1000 MHz).

As can be seen from the table, all samples synthesized had a near-100 quality factor as

the minimal outcome. Unfortunately, this does not have a direct correlation with

particle size. This is thought to be mostly due to possible instrumental error. Given that

the imaginary component for all three is very small (e.g. ~0.01), the instrument may have

issues with accurately measuring permeability values very near zero.

Table 3: Quality factors for differing quantities of Fe

Quality Factors

Quality Factor at 600- Retest, Quality Factor


Fraction Fe 1000 MHz 600-1000 MHz

1 146 951

0.75 322 317

0.5 131 102

Further, it is instructive to look at a set of permeability readings across frequency.

Specifically, the permeability readings for the standard KBr toroid (no Fe load) were

chosen as an example of a typical reading. From these, an understanding of the

limitations of the instrument can be gained.For the imaginary part in Figure 17, the

permeability is very nearly zero, as would be expected specifically for the standard.

However, a large degree of noise is evident up until roughly 100 MHz. Since these high

noise, lower frequency regions are not in the region of interest (1 GHz), noise is not a

major issue in the present study.


41

Pure KBr standard, imaginary part

Permeability, Imaginary Part


1.50E-02

5.00E-03

-5.00E-03
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
-1.50E-02

-2.50E-02

-3.50E-02
Frequency (Hz)

Figure 8: Pure KBr standard permeability reading, imaginary part

In Figure 18, the real part of permeability trends slightly up as frequency increases. This

may be explained as instrument error. Again, there is a large amount of noise up to

roughly 100 MHz. The trends found for the standard tend to apply for the remaining

samples as well. For the remaining readings, refer to Appendix B.

Pure KBr standard, Real Part


1.03E+00
Permeability, Real Part

1.02E+00
1.01E+00
1.00E+00
9.90E-01
9.80E-01
9.70E-01
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)

Figure 9: Pure KBr standard permeability reading, real part

The presence of a high amount of noise at lower frequencies is the greatest problem for

obtaining a clear picture of how realistic the magnetic readings are. The material is

known to be magnetic. Therefore, a large real part of permeability is expected at lower

frequencies. Unfortunately, no clear way of fixing this issue has presently been found.
42

4.4.4 KBr Composite Samples Re-tested (5 Weeks Later)

To ensure that all experimental work was reproducible, and that gradual oxidation over

time was not a major concern, the KBr composites made were re-tested. The real and

imaginary components of permeability for both the original test and re-test of the KBr

sample are provided below in Figure 19 and Figure 20.

KBr, Real Part Permeability


1.02E+00
1.02E+00
1.01E+00
1.01E+00
1.00E+00
9.95E-01
9.90E-01
9.85E-01
9.80E-01
9.75E-01
9.70E-01
9.65E-01
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08

KBr Permeability Re-test, Real Part Original KBr Permeability Test, Real Part

Figure 10: KBr Real Part Permeability Values


43

KBr, Imaginary Part Permeability


7.00E-03
6.00E-03
5.00E-03
4.00E-03
3.00E-03
2.00E-03
1.00E-03
0.00E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08

KBr Permeability Re-test, Imaginary Part


Original KBr Permeability Test, Imaginary Part

Figure 20: KBr Imaginary Part Permeability Values

Here, the real part of permeability is very close to one in both cases, as would be

expected for a non-magnetic material. Further, the imaginary component in both cases

is very close to zero, which is also expected.

The three KBr iron composites were also re-tested. The corresponding permeability

values are compared in Figure 21 and Figure 22 below. Here, the 1x Fe sample

encounters significant differences in permeability values. Its value falls to values similar

to the KBr sample for its real component, indicating that over time it became

nonmagnetic. This could be an effect of oxidation. However, the other two samples do

not encounter this same effect, maintaining their magnetic properties. This is reflected

by very similar quality factors between the two measurements. For quality factors for

all re-tests, see Appendix C.


