133
Rational Synthesis of Nanomagnets and Magnetic Nanocomposites for Permanent Magnet Applications By Bo Shen B.Sc., Nankai University, 2013 M.A., Brown University, 2015 A Dissertation Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy in the Department of Chemistry at Brown University Providence, Rhode Island May 2019

Rational Synthesis of Nanomagnets and Magnetic

  • Upload
    others

  • View
    21

  • Download
    0

Embed Size (px)

Citation preview

Rational Synthesis of Nanomagnets and Magnetic Nanocomposites

for Permanent Magnet Applications

By

Bo Shen

B.Sc., Nankai University, 2013

M.A., Brown University, 2015

A Dissertation Submitted in Partial Fulfillment of the

Requirements for the Degree of Doctor of Philosophy

in the Department of Chemistry at Brown University

Providence, Rhode Island

May 2019

© Copyright 2019

by

Bo Shen

iii

This dissertation by Bo Shen is accepted in its present form

by the Department of Chemistry as satisfying the dissertation requirement

for the degree of Doctor of Philosophy.

Date ______________ _____________________

Shouheng Sun, Advisor

Recommended to the Graduate Council by

Date ______________ _____________________

Eunsuk Kim, Reader

Date ______________ _____________________

Lai-Sheng Wang, Reader

Approved by the Graduate Council

Date ______________ _____________________

Andrew G. Campbell, Dean of the Graduate School

iv

Curriculum Vitae

Bo Shen was born on June 15, 1990, in Zhangjiakou, Hebei, China. He grew up in

Tianjin and studied in Nankai University (Tianjin, China) from 2009 to 2013,

graduating with obtaining B.Sc. degree in Chemistry. In 2013, he was admitted with a

fellowship in the graduate program of the Department of Chemistry at Brown

University, pursuing the degree of Doctor of Philosophy in Chemistry. During this time,

he worked as research and teaching assistants. Since January 2014, he has been focusing

on the rational synthesis of nanomagnets and magnetic nanocomposites for permanent

magnet applications under the supervision of Professor Shouheng Sun. He has 12

papers published thus far in peer-reviewed journals and 1 patent in pending.

v

Publications

[13] Bo Shen, Chao Yu, Scott K. McCall, Zhouyang Yin and Shouheng Sun*, “A

general way to Synthesize Sm based Nanomagnets” 2018, in preparation.

[12] Bo Shen, Chao Yu, Dong Su, Zhouyang Yin, Junrui Li, Zheng Xi and Shouheng

Sun*, “A novel Approach to Anisotropic SmCo5 Nanomagnets” Nanoscale, 2018, 10,

8735-8740.

[11] Bo Shen, Adriana Mendoza-Garcia, Sarah E. Baker, Scott K. McCall, Chao Yu,

Liheng Wu and Shouheng Sun*, “Stabilizing Fe Nanoparticles in the SmCo5 Matrix”

Nano Lett., 2017, 17, 5695–5698.

[10] Junrui Li, Zheng Xi, Jacob S. Spendelow, Paul N. Duchesne, Dong Su, Qing Li,

Chao Yu, Zhouyang Yin, Bo Shen, Yu Seung Kim, Peng Zhang and Shouheng Sun*,

“Ordered Intermetallic Core/Shell FePt/Pt Nanoparticle with Atomic Layers of Pt as

Highly Active and Durable Oxygen Reduction Catalyst Utilized for Fuel Cells” J. Am.

Chem. Soc., 2018, 140, 2926–2932.

[9] Chao Yu, Xuefeng Guo, Mengqi Shen, Bo Shen, Michelle Muzzio, Zhouyang

Yin, Qing Li, Zheng Xi, Junrui Li, Christopher T. Seto* and Shouheng Sun*,

“Maximizing the Catalytic Activity of Nanoparticles through Monolayer Assembly on

Nitrogen-Doped Graphene” Angew. Chem. Int. Ed. 2018, 57, 451-455.

[8] Chao Yu, Xuefeng Guo, Zheng Xi, Michelle Muzzio, Zhouyang Yin, Bo Shen,

Junrui Li, Christopher T. Seto* and Shouheng Sun*, “AgPd Nanoparticles Deposited

on WO2.72 Nanorods as an Efficient Catalyst for One-Pot Conversion of

Nitrophenol/Nitroacetophenone into Benzoxazole/Quinazoline” J. Am. Chem. Soc.

vi

2017, 139, 5712-5715.

[7] Qing Li*, Jiaju Fu, Wenlei Zhu, Zhengzheng Chen, Bo Shen, Liheng Wu, Zheng

Xi, Tanyuan Wang, Gang Lu, Jun-jie Zhu and Shouheng Sun*, “Tuning Sn-Catalysis

for Electrochemical Reduction of CO2 to CO via the Core/Shell Cu/SnO2 Structure”

J. Am. Chem. Soc. 2017, 139, 4290-4293.

[6] Guangming Jiang, Huiyuan Zhu*, Xu Zhang, Bo Shen, Liheng Wu, Sen Zhang,

Gang Lu, Zhongbiao Wu* and Shouheng Sun*, “Core/Shell Face-Centered

Tetragonal FePd/Pd Nanoparticles as an Efficient Non-Pt Catalyst for the Oxygen

Reduction Reaction” ACS Nano, 2015, 9, 11014–11022.

[5] Liheng Wu, Bo Shen, Shouheng Sun*, “Synthesis and Assembly of Barium-

doped Iron Oxide Nanoparticles and Nanomagnets” Nanoscale, 2015,7, 16165-16169

[4] Liheng Wu, Qing Li, Cheng Hao Wu, Huiyuan Zhu, Adriana Mendoza-Garcia, Bo

Shen, Jinghua Guo and Shouheng Sun*, “Stable Cobalt Nanoparticles and Their

Monolayer Array as an Efficient Electrocatalyst for Oxygen Evolution Reaction” J.

Am. Chem. Soc., 2015, 137, 7071–7074

[3] Bo Shen, Peng-Fei Shi, Yin-Ling Hou, Fan-Fan Wan, Dong-Liang Gao and Bin

Zhao*, “Structural diversity and magnetic properties of five copper-organic

frameworks containing one-, two-, and three-types of organic ligands” Dalton Trans.,

2013, 42, 3455-3463.

[2] Yin-Ling Hou, Gang Xiong, Bo Shen, Bin Zhao*, Zhi Chen and Jian-Zhong Cui*,

“Structures, luminescent and magnetic properties of six lanthanide–organic

frameworks: observation of slow magnetic relaxation behavior in the DyIII

vii

compound” Dalton Trans., 2013, 42, 3587-3596.

[1] Peng-Fei Shi, Zhi Chen, Gang Xiong, Bo Shen, Jing-Zhe Sun, Peng Cheng* and

Bin Zhao* “Structures, Luminescence, and Magnetic Properties of Several Three-

Dimensional Lanthanide–Organic Frameworks Comprising 4-Carboxyphenoxy

Acetic Acid” Cryst. Growth Des., 2012, 12, 5203–5210.

viii

Acknowledgements

To be a chemist is my dream. And the dream is becoming more and more realistic

in my PhD career at Brown University. Looking back to the five years, I would like to

thank many wonderful people who has made my Ph.D. study colorful and meaningful.

First of all, my greatest thanks are definitely given to my research advisor, Prof.

Shouheng Sun. Before coming to Brown, I was deeply attracted by his excellent

research in magnetic nanomaterials. In my first semester, I chose his course of

Nanoscale Materials CHEM 1700. His lecture is always vivid and his logic flow is

always scientific, with clear concepts and systematical summary. He opened a new gate

of the amazing nano-world to me. In the second semester, I was lucky enough to join

his lab and started my research career at Brown University. In the past five years, I

learned a lot from not only the scientific method to research but also his noble

personality, like his endless passion to science, his critical thinking and his strict

training on logical conversation for the future career. He is not only my Ph.D. advisor,

but also like a father or friend. He would directly point out my drawbacks on research.

When I felt frustrated, he would warmly encourage me with applauding for my progress.

In these year, I gradually understand what the quality of a Ph.D. should have. I cannot

think of how my Ph.D. career would be without his guidance and support. The time and

experience I worked with him is a great treasure for my future career.

I am also grateful to my committee members, Prof. Eunsuk Kim and Prof. Lai-

Sheng Wang. During these years, they are always helpful to provide me valuable

suggestions and shared their valuable time on my RPD, ORP and defense. Thanks go

ix

to Prof. J. William Suggs, for his encouragement and help of my class during my first

year at Brown. I also appreciate Dr. Li-Qiong Wang, for her great help in my teaching

career and daily life.

I feel lucky enough to have the greatest collaborators and group members. Without

their help, it would be very tough to finish my research. Thank my close cooperators

Dr. Scott McCall, Dr. Sarah Baker and Dr. Alexander Baker in Lawrence Livermore

National Laboratory for magnetic properties measurement. Thank Dr. Dong Su at

Brookhaven National Laboratory for his work on STEM analysis of my samples. My

appreciations also deliver to Dr. Anthony McCormick for his help on SEM and TEM

operation, Dr. Paul Waltz and Dr. Garces Hector for XRD operation in the Department

of Engineering at Brown. Thanks to Prof. Gang Xiao and his student Wenzhe Chen and

Lijuan Qian for magnetic properties measurement in Physics Department, also Dr.

Joseph Orchardo in the Department of Geological Science at Brown for the help with

ICP measurement. Thank to Kenneth Talbot and Randy Goulet for mechanical

instrument making. Thanks go to my excellent group members, Dr. Chao Yu, my best

friend and closest cooperator in magnetic projects and catalysis projects, Dr. Adriana

Mendoza-Garcia and Dr. Liheng Wu for training me nanoparticle synthesis and

magnetic characterization when I was a green-hand. A special acknowledgement goes

to Dr. Qing Li, Dr. Sen Zhang, Dr. Huiyuan Zhu, Dr. Guangming Jiang, Junrui Li,

Zhouyang Yin, Jiaju Fu, Hu Liu for their assistance in experiment and a lot of precious

time we had together. Thanks to my roommate and lab partner Dr. Zheng Xi for the

help not only in research but also in daily life. I also like to thank Dr. Wenlei Zhu, Dr.

x

Hongyi Zhang, Yuyang Li, Michelle Muzzio, Mengqi Shen, Honghong Lin, Kecheng

Wei, Joshua Dunn, and all other Sun group members.

Finally, I would thank my parents for their endless love and support in my life. I

love you!

xi

To my family

xii

Abstract of “Rational Synthesis of Nanomagnets and Magnetic Nanocomposite for

Permanent Magnet Applications” by BO SHEN, Ph. D., Brown University, May 2019.

In the past two decades, the synthesis of magnetic nanoparticles (NPs) has been

intensively explored for both fundamental scientific research and industrial applications.

Different from the bulk sintered or bonded magnet, magnetic NPs show unique

magnetic properties, which permits to adjust their magnetism by systematic nanoscale

engineering. This thesis focuses on the synthesis of permanent nanomagnets, as well as

magnetic hard-soft phase exchange-coupled nanocomposite for their applications in

energy store and convention as permanent magnets.

The traditional bulk permanent magnet with the largest magnetic energy product

is NdFeB. However, the Curie temperature is low and it cannot be used above 200 oC.

SmCo alloy, a class of hard magnets for NdFeB substitution, shows a large coercivity

and high Curie temperature. But the relative low moment limits its usage. To solve the

problem, SmCo need to exchange-coupled with soft magnet like Fe to form

nanocomposite. The traditional way is to mix SmCo and Fe NPs together and anneal it.

This method causes Fe NPs diffusing into SmCo matrix to form SmCoFe alloy,

decreasing their magnetic property. We coated the pre-synthesized Fe NPs with SiO2

and assembled the Fe/SiO2 NPs with Sm−Co−OH. After reductive annealing at 850 °C

in the presence of Ca, we obtain SmCo5−Fe/SiO2 composites. Following aqueous

NaOH washing and compaction, we produced exchange-coupled SmCo5-Fe

nanocomposites with Fe NPs controlled at 12 nm.

xiii

Another challenge in developing nanostructured SmCo5 magnets is to control the

nanoscale dimensions of SmCo5 with large magnetic coercivity. I developed a novel

strategy to synthesize anisotropic SmCo5 nanoplates. This method involves the pre-

synthesis of 125 x 12 nm Sm(OH)3 nanorods and self-assembly of these nanorods and

10 nm Co NPs into Sm(OH)3-Co nanocomposites. After a CaO protection coating and

a reductive annealing process, 125 x 10 nm SmCo5 nanoplates are obtained, which can

be dispersible in ethanol, allowing the alignment in epoxy resin under a magnetic field.

The aligned SmCo5 nanoplates show a square hysteresis behavior with room

temperature coercivity reaching 30.1 kOe, which is among the highest values ever

reported for SmCo5.

The third challenge in the rare-earth magnet studies is the difficulty to extend a

method to prepare different types of rare-earth nanomagnets., I developed a general

chemical approach to SmCo- and SmFeN-based NPs. Using Co(acac)3 decomposition

in oleylamine, SmCo-O NPs were obtained which can be further coated with CaO and

reduced with Ca at 850 °C to form SmCo5 in the size range of 50-200 nm. The 200 nm

SmCo5 NPs can be dispersed in ethanol, and magnetically aligned in a polymer matrix

or compacted into pellet with the largest coercivity of 5 T and energy product of 16.8

MGOe, the highest values ever reported for chemically synthesized SmCo5. The

synthesis can be extended to synthesize Sm2Co17 by composition control, or even

Sm2Fe17N3 NPs (the 100 nm Sm2Fe17N3 NPs have the highest Hc (>1.3T) and Ms (>120

emu/g)). These high performance SmCo and SmFeN NPs are an important class of

magnetic building blocks for the fabrication of magnetic devices and of high

xiv

performance nanocomposite magnets.