44

1x Fe Permeability, Real Part


1.10E+00
Permeability, real part
1.05E+00

1.00E+00

9.50E-01

9.00E-01
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

1x Fe, test values 5 weeks later 1x Fe, original test values

Figure 21: 1x Fe, real part permeability as a function of frequency

1x Fe Permeability, Imaginary Part


4.00E-02
3.00E-02
2.00E-02
1.00E-02
0.00E+00
1.00E+07
-1.00E-02 2.10E+08 4.10E+08 6.10E+08 8.10E+08
-2.00E-02
-3.00E-02
-4.00E-02
Frequency (Hz)

1x Fe, test values 5 weeks later 1x Fe, original test values

Figure 22: 1x Fe, imaginary part of permeability measurement as a function of frequency

With a basic understanding of oxidation effects on the samples, instrument repeatability

was another factor that needed consideration when determining whether measurement

values are accurate.


45

4.4.5 Assessment of Instrument Repeatability

In an attempt to quantify variance after observing the considerable noise for the lower

scanning frequencies and the overall relatively low permeability readings for the real

component of samples, repeatability of measurement values for the instrument was

addressed. Given below in Figure 23 and Figure 24 is the KBr sample, re-measured five

times in succession (with the instrument being re-calibrated for each measurement).

KBr, real part permeability for five replicates


1.025
1.02
1.015
Real Part Permeability

1.01
1.005
1
0.995
0.99
0.985
0.98
0.975
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

KBr, 1st replicate KBr, 2nd replicate KBr, 3rd replicate


KBr, 4th replicate KBr, 5th replicate

Figure 23: KBr real part permeability for five replicates


46

KBr , imaginary part permeability for five


replicates
0.025
0.02
Imaginary part permeability

0.015
0.01
0.005
0
1.00E+07
-0.005 2.10E+08 4.10E+08 6.10E+08 8.10E+08
-0.01
-0.015
-0.02
-0.025
-0.03
Frequency (Hz)

KBr, 1st replicate KBr, 2nd replicate KBr, 3rd replicate


KBr, 4th replicate KBr, 5th replicate

Figure 24: KBr imaginary part permeability for five replicates

As can be seen from the values, the real part of permeability varies very little, whereas

the imaginary part varies substantially in the lower frequencies (up to about 200 MHz),

while still having some deviations in the higher frequency region. This is quantitatively

represented by the standard deviations given in table 4 below, which is taken from the

average permeability value of the 600 MHz to 1 GHz values for each sample.

Table 4: KBr real and imaginary part averages and standard deviation between measurements
KBr
real imag
5th 1.00E+00 -1.29E-03
4th 1.00E+00 9.35E-04
3rd 1.00E+00 1.44E-04
2nd 1.00E+00 -3.39E-04
1st 1.00E+00 2.89E-04
average 1.00E+00 -5.31E-05
stdev 4.62E-04 8.30E-04
47

Repeatability measurements were then taken for the 1x Fe sample. This sample was

characteristically very similar to the KBr sample. Figures 25 and 26 below indicate the

reproducibility for the readings of the real and imaginary components of permeability.

As can be seen, the differences are very similar to those found for the KBr samples.

1x Fe real part permeability for 3 replicates


1.04
1.03
Real part permeability

1.02
1.01
1
0.99
0.98
0.97
0.96
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

1x Fe, 1st replicate 1x Fe, 2nd replicate 1x Fe, 3rd replicate

Figure 11: 1x Fe real part permeability for 3 replicates


48

1x Fe, imaginary part permeability for 3


replicates
4.00E-02

3.00E-02
Imaginary part permeability

2.00E-02

1.00E-02

0.00E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
-1.00E-02

-2.00E-02

-3.00E-02

-4.00E-02
Frequency (Hz)

1x Fe, 1st replicate 1x Fe, 2nd replicate 1x Fe, 3rd replicate

Figure 12: 1x Fe imaginary part permeability for 3 replicates

Table 5 below provides a quantitative approximation of standard deviation using the

same analytical process described for the KBr sample. The variation for the real

component of permeability is much smaller in magnitude compared to the reading (by a

factor of ~2000). By comparison, the imaginary component has a standard deviation of

the same order of magnitude as both the Fe and the KBr sample measurements. This

would indicate that the reading of the imaginary component of permeability is the less

reliable of the two.