My research further extended to non-rare earth magnetic NPs, such as hexagonal

BaFeO NPs. These NPs were prepared by annealing of barium doped iron oxide NPs

at 700 °C in air. They are ferromagnetic with room temperature Hc reaching 5260 Oe

and Ms at 54 emu g−1. I developed a self-assembly method to allow these BaFeO NPs

to form 2D magnetic arrays, which may serve as a unique model system for

nanomagnetic applications.

xv

Table of Contents

Chapter 1. Introduction to Nanomaterials, Magnetism and Magnetic Nanoparticle

Applications.……………….………………………….….……………………..…....1

1.1 General Introduction to Nanomaterials….…………….……...….….…...…..2

1.2 Introduction to Nanomagnetism …………………….…….……………..…..5

1.2.1 Classification of Magnetism .................................................................. 5

1.2.2. Size, Shape, Structure and Temperature Effect of Ferromagnetic

Nanoparticles…………………………………………………………………6

1.2.3. Applications of Ferromagnetic Nanoparticles .................................... 12

References…………………………………...…………………………………...20

Chapter 2. Synthesis and Characterization of Magnetic Nanoparticles………….24

2.1 Chemical Synthesis of Monodisperse Nanoparticles...………………....…….25

2.1.1 Nanoparticles Growth Mechanism ………………..……………………25

2.1.2 Experiment Setup …………………...……………………...…………..27

2.1.3 Nanoparticle Collection and Purification …………………..…………..29

2.2 Nanoparticle Characterization………………………….………….…. ……...30

2.2.1 Transmission Electron Microscopy (TEM)……...……………………...30

2.2.2 Scanning Electron Microscopy (SEM) …………………...…………….31

2.2.3 Scanning Transmission Electron Microscopy (STEM)…………..…….31

2.2.4 X-ray Powder Diffraction Pattern (XRD)…………………………........32

2.2.5 Inductive Coupled Plasma - Atomic Emission Spectroscopy (ICP-AES)32

2.2.6 Magnetic Measurements………………………….…………………….32

xvi

References………………………………………………………………………...34

Chapter 3. Stabilization of Fe Nanoparticles in SmCo5 matrix to Synthesize

SmCo5-Fe Nanocomposites………………..…….……………………36

3.1 Introduction……………………………………………………………….…...37

3.2 Experimental Details……………………………………………….………….39

3.3 Results and Discussion…………………………………….………….……….41

3.3.1 Synthesis and Characterization of Fe/SiO2 Nanoparticles …….………41

3.3.2 Synthesis and Characterization of SmCo5 Hard Magnet………...……..43

3.3.3 Embedding Fe Nanoparticles into SmCo5 Matrix for Nanocomposite

Fabrication……………………………………………………………...45

3.4 Conclusion……………………………………………………….………........49

References……………………………………………………….………………...51

Chapter 4. Synthetic of Anisotropic SmCo5 Nanoplates as Hard Nanomagnets…54

4.1 Introduction…………………………………………………….……………...55

4.2 Experimental Details……………………………………………………….….56

4.3 Results and Discussion………………………………………………………..60

4.3.1 Synthesis of Sm(OH)3-Co Nanocomposites……….…………...………60

4.3.2 Synthesis of SmCo5 Nanoplates……….……………………..………..61

4.3.3 Alignment of SmCo5 Nanoplates in Polymer…………………............66

4.4 Conclusion………………………………………………………………….....69

References…………………………………………………………………………71

Chapter 5. A General Method to Synthesize Anisotropic Sm-based Nanomagnets

xvii

with Ultra-large Coercivity…………………………………..….……..74

5.1 Introduction ………………………………………………….………………..75

5.2 Experimental Details…………………………………………………………..76

5.3 Results and Discussion………………………………………….….………….78

5.3.1 Synthesis of SmCo5 Nanoparticles with Size Control…………….……78

5.3.2 Alignment of SmCo5 in Polymer Matrix and Compaction of SmCo5 to

Pellet………………………………………….………………………….83

5.3.3 Synthesis of Sm2Co17 Nanoparticles and Sm2Fe17N3 Nanoparticles…...86

5.4 Conclusion………………………………………………………………..........89

References…………………………………………………………………………91

Chapter 6. Synthesis and Self-Assembly of Non-rare Earth Permanent

Nanomagnets……………………….…………………………………...94

6.1 Introduction………………………………………………………….…….......95

6.2 Experimental Details……………………………………………….……….....96

6.3 Results and Discussion…………………………………………………………98

6.3.1 Synthesis of Ba doped Iron Oxide (BaFeO) Nanoparticles with

Composition Control …………98

6.3.2 Self-assembly of BaFeO Nanoparticles……..…………..……….…....103

6.4 Conclusion…………………………………………………………………....105

References…………………………………………………………….……….106

xviii

List of Figures

Figure 1-1. Illustration of various objects in nanometer (nm)………………………...3

Figure 1-2. The relationship between the number of atoms in cluster nanoparticles

and the percentage of surface atoms…………………………………………………...4

Figure 1-3. Schematic illustrating the arrangements of magnetic moment for five

different types of materials in the absence or presence of an external magnetic field…6

Figure 1-4. (a) Schematic illustration of the hysteresis loops of ferromagnetic NPs and

(b) superparamagnetic NPs………………………….…………………………………8

Figure 1-5. Schematic illustration of size-dependent Hc of a ferromagnetic particle...8

Figure 1-6. Schematics of the local structures of (a) fcc-FePt and (b) fct-FePt………10

Figure 1-7. Hysteresis loops (a) unaligned and (b) aligned Co NRs. TEM images of (c)

unaligned and (d) aligned Co NRs................................................................................11

Figure 1-8. Illustration of temperature effect to magnetic NPs. The double well

potential shows the energy versus the orientation of the moment of magnetic NPs

without external field………………………………………………………………...12

Figure 1-9. Converting M-H hysteresis loop to B-H hysteresis loop………………..13

Figure 1-10. Magnetic characterization of (a) non-exchange-coupled system and (b)

well exchange-coupled system in magnetic soft and hard composites………………14

Figure 1-11. The fabrication process of fct-FePt/Fe3Pt magnetic nanocomposite…..15

Figure 1-12. Phase diagram of SmCo alloy………………….……………………....17

Figure 1-13. A hexagonal unit cell of SmCo5……………….……………………….17

Figure 2-1. (a) The process of the La Mer model for NPs formation. (b) a characteristic

xix

experimental setup for the organic phase solution synthesis…………………………26

Figure 2-2. Photograph is the typical setup for the organic solution synthesis……....28

Figure 2-3. Photograph for the furnace for high temperature annealing…………….29

Figure 2-4. Photograph for the synthesized green SmCoO NPs in hexane………….30

Figure 2-5. Photograph for the magnetic measurement setup: vibrating sample

magnetometry and physical property measurement system………………………….33

Figure 3-1. (a) TEM image of the as-synthesized 12 nm Fe NPs; (b) XRD pattern of

the as-synthesized 12 nm Fe NPs, showing the typical pattern that matches with the

standard bcc-Fe pattern; (c) TEM image of the 12 nm Fe NPs coated with 7 nm thick

SiO2 shell………………………………………………………………….………….42

Figure 3-2. TEM image of the Sm(OH)3 nanorods (a) and Co(OH)2 nanoplates (b); (c)

XRD of the SmCo5 powder obtained from our chemical synthesis (black curve) and

from the standard pattern (red lines, JPCDS No. 65-8981); (d) Hysteresis loop of the

SmCo5 powder measured at 300 K…………………………………………………...44

Figure 3-3. (a) XRD of hexagonal crystalline Co(OH)2 nanoplate precipitation and

standard pattern of Co(OH)2. (b) XRD of 60nm x15nm crystalline Sm(OH)3 nanorods

and standard pattern of Sm(OH)3……………………………………….……………45

Figure 3-4. (a) XRD patterns of SmCo5-Fe(x wt%) composite with x = 0, 5, 10 and 20.

(b) HAADF-STEM image and (c) elemental mapping of the SmCo5-Fe(10 wt%)

composite. Note: the overall Fe NP content is in 10 wt%, but the image shows an area

enriched with Fe NPs………………………………………………………………..47

Figure 3-5. (a) Hysteresis loops of the nanocomposites of SmCo5-Fe(x wt%) (x = 0-20)

xx

nanocomposites at 300 K. Inset: the change of Hc and Ms with the different Fe NP

contents in the SmCo5-Fe nanocomposites; (b) Hysteresis loops of the nanocomposite

of SmCo5-Fe(10 wt%) before (black) and after (red) 1.5 GPa compaction at 300 K..48

Figure 3-6. A photograph of compressed SmCo5-Fe nanocomposite……………….49

Figure 3-7. Hysteresis loops of the nanocomposites of SmCo5 + 20 wt. % Fe

nanocomposites before and after 1.5 GPa press at 300K. The Ms increases from 78.6

emu/g to 82.7 emu/g. Coercivity decreases from 11.2kOe to 8.1kOe……………….49

Figure 4-1. Schematic illustration of the synthesis of anisotropic SmCo5 nanoplates by

self-assembly of Sm(OH)3 NRs and Co NPs, followed by CaO coating and reductive

annealing…………………………………………………………………………..…56

Figure 4-2. (a) TEM image of the as-synthesized 10 nm Co NPs. (b) XRD of the as-

prepared Co NPs and the standard fcc-Co. (c) TEM image of 125 x 12 nm Sm(OH)3

NRs. (d) XRD of the as-prepared Sm(OH)3 NRs and the standard pattern of Sm(OH)3

......................................................................................................................................61

Figure 4-3 (a) TEM image of Sm(OH)3-Co nanocomposite with Sm:Co =1:4.5 (molar

ratio). (b) TEM image of Sm(OH)3-Co nanocomposite embedded in CaO matrix. (c)

TEM image of Sm(OH)3-Co nanocomposite obtained 10 min after the annealing. (d)

TEM image of the as-synthesized SmCo5 nanoplates. (e) HAADF-STEM and elemental

mapping of the SmCo5 nanoplates, showing the formation of uniform alloy structure

within each nanoplate…………………………………………………………………63

Figure 4-4. (a) TEM image of Sm(OH)3 NRs embedded in Co(OH)2 matrix. (b) TEM

image of SmCo5, showing no specific shape feature…………………………………64

xxi

Figure 4-5. (a) HRTEM image of a part of one SmCo5 nanoplate (planar view). (b) Fast

Fourier transform pattern of (a). (c) Simulated SAED pattern of hexagonal SmCo5

projected along the c-axis. (d) A fraction of HRTEM imaging area in showing the

arrangement of Sm and Co atoms. (e) Modeled hexagonal SmCo5 structure projected

along the c-axis. (f) HRTEM image of the side-view of a SmCo5 nanoplate. (g) Modeled

SmCo5 structure projected along [1, -1, 0]……………………………………………65

Figure 4-6. (a) XRD of the as-synthesized SmCo5 nanoplate powder (black curve) and

the standard pattern of D2d structure SmCo5 (red lines, JPCDS No. 65-8981). (b)

hysteresis loop of the as-synthesized SmCo5 nanoplate powder measured at 300 K…66

Figure 4-7. XRD pattern of SmCo5 nanoplates obtained from their ethanol dispersion

after ethanol evaporation under a 20 kOe field………………………………………68

Figure 4-8 (a) Schematic illustration of SmCo5 nanoplate alignment in resin along the

magnetic field direction for TEM and XRD characterizations. (b) TEM image of the

aligned SmCo5 nanoplates embedded in resin. (c) XRD patterns of the non-aligned

SmCo5 and the aligned SmCo5 nanoplates (red curve). (d) Room temperature hysteresis

loops of the aligned SmCo5 nanoplates measured along the c axis and perpendicular to

the c axis………………………………………………………………………………68

Figure 5-1. TEM images of as-synthesized (a) 60 nm (b) 110 nm (c) 220 nm SmCoO

flower-liked NPs. (d) XRD patterns of SmCoO NPs with different sizes and standard

CoO pattern (JPCDS No. 80-0075). (e) HADDF-STEM image and elemental mapping

of Sm (red), Co (blue) and O (green)………………………………………………...80

Figure 5-2. TEM image of 100 nm SmCoO NPs in CaO matrix coating……………82

xxii

Figure 5-3. TEM image of 100 nm SmCoO NPs after 15 min annealing at 850 °C…82

Figure 5-4. TEM images of annealed (a) 50 nm (b) 100 nm (c) 200 nm polyhedral

SmCo5 NPs. (d) HRTEM of a part of a 100 nm SmCo5 particle. (e) HADDF-STEM

image of a 100 nm SmCo5 particle and elemental mapping of Sm (red) and Co (blue),

showing uniform elemental distribution. (f) XRD patterns of SmCo5 NPs and standard

SmCo5 pattern (JPCDS No. 65-8981). Non-aligned hysteresis loops of (g) 50 nm (h)

100 nm and (i) 200 nm SmCo5 NPs at 300 K………………………………………..83

Figure 5-5. (a) Hysteresis loops of 50 nm, 100 nm and 200 nm SmCo5 NPs after

external field alignment with PEG at 300 K. (b) A picture of compacted SmCo5

nanomagnet. (c) SEM of the SmCo5 nanomagnet after compaction. (d) Hysteresis loops

of compacted 200 nm SmCo5 nanomagnet at 300 K…………………………..…….85

Figure 5-6. B-H hysteresis loops of aligned 200 nm SmCo5 NPs at 300 K…………85

Figure 5-7. (a) TEM image of 120 nm SmCo8.5O NPs. (b) TEM image of 100 nm

Sm2Co17 NPs. (c) XRD patterns of Sm2Co17 NPs and standard hexagonal Sm2Co17

pattern. (d) Hysteresis loop of unaligned and aligned as-synthesized Sm2Co17 NPs at

300 K……………………………………………………………………..…………..87

Figure 5-8. (a) TEM image of a 120 nm SmFeO nanocubes (b) XRD of as-prepared

Sm2Fe17 NPs (black curve) and the standard pattern of rhombohedral structure Sm2Fe17

(red lines, JPCDS No. 01-074-7186). (c) Hysteresis loops of as-prepared 100 nm

Sm2Fe17 NPs at 300 K. (d) TEM of 100 nm Sm2Fe17N3 NPs. (e) XRD of as-prepared

Sm2Fe17N3 NPs (black curve) and the standard pattern of rhombohedral structure

Sm2Fe17N3 (red lines, JPCDS No. 00-048-1790). (f) Hysteresis loops of unaligned

xxiii

(black) and aligned (red) nitrogenized Sm2Fe17N3 NPs at 300 K……………………88

Figure 5-9. XRD of Sm2Fe17 NPs annealed with melamine at 650 oC for 2h (black

curve). The product matches well to standard SmN (red lines) and standard bcc-Fe (blue

lines)………………………………………………………………………………….89

Figure 6-1. (a) TEM image of the as-synthesized Ba0.04–Fe–O NPs. (b) HR-TEM image

of a representative Ba0.04–Fe–O NP. (c) TEM image of the as-synthesized Ba0.082–Fe–O

NPs……………………………………………………………………………….100

Figure 6-2. (A) XRD patterns and (B) room temperature hysteresis loops of the Ba0.04–

Fe–O NPs before and after O2 annealing treatment. (C) XRD patterns and (D) room

temperature hysteresis loops of the Ba–Fe–O NPs with different Ba compositions after

annealing in O2 at 700 °C for 1 h………………………………………………..….102

Figure 6-3. (A) TEM image of the monolayer assembly of Ba0.082–Fe–O NPs. (B) SEM

image of the monolayer assembly deposited on a Si substrate. (C) SEM image of the

monolayer assembly after annealing in O2 at 700 °C for 1 h. (D) SEM images of the

multilayer assembly of Ba0.082-Fe-O NPs deposited on a Si substrate by the drop-casting

method. (E) SEM image of the multilayer assembly after annealing in O2 at 700 °C for

1 h. (F) Room temperature hysteresis loops of the multilayer assembly after annealing

in O2 at 700 °C for 1 h……………………………………………………………….104

xxiv

List of Tables and Schemes

Table 1-1. Dc (Dsd) and superparamagnetic critical size Ds (Dsp) values of common

magnetic materials…………………………………………………………………….8

Table 1-2. A list of parameters of common hard magnetic materials…………………15

Scheme 3-1. Schematic illustration of the synthesis of SmCo5-Fe nanocomposite by

assembling Sm(OH)3 nanorods, Co(OH)2 nanoplates and Fe/SiO2 NPs, followed by

reductive annealing, NaOH solution washing and compaction……………………...38

Table 4-1. A list of SmCo5 made by chemical method. The theoretical calculated

(BH)max for perfect SmCo5 is 28.6 MGOe.

Table 6-1. Experimental conditions for synthesizing Ba–Fe–O NPs with different Ba

compositions……………………………………………………………………...…101

1

Chapter 1

Introduction to Nanomaterials, Magnetism and Magnetic

Nanoparticle Applications

2

1.1 General Introduction to Nanomaterials

The original concept of “Nanomaterial” was first discussed by the famous

American theoretical physicist Richard Feynman, giving a talk entitled “There are

plenty of room at the bottom” which describes that the molecular machines can be built

with atomic precision at the California Institute of Technology in 1959.1 In 1974,

Japanese scientist Norio Taniguchi firstly used the word “Nanotechnology” in his paper

to describe semiconductor processes with precise control at nanometer level,2 and

American engineer Kim Eric Drexler extended the concept to molecular

nanotechnology3.

The modern nanotechnology made a real breakthrough in 1981, when Gerd Binnig

and Heinrich Rohrer at IBM invented the scanning tunneling microscopy (STM). The

characterization instrument makes it possible to observe and operate materials in

individual atom scale. Since then, nanotechnology has been a popular area not only for

fundamental research but also for practical applications.4-10 One nanometer (nm) is one

billionth (10-9) of a meter (m). The scale of nanometer can be straightly shown in Figure

1-1[4] For example, the size of a tennis ball is about 108 nm (ten to the eighth); a

biological cell is in a range of 104 ~ 105 nm; protein, DNA and virus are in a range of 1

~ 102 nm. Typically, nanoparticles (NPs) is a class of particles between 1 ~ 1000 nm in

size.