Table 5: 1x Fe real and imaginary part averages and standard deviation between measurements

1x Fe

real imag
49

3rd 1.01E+00 9.13E-04

2nd 1.01E+00 2.22E-03

1st 1.01E+00 2.57E-03

average 1.01E+00 1.90E-03

stdev 1.94E-03 8.72E-04

4.4.6 Oxidized KBr Composite Data

One sample of oxidized iron was made into a 90 weight percent composite with KBr as

the matrix medium. An oxide material for characterization was made as detailed by the

procedure for making iron oxides section in the experimental section. The results are

summarized in Figure 27 and Figure 28, showing XRD and permeability data for the

oxide sample. Here, the XRD peaks are consistent with Fe3O4. Also, it should be noted

that the average quality factor at high frequencies is about 30, which does not compare

favorably against the typical quality factor of 100 for all iron composite samples. This

indicates that the creation of zero valent iron was pivotal to getting a higher quality

factor.
50

Oxidized Fe peaks
600
Fe3O4 Fe3O4 Fe3O4 Fe3O4 Fe3O4 Fe3O4
500

400
Intensity

300

200

100

0
20 30 40 50 60 70 80
2-theta

Figure 13: Oxidized iron XRD peaks for annealed composite sample, with Fe3O4 peaks indicated.

Iron Oxide from Fe (II), annealed


100
90
80
70
Quality Factor

60
50
40
30
20
10
0
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

Figure 14: Quality Factor values for iron oxide composite

Also provided in Figure 29 and Figure 30 is the real and imaginary parts of permeability.

While the real component is slightly higher at higher frequencies than the typical Fe

sample (~1.25), the imaginary part is substantially higher (0.05). This is what leads to

the quality factor that is significantly less than the typical value of 100.
51

Iron Oxide, Imaginary Part


Permeability
8.00E-02
Permeability, Imaginary Part

6.00E-02

4.00E-02

2.00E-02

0.00E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
-2.00E-02
Frequency (Hz)

Figure 15: Iron oxide magnetic measurement, imaginary part of permeability

Iron Oxide, Real Part Permeability


1.28E+00
1.27E+00
Permeability, Real Part

1.26E+00
1.25E+00
1.24E+00
1.23E+00
1.22E+00
1.21E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

Figure 30: Iron oxide magnetic measurement, real part of permeability


52

5. Conclusions and Future Work

Initial XRD results of the zero valent iron nanoparticle synthesis process confirmed that

zero valent iron was formed, and indicated that annealing would be important to

achieve the sample crystallinity necessary to get XRD readings. Particle size

characterization was determined via XRD. From the XRD data, particle size was found to

vary between ~25 nm to ~90nm, depending on the concentration of iron in solution. At

the lower end of iron concentration, the synthesized ~25 nm average size is close to the

desired 10-20 nm particle size range. As the concentration of iron decreased, the

particle size decreased.

Quality factors were determined by dividing the real part by the imaginary part of

permeability for each sample. These factors were higher for the zero valent iron

nanoparticle composite materials than for both the NiFe2O4 and Fe3O4 composite

materials. The zero valent iron composites operated at a typical quality factor of about

100 or higher, whereas the iron oxide composite sample had a quality factor of about 30

and the NiFe2O4 composite samples had quality factors of less than fifty. Epoxy and KBr

were used as matrix materials for composite samples, but other alternatives could be

implemented to possibly contribute to a higher degree of passivation. A polymer-type

coating in the synthesis process could possibly achieve this. Further, one possible goal

for future work would be to achieve a higher degree of iron nanoparticle sample purity.

While the cause of impurities is unclear, it could be chemical interactions of iron with

CTAB or some other component. Experimenting with pH control may guarantee that Fe
53

hydroxides and oxides are not forming in solution, and at least partially mitigate

currently encountered purity issues.

Also, a high degree of oxidation was experienced for all samples. For this reason,

particle sizing took place for particles that were partially oxidated in addition to zero

valent iron particles. While zero valent iron was present in the samples, logistically, it

made more sense to size the mixed materials to simplify synthetic work. Regardless, the

sizing with Winfit can only be taken to be an approximation, and, as such, the valuable

information obtained from the set of experiments is that size control can be achieved by

adjusting Fe concentration.

TEM results also provide some insight into particle sizes, with many falling into the 10-

20nm regime considered to be useful. The XRD sizing results were bigger, but for

smaller concentrations of Fe, sample average particle size fell close to the desired size

regime (~25 nm). Given that the particles were often obscured in aggregate clusters in

the TEM images, it is very possible that XRD was the more accurate sizing method of the

two.

Finally, the magnetic composites made do not exhibit high permeability values at high

frequencies based on instrument readings. However, the commercially supplied nickel

ferrite material is known to be highly magnetic (e.g. ~1500 at low frequencies). Here,

instrument real part permeability readings between one and two exist for the entire

range of frequency values (from 107 to 109 Hz). Therefore, the instrument readings

should be considered with some skepticism. A calibration factor in the future may help

to remedy this problem. Further, this could be due to the presence of NaBr in the
54

material. Because of the lack of variation in permeability values, no conclusions can be

made as to whether a KBr or epoxy composite matrix is better. Regardless, these

materials did provide high quality factors and could have practical uses in an inductor

application. In the future, the use of an instrument more sensitive to lower magnetic

permeability readings at high frequencies may be best to ensure that the results

obtained in this thesis are accurate. Additionally, this may help to reduce the

uncertainty currently experienced as to the accuracy of magnetic measurement results

due to high amounts of noise at lower frequencies below 200 MHz.


55

Appendix A: Quality Factors for Epoxy Composites

Please note: all samples labeled “aqueous prep.” followed a procedure as follows: 1.5

grams ferrous ammonium sulfate + 1.32 grams CTAB + 75 mL DIW, with 0.3 grams

sodium borohydride in 15 mL DIW added to it.

Quality Factor: 20 wgt% Fe Large


Toroid, aqueous prep, 25-pt average
(with field)
8.00E+01
Quality Factor

6.00E+01
4.00E+01
2.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: small toroid 35 wgt%


Fe sample 25-pt boxcar average,
aqeuous prep.
1.00E+03
Quality Factor

5.00E+02

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
56

Quality Factor: 33 wgt% Fe small


toroid 25-pt boxcar average, aqueous
prep.
1.00E+03
Quality Factor

5.00E+02

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 Large Toroid


20wgt% (field)
1.00E+02
Quality Factor

5.00E+01

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: 35wgt% NiFe2O4 small


toroid
5.00E+02
4.00E+02
3.00E+02
Quality Factor

2.00E+02
1.00E+02
0.00E+00
1.00E+06
-1.00E+02 2.01E+08 4.01E+08 6.01E+08 8.01E+08
-2.00E+02
-3.00E+02
Frequency (Hz)
57

Quality Factor: 35 wgt% Permalloy


small toroid 25-pt average
1.00E+03
Quality Factor

8.00E+02
6.00E+02
4.00E+02
2.00E+02
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: Permalloy small


toroid 30 wgt% 25-pt average
1.00E+03
Quality Factor

5.00E+02

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: 0.75x surfactant 10


wgt% Fe, 25-pt average
3.00E+02
2.50E+02
Quality Factor

2.00E+02
1.50E+02
1.00E+02
5.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
58

Quality Factor: 1.5x Surfactant


10wgt% Fe 25-pt average
2.00E+03
Quality Factor

1.50E+03
1.00E+03
5.00E+02
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: 1.5x surfactant,


35wgt% replicate 1 25-pt average
6.00E+02
Quality Factor

4.00E+02
2.00E+02
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: 1.5x surfactant


replicate 2, 35wgt% Fe 25-pt average
6.00E+03
Quality Factor

4.00E+03

2.00E+03

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
59

Quality Factor: 1.5x surfactant


replicate 3, 35wgt% Fe 25-pt average
5.00E+02
Quality Factor

4.00E+02
3.00E+02
2.00E+02
1.00E+02
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: 35wgt% NiFe2O4


epoxy Composite large toroid
4.00E+01
Quality Factor

3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

10 wgt% NiFe2O4, large toroid (no


field)
1.20E+02
1.00E+02
Quality Factor

8.00E+01
6.00E+01
4.00E+01
2.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
60