3

Figure 1-1. Illustration of various objects in nanometer (nm).

Nanomaterials, compared with their bulk counterparts, show very different

physical and chemical properties. The reason is called “size-dependent effect”. The

effect directly leads to dramatic change in surface area of materials. Here is a simple

example. Imagining a solid cubic material with the length of 1 cm, its corresponding

surface area is 6 cm2. If the cube is divided into 1 mm small cube, 1000 small cubes

can be obtained, and the total surface area is 60 cm2. If the 1 cm cube further divided

into 1 nm nanocubes, the total cube number is 10,000,000 and the total surface area

increases to 60,000,000 cm2. The size effect in material can also affect the percentage

of surface atoms. For example, in Figure 1-2, if the atoms in a cluster decreases from

561 to 13, the surface atoms percentage increases greatly from 45% to 92%.11

4

Figure 1-2. The relationship between the number of atoms in cluster nanoparticles and

the percentage of surface atoms.

The physical and chemical differences between the surface atoms and the inner

ones are totally different. The inner atoms, due to their fully coordinated structure,

perform similar properties as bulk materials. Nevertheless, the unsaturated surface

atoms are much less stable than the inner ones and owns a high surface energy. The

highly active surface atoms provide large space and numerous sites for chemical

reaction with different kinetic mechanism, which is very important for catalysis

application of NPs.12-15 Also, the size effect also be found in nanoscale semiconductors

(known as quantum dots). The size of quantum dots directly affects the band gap, and

as a result, the optical properties can be rationally adjusted and show great potential in

the fields of biological labeling, photovoltaics and solid-state lighting.16-19 Apart from

the catalysis and optical applications mentioned above, the magnetic application of NP

is also one of the most noteworthy and promising field in the current technology, which

5

will be discussed in detail below.

1.2 Introduction to Nanomagnetism

1.2.1 Classification of Magnetism

Magnets have been widely used over thousands of years. In modern history, the

applications of magnetic materials can be found in electronic and magnetic devices,

such as motors, computers, medical equipment and so on. The magnetism is originally

from the electron magnetic moment, which is result of the orbital and spin

magnetization coupling of an electron. According to the interaction of atomic magnetic

moments, materials can be classified into diamagnetic, paramagnetic, ferromagnetic,

ferrimagnetic, and antiferromagnetic.20-22

Figure 1-3 shows schematic diagrams of these five different states. Diamagnetic

materials are consisted of atoms which have no single electrons. Their paired electrons

have no net magnetic moments in the absence of magnetic field. In paramagnetic

materials, some single electrons partially fill the orbitals. In the absence of external field,

the spins are randomly oriented due to thermal fluctuation. If an external magnetic field

applied, those atomic magnetic moments will align in the directions of the field in some

extent, resulting in a net but very weak magnetization. The ferromagnetic materials

show a totally different magnetism. The strong interaction of the atomic moment,

produced by electronic exchange force, lead to a parallel alignment of the atomic

moments. The ferromagnetic materials show a large net magnetization even without

external field alignment. The ferrimagnetic materials own two magnetic sublattices.

6

These two set of sublattices are antiparallelly aligned, which caused by super-exchange

coupling effect. However, the magnetic moment of one set is stronger than another set

and an existing net magnetic moment exist, showing similar magnetic behavior to

ferromagnetism. Antiferromagnetic materials also have two sublattices, antiparallelly

aligned to each other. However, they have the same magnetic moments and cancel each

other, resulting in no net magnetization. Generally, only ferromagnetic or ferrimagnetic

materials are called magnetic materials.

Figure 1-3. Schematic illustrating the arrangements of magnetic moment for five

different types of materials in the absence or presence of an external magnetic field.

1.2.2 Size, Shape, Structure and Temperature Effect of Ferromagnetic NPs

The behavior of magnetic NPs can be described by hysteresis loop, as shown in

Figure 1-4 (a). In the absence of external magnetic field, the NPs are randomly aligned,

and the total moment is 0. If an external field applied, the interaction of magnetic NPs

and field can align the moment of NPs to the field direction. If the external field is

7

strong enough, all the particles will turn to field direction and the corresponding

moment reaches the maximum value, which is defined as saturation moment (Ms). The

total moment will also decline with decreasing the external field. But when the field

reduces to 0, the magnetic NPs can still hold an extent of moment, which is called

remnant moment (Mr). To fully demagnetize the NPs, a reverse field should be added

to the point where the total moment is 0, the value of which is named as coercivity (Hc).

The magnetic properties, especially Hc, are strongly size-dependent.23,24 As shown in

Figure 1-5, if the size of magnetic particle is in single domain (SD) range, the moment

doesn’t change direction across the particle. The coercivity increases with growing size

and can reach to a maximum value and the size is called critical size (Dc). The Dc values

for a spherical magnetic NP can be roughly calculated as Dc ≈36√𝐴𝐾

𝜇0𝑀𝑠2 , where A is the

exchange constant, K is the anisotropy constant which stands for the energy per unit

volume required to flip moment direction, μ0 is the vacuum permeability, and Ms is the

saturation moment. The Dc values are usually in the range of 5-1000 nm. The Dc values

of some commercial used magnetic materials are listed in Table 1. If the particle size is

larger than Dc, multi-domains (MD) exist in the particle. Because of the domain-domain

interaction, a moderately low external field is needed to change the magnetization. And

the Hc will decrease with increasing the size. On the other hand, if the particle is

extremely small, the thermal fluctuation effect become obvious and cause the moment

of NPs to flip direction. As a result, the ferromagnetic NPs turn to superparamagnetic

(Figure 1-4 (b)). Once the external field is removed, the magnetic NPs will be randomly

aligned leaving no Mr and Hc.

8

Figure 1-4. (a) Schematic illustration of the hysteresis loops of ferromagnetic NPs and

(b) superparamagnetic NPs.

Figure 1-5. Schematic illustration of size-dependent Hc of a ferromagnetic particle.

Table 1-1. Dc (Dsd) and superparamagnetic critical size Ds (Dsp) values of common

magnetic materials.

9

The structure and shape effect of magnetic NPs are also very important. Crystal

structure directly influences intrinsic spin−orbital moment interaction, which can be

expressed as magnetocrystalline anisotropy. The value of magnetocrystalline anisotropy

constant (K) determine the coercivity of magnetic NPs. Theoretically, if the material

owns a large intrinsic anisotropy constant, the material can obtain a large coercivity. For

example, the equal-atomic FePt alloy present two distinctive crystal structures (Figure

1-6). One shows face-centered cubic (fcc) structure (Figure 1-6a), in which the Fe and

Pt atoms randomly occupy the lattice points, forming a solid solution. The fcc-FePt is

magnetically isotropic, displaying superparamagnetic property. The other shows

chemically ordered face-centered-tetragonal (fct) structure (Figure 1-6b). The Fe and

Pt atom layers stacks alternatively along the [001] direction.25,26 Because of the strong

coupling of 3d orbital from Fe and 5d orbital from Pt, the fct-FePt has a large anisotropy

constant K (up to 7 × 106 J/m3) and are strongly ferromagnetic with a coercivity larger

than 3 T.

10

Figure 1-6. Schematics of the local structures of (a) fcc-FePt and (b) fct-FePt.

The shape of magnetic NPs also affects the magnetic properties. If a NP is sphere-

shaped, there will be have no shape anisotropy and the magnetic property has no change

in different directions. But if magnetic NPs are prepared in specific shape (rods, plates),

their magnetic properties measured from the magnetic easy and hard direction are

different. Generally, the hysteresis loop obtained from magnetic easy axis owns a high

Mr, a large Hc and a better squareness compared with that from hard axis. For example,

the Co NPs less than 15 nm usually shows a coercivity less than 1 kOe. However, the

Co nanorods (NRs) with the diameter of 15 nm and length-width ratio of 10, after

magnetic alignment, own a Hc of 4.5 kOe along magnetic easy direction at room

temperature (Figure 1-7).27 The shape anisotropy effect is very important to fabricate

magnetic media in high density tape recording.

11

Figure 1-7. Hysteresis loops (a) unaligned and (b) aligned Co NRs. TEM images of (c)

unaligned and (d) aligned Co NRs. Copyright 2009 Wiley.

The magnetic properties are also temperature-dependent. If the temperature of

magnetic material is above a critical temperature, the NPs will lose ferromagnetism and

transits to superparamagnetic. The critical temperature is called Curie temperature (Tc).

The mechanism can be illustrated in Figure 1-8, which shows the energy barrier

between the “up” and “down” moment in a NP. The energy needed to turn the moment

to an inverse direction is KV. If temperature increases, the thermal energy kBT (kB is

Boltzmann constant) would become stronger. If the temperature is above Tc, kBT will

overcome the energy barrier, causing a randomization of moment in NPs. Tc is an

intrinsic property which determines the upper limit temperature of ferromagnetic

material application. Generally, most magnetic materials have a negative temperature

coefficient, which means the coercivity of magnetic NPs will decrease with increasing

12

temperature. But there are still a few materials (like MnBi) own a positive temperature

coefficient and their coercivity will increase with increasing temperature until reaching

to Tc.

Figure 1-8. Illustration of temperature effect to magnetic NPs. The double well

potential shows the energy versus the orientation of the moment of magnetic NPs

without external field.

1.2.3. Applications of Ferromagnetic NPs

Ferromagnetic NPs have a wide application as recording media in hard disk drive

(HDD), in biochemistry and electrochemistry catalysis. Among them, one of the most

important and promising application is to fabricate permanent magnets for energy

storage and conversion, as that in the direct-current motors and wind turbines.28-39 The

permanent magnet can keep a high Mr and generate a magnetic field after magnetization.

Also, the large Hc of permanent magnet can stabilize the magnetic field for long-time

use. To evaluate the magnetic energy storage capacity of a permanent magnet, a figure-

of-merit is introduced as the maximum energy product (BH)max. To calculate the value

13

of (BH)max, first the M-H hysteresis loop was converted to B-H hysteresis loop with

an equation of B = H + 4πM (Figure 1-9), where B is called magnetic induction. The

(BH)max corresponds to the area of the largest rectangle in the second quadrant of the

B-H hysteresis loop. The unit is kJ/m3 (SI) or MGOe (GCS). Therefore, to obtain the

optimum magnetic property, the material should own both large coercivity and high

moment.

Figure 1-9. Converting M-H hysteresis loop to B-H hysteresis loop.

However, traditional magnetic materials, such as Fe and Co, have a high moment

but a small coercivity (less than 1000 Oe at room temperature), which we call them as

“soft magnet”.40,41 While another type materials called “hard magnet”, like NdFeB,

SmCo5 and fct-FePt, have a large coercivity (larger than 1000 Oe) but relatively low

moment.42-45 If the advantages of both hard and soft magnet can be combined in one

material, the (BH)max will be enhanced. And exchange coupling provides a promising

method to achieve the goal. The illustration of an exchange-coupled magnetic system

is shown in Figure 1-10. The requirements for effective exchange coupling between the

two phases are very strict.46,47 First of all, the hard magnet and soft magnet much be in

close contact with each other. Secondly and the most importantly, the size of the soft

phase must be small enough (~10 nm). If the requirement is not satisfied, the composite

14

will be decoupled and the hysteresis loop will show magnetic two-phase behavior,

which decreases the (BH)max (Figure 1-10A). Just the interface between the hard and

soft magnets is effective coupled, the majority of the soft would be easily magnetized

and demagnetized. Only with appropriate size of the soft phase will the composite be

effectively exchange couples with enhanced energy storage (Figure 1-10B).

Figure 1-10. Magnetic characterization of (a) non-exchange-coupled system and (b)

well exchange-coupled system in magnetic soft and hard composites.

A successful exchange-coupled nanocomposite of fct-FePt/Fe3Pt has been studied,

which gave scientist a viable way to fabricate magnetic nanocomposite.48 In this work

(Figure 1-11), a self-assembly of FePt and Fe3O4 NPs was annealed under H2 at 650 oC.

After annealing, the fcc-FePt was converted to magnetic hard fct-FePt and Fe3O4 was

reduced to magnetic soft Fe3Pt. The fct-FePt/Fe3Pt nanocomposite showed a smooth

hysteresis loop with enhanced (BH)max of 20.1 MGOe, which is 50 % higher than

15

single hard phase fct-FePt (13 MGOe).

Figure 1-11. The fabrication process of fct-FePt/Fe3Pt magnetic nanocomposite.

Copyright 2002 Nature Publishing Group.

To maximize the (BH)max in magnetic nanocomposite, it is very important to

choose appropriate soft and hard segments. For soft magnetic, the candidates are always

Fe, Co or FeCo alloys with high moment. For the common strong hard phase candidates

(the type and magnetic parameters are shown in Table 1-2),49 if we only considering

(BH)max, the NdFeB is the best one. However, the NdFeB shows a relative low Hc and

a low Tc, hence NdFeB cannot be used as permanent magnet above 200 oC in industry.

Another type of rare-earth hard magnet, SmCo, would be a good substitute of NdFeB.50

For example, the SmCo5 shows a large magnetocrystalline anisotropy constant (K=107

J m-3) and Tc (747 oC), which means it can still keep large coercivity at high temperature.

Therefore, in my Ph.D study I mainly focus on SmCo NPs synthesis and SmCo-Fe

nanocomposite fabrication.

Table 1-2. A list of parameters of common hard magnetic materials.

16

The group of SmCo magnets has been developed since 1970s and exist in several

phases (Figure 1-12).51 Among them, two most important alloys, SmCo5 and Sm2Co17,

shows superior magnetic property. The SmCo5 owns the greatest K value in known

materials up till now. The Sm2Co17 magnets, though showing smaller K value, have

higher saturation moment than the SmCo5. Both have higher Tc than Nd-based magnets

and are widely used in industrial production. If they combine with soft magnets though

exchange-coupling, the Ms will be increase and further improve the (BH)max to form

a magnetic superior nanocomposite. The crystal unit cell of SmCo5 is shown in Figure

1-13. This material adopts a hexagonal CaCu5 structure with Co layers and Sm + Co

layers present alternatively along the c-axis. The c-axis of the lattice is also magnetic

easy direction of SmCo5. The strong ferromagnetic property results from the parallel

coupling of delocalized Co 3d and localized Sm 4f moments.

17

Figure 1-12. Phase diagram of SmCo alloy (Highlighted in red and green are the two

important hard magnetic phases: SmCo5 and Sm2Co17, respectively).

Figure 1-13. A hexagonal unit cell of SmCo5.

SmCo5 NPs and SmCo-Fe nanocomposite are normally synthesized by physical

methods like high energy ball milling, high temperature melt-spinning or spurting into

films. The physical method is easy to conduct, but very hard to control the size and

shape of particles, which leads to a poor magnetic property.52-54 Therefore, the first

problem need to be solved is to synthesis magnetic composite with controllable size.

Chapter 2 introduces the experimental instruments I used to synthesize and

characterize SmCo/SmCo-Fe nanomagnet.

To make an effective exchange-coupled SmCo-Fe nanocomposite, an annealing

18

process is required. However, if SmCo and Fe NPs were simply mixed together and

annealed directly, Fe will diffuse into SmCo matrix to form SmCoFe alloy, not SmCo-

Fe composite, which will decline the magnetic properties. To overcome the difficulty,

Chapter 3 describes a new method to stabilize Fe NPs in SmCo matrix. In the work,

we first synthesize Fe/SiO2 core-shell structure NPs. SiO2 coating can prevent Fe NPs

diffusion or aggregation at high temperature, which can be removed by base and SmCo-

Fe nanocomposite with adjustable magnetic properties can be obtained.

The direct annealing reduction method of SmCo synthesis without control leading

to a low magnetic performance of SmCo5. To solve the problem, Chapter 4

demonstrates a self-assembly method to synthesize anisotropic SmCo5 nanoplates. The

SmCo5 nanoplates can be magnetically aligned and show an obverse magnetic

anisotropy with a Hc reaching to 30 kOe along the alignment direction.

To obtain versatile SmCo NPs with high yield, Chapter 5 describes a general

chemical method to synthesize hard magnets SmCo NPs with tunable sizes (50, 100 and

200 nm) and composition (SmCo5, Sm2Co17), by reducing different sizes of SmCoO

NPs in CaO matrix. The largest coercivity of 200 nm SmCo5 NPs can reach to 50 kOe

after alignment in polymer. This method can also be applied to SmFeN NPs synthesis

and hence stands for a general method of Sm based nanomagnet synthesis.

Non-rare earth magnet is also a class of important permanent magnet due to its

easy availability and low cost. The non-rare earth permanent magnet holds roughly half

of the magnet market and widely used in our daily life. For example, barium ferrite,

work as a recording media in hard drive device due to its comparable large coercivity,

19

chemical stability and relatively cheap price. Chapter 6 discusses the efforts to

synthesize and self-assemble BaFeO NPs for permanent magnet production.

20

References:

1 R. P. Feynman, Eng. Sci., 1959, 23, 22.

2 N. Taniguchi, Proc. Intl. Conf. Prod. Eng. Tokyo, Part II, Japan Society of Precision

Engineering, 1974, 5–10.

3 K.E. Drexler, Nanosystems: Molecular machinery, manufacturing, and computation.

New York: John Wiley & Sons, Inc, 1992.

4 B. D. Gates, Q. B. Xu, M. Stewart, D. Ryan, C. G. Willson, G. M. Whitesides, Chem.

Rev. 2005, 105, 1171-1196.

5 N. L. Rosi, C. A. Mirkin, Chem. Rev. 2005, 105, 1547-1562.

6 E. Boisselier, D. Astruc, Chem. Soc. Rev. 2009, 38, 1759-1782.

7 B. L. Cushing, V. L. Kolesnichenko, C. J. O'Connor, Chem. Rev. 2004, 104, 3893-

3946.

8 M. Haase, H. Schafer, Angew. Chem. Int. Ed. 2011, 50, 5808-5829.

9 A. H. Lu, E. L. Salabas, F. Schuth, Angew. Chem. Int. Ed. 2007, 46, 1222-1244.

10 D. V. Talapin, J.-S. Lee, M. V. Kovalenko, E. V. Shevchenko, Chem. Rev. 2010, 110,

389-458.

11 J. P. Wilcoxon, B. L. Abrams, Chem. Soc. Rev. 2006, 35, 1162-1194.

12 Y. J. Wang, D. P. Wilkinson, J. J. Zhang, Chem. Rev. 2011, 111, 7625-7651.

13 S. Zhang, Y. Z. Hao, D. Su, V. V. T. Doan-Nguyen, Y. T. Wu, J. Li, S. H. Sun, C. B.

Murray, J. Am. Chem. Soc. 2014, 136, 15921-15924.

14 Y. J. Kang, J. Snyder, M. F. Chi, D. G. Li, K. L. More, N. M. Markovic, V. R.

Stamenkovic, Nano Lett. 2014, 14, 6361-6367.

21

15 C. Wang, D. van der Vliet, K. L. More, N. J. Zaluzec, S. Peng, S. H. Sun, H. Daimon,

G. F. Wang, J. Greeley, J. Pearson, A. P. Paulikas, G. Karapetrov, D. Strmcnik, N. M.

Markovic, V. R. Stamenkovic, Nano Lett. 2011, 11, 919-926.

16 D. Ioannou, D. K. Griffin, Nano Rev. 2010, 1, 5117

17 I. L. Medintz, H. T. Uyeda, E. R. Goldman, H. Mattoussi, Nat. Mater. 2005, 4, 435-

446.

18 A. M. Smith, S. M. Nie, Acc. Chem. Res. 2010, 43, 190-200.

19 O. Chen, J. Zhao, V. P. Chauhan, J. Cui, C. Wong, D. K. Harris, H. Wei, H. S. Han,

D. Fukumura, R. K. Jain, M. G. Bawendi, Nat. Mater. 2013, 12, 445-451.

20 R. C. O’Handley, Modern Magnetic Materials: Principles and Applications, Wiley,

New York 2000.

21 N. Spaldin, Magnetic Materials: Fundamentals and Device Applications,

Cambridge University Press, Cambridge, UK 2003.

22 D. Weller, A. Moser, L. Folks, M. E. Best, W. Lee, M. F. Toney, M. Schwickert, J.

U. Thiele, M. F. Doerner, IEEE Trans. Magn. 2000, 36, 10.

23 Y. W. Jun, J. S. Choi, J. Cheon, Chem. Commun. 2007, 1203-1214.

24 B. M. Moskowitz, Hitchhiker's guide to magnetism 1991. 29

25 O. Gutfleisch, J. Lyubina, K. H. Muller, L. Schultz, Adv. Eng. Mater. 2005, 7, 208.

26 D. E. Laughlin, K. Srinivasan, M. Tanase, L. Wang, Scr. Mater. 2005, 53, 383.

27 Y. Soumare, C. Garcia, T. Maurer, G. Chaboussant, F. Fievet, J. Y. Piquemal, G. Viau,

Adv. Funct. Mater. 2009, 19, 1971.

28 J. H. Lee, Y. M. Huh, Y. Jun, J. Seo, J. Jang, H. T. Song, S. Kim, E. J. Cho, H. G.

22

Yoon, J. S. Suh, J. Cheon, Nat. Med. 2007, 13, 95-99.

29 H. B. Na, J. H. Lee, K. J. An, Y. I. Park, M. Park, I. S. Lee, D. H. Nam, S. T. Kim, S.

H. Kim, S. W. Kim, K. H. Lim, K. S. Kim, S. O. Kim, T. Hyeon, Angew. Chem. Int. Ed.

2007, 46, 5397-5401.

30 J. H. Gao, G. L. Liang, J. S. Cheung, Y. Pan, Y. Kuang, F. Zhao, B. Zhang, X. X.

Zhang, E. X. Wu, B. Xu, J. Am. Chem. Soc. 2008, 130, 11828-11833.

31 K. Cheng, S. Peng, C. J. Xu, S. H. Sun, J. Am. Chem. Soc. 2009, 131, 10637-10644.

32 J. T. Jang, H. Nah, J. H. Lee, S. H. Moon, M. G. Kim, J. Cheon, Angew. Chem. Int.

Ed. 2009, 48, 1234-1238.

33 T. Hyeon, J. E. Lee, N. Lee, H. Kim, J. Kim, S. H. Choi, J. H. Kim, T. Kim, I. C.

Song, S. P. Park, W. K. Moon, J. Am. Chem. Soc. 2010, 132, 552-557.

34 D. Ho, X. Sun, S. Sun, Acc. Chem. Res. 2011, 44, 875-882.

35 S. Peng, S. H. Sun, Angew. Chem. Int. Ed. 2007, 46, 4155-4158.

36 S. Peng, C. Wang, J. Xie, S. H. Sun, J. Am. Chem. Soc. 2006, 128, 10676-10677.

37 C. J. Xu, B. D. Wang, S. H. Sun, J. Am. Chem. Soc. 2009, 131, 4216-4217.

38 S. H. Sun, L. L. Lacroix, L. M., N. F. Huls, D. Ho, X. L. Sun, K. Cheng, Nano Lett.

2011, 11, 1641-1645.

39 C. Wang, Y. J. Wei, H. Y. Jiang, S. H. Sun, Nano Lett. 2009, 9, 4544-4547.

40 L. M. Lacroix, N. F. Huls, D. Ho, X. L. Sun, K. Cheng, S. H. Sun, Nano Lett. 2011,

11, 1641-1645.

41 C. Wang, S. Peng, L. M. Lacroix, S. H. Sun, Nano Res. 2009, 2, 380-385.

42 Y. L. Hou, Z. C. Xu, S. Peng, C. B. Rong, J. P. Liu, S. H. Sun, Adv. Mater. 2007, 19,

23

3349-3352. 30

43 S. H. Sun, C. B. Murray, D. Weller, L. Folks, A. Moser, Science 2000, 287, 1989-

1992.

44 J. M. Kim, C. B. Rong, J. P. Liu, S. H. Sun, Adv. Mater. 2009, 21, 906-909.

45 J. Kim, C. Rong, Y. Lee, J. P. Liu, S. Sun, Chem. Mater. 2008, 7242–7245.

46 E. F. Kneller, R. Hawig, IEEE Trans. Magn. 1991, 27, 3588-3600.

47 R. Skomski, J. M. D. Coey, Phys. Rev. B 1993, 48, 15812-15816.

48 H. Zeng, J. Li, J. P. Liu, Z. L. Wang, S. H. Sun, Nature 2002, 420, 395-398.

49 R. Skomski, J. M. D. Coey Permanent Magnetism 1998.

50 H. A. Leupold, F. Rothwarf, J.T. Breslin, J. J. Winter, A. Tauber, D. I. Paul, J

Appl Phys 1982, 53, 2392.

51 H. Okamoto, J Phase Equilib 1999, 20, 535.

52 J. Ding, P. G. McCormick, R. Street, J. Alloys Compd. 1995, 228, 102.

53 S. K. Chen, J. L. Tsai, T. S. Chin, J. Appl. Phys. 1996, 79, 5964.

54 E. M. Kirkpatrick, D. L. Leslie-Pelecky, J. Appl. Phys. 2000, 87, 6734.

24

Chapter 2

Synthesis and characterization of magnetic nanomaterials

25

2.1 Chemical Synthesis of Monodisperse NPs

2.1.1 Mechanism of Monodisperse NPs Growth

To prepare NPs with controlled size, shape and composition is a key step for their

application. Two general ways to prepare NPs are called as “top-down” and “bottom-

up” methods. The “top-down” method means the nanomaterials are crushed from bulk

size materials, which is often refer to physical methods, like mechanical ball milling,

lithography patterning, melt-spinning etc.1,2 The “top-down” method is a facile way but

usually lack of control of the producing monodisperse NPs (usually the size is from

several nanometers to micrometers). The “bottom-up” method refers to chemically

synthesize nanomaterials in atom level. This method includes chemical vapor

deposition (CVD),3 aqueous sol-gel process,4 microemulsion process5 and

hydrothermal synthesis6 and organic solution phase synthesis. The different synthesis

methods own specific advantages. For example, aqueous sol-gel synthesis is easy to

operate and can give a high yield of product; The organic solution phase approach can

precisely control monodisperse NPs in 0.1 nm scale with various shapes7-12 and versatile

structures.13-16 And herein, I mainly focused on the organic solution phase synthesis to

prepare monodisperse NPs and nanocomposites for magnetic applications.

The growth mechanism of NPs can be explained by “La Mer Model” in organic

solution phase synthesis process, which can be illustrated in Figure 2-1a17. In the

process, a vital concentration of precursor is called critical supersaturation or nucleation

threshold. If the concentration of precursor is lower than this level, no nucleation

appears and no NPs form. When the concertation of the precursors is higher than critical

26

supersaturation, large number of nuclei can be formed spontaneously. The spontaneous

nucleation rapidly decreased the concentration of precursors below the critical level in

the solution and hence no further nuclei can be formed. The existing nuclei will grow

into particles by aging at high temperature, which is called Ostwald Ripening process.

In the process, small nuclei will dissolute due to their high surface energy, and the

material will then redeposite on large NPs. The average size will increase with a

reimbursing reduction of the nuclei number.

Figure 2-1. (a) The process of the La Mer model for NPs formation. (b) a characteristic

experimental setup for the organic phase solution synthesis.

Figure 2-1b shows the experimental setup used to prepare NPs in organic solution.

The organic solvents with high boiling point (usually above 200 oC) are often used to

provide a high temperature environment for precursor decomposition. The common

solvents are like 1-octadecene (ODE, bp. 315 oC), benzyl ether (BE, bp. 298 oC),

1,2,3,4-tetrahydronaphthalene (tetralin, bp. 208 oC). In the reaction, usually there are

27

two mechanisms of decomposition of precursors.18-21 One is “hot-injection”, in which

the precursors are fast injected into the pre-heated reactor to reach critical

supersaturation. Another way is “heating-up” the precursors from low temperature to

their decomposition temperature, in which the nucleation process can be controlled by

the heating rate. However, due to high surface energy of NPs, a major problem in

organic solution synthesis is particle aggregation. To stabilize the NPs in the synthesis

process, selected surfactant(s) like oleylamine (OAm) and oleic acid (OAc), should also

be added in the reaction system. The surfactants coating around the NPs will cause steric

repulsion, which can make NPs dispersed in organic solvent for further use.

2.1.2 General Synthesis Setup

A typical synthesis setup I used in our lab is shown in Figure 2-2. Inert gases (N2,

Ar) can be introduced through the Schlenk line to exclude oxygen and water in the flask.

In the synthesis process, the precursors and organic solvent in the flask is stirred by a

magnetic bar. The four-neck flask can be heated by a heating mantle. One neck of the

flask needs to connect to the temperature controller (the thermal couple) for monitoring

the reaction temperature. One is connected to a gas inlet of the Schlenk line and another

is connected to the gas outlet trap, which can collect low boiling point impurity and

byproduct from the reaction. The extra neck is usually covered with a rubber stopper in

case of any chemicals needed to be injected into the reaction system during the reaction.

28

Figure 2-2. Photograph is the typical setup for the organic solution synthesis used in

our lab.

The chemicals used in my synthesis such as the organic solvent, metal salts,

surfactants, and reductive agents were purchased from Strem Chemicals (e.g. metal

acetylacetonate, metallic calcium) or Sigma Aldrich (e.g. iron carbonyl and cobalt

carbonyl) without further purification.

As shown in Figure 1-12, SmCo5 is thermodynamic stable above 820 oC.

Therefore, a high temperature annealing setup is needed. In our lab, we use the furnace

to anneal the sample (Figure 2-3). The quartz or ceramic annealing tubes connecting

with gas inlet and outlet can fill with inert gas (N2, Ar) and reductive gas (forming gas,

95%Ar + 5% H2). The samples in a ceramic or stainless-iron annealing boat are located

at the center of the tube. The temperature, heating speed and time can be automatically

controlled by the program of the furnace.

29

Figure 2-3. Photograph for the furnace for high temperature annealing.

2.1.3 NPs Collection and Purification

When the reaction finished, the products and organic solvent are first transferred

from flask centrifugation tubes. Then certain precipitant is added into the centrifugation.

The precipitants are usually polar solvent such as ethanol, acetone, isopropanol or their

mixture. The as-prepared NPs are coated with a layer of oleic acid and/or oleylamine,

so the hydrophobic end cannot be stably dispersed in polar solvent environment and

will precipitate from the solvent. After the precipitation, the NPs are collected by the

Beckman Coulter Allegra® 64R Centrifuge. Then the NPs can be easily separated by

pouring out the solvent. The remaining NPs attaching on the wall of centrifuge tubes

can be re-dispersed in non-polar solvents such as hexane and toluene. The washing

process should be repeated 2 ~ 3 times to clean the surface of NPs and remove extra

surfactants in the solvent. Here we should also mention that the centrifuge can also be

used for size selection if different sizes of NPs formed. By carefully choosing the speed

and time in the centrifuge process, different sizes NPs can be separated. The final

washed NPs can be dispersed in hexane again for the further use. Figure 2-4 shows the

30

as-synthesized dark green SmCoO NPs dispersion in hexane. The dispersion usually

stable and can be kept in ambient environment. However, some non-noble metal NPs

(e.g. Fe and Co NPs) and alloys (SmCo5) are not compatible with oxygen which can be

gradually oxidized in air. These NPs should be stored in glove box full of Ar for long-

term use.

Figure 2-4. Photograph for the synthesized green SmCoO NPs in hexane.

2.2 NPs Characterization

Various of characterization techniques are needed to explore the properties of the

as-prepared NPs. The typical characterization methods I used during my Ph. D. study

are listed below.