Quality Factor: NiFe2O4 50wgt% small


toroid (with field)
4.00E+01
Quality Factor

3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 50 wgt%


small toroid (no field)
5.00E+01
Quality Factor

4.00E+01
3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 35wgt% (with


field) small toroid
5.00E+01
Quality Factor

4.00E+01
3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
61

Quality Factor: NiFe2O4 35wgt% (no


field) small toroid
4.00E+01
Quality Factor

3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 20wgt% small


toroid (no field)
1.00E+03
Quality Factor

5.00E+02

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 50wgt% large


toroid (with field) rep. 2
5.00E+01
Quality Factor

4.00E+01
3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
62

Quality Factor: NiFe2O4 50wgt% large


toroid (with field)
5.00E+01
Quality Factor

4.00E+01
3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality factor: NiFe2O4 50 wgt% large


toroid rep. 2 (no field)
4.00E+01
Quality Factor

3.00E+01

2.00E+01

1.00E+01

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality factor: NiFe2O4 50 wgt% (no


field) large toroid
5.00E+01
4.00E+01
Quality Factor

3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
63

Quality Factor: 35 wgt% NiFe2O4, rep.


2 (no field)
4.00E+01
Quality Factor

3.00E+01
2.00E+01
1.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 25wgt% (with


field) large toroid
1.00E+02
Quality Factor

8.00E+01
6.00E+01
4.00E+01
2.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)

Quality Factor: NiFe2O4 Large Toroid


25 wgt% (no field)
1.00E+02

8.00E+01
Quality Factor

6.00E+01

4.00E+01

2.00E+01

0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
64

10wgt% NiFe2O4, large toroid (with


field)
1.00E+02
8.00E+01
Quality Factor

6.00E+01
4.00E+01
2.00E+01
0.00E+00
1.00E+06 2.01E+08 4.01E+08 6.01E+08 8.01E+08
Frequency (Hz)
65

Appendix B: Magnetic Composite Test Results, KBr

1.0x Fe, Imaginary component


7.00E-02
Permittivity, imaginary

6.00E-02
5.00E-02
component

4.00E-02
3.00E-02
2.00E-02
1.00E-02
0.00E+00
0.00E+00 5.00E+08 1.00E+09
Frequency (Hz)

1.0x Fe, Real Component


1.10E+00
Permittivity, Real

1.05E+00
Component

1.00E+00
9.50E-01
9.00E-01
0.00E+00 5.00E+08 1.00E+09
Frequency (Hz)

0.75x Fe, Imaginary Component


4.00E-02
Permittivity (Imaginary

3.00E-02
2.00E-02
Component)

1.00E-02
0.00E+00
0.00E+00
-1.00E-02 5.00E+08 1.00E+09
-2.00E-02
-3.00E-02
Frequency (Hz)
66

0.75x Fe, Real Component


1.10E+00
Permittivity (Real Component) 1.08E+00
1.06E+00
1.04E+00
1.02E+00
1.00E+00
9.80E-01
9.60E-01
0.00E+00 5.00E+08 1.00E+09
Frequency (Hz)

0.5x Fe, Imaginary Component


4.00E-02
Permittivity, Imaginary

3.00E-02
Component

2.00E-02
1.00E-02
0.00E+00
0.00E+00
-1.00E-02 5.00E+08 1.00E+09
-2.00E-02
Frequency (Hz)

0.5x Fe, Real Component


1.08E+00
Permittivity (Real Component)

1.07E+00
1.06E+00
1.05E+00
1.04E+00
1.03E+00
1.02E+00
1.01E+00
1.00E+00
0.00E+00 5.00E+08 1.00E+09
Frequency (Hz)
67

Appendix C: Quality Factor Graphs, KBr

Using 25-point boxcar moving average:

Quality Factor, KBr


1.00E+04

8.00E+03
Quality factor

6.00E+03

4.00E+03

2.00E+03

0.00E+00
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)

Quality Factor, KBr re-test


1.40E+04
1.20E+04
Quality Factor

1.00E+04
8.00E+03
6.00E+03
4.00E+03
2.00E+03
0.00E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)
68

Quality Factor, 0.5x Fe


6.00E+02
Quality Factor 5.00E+02
4.00E+02
3.00E+02
2.00E+02
1.00E+02
0.00E+00
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)

0.5x Fe result retest


180
160
140
120
100
80
60
40
20
0
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08

Quality Factor, 0.75x Fe


6.00E+02
5.00E+02
Quality Factor

4.00E+02
3.00E+02
2.00E+02
1.00E+02
0.00E+00
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)
69