2.2.1 Transmission Electron Microscopy (TEM)

Transmission Electron Microscopy (TEM) is the most common equipment applied

to illustrate the shape and morphology of as-synthesized NPs. The TEM images were

recorded from a Philips CM20 Microscopy with an operating voltage of 200 kV at

Brown University. The sample can be prepared by depositing a droplet of NP dispersion

on a carbon coated Cu TEM grid (Ted Pella) for TEM analysis. High-Resolution

31

Transmission Electron Microscopy (HRTEM) is a more powerful instrument, which

allows direct observation of crystal lattice of the NPs. HRTEM images were collected

by using a JEOL 2010 with an accelerating voltage of 200 kV. The sample preparation

process is the same as that in regular TEM test.

2.2.2. Scanning Electron Microscopy (SEM)

Scanning electron microscopy (SEM) is another important instrument to

characterize the size and morphology of NPs. Compared with TEM, the SEM has a

lower resolution, but it can be applied to bulk material characterization. SEM images

were obtained on a LEO 1530 microscope at an accelerating voltage of 10 kV. To make

a n SEM sample, the material can be deposit on a conductive substrate such as Si wafer.

Besides the morphology analysis, the SEM can perform elemental analysis with the

help of the equipped energy-dispersive X-ray (EDX) spectroscopy. This EDX can

analyze X-ray emission spectrum of the elements and thus work as an instrument to

check elements existence in the specimen.

2.2.3 Scanning Transmission Electron Microscopy (STEM)

Scanning transmission electron microscopy (STEM) is a type of advanced TEM.

The difference is in STEM the electron beam can focus on a tiny spot (0.05 ~ 0.2 nm)

and then scanned over the sample in a raster. STEM can be suitable for unique analytical

techniques such as high-angle annular dark-field imaging (HAADF), elemental

mapping analysis, and electron energy loss spectroscopy (EELS). My STEM analysis

32

work was done on a Hitachi HD2700C (200 kV) with a probe aberration corrector, at

the Center for Functional Nanomaterials, Brookhaven National Lab.

2.2.4 X-Ray Diffraction (XRD)

X-ray diffraction (XRD) technique is a very critical tool to determine the crystal

structure and chemical composition of as-prepared NPs. For magnetic NPs, the XRD

can even characterize the magnetic NPs alignment direction. The XRD patterns are

collected by a Bruker AXS D8-Advanced diffractometer with Cu Kα radiation (λ =

1.5418 Å). The XRD sample can be prepared by drying NP dispersion or depositing

powders on the glass or silicon slide. Moreover, using Scherrer equation, the size of the

NPs can be roughly determined.

2.2.5 Inductively Coupled Plasma Atomic Emission Spectroscopy (ICP-AES)

Inductively coupled plasma atomic emission spectroscopy (ICP-AES) is a precise

measurement technique for quantitative detection of chemical elements. In my work,

all ICP-AES measurement was recorded by a JY2000 Ultrace ICP atomic emission

spectrometer equipped with a JYAS 421 autosampler and 2400 g/mm holographic

grating at Brown University. To make samples, tiny amount of NPs were totally

dissolved in 2 ml aqua regia (VHNO3 : VHCl = 1 : 3) and dried by heating. Subsequently,

2 % HNO3 was added to dissolve the precipitate for ICP-AES test.

2.2.6. Magnetic Measurements

33

Magnetic properties of NPs were evaluated by a vibrating sample magnetometry

(VSM, LakeShore 7404) with a maximum field of 1.5 T in our group. The magnetic

NPs were dried and then transferred into a round-bottled capsule (0.3 mL, Electron

Microscopy Science), and a piece of cotton is pressed firmly into the capsule on top to

fix the magnetic sample. The capsule was placed at the center of the external field to

measure its hysteresis loop. Also, the NPs dispersion can be deposited onto Si substrates.

After solvent evaporation, the NPs form a film and the magnetic property can be

measured in-plane and out-of-plane. The VSM can measure the hysteresis loops of Fe,

Co and ferrite, but the field is not strong enough to measure rare-earth containing alloy

like SmCo5. In this case, a physical property measurement system (PPMS) with a field

up to 9 T from Physics Department at Brown and a PPMS with a field up to 14 T in

Lawrence Livermore National Laboratory (LLNL) was used (Figure 2-5).

Figure 2-5. Photograph for the magnetic measurement setup: vibrating sample

magnetometry (left) and physical property measurement system (right).

34

References:

1 A. Calka, D. Wexler, Nature 2002, 419, 147.

2 Y. Zhang, L. Zhang, C. Zhou, Acc. Chem. Res., 2013, 46, 2329.

3 T. P. Yadav, R. M. Yadav, D. P. Singh, Nanoscience and Nanotechnology 2012, 2,

22.

4 J. Yang, D. Yan, T. S. Jones, Chem. Rev. 2015, 115, 5570.

5 S. Peng, Y. Lee, C. Wang, H. Yin, S. Dai, S. Sun, Nano Res. 2008, 1, 229.

6 B. L. Cushing, V. L. Kolesnichenko, C. J. O'Connor, Chem. Rev. 2004, 104, 3893.

7 W. Zhu, R. Michalsky, O. Metin, H. Lv, S. Guo, C. J. Wright, X. Sun, A. A. Peterson,

S. Sun, J. Am. Chem. Soc. 2013, 135, 16833.

8 W. Zhu, Y. Zhang, H. Zhang, H. Lv, Q. Li, R. Michalsky, A. A. Peterson, S. Sun, J.

Am. Chem. Soc. 2014, 136, 16132.

9 C. Wang, H. Daimon, T. Onodera, T. Koda, S. Sun, Angew. Chem. Int. Ed. 2008, 47,

3588.

10 N. Yang, Z. Zhang, B. Chen, Y. Huang, J. Chen, Z. Lai, Y. Chen, M. Sindoro, A.

Wang, H. Cheng, Z. Fan, X. Liu, B. Li, Y. Zong, L. Gu, H. Zhang, Adv. Mater. 2017,

29, 1700769.

11 M. Grzelczak, J. Pérez-Juste, P. Mulvaney, L. M. Liz-Marzán, Chem. Soc. Rev. 2008,

37, 1783.

12 C. Chen, Y. Kang, Z. Huo, Z. Zhu, W. Huang, H. L. Xin, J. D. Snyder, D. Li, J. A.

Herron, M. Mavrikakis, M. Chi, K. L.More, Y. Li, N. M. Markovic, G. A. Somorjai, P.

Yang, V. R. Stamenkovic, Science 2014, 343, 1339.

35

13 S. Zhang, Y. Hao, D. Su, V. V. T. Doan-Nguyen, Y. Wu, J. Li, S. Sun, C. B. Murray,

J. Am. Chem. Soc. 2014, 136, 15921.

14 W. Chen, R. Yu, L. Li, A. Wang, Q. Peng, Y. Li, Angew. Chem. Int. Ed. 2010, 49,

2917.

15 X. Sun, S. Guo, C. Chung, W. Zhu, S. Sun, Adv. Mater. 2013, 25, 132.

16 S. Sun, C. B. Murray, D. Weller, L. Folks, A. Moser, Science, 2000, 287, 1989.

17 C. B. Murray, C. R. Kagan, M. G. Bawendi, Annu. Rev. Mater. Sci. 2000, 30, 545.

18 J. Park, J. Joo, S. G. Kwon, Y. Jang, T. Hyeon, Angew. Chem. Int. Ed. 2007, 46,

4630.

19 J. Park, K. J. An, Y. S. Hwang, J. G. Park, H. J. Noh, J. Y. Kim, J. H. Park, N. M.

Hwang, T. Hyeon, Nat. Mater. 2004, 3, 891.

20 S. H. Sun, H. Zeng, J. Am. Chem. Soc. 2002, 124, 8204.

21 S. H. Sun, H. Zeng, D. B. Robinson, S. Raoux, P. M. Rice, S. X. Wang, G. X. Li, J.

Am. Chem. Soc. 2004, 126, 273.

36

Chapter 3

Stabilizing Fe Nanoparticles in the SmCo5 Matrix to Synthesize

SmCo5-Fe nanocomposite

Reprinted with permission from Nano Lett. 2017, 17, 5695-5698. Copyright ©2017

American Chemical Society.

37

3.1 Introduction:

Embedding a nanoscale soft magnetic phase into a hard-magnetic matrix is a key

step to develop exchange-spring nanocomposites with optimum energy product and a

reduced demand for critical rare earth elements.1-5 Such nanocomposites can show

magnetic performances that are superior to the corresponding single component hard

magnets, and serve as a new class of high performance permanent magnets for

applications in device miniaturization and in efficient energy conversions.

Conventional high performance permanent magnets are made of rare-earth metal alloys

based on NdFeB or SmCo, among which SmCo magnets are important for high

temperature applications due to their high Curie temperatures (747oC) and large

magnetocrystalline anisotropy constant (up to Ku = 1.7 x 108 erg cm-3 for the hcp-

SmCo5).6-11 Compared with NdFeB, the SmCo magnets have lower magnetization (M)

values, limiting their energy density they can store.12 An obvious solution to enhance a

SmCo magnet performance is to increase its M value, which may be achieved by

incorporating a high M nanoscale soft phase in its matrix, forming an exchange-coupled

composite.13,14 This has led to the extensive efforts in developing a proper method to

prepare such magnetic nanocomposites, including melt-spinning into ribbons,15,16

mechanically ball-milling into powders17,18, and sputtering into thin films.7,19,20 To

better control the size of the soft phase in the composite structure, solution phase

chemical synthesis methods are also tested.13, 21-23 Despite these efforts, it is still

extremely difficult to maintain the size of the soft phase in the composites due to the

harsh reductive annealing conditions required for the formation of SmCo5 alloy

38

structure. This annealing often induces an uncontrolled diffusion of the soft phase into

the hard phase, forming an alloy which destroys the desired exchange-coupling and

degrades the magnetic performance.

In the process of testing to stabilize FePt nanoparticles (NPs) during a high

temperature annealing condition, a robust inorganic coating layer, such as MgO or SiO2,

was applied to prevent NP sintering at temperatures as high as 800oC.24,25 MgO was

removed by acid washing while SiO2 was dissolved with a base to give well-dispersed

hard magnetic FePt NPs. Since the acid washing process to remove MgO is

incompatible with the condition used to stabilize Fe NPs, we tested to use the SiO2

coating to stabilize Fe NPs. We found that Fe NPs were indeed stabilized even in the

condition leading to the reductive conversion of SmCo-OH to SmCo. Herein, we report

our chemical approach to SmCo5-Fe nanocomposites with controlled Fe NP size. The

synthesis process, illustrated in Scheme 3-1, involves the precipitation of Sm(OH)3 and

Co(OH)2 and mixing the hydroxides with Fe/SiO2 followed by 850oC annealing in the

presence of calcium (Ca), in which SmCo-OH is reduced and converted to SmCo5 and

Fe/SiO2 core/shell structure is preserved. Removal of SiO2 by NaOH solution gives

SmCo5-Fe nanocomposites with 12 nm Fe NPs embedded in a SmCo5-matrix, which

shows tunable magnetic properties.

39

Scheme 3-1. Schematic illustration of the synthesis of SmCo5-Fe nanocomposite by

assembling Sm(OH)3 nanorods, Co(OH)2 nanoplates and Fe/SiO2 NPs, followed by

reductive annealing, NaOH solution washing and compaction.

3.2 Experimental details:

Chemicals: The syntheses were carried out using standard airless procedures and

commercially available reagents. All the following materials are commercially

available. Samarium chloride (99%), cobalt (II) chloride (98%), tetraethyl orthosilicate

(TEOS, 98%) were purchased from Strem Chemicals. Other were purchased from

Aldrich: iron pentacarbonyl (Fe(CO)5), sodium hydroxide (98%) 1-octadecene (ODE,

90%), oleylamine (OAm, 70%), hexadecylamine (HDA, 90%), polyoxyethylene (5)

nonylphenylether (Igepal CO-520) and HCl in diethylether (2.0 M).

Synthesis of HDA•HCl: An excess amount HCl in diethylether (6 mL, 2.0 M) was

dropwisely added to a solution of 10 mmol of HDA (2.44 g) in 100 mL of hexanes.

Then the white precipitate was formed and the solution was cooled in an ice bath. After

2h, the reaction mixture was warmed up to room temperature and was stirred for another

2 h. The precipitation was separated from the solution by centrifugation and was washed

for 3 times with hexanes. After dried in air, 2.1 g (78% yield) of HDA•HCl was obtained.

Synthesis of 12 nm Fe crystalline NPs: Typically, a mixture of 20ml ODE, 2ml

OAm and 0.28g HDA•HCl in a four-neck flask was heated to 120 oC under Ar flow for

1 h. Then it was further heated up to 180 oC. Under the blanket of Ar, 0.45 mL of

Fe(CO)5 was injected and the mixture was keep at 180 oC for another 30 min. The

40

mixture was then cooled down to room temperature and black material was collected

by the addition of ethanol and subsequent centrifugation (8500 rpm, 8 min) for 3 times.

The obtained black NPs were re-dispersed in cyclohexane.

Synthesis of Silica-coated NPs: In a flask, 1 mL IGEPAL-520 and 40 mL

cyclohexane were mixed under magnetic stirring. 20 mg of the Fe NPs was added to

the mixture and further stirred for 1h. Then 0.4 mL of tetraethyl orthosilicate and 0.4

mL of ammonium hydroxide (28%) were added into the solution. After 5 h, the mixture

was collected and precipitated by adding ethanol (20 mL) followed by centrifugation.

The precipitate was washed twice with ethanol (20 mL) and hexane (30 mL) and then

dried for further annealing.

Synthesis of Co(OH)2 and Sm(OH)3: 0.650g CoCl2 was dissolved in 40ml

deionized water. The aqueous suspension was stirred gently for 15 min to achieve good

homogeneity. Then the solution was heated to 100oC and 10ml of a 2M NaOH aqueous

solution was added dropwise into the solution. After leaving the reaction to reflux at

100 oC for 5 h, the nanoparticles were cooled down to room temperature and collected

by centrifugation. The brownish precipitates Co(OH)2 were filtered off, washed with

deionized water and dried at room temperature. The Sm(OH)3 synthesis procedure was

nearly the same except the initial chemical was 0.420g SmCl3 instead of CoCl2.

Synthesis of SmCo5-Fe nanocomposites: To synthesis SmCo5 + 10wt% Fe

nanocomposite, 0.149g Co(OH)2, 0.081g Sm(OH)3 and 0.06g Fe/SiO2 were ground

together and transferred onto 0.35g metallic Ca layer in the stainless-steel boat. Then

the mixture was annealed in argon at 850 oC for 30min at a rate of 25 oC min–1. After

41

being cooled down to room temperature, the powder was washed with deionized water

in argon atmosphere. Then 20ml NaOH solution (10M) was used to wash residual SiO2

in the composite at 60oC under sonication. After SiO2 removal. the particles were

compacted under 1.5 GPa for 24 h at 300 K in a piston cylinder apparatus. The amount

of Fe/SiO2 can be adjusted to make different ratio SmCo-Fe nanocomposites.

Characterization: TEM images were obtained by a Philips CM 20 operating at

200kV. High-angle annular dark-field scanning transmission electron microscopy

(HAADF-STEM) and EDS mapping were performed with an FEI TitanX to

characterize the elemental distribution of the nanocomposite. Powder XRD patterns of

the samples were recorded on a Bruker AXS D8-Advanced diffractometer with CuKa

radiation (λ= 1.5418 Å). The Sm/Co/Fe composition was determined by elemental

analysis using a JY2000 Ultrace ICP Atomic Emission Spectrometer. Magnetic

properties were measured using a Physical Property Measurement System (PPMS)

under a maximum applied field of 70 kOe.

3.3 Result and Discussion:

3.3.1 Synthesis and Characterization of Fe/SiO2 NPs

SmCo5 has a single domain size of 100-300 nm and domain wall width about 5-6

nm.22,26 Based on the commonly accepted model used to form effective exchange-

coupling between a hard and a soft phase, the soft phase should be around or below 12

nm to remain well-coupled to the SmCo5 phase.27 On the other hand, to maximize the

magnetization value, the soft NPs should be as large as possible so that the detrimental

42

surface effects of the NPs on the M enhancement can be minimized. With these

considerations in mind, we chose to synthesize monodisperse 12 nm Fe NPs and used

these NPs to demonstrate the new strategy leading to the formation of SmCo5-Fe with

the Fe NP morphology preserved. We prepared the Fe NPs by the decomposition of

Fe(CO)5 in the presence of oleyamine and hexadecylammonium chloride (HDA·HCl)

at 180oC.28 Figure 3-1a shows a transmission electron microscopy (TEM) image of the

12 nm Fe NPs (a thin layer oxide around each Fe NP is due to natural oxidation). The

Fe NPs have a bcc-structure (Figure 3-1b). We coated the Fe NPs with SiO2 by

controlled hydrolysis and condensation of tetraethyl orthosilicate (TEOS), in which

TEOS was hydrolyzed in the presence of ammonia to form a layer of SiO2 around each

Fe NP. The coating thickness was adjusted by the reaction time. Here we chose the 7

nm coating obtained from 5 h reaction (Figure 3-1c).

Figure 3-1. (a) TEM image of the as-synthesized 12 nm Fe NPs; (b) XRD pattern of

43

the as-synthesized 12 nm Fe NPs, showing the typical pattern that matches with the

standard bcc-Fe pattern; (c) TEM image of the 12 nm Fe NPs coated with 7 nm thick

SiO2 shell.

3.3.2 Synthesis and Characterization of SmCo5 Hard Magnet

The common bulk SmCo5 ingots are made of micro-structured SmCo5 with an

average diameter larger than 1000 nm. Their coercivities are between 1-5 kOe and Ms

around 50-60emu/g.29,30 The small coercivity values results from the large micrometer

grain sizes that cause the formation of multi-domains within the SmCo5 structure. To

increase Hc, SmCo5 should be prepared in less than 100-300 nm sizes. The direct

synthesis of nanostructured SmCo5 using the solution chemistry is challenging due to

the difficulty in co-reducing the Co2+ and Sm3+ in solution and in stabilizing the pre-

formed SmCo5 NPs against their fast oxidation.31 Nanostructured SmCo alloys are

typically synthesized by reductive annealing of SmCo-oxides at high temperatures, 32,33

similar to the commercial fabrication of SmCo magnets by high temperature reduction

of Sm-oxide and Co-oxide by Ca. In our current test, we first prepared nanostructured

Sm(OH)3 and Co(OH)2 by adding 2 M NaOH solution dropwise to an aqueous solution

of SmCl3 or CoCl2 at 100oC. Refluxing the reaction mixture for 5 h yielded Sm(OH)3

nanorods (60 nm 15 nm) (Figure 3-2a) or Co(OH)2 nanoplates (Figure 3-2b)

respectively. The hydroxide structures were further confirmed by XRD (Figure 3-3).

We used Ca to reduce the hydroxide mixture of Sm(OH)3 and Co(OH)2 at the molar

ratio of 1/4 at 850oC under an Ar atmosphere and obtained well-crystallized SmCo5

44

(Figure 3-2c) with the average size estimated by Scherrer’s formula to be 68nm. This

SmCo5 powder shows strong ferromagnetism at room temperature with Ms = 42 emu/g

and Hc = 20.1 kOe (Figure 3-2d). This coercivity value is one of the largest compared

to other SmCo5 prepared by chemical methods.10,13,14,23,34

Figure 3-2. TEM image of the Sm(OH)3 nanorods (a) and Co(OH)2 nanoplates (b); (c)

XRD of the SmCo5 powder obtained from our chemical synthesis (black curve) and

from the standard pattern (red lines, JPCDS No. 65-8981); (d) Hysteresis loop of the

SmCo5 powder measured at 300 K.

45

Figure 3-3. (a) XRD of hexagonal crystalline Co(OH)2 nanoplate precipitation (black)

and standard pattern of Co(OH)2 (red lines, JPCDS No. 89-8616). (b) XRD of 60nm

x15nm crystalline Sm(OH)3 nanorods (black) and standard pattern of Sm(OH)3 (red

lines, JPCDS No. 83-2036).

3.3.3 Embedding of Fe NPs into SmCo5 Matrix for Nanocomposite Fabrication

To obtain the SmCo-Fe composite, we mixed the Sm(OH)3 nanorods, Co(OH)2

nanoplates, and Fe/SiO2 NPs in ethanol under sonication. After decanting ethanol and

drying the powder under air, we ground the powder mixture together with Ca under an

Ar atmosphere and annealed the powder mixture at 850oC for 30 min (the optimum

annealing condition we obtained after a series of annealing tests from 800-900 C for

0.5 – 2 h.). Once cooled to room temperature, the powder was washed with distilled

water under argon to dissolve CaO and any unconsumed reactants (note that it is

important to prevent CO2 from presence in this process to avoid the formation of CaCO3

that is not water-soluble.). Then the powder was immersed in the pre-heated (60C) 10

M NaOH solution under sonication to remove the residual SiO2 in the composite. The

powder was further washed with water and ethanol and dried under vacuum at room

46

temperature. The Sm/Co/Fe composition in the composite was analyzed by inductively

coupled plasma-atomic emission spectroscopy (ICP-AES). SmCo5 was obtained from

the 1/4 SmCl3/CoCl2 precursors, indicating a small amount of Sm lost during the

annealing and/or subsequent washing processes.10 The Fe composition was carried over

to the final product.

Figure 3-4a shows the XRD patterns of different SmCo5-Fe composites prepared

from the reductive annealing. The crystal structure of the SmCo5 can be indexed with

the standard hcp-SmCo5. The more important part is that the bcc-Fe NP structure is

preserved, and the relative intensity of the characteristic bcc-Fe peaks increases with

the increasing Fe content in the composite, which indicates that Fe NPs survive in the

annealing procedure without obvious sign of diffusion into SmCo5 phase. The

morphology of the Fe NPs in the SmCo5-Fe composite was further characterized by

high angle annular dark field scanning TEM (HAADF-STEM) analysis (Figure 3-4b).

The brighter particles embedded inside the relatively dark background. EDX elemental

mapping (Figure 3-4c) shows the red dots with an average size of 12-13 nm

representing Fe NPs. Both XRD and TEM analyses show that after annealing, the Fe

NPs were well protected in the SmCo5 matrix and showed no obvious sign of

aggregation/sintering.

47

Figure 3-4. (a) XRD patterns of SmCo5-Fe(x wt%) composite with x = 0, 5, 10 and 20.

(b) HAADF-STEM image and (c) elemental mapping of the SmCo5-Fe(10 wt%)

composite. Note: the overall Fe NP content is in 10 wt%, but the image shows an area

enriched with Fe NPs.

Figure 3-5a shows room temperature magnetic hysteresis loops of the SmCo5-Fe

composites with different Fe NP weight percentages (wt%). It shows incorporation of

Fe NPs into ferromagnetic SmCo5 matrix changes both Hc and Ms of the composites.

Ms monotonically increases from 42.5 emu/g for the pure SmCo5 to 77.6 emu/g for the

SmCo5-Fe(20 wt%) nanocomposite, while Hc decreases from 20.1 to 11.2 kOe (inset of

Figure 3-5a). We should note that the pure SmCo5 shows a single-phase behavior while

the composites have kinks in their demagnetization curves. To prove that the kinks arise

from the loose Fe NP packing within the SmCo5 matrix, not from the “overdose” of the

Fe NPs in the composites, we compacted the powders in piston cylinder apparatus at

room temperature (high temperature may lead to the uncontrolled grain growth.8,35

48

Figure 3-6). Figure 3-5b shows the magnetic properties of the SmCo5-Fe (10 wt%)

pressed under 1.5 GPa for 24 h at 300 K. After compaction, the nanocomposite shows

a near single-phase magnetic behavior with its Ms increasing from 61.5 emu/g to 63.9

emu/g but Hc deceasing from 13.2 kOe to 10.5 kOe. SmCo5-Fe (20 wt%)

nanocomposite was also pressed in the same condition and its M-H behavior (Figure

3-7) becomes similar to what is shown in the compressed SmCo5-Fe (10 wt%). These

improved magnetization reversal behaviors of the compressed composites are indicative

of the enhanced interaction between SmCo5 and Fe. We may conclude that compression

does help to establish the desired exchange-coupling between SmCo5 and 12 nm Fe NPs

in the SmCo5-Fe structure. More studies on homogeneous distribution of Fe NPs into

SmCo5 matrix and the control of SmCo5 phase in more uniform nanoscale dimensions

in the SmCo5-Fe nanocomposites are underway.

Figure 3-5. (a) Hysteresis loops of the nanocomposites of SmCo5-Fe(x wt%) (x = 0–20)

nanocomposites at 300 K. Inset: the change of Hc and Ms with the different Fe NP

contents in the SmCo5-Fe nanocomposites; (b) Hysteresis loops of the nanocomposite

of SmCo5-Fe(10 wt%) before (black) and after (red) 1.5 GPa compaction at 300 K.

49

Figure 3-6. A photograph of cylindrical form, compressed SmCo5-Fe nanocomposite.

Figure 3-7. Hysteresis loops of the nanocomposites of SmCo5 + 20 wt. % Fe

nanocomposites before and after 1.5 GPa press at 300K. The Ms increases from 78.6

emu/g to 82.7 emu/g. Coercivity decreases from 11.2kOe to 8.1kOe.

3.4 Conclusion

In summary, we have reported a new strategy to stabilize Fe NPs in the high

temperature (850C) reductive annealing condition that leads to the reduction of

Sm(OH)3 and Co(OH)2 to hard magnetic SmCo5. The Fe/SiO2 NPs are mixed with

Sm(OH)3 nanorods and Co(OH)2 nanoplates in ethanol under sonication, forming a

composite of Sm(OH)3-Co(OH)2-Fe/SiO2. Upon high temperature (850 C) reduction

50

by Ca, the hydroxides are converted into SmCo5 powder with Fe NPs staying intact in

the SiO2 enclosure. Washing with 10 M NaOH solution, water and ethanol removes the

SiO2 coating around each Fe NP, giving the SmCo5-Fe nanocomposite with the Fe NP

morphology preserved and Fe content tunable (up to 20 wt% in this paper). The Fe NPs

and SmCo5 are not strongly coupled in the loose powder but their interaction can be

enhanced by compaction under 1.5 GPa at room temperature. As a result, the SmCo5-

Fe composites show the one-phase behavior and their magnetic properties are tunable

by wt% of the Fe NPs. Work on controlled syntheses of nanostructured SmCo and

SmCo-M (M = Fe, Co, or FeCo) are on the way to obtain the optimum magnetic

performance from these isotropic composites. The Sm(OH)3 nanorods will be further

explored as the starting precursor to prepare anisotropic SmCo or SmCo-M for high

performance permanent magnet applications.

51

References:

(1) Skomski, R.; Coey, J. M. D. Phys. Rev. B 1993, 48, 15812-15816.

(2) Zeng, H.; Li, J.; Liu, J. P.; Wang, Z. L.; Sun, S. H. Nature 2002, 420, 395-398.

(3) Liu, F.; Hou, Y.; Gao, S. Chem. Soc. Rev. 2014, 43, 8098-8113.

(4) Zeng, H.; Li, J.; Wang, Z.; Liu, J.; Sun, S. Nano lett. 2004, 4, 187-190.

(5) Kneller, E. F.; Hawig, R. IEEE Trans. Magn. 1991, 27, 3588−3600

(6) Cui, W. B.; Takahashi, Y. K.; Hono, K. Adv. Mater. 2012, 24, 6530-6535.

(7) Sawatzki, S.; Heller, R.; Mickel, C.; Seifert, M.; Schultz, L.; Neu, V. J. Appl. Phys.

2011, 109, 123922.

(8) Rong, C.; Zhang, Y.; Poudyal, N.; Szlufarska, I.; Hebert, R. J.; Kramer, M. J.; Liu,

J. P. J Mater. Sci. 2011, 46, 6065-6074.

(9) Li, H.; Li, X.; Guo, D.; Lou, L.; Li, W.; Zhang, X. Nano lett. 2016, 16, 5631-5638.

(10) Hou, Y.; Xu, Z.; Peng, S.; Rong, C.; Liu, J. P.; Sun, S. Adv. Mater. 2007, 19, 3349-

3352.

(11) Cui, B. Z.; Gabay, A. M.; Li, W. F.; Marinescu, M.; Liu, J. F.; Hadjipanayis, G. C.

J. Appl. Phys. 2010, 107, 09A721.

(12) Poudyal, N.; Ping Liu, J. J. Phys. D: Appl. Phys. 2013, 46, 043001.

(13) Hou, Y.; Sun, S.; Rong, C.; Liu, J. P. Appl. Phys. Lett. 2007, 91, 153117.

(14) Ma, Z.; Zhang, T.; Jiang, C. RSC Adv. 2015, 5, 89128-89132.

(15) Panagiotopoulos, I.; Withanawasam, L.; Hadjipanayis, G. J. Magn. Magn. Mater.

1996, 152, 353-358.

(16) Zhang, X.; Guan, Y.; Yang, L.; Zhang, J. Appl. Phys. Lett. 2001, 79, 2426-2428.

52

(17) Rai, B. K.; Mishra, S. R. J. Magn. Magn. Mater. 2013, 344, 211-216.

(18) Li, X.; Lou, L.; Song, W.; Zhang, Q.; Huang, G.; Hua, Y.; Zhang, H.; Xiao, J.;

Wen, B.; Zhang, X. Nano lett. 2017, 17, 2985–2993.

(19) Fullerton, E. E.; Jiang, J.; Bader, S. J Magn. Magn. Mater. 1999, 200, 392-404.

(20) Zhang, J.; Takahashi, Y.; Gopalan, R.; Hono, K. Appl. Phys. Lett. 2005, 86, 122509.

(21) Wu, L.; Mendoza-Garcia, A.; Li, Q.; Sun, S. Chem. Rev. 2016, 116, 10473-10512.

(22) Yang, C.; Jia, L.; Wang, S.; Gao, C.; Shi, D.; Hou, Y.; Gao, S. Sci. Rep. 2013, 3,

3542.

(23) Chaubey, G. S.; Poudyal, N.; Liu, Y.; Rong, C.; Liu, J. P. J. Alloys Compd. 2011,

509, 2132-2136.

(24) Li, Q.; Wu, L.; Wu, G.; Su, D.; Lv, H.; Zhang, S.; Zhu, W.; Casimir, A.; Zhu, H.;

Mendoza-Garcia, A.; Sun, S. Nano lett. 2015, 15, 2468-2473.

(25) Lee, D. C.; Mikulec, F. V.; Pelaez, J. M.; Koo, B.; Korgel, B. A. J. Phys. Chem. B

2006, 110, 11160-11166.

(26) De Campos, M. F.; Landgraf, F. J. G. Mater. Sci. Forum 2005, 498, 129-133.

(27) Rong, C.; Zhang, Y.; Poudyal, N.; Xiong, X.; Kramer, M. J.; Liu, J. P. Appl. Phys.

Lett. 2010, 96, 102513.

(28) Lacroix, L. M.; Huls, N. F.; Ho, D.; Sun, X.; Cheng, K.; Sun, S. Nano lett. 2011,

11, 1641-1645.

(29) Chen, C.; Knutson1, S.; Shen, Y.; Wheeler, R.; Horwath, J, and Barnes, P. Appl.

Phys. Lett. 2011, 99, 012504

(30) Knutson1, S.; Shen, Y.; Horwath, C.; Barnes, P.; Chen, C. J. Appl. Phys. 2011,

53

109, 07A762

(31) Chinnasamy, C. N.; Huang, J. Y.; Lewis, L. H.; Latha, B.; Vittoria, C.; Harris, V.