0.75x Fe, retest


1.00E+03
9.00E+02
8.00E+02
7.00E+02
Quality Factor

6.00E+02
5.00E+02
4.00E+02
3.00E+02
2.00E+02
1.00E+02
0.00E+00
1.00E+07 2.10E+08 4.10E+08 6.10E+08 8.10E+08
Frequency (Hz)

Quality Factor, 1.0x Fe


6.00E+02
5.00E+02
Quality Factor

4.00E+02
3.00E+02
2.00E+02
1.00E+02
0.00E+00
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)

1x Fe sample Result, retest


10,000
9,000
8,000
7,000
Quality Factor

6,000
5,000
4,000
3,000
2,000
1,000
0
0.00E+00 2.00E+08 4.00E+08 6.00E+08 8.00E+08 1.00E+09
Frequency (Hz)
70

Appendix D: XRD Peaks

All peaks for Fe2O3, Fe3O4, and alpha-Fe are taken from Xiaomin et al.[1-3]
71

Appendix E: XRD Results

0.5x Fe, 1x surf. rep. 3, 5-pt


3500
3000
2500
2000
Intensity

1500
1000
500
0
-500 25 35 45 55 65
-1000
2-theta

0.10x Fe, 1x surf. rep. 1, 5-pt


3000

2000
Intensity

1000

0
25 35 45 55 65
-1000
2-theta

0.1x Fe, 1x surf. rep. 2, 5-pt


3000
2500
2000
1500
Intensity

1000
500
0
-500 25 30 35 40 45 50 55 60 65 70
-1000
2-theta
72

0.10x Fe, 1x surf., rep.3, 5-pt


4000
3500
3000
2500
Intensity

2000
1500
1000
500
0
-500 25 35 45 55 65
2-theta

0.25x Fe, 1x surf., rep. 1, 5-pt


7000
6000
5000
4000
Intensity

3000
2000
1000
0
-1000 25 30 35 40 45 50 55 60 65 70
2-theta

0.25x Fe, 1x surf., rep.2, 5-pt


7000
6000
5000
4000
Intensity

3000
2000
1000
0
-1000 25 30 35 40 45 50 55 60 65 70
2-theta
73

0.25x Fe, 1x surf., rep. 3, 5-pt


8000
7000
6000
5000
Intensity

4000
3000
2000
1000
0
-1000 25 30 35 40 45 50 55 60 65 70
2-theta

1x Fe, 1x surf., rep. 3, 5-pt


3500
3000
2500
2000
Intensity

1500
1000
500
0
-500 25 30 35 40 45 50 55 60 65 70
2-theta

0.75x Fe, 1x surf., rep. 3, 5-pt


3000
2500
2000
Intensity

1500
1000
500
0
25 30 35 40 45 50 55 60 65 70
-500
2-theta
74

0.75x surf., 0.5x Fe, rep. 2, 5-pt


1500

1000
Intensity

500

0
25 35 45 55 65
-500
2-theta

0.75x surf., 0.5x Fe, rep. 1, 5-pt


1200
1000
800
600
Intensity

400
200
0
-200 25 30 35 40 45 50 55 60 65 70
-400
-600
2-theta

0.75x Fe, 1.0x surf., rep. 1, 5-pt


2000

1500
Intensity

1000

500

0
25 30 35 40 45 50 55 60 65 70
-500
2-theta
75

0.75x Fe, 1x surf., rep. 2, 5-pt


1400
1200
1000
800
Intensity

600
400
200
0
-200 25 35 45 55 65
2-theta

1.0x surf., 0.5x Fe, rep. 1, 5-pt


5000

4000
Intensity

3000

2000

1000

0
25 35 45 55 65
2-theta

1.0x surf., 0.5x Fe, rep. 2, 5-pt


2000

1500

1000
Intensity

500

0
25 30 35 40 45 50 55 60 65 70
-500
2-theta
76

1.25x surf., 0.5x Fe, rep. 1, 5-pt


2000

1500

1000
Intensity

500

0
25 30 35 40 45 50 55 60 65 70
-500

-1000
2-theta

1.25x surf., 0.5x Fe, rep. 2, 5-pt


2000

1500
Intensity

1000

500

0
25 35 45 55 65
-500
2-theta
77

Works Cited

[1] H. Greve, et al., "Nanostructured magnetic Fe-Ni-Co/Teflon multilayers for high-

frequency applications in the gigahertz range," Applied Physics Letters, vol. 89,

pp. -, Dec 11 2006.

[2] D. R. Lide, CRC Handbook of Chemistry and Physics, 82nd ed., CRC Press, Boca

Raton, FL, 2001.

[3] D. L. Huber, "Synthesis, properties, and applications of iron nanoparticles,"

Small, vol. 1, pp. 482-501, May 2005.

[4] F. Sato, et al., "All-in-one package ultracompact micropower module using thin-

film inductor," Ieee Transactions on Magnetics, vol. 40, pp. 2029-2031, Jul 2004.

[5] P. Galle, et al., “Ultra-compact power conversion based on a CMOS-compatible

microfabricated

power inductor with minimized core losses,” Proc. 57th Electronic Components

and Technology Conference, 2007. ECTC ‘07.

[6] S.C.O. Mathuna, et al., “Magnetics on silicon: an enabling technology for power

supply on chip,” IEEE Trans. Power Electronics 2005, 20, pp. 585-592.

[7] A. Ludwig, et al., “Magnetoelastic thin films for high-frequency applications,”

IEEE Trans. Magn. 2001, 37, p.2690

[8] V. Korenivski and R. B. van Dover, “Design of high frequency inductors based on

magnetic films,” IEEE Trans. Magn., 1998, 34, 1375-1377.

[9] B. D. Cullity and C.D. Graham, Introduction to Magnetic Materials, 2 nd ed., IEEE

Press, Piscataway, NJ, 2009, pp. 18-19, 360-361, 465.


78

[10] S. Chikazumi and S. Charap, Physics of Magnetism, New York: John Wiley & Sons,

Inc., p. 17

[11] H. Nalwa, Handbook on Thin Film Materials, vol. 5, Academic Press, San Diego,

CA, 2002, p. 504.

[12] L.-M. Lacroix et al., “Ultrasmall iron nanoparticles: Effect of size reduction on

anisotropy and magnetization,” Journal of Applied Physics, vol. 103, 2008 p.

07D521.

[13] M. Getzlaff, Fundamentals of Magnetism, Springer, New York: 2008,

pp. 133-140.

[14] W. L. Pei, et al., "Controlled monodisperse Fe nanoparticles synthesized by

chemical method," Ieee Transactions on Magnetics, vol. 41, pp. 3391-3393, Oct

2005.

[15] M. N. Nadagouda and R. S. Varma, "A greener synthesis of core (Fe, Cu)-shell

(An, Pt, Pd, and Ag) nanocrystals using aqueous vitamin C," Crystal Growth &

Design, vol. 7, pp. 2582-2587, Dec 2007.

[16] C. S. Kuroda, et al., ""Ferrite route" preparation of iron nanoparticles and their

encapsulation in styrene-GMA co-polymer for biomedical applications," Ieee

Transactions on Magnetics, vol. 42, pp. 3569-3571, Oct 2006.

[17] K. C. Huang and S. H. Ehrman, "Synthesis of iron nanoparticles via chemical

reduction with palladium ion seeds," Langmuir, vol. 23, pp. 1419-1426, Jan 30

2007.

[18] I. Capek, "Preparation of metal nanoparticles in water-in-oil (w/o)


79

microemulsions," Advances in Colloid and Interface Science, vol. 110, pp. 49-74,

Jun 30 2004.

[19] C. Y. Wang, et al., "The preparation, surface modification and characterization of

metallic alpha-Fe nanoparticles (vol 60, pg 223, 1999)," Materials Science and

Engineering B-Solid State Materials for Advanced Technology, vol. 67, pp. 153-

153, Dec 15 1999.

[20] F. Li, et al., "Microemulsion and solution approaches to nanoparticle iron

production for degradation of trichloroethylene," Colloids and Surfaces a-

Physicochemical and Engineering Aspects, vol. 223, pp. 103-112, Aug 21 2003.

[21] T. Kinoshita, et al., "Magnetic evaluation of nanostructure of gold-iron composite

particles synthesized by a reverse micelle method," Journal of Alloys and

Compounds, vol. 359, pp. 46-50, Sep 22 2003.