G. Appl. Phys. Lett. 2008, 93, 032505.

(32) Ma, Z.; Zhang, T.; Jiang, C. Chem. Eng. J. 2015, 264, 610-616.

(33) Zheng, L.; Cui, B.; Zhao, L.; Li, W.; Zhu, M.; Hadjipanayis, G. C. J. Nanopart.

Res. 2012, 14, 1129.

(34) Ma, Z.; Yang, S.; Zhang, T.; Jiang, C. Chem. Eng. J. 2016, 304, 993-999.

(35) Rong, C.; Poudyal, N.; Liu, X.; Zhang, Y.; Kramer, M.; Liu, J. P. Appl. Phys. Lett.

2012, 101, 152401.

54

Chapter 4

A New Strategy to Synthesize Anisotropic SmCo5 Nanomagnets

Reprinted with permission from Nanoscale. 2018, 10, 8735-8740. Copyright © The

Royal Society of Chemistry 2018.

55

4.1 Introduction

Developing nanomagnets that contain rare-earth metal alloys is an important step

to maximize their magnetic performance for miniaturization of magnetic and electronic

devices.1-5 Among two classes of well-known rare earth magnets of Nd-Fe-B and Sm-

Co, SmCo alloys, such as SmCo5, are especially sought after for high temperature

applications due to their large magnetocrystalline anisotropy constant (1.7 × 108

erg/cm3) and high Curie temperature (747 C).6-10 Compared to isotropic SmCo5

nanomagents, anisotropic SmCo5 ones attract even more attention due to their square

hysteresis behaviors and their capability of storing high magnetic energy densities.11-12

However, fabrication of anisotropic SmCo5 is a very challenging goal to reach thus far

due to the difficulty in controlling the texture of nanosized SmCo5 and the fast oxidation

of Sm at the nanoscale. Previously, attempts to make anisotropic SmCo5 by using ball

milling, spark sintering, and spin melting, often yield microstructured magnets without

showing the desired enhancement in magnetics.13-16 Among these work, the best SmCo5

made from ball milling following by annealing shows a coercivity of 41.5 kOe14, but

the size is around 280-400 nm, which is hard to be dispersed in solution and shows two

phase behaviour hysteresis. Solution phase based chemical reduction methods have also

been explored to prepare nanostructured SmCo5, but they only lead to the formation of

shape-isotropic SmCo5.17-21

Herein, we report a new strategy to synthesize anisotropic SmCo5 nanoplates that

can be aligned magneically to show large magnetic coercivity. The method, illustrated

in Figure 4-1, includes the preparation of a nanocomposite of Sm(OH)3 nanorods (NRs)

56

and Co nanoparticles (NPs) via self-assembly, the coating of this nanocomposite with

a protective CaO layer, and high temperature (850 C) reduction of Sm(OH)3-Co

nanocomposite to obtain SmCo5 nanoplates. With the Co NP size fixed at 10 nm, the

right Sm/Co ratio is realized by controlling the NR dimension. These nanoplates (125

nm × 10 nm) could be suspended in ethanol and further mixed in epoxy resin. Under an

external magnetic field of 20 kOe, these nanoplates can be aligned with the plates

stacking along their crystallographic c-direction. The anisotropic SmCo5 nanoplate

assembly in epoxy resin has a square hysteresis behavior with its room temperature

coercivity reaching 30.1 kOe, which is among the highest values ever reported for

nanostructured SmCo5. Our synthesis offers a promising new approach to the

fabrication of anisotropic SmCo5 nanomagnets for high performance permanent

magnetic applications.

Figure 4-1. Schematic illustration of the synthesis of anisotropic SmCo5 nanoplates by

self-assembly of Sm(OH)3 NRs and Co NPs, followed by CaO coating and reductive

annealing.

4.2 Experimental Details

Chemicals: The syntheses were carried out using standard airless procedures and

commercially available reagents. Samarium chloride (99%), calcium acetylacetonate

57

(Ca(acac)2, 98%) and metallic Ca (99%) were purchased from Strem Chemicals.

Hexadecyltrimethylammonium hydroxide (HTMA-OH, 25% in methanol) was

purchased from TGI America. Cobalt carbonyl (Co2(CO)8), sodium hydroxide (98%),

tetralin (1,2,3,4 - tetrahydronaphthalen, 99%), 1-octadecene (ODE, 90%), oleylamine

(OAm, 70%), dioctylamine (98%), oleic acid (90%) were from Sigma-Aldrich.

Synthesis of Sm(OH)3 NRs: 0.42 g SmCl3 was dissolved in a solution of 40 mL

deionized water and 10 mL ethanol. The aqueous solution was heated to 90 °C and 10

mL of 2 M NaOH aqueous solution was added dropwise. After refluxed at 90 °C for 5

h, the solution was cooled to room temperature and the product was collected by

centrifugation (8000 rpm, 8 min). The white precipitate, Sm(OH)3 NRs, was further

washed with deionized water and dried at room temperature.

Synthesis of Co NPs: A mixture of 17 mL tetralin, 0.35 mL of oleic acid and 0.5

mL of dioctylamine in a four-neck flask was heated to 120 °C under argon flow for 1 h.

Then the solution was heated up to 210 °C. Under a blanket of argon, a solution of 0.27

g Co2(CO)8 dissolved in 3 mL tetralin was injected and the solution was kept at 210 °C

for 30 min. The solution was then cooled to room temperature and black NPs were

collected by the addition of ethanol and subsequent centrifugation (8500 rpm, 8 min).

The solid product was dispersed in hexane (15 mL) and precipitated by adding ethanol

(20 mL) and by centrifugation. The Co NPs were re-dispersed in hexane for further use.

Synthesis of Sm(OH)3-Co composite: To prepare Sm(OH)3-Co composite, 0.02 g

of Sm(OH)3 NRs were suspended in 20 mL hexane under sonication. 0.027 g of Co NPs

dispersed in 10 mL of hexane was added to the Sm(OH)3 suspension dropwise under

58

sonication. After 3 h of sonication, hexane dispersion was obtained and the Sm(OH)3-

Co composite was collected by adding ethanol (20 mL) and centrifugation (8500 rpm,

8 min). The solid product was measured by ICP-AES to have a Sm/Co mass ratio of

1:4.5 and was re-dispersed in hexane for further uses.

Coating Sm(OH)3-Co composite with CaO: 0.5 g Ca(acac)2, 20 mL ODE, 1 mL

oleic acid and 1 mL oleylamine were mixed in a flask under magnetic stirring. The

solution was heated to 100 °C under argon and kept at this temperature for 30 min. Then

0.1 g Sm(OH)3-Co (prepared from repeated synthesis due to the amount of hexane (30

mL) used in each synthesis) was added to the solution. After 10 min, 8 mL methanol

solution of HTMA-OH was added into the solution dropwise. The mixture was kept at

100 °C for 30 min to remove methanol. Then the solution was heated to 150 °C and

kept at 150 °C for 1 h before it was cooled to room temperature. The Sm(OH)3-Co/CaO

composite was precipitated by adding ethanol (20 mL) followed by centrifugation (8500

rpm, 8 min). The precipitate was washed twice with ethanol (2 x 40 mL) and hexane (2

x 10 mL) and then dried for further annealing. As we saw no loss of Sm and Co during

the coating process, we should still have 0.1 g of Sm(OH)3-Co in the CaO matrix.

Synthesis of anisotropic SmCo5 nanoplates: All processes were done under argon.

0.3 g Sm(OH)3-Co embedded in CaO was ground with 0.3 g metallic Ca in the stainless-

steel boat. Then the mixture was heated under an argon atmosphere to 850 °C at a rate

of 25 °C/min and kept at 850 °C for 30 min. After cooled to room temperature, the

powder was washed with deionized water to remove CaO and excess Ca. Then the

powder was milled in 10 mL ethanol in the presence of 0.1 mL oleic acid for 1 h to form

59

a dispersion (note: in the synthesis of Co NPs described above, oleate-coated Co NPs

were dispersed in hexane, but here, SmCo5 nanoplates were dispersible in ethanol,

suggesting the formation of a bilayer coating of oleate with hydrocarbon chains

intercalated.). The undispersed was removed by centrifugation (100 rpm, 10 seconds),

and the product in the dispersion was collected by a bar magnet and dried, giving 0.08

g SmCo5 nanoplates for further uses.

Embedding SmCo5 nanoplates in epoxy resin and magnetic alignment: All

processes were done under argon. First, 0.2 g epoxy resin was dissolved in 2 mL ethanol

to form a clear solution. Then 0.2 g SmCo5 nanoplates (obtained from the repeated

syntheses due to the volume constraint of the stainless-steel boat we used) in 2 mL

ethanol dispersion was added dropwise into the resin solution under sonication to obtain

a homogenous SmCo5-resin solution. After ethanol evaporation, the resin gel was

pasted on the surface of a TEM grid or a silicon substrate. The TEM grid or silicon

substrate were put in a 20 kOe field until the resin was solidified. For the TEM grid, the

external magnetic field was set parallel to the grid surface. For the silicon substrate, the

external magnetic field was set perpendicular to the substrate surface for XRD and

magnetic property measurements.

Characterization: TEM images and High-resolution TEM (HRTEM) images were

obtained on a JEOL 2010 TEM at 200 kV. High-angle annular dark-field scanning

transmission electron microscopy (HAADF-STEM) and STEM-electron energy-loss

spectroscopy (STEM-EELS) elemental mapping were collected from a Hitachi

HD2700C (200 kV) to characterize the elemental distribution of the nanoplates. Powder

60

X-ray diffraction (XRD) patterns of the SmCo5 nanoplates were recorded on a Bruker

AXS D8-Advanced diffractometer with CuKa radiation (λ = 1.5418 Å). The Sm/Co

composition was determined by elemental analysis using a JY2000 Ultrace Inductively

coupled plasma-atomic emission spectroscopy (ICP-AES). Magnetic properties were

measured on a Physical Property Measurement System (PPMS) under a maximum

applied field of 90 kOe.

4.3 Result and Discussion

4.3.1 Synthesis of Sm(OH)3-Co Nanocomposite

Monodisperesed 10 (± 1) nm Co NPs were obtained by decomposition of Co2(CO)8

in tetralin solution of dioctylamine and oleic acid (Figure 4-2a).22 The XRD shows the

Co NPs have a crystalline face centered cubic (fcc) structure (Figure 4-2b). Separately,

Sm(OH)3 NRs were prepared in aqueous solution. It was important here to control NR

aspect ratio to obtain the correct Sm/Co ratio. For example, if Sm(OH)3 NRs were

prepared by precipitating aqueous solution of SmCl3 with 2 M NaOH at 100 C as we

described previously,18 we could only obtain 60 x 15 nm Sm(OH)3 NRs that were

unsuitable for the formation of Sm(OH)3-Co composites with the right 1/5 Sm/Co ratio.

In the current synthesis, we still used 2 M NaOH to precipitate SmCl3, but the reaction

was controlled in a mixture solution (40 ml deionized water and 10 ml ethanol) at 90

C (refluxing) for 5 h, which yielded 125 (± 25) nm × 12 (± 3) nm Sm(OH)3 NRs

(Figure 4-2c). The hexagonal Sm(OH)3 structure was confirmed by X-ray diffraction

(XRD) (Figure 4-2d). The Sm(OH)3 NRs obtained from our current new synthesis have

61

a higher aspect ratio (about 10), allowing to accomodate more Co NPs to reach the

desired Sm/Co ratio close to 1/5.

Figure 4-2 (a) TEM image of the as-synthesized 10 nm Co NPs. (b) XRD of the as-

prepared Co NPs (black curve) and the standard fcc-Co (red lines, JPCDS No. 15-0806).

(c) TEM image of 125 x 12 nm Sm(OH)3 NRs. (d) XRD of the as-prepared Sm(OH)3

NRs (black curve) and the standard pattern of Sm(OH)3 (red lines, JPCDS No. 83-2036).

4.3.2 Synthesis and Charactorization of SmCo5 Nanoplates

The hexane dispersion of Co NPs were added dropwise to a suspension of

Sm(OH)3 NRs in hexane at a controlled molar ratio (Sm:Co = 1:4.5 for the synthesis of

SmCo5). After 3 hours of sonication, the Sm(OH)3-Co nanocomposite was obtained

(Figure 4-3a). The rod-like nanocomposite work as the precursor for the reductive

annealing. We should emphasize that the anisotropic feature of Sm(OH)3 NRs and the

62

Co NP attachment to the NRs is essential for the formation of SmCo5 nanoplates from

the next step reduction procedure. Other combinations of Sm(OH)3 NRs and Co(OH)2,

including Co(OH)2 matrix coating over Sm(OH)3 NRs, could only yield shape-isotropic

SmCo5 without good control on nanostructures (Figure 4-4). Nanostructured SmCo5

was obtained by high temperature (850 C) annealing of Sm(OH)3-Co nanocomposite

embedded in CaO in the presence of Ca. Here CaO was specifically chosen for the

SmCo5 stabilization need because of following benefits: 1) CaO is thermally very stable,

having a high melting point (above 2500 °C); 2) CaO is compatible with the Ca

reduction process, which also leads to the formation of CaO; 3) CaO can be removed

easily by water washing, facilitating SmCo5 product purification. In the experiment, we

mixed the nanocomposite with Ca(acac)2 and HTMA-OH in 1-octadecene and heated

the solution at 150 C for 1 h to allow the decomposition of Ca(acac)2 to CaO. Figure

4-3b shows a representative TEM image of the Sm(OH)3-Co nanocomposite embedded

in the CaO matrix. ICP-AES analysis confirmed that the CaO coating process had no

effect on Sm/Co composition. After drying the powder in air, we ground the Sm(OH)3-

Co/CaO nanocomposite with Ca under an argon atmosphere and annealed the mixture

at 850 C. We monitored the annealing process by sampling and characterizing a small

amount of the annealed product by TEM at different times. 10 min after the annealing,

Co NPs started to diffuse into the Sm(OH)3 NRs (Figure 4-3c). After 30 min annealing,

SmCo5 nanoplates were formed. Once cooled to room temperature, the powder was

washed with distilled water under argon to remove CaO and any unconsumed reactants.

After washing, the powder was ground in ethanol and formed a dispersion. A gentle

63

centrifugation (100 rpm) was applied to remove a small amount of precipitate from the

dispersion, then the product was collected from the dispersion by a bar magnet. The

product had a final Sm/Co composition of 1:5 as analyzed by ICP-AES, which is

reduced from the original 1:4.5, suggesting a small Sm loss during the annealing and

washing processes. Figure 4-3d shows the TEM image of the as-synthesized SmCo5

nanoplates with their hexagon-like lateral dimension in 125 ± 25 nm and thickness

around 10 ± 5 nm. The chemical composition of the nanoplates was further

characterized with HAADF-STEM analysis and STEM-EELS elemental mapping

(Figure 4-3e). The elemental distribution confirms the presence of Sm (red) and Co

(green) across the nanoplate, and the combined (green and red) image shows Sm and

Co elements are homogeneously distributed in the nanoplates, indicating that the

nanoplates have a uniform SmCo5 alloy structure.

Figure 4-3 (a) TEM image of Sm(OH)3-Co nanocomposite with Sm:Co =1:4.5 (molar

64

ratio). (b) TEM image of Sm(OH)3-Co nanocomposite embedded in CaO matrix. (c)

TEM image of Sm(OH)3-Co nanocomposite obtained 10 min after the annealing. (d)

TEM image of the as-synthesized SmCo5 nanoplates. (e) HAADF-STEM and elemental

mapping of the SmCo5 nanoplates, showing the formation of uniform alloy structure

within each nanoplate.

Figure 4-4. (a) TEM image of Sm(OH)3 NRs embedded in Co(OH)2 matrix. (b) TEM

image of SmCo5 obtained after Ca reduction of (a), showing no specific shape feature.