[22] S. J. Cho, et al., "Growth mechanisms and oxidation resistance of gold-coated

iron nanoparticles," Chemistry of Materials, vol. 17, pp. 3181-3186, Jun 14 2005.

[23] J. Lin, et al., "Gold-coated iron (Fe@Au) nanoparticles: Synthesis,

characterization, and magnetic field-induced self-assembly," Journal of Solid

State Chemistry, vol. 159, pp. 26-31, Jun 2001.

[24] P. Ayyub, et al., "Microstructure of the Ctab-Butanol-Octane-Water

Microemulsion System - Effect of Dissolved Salts," Journal of the Chemical

Society-Faraday Transactions, vol. 89, pp. 3585-3589, Oct 7 1993. Permanent

URL link: http://dx.doi.org/10.1039/FT9938903585


80

[25] O. Margeat, et al., "Synthesis of iron nanoparticles: Size effects, shape control

and organisation," Progress in Solid State Chemistry, vol. 33, pp. 71-79, 2005.

[26] C. T. Seip and C. J. O'Connor, "The fabrication and organization of self-assembled

metallic nanoparticles formed in reverse micelles," Nanostructured Materials,

vol. 12, pp. 183-186, Jul 1999.

[27] S. J. Cho, et al., "Characterization and magnetic properties of core/shell

structured Fe/Au nanoparticles," Journal of Applied Physics, vol. 95, pp. 6804-

6806, Jun 1 2004.

[28] I. Capek, "Preparation of metal nanoparticles in water-in-oil (w/o)

microemulsions," Advances in Colloid and Interface Science, vol. 110, pp. 49-74,

Jun 30 2004.

[29] E. E. Carpenter, et al., "Passivated iron as core-shell nanoparticles," Chemistry of

Materials, vol. 15, pp. 3245-3246, Aug 26 2003.

[30] F. Faupel, et al., "Functional Polymer Nanocomposites," Polymers & Polymer

Composites, vol. 16, pp. 471-481, 2008.

[31] S. I. Nikitenko, et al., "Synthesis of air-stable iron-iron carbide nanocrystalline

particles showing very high saturation magnetization," Ieee Transactions on

Magnetics, vol. 38, pp. 2592-2594, Sep 2002.

[32] N. A. D. Burke, et al., "Magnetic nanocomposites: Preparation and

characterization of polymer-coated iron nanoparticles," Chemistry of Materials,

vol. 14, pp. 4752-4761, Nov 2002.


81

[33] H. Shokrollahi and K. Janghorban, "Soft magnetic composite materials (SMCs),"

Journal of Materials Processing Technology, vol. 189, pp. 1-12, Jul 6 2007.

[34] J. Wagner, et al., "Microfluidic generation of metal nanoparticles by borohydride

reduction," Chemical Engineering Journal, vol. 135, pp. S104-S109, Jan 15 2008.

[35] T. Calliot, et al., "Microwave hydrothermal flash synthesis of nanocomposites Fe-

Co alloy/cobalt ferrite," Journal of Solid State Chemistry, vol. 177, pp. 3843-3848,

Oct 2004.

[36] S. I. Nikitenko, et al., "Synthesis of highly magnetic, air-stable iron iron carbide

nanocrystalline particles by using power ultrasound," Angewandte Chemie-

International Edition, vol. 40, pp. 4447-+, 2001.

[37] G. Cao, “Nanostructures & Nanomaterials,” pp. 63-67, Imperial College Press,

Covent Garden, London, 2008.

[38] E. M. A. Jamal, et al., "Synthesis of nickel-rubber nanocomposites and evaluation

of their dielectric properties," Materials Science and Engineering B-Advanced

Functional Solid-State Materials, vol. 156, pp. 24-31, Jan 25 2009.

[39] E. Bayramli, et al., "Powder metal development for electrical motor

applications," Journal of Materials Processing Technology, vol. 161, pp. 83-88,

Apr 10 2005.

[40] T. J. Fiske, et al., "Percolation in magnetic composites," Journal of Materials

Science, vol. 32, pp. 5551-5560, Oct 15 1997.

[41] Klux and Alexander, X-ray Diffraction Procedures for Polycrystalline and

Amorphous Materials, New York: John Wiley & Sons, Inc., 1954, pp. 511-513
82

S-ar putea să vă placă și