The detailed structure of the nanoplate was analyzed by HRTEM image. Figure

4-5a is a plane-view HRTEM image of a representative nanoplate. The distance of the

lattice fringe was measured to be 2.15 Å that is close to the lattice spacing of (200)

planes of the hexagonal SmCo5 (2.16 Å). The fast Fourier transform (Figure 4-5b)

pattern obtained from Figure 4-5a matches with the simulated electron diffraction

pattern (Figure 4-5c) along [001] zone axis of a hexagonal phase of the SmCo5

(P6/mmm). The atomic arrangement of the nanoplate revealed by HRTEM (Figure 4-

5d) shows the same Sm, Co periodicity as that from an atom model built along [001]

zone axis of SmCo5 crystal lattice (Figure 4-5e). All these analyses support that the c-

axis of the SmCo5 nanoplate is perpendicular to the hexagonal plane. HRTEM image

65

of the side-view of a nanoplate (Figure 4-5f) show two kinds of lattice fringes with

their interfringe distances at 2.48 Å and 1.99 Å, which correspond to lattice spacing of

(110) planes (2.49 Å) and (002) planes (1.98 Å) of SmCo5 respectively. Such image

agrees well with the simulated atom model along [1, -1, 0] zone axis of SmCo5 crystal

lattice (Figure 4-5g). The further supports that the c-axis is perpendicular to the plane

of the SmCo5 nanoplate.

Figure 4-5. (a) HRTEM image of a part of one SmCo5 nanoplate (planar view). (b) Fast

Fourier transform pattern of (a). (c) Simulated SAED pattern of hexagonal SmCo5

projected along the c-axis. (d) A fraction of HRTEM imaging area in showing the

arrangement of Sm and Co atoms. (e) Modeled hexagonal SmCo5 structure projected

along the c-axis. (f) HRTEM image of the side-view of a SmCo5 nanoplate. (g) Modeled

SmCo5 structure projected along [1, -1, 0].

XRD peaks of the nanoplate powder confirm that the nanoplates have the

crystalline hexagonal D2d structure of SmCo5 (Figure 4-6a). The pattern intensity fits

66

well with the standard SmCo5, indicating that the nanoplates in the powder form have

no preferred texture. The powder was strongly ferromagnetic at room temperature with

its coercivity (Hc) and saturation magnetization (Ms) values at 25.3 kOe and 52.5 emu/g,

respectively (Figure 4-6b). The Ms value is reduced from the bulk SmCo5 value (99

emu/g) due to surface coating of oleate (for forming the nanoplate dispersion) and the

related nanoscale surface effects, which is consistent with what have been observed on

nanostructured SmCo5.23, 24

Figure 4-6. (a) XRD of the as-synthesized SmCo5 nanoplate powder (black curve) and

the standard pattern of D2d structure SmCo5 (red lines, JPCDS No. 65-8981). (b)

hysteresis loop of the as-synthesized SmCo5 nanoplate powder measured at 300 K.

4.3.3 Alignment of SmCo5 Nanoplates in Polymer

The dispersible SmCo5 nanoplates in ethanol made it possible to align them under

a magnetic field. To demonstrate this point, we first tested the assembly during ethanol

evaporation in a 20 kOe magnetic field, and found that the sample was only partially

aligned (Figure 4-7). We then mixed the SmCo5 nanoplates and epoxy resin in ethanol

at the mass ratio of 1:1. After ethanol evaporation, the SmCo5 nanoplates were

67

embedded in epoxy resin, and aligned in the same 20 kOe field before the resin was

hardened, as indicated in Figure 4-8a. As the c-axis of each of the nanoplates is also its

magnetic easy axis direction, the aligned SmCo5 nanoplates should stack face-to-face

along the field direction. From the TEM images (Figure 4-8b), we can see that the

SmCo5 nanoplates are aligned along the external field direction with a face-to-face

arrangement. Furthermore, we measured the XRD diffraction patterns and used the

relative intensity ratio between the diffraction peaks of (002) and (111), I(002)/I(111), to

measure the alignment factor (The intensity ratio is 0.26 for an isotropic SmCo5

sample).17 As shown in Figure 4-8c, the non-aligned SmCo5 nanoplates give the

I(002)/I(111) ratio of 0.4, suggesting that there is some degree of alignment in the powder

product due likely to the nanoplate shape effect. After magnetic field alignment, the

(111) peak nearly disappears and the I(002)/I(111) ratio increases to 20, indicating a strong

texture with the SmCo5 nanoplates parallel to the substrate. These aligned nanoplates

show obvious anisotropic magnetic hysteresis behavior at room temperature - the out-

of-plane loop is square (Hc = 30.1 kOe and Ms = 66.1 emu/g) while the in-plane one is

minor, showing no Ms at a 90 kOe field (Figure 4-8d). The measured Hc (30.1 kOe)

from the aligned nanoplate assembly is among the highest values ever reported for

nanostructured SmCo5.17-21, 23-26 The magnetic alignment factor can be quantitatively

measured by the remanence ratio (mr), which is defined as Mr/Ms, (Mr is the remanence

along the aligned direction).27-29 For a group of Stoner-Wohlfarth type particles, this

remanence ratio is 0.5 for the randomly oriented particles but 1 for a perfectly aligned

particle assembly.30 Practically, the larger the mr, the better the magnetic alignment.

68

The loop from the aligned SmCo5 nanoplates in Figure 4-8d has mr = 0.92, which is

among the highest values ever reported,8,11,12,30 indicating a high anisotropic order of

the SmCo5 nanoplates in resin.

Figure 4-7. XRD pattern of SmCo5 nanoplates obtained from their ethanol dispersion

after ethanol evaporation under a 20 kOe field.

`

Figure 4-8 (a) Schematic illustration of SmCo5 nanoplate alignment in resin along the

magnetic field direction for TEM and XRD characterizations. (b) TEM image of the

69

aligned SmCo5 nanoplates embedded in resin. (c) XRD patterns of the non-aligned

SmCo5 (black curve) and the aligned SmCo5 nanoplates (red curve). (d) Room

temperature hysteresis loops of the aligned SmCo5 nanoplates measured along the c axis

(black curve) and perpendicular to the c axis (red curve).

4.4 Conclusion

In summary, we have reported a novel method to synthesize dispersible SmCo5

nanoplates and to align them in resin to obtain anisotropic nanoplate assemblies. The

key process is first to assemble 10 nm Co NPs along 125 nm × 12 nm Sm(OH)3 NRs

and then to embed the Sm(OH)3-Co nanocomposite in CaO matrix for high temperature

(850 C) annealing in the presence Ca. The CaO coating ensures Co diffusion and

alloying with Sm in the annealing condition, and the NR shape facilitates the formation

of 125 × 10 nm SmCo5 nanoplates. An important feature of these nanoplates is that they

can be dispersed in ethanol and therefore, be assembled in resin under a magnetic field

to allow the SmCo5 nanoplates to stack face-to-face, establishing the desired anisotropic

texture and magnetic alignment. The aligned anisotropic SmCo5 nanoplates have a

square hysteresis loop with a room temperature Hc of 30.1 kOe that is among the largest

coercivity values ever reported for SmCo5. With the Sm(OH)3 NR dimension and Co

NP size controls, SmCo5 nanoplate dimensions should in principle be further tuned to

achieve optimum magnetic performance. The dispersion nature of these nanoplates

should also allow assembly of SmCo5 with other high moment magnetic NPs of Co, Fe,

or FeCo, making it possible to fabricate anisotropic SmCo5-M (M = Co, Fe, or FeCo)

70

exchange-coupled nanocomposites with optimum magnetics for high performance

permanent magnet applications.

71

References:

1 R. Skomski, J. M. D. Coey, Phys. Rev. B 1993, 48, 15812-15816.

2 F. Liu, Y. Hou, S. Gao, Chem. Soc. Rev. 2014, 43, 8098-8113.

3 D. J. Sellmyer, Nature 2002, 420, 374-375

4 W. B. Cui, Y. K. Takahashi, K. Hono, Adv. Mater. 2012, 24, 6530-6535.

5 M. Yue, X. Zhang, J. P. Liu, Nanoscale 2017, 9, 3674-3697.

6 S. Sun, C. B. Murray, D. Weller, L. Folks, A. Moser, Science 2000, 287, 1989-1992.

7 Z. Zhang, X. Song, Y. Qiao, W Xu, J. Zhang, M. Seyring, M. Rettenmayr, Nanoscale,

2013, 5, 2279-2284

8 S. Sawatzki, R. Heller, C. Mickel, M. Seifert, L. Schultz, V. Neu, J. Appl. Phys. 2011,

109, 123922.

9 C. H. Chen, S. J. Knutson, Y. Shen, R. A. Wheeler, J. C. Horwath, P. N. Barnes, Appl.

Phys. Lett. 2011, 99, 012504.

10 M. Seyring, X. Song, Z. Zhang, M. Rettenmayra, Nanoscale, 2015, 7, 12126-12132

11 W. L. Zuo, X. Zhao, J. F. Xiong, M. Zhang, T. Y. Zhao, F. X. Hu, J. R. Sun, B. G.

Shen, Sci. Rep. 2015, 5, 13117

12 C. B. Rong, V. V. Nguyen, J. P. Liu, J. Appl. Phys. 2010, 107, 09A717.

13 L. Zheng, B. Cui, N. G. Akdogan, W. Li, G. C. Hadjipanayis, J. Alloys. Compd.

2010, 504, 391-394.

14 A. M. Gabay, X. C. Hu, G. C. Hadjipanayis, J. Magn. Magn. Mater. 2014, 368, 75-

81.

15 M. Yue, J. H. Zuo, W. Q. Liu, W. C. Lv, D. T. Zhang, J. X. Zhang, Z. H. Guo, W.

72

Li, J. Appl. Phys. 2011, 109, 07A711.

16 A. R. Yan, W. Y. Zhang, H. W. Zhang, B. G. Shen, J. Magn. Magn. Mater. 2000,

210, L10-L14

17 Y. Hou, Z. Xu, S. Peng, C. B. Rong, J. P. Liu, S. Sun, Adv. Mater. 2007, 19, 3349-

3352.

18 B. Shen, A. Mendoza-Garcia, S. E. Baker, S. K. McCall, C. Yu, L. Wu, S. Sun, Nano

Lett. 2017, 17, 5695-5698.

19 Z. Ma, T. Zhang, C. Jiang, Chem. Eng. J. 2015, 264, 610-616.

20 C. Yang, L. Jia, S. Wang, C. Gao, D. Shi, Y. Hou, S. Gao, Sci. Rep. 2013, 3, 3542.

21 G. S. Chaubey, N. Poudyal, Y. Liu, C. B. Rong, J. P. Liu, J. Alloys. Compd. 2011,

509, 2132-2136.

22 L. Wu, Q. Li, C. H. Wu, H. Zhu, A. Mendoza-Garcia, B. Shen, J. Guo, S. Sun, J.

Am. Chem. Soc. 2015, 137, 7071-7074.

23 H. W. Zhang, S. Peng, C. B. Rong, J. P. Liu, Y. Zhang, M. J. Kramer, S. H. Sun, J.

Mater. Chem. 2011, 21, 16873-16876.

24 Z. Ma, T. Zhang, C. Jiang, RSC Adv. 2015, 5, 89128-89132.

25 D. W. Hu, M. Yue, J. H. Zuo, R. Pan, D. T. Zhang, W. Q. Liu, J. X. Zhang, Z. H.

Guo, W. Li, J. Alloys Compd. 2012, 538, 173-176.

26 Z. Ma, S. Yang, T. Zhang, C. Jiang, Chem. Eng. J. 2016, 304, 993-999.

27 F. Wang, H. Wei, L. Liu, H. Yang, J. Zhang, J. Du, W. Xia, A. Yan, J. P. Liu, J.

Appl. Phys. 2015, 117, 17D142.

28 X. Li, L. Lou, W. Song, Q. Zhang, G. Huang, Y. Hua, H. Zhang, J. Xiao, B. Wen,

73

X. Zhang, Nano Lett. 2017, 17, 2985–2993.

29 B. Balasubramanian, B. Das, R. Skomski, W. Y. Zhang, D. J. Sellmyer, Adv. Mater.

2013, 25, 6090-6093.

30 L. Liu, S. Zhang, J. Zhang, J. P. Liu, W. Xia, J. Du, A. Yan, J. Yi, W. Li, Z. Guo, J.

Magn. Magn. Mater. 2015, 374, 108-115.

74

Chapter 5

A General Method to Synthesize Anisotropic Sm-based

Nanomagnets with Ultra-large Coercivity

75

5.1 Introduction

Synthesis of magnetically “hard” nanoparticles (NPs) with room temperature

coercivity larger than 1 T is an essential step to developing ultra-strong magnets and

magnetic devices for broad information storage,1 electronic,2 medical3 and green

energy4 applications. Past studies have demonstrated the possibility of preparing

monodisperse magnetic NPs with large coercivity, but these NPs are mostly Pt-alloy

based5-7 and have very limited scale-up application potentials. Rare-earth metal (REM)

based alloys, such as NdFeB, SmCo, and SmFeN alloys, have magnetic characteristics

that are similar or even superior to the Pt-alloy systems and are the materials of choice

in making hard magnetic NPs with enhanced magnetic performance.8-14 However, to

prepare these REM alloy NPs has been extremely challenging due to the high negative

reductive potentials of REM cations, high REM atom reactivity, and the difficulty in

controlling the reduction chemistry between a REM salt and a transition metal (Co or

Fe) one to form a uniform alloy structure.15-21 Here we developed a general chemical

approach leading to the formation of SmCo5 NPs (50-200 nm) (85% yield) that are

dispersible in ethanol and mixable with polyethylene glycol (PEG). The NPs can be

compacted into a solid pellet or embedded in the PEG matrix with their magnetic easy

axis aligned and room temperature coercivities reaching up to 5 T, the hardest magnetic

NPs ever reported thus far.22-27 Our synthesis can be further extended to prepare other

REM alloy NPs, such as Sm2Co17 (100 nm) and Sm2Fe17N3 (100 nm) NPs, all with

room temperature coercivities larger than 1.3 T. It overcomes the known problems

observed from previous synthetic approaches on low yield, wide NP size distribution,

76

and NP sintering, and provides a general approach to hard magnetic REM alloy NPs for

various magnetic applications.

5.2 Experimental Details:

Chemicals: The syntheses were all carried out in argon atmosphere. All the

following materials are commercially available. samarium(III) acetylacetonate hydrate

(Sm(acac)3, 99%), cobalt(III) acetylacetonate (Co(acac)3, 97%), iron(III)

acetylacetonate (Fe(acac)3, 97%), calcium acetylacetonate hydrate (Ca(acac)2, 99%),

melamine (99%), 1-octadecene (ODE, 90%), oleylamine (OAm, 70%), oleic acid (OAc,

90%), and polyethylene glycol (PEG, m.w.= 3350) were purchased from Aldrich. Ca

granules (99%) was purchased from Strem Chemicals. Hexadecyltrimethylammonium

hydroxide (HTMA-OH, 25% in methanol) was purchased from TGI America.

Synthesis of SmCo-O NPs: Here we take 110 nm SmCo-O NPs synthesis process

as an example. Typically, a mixture of Co(acac)3 (0.36 g), Sm(acac)3 (0.1 g) and OAm

(20 ml) in a four-neck flask was heated to 120 oC under Ar flow for 1 h. Then it was

further heated up to 230 °C with a rate of 10 °C/min. and kept at this temperature for

another 3 h. The mixture was then cooled down to room temperature and green NPs

was collected by the addition of 40 ml ethanol and subsequent centrifugation (8500 rpm,

8 min) for 3 times. The obtained NPs were redispersed in hexane. If we decreased the

Co(acac)3 to 0.18 g and Sm(acac)3 to 0.05 g, 60 nm SmCo-O NPs can be synthesized.

If the amount of Co(acac)3, Sm(acac)3 doubled at initial stage. 220 nm SmCo-O NPs

77

can be obtained.

Synthesis of CaO-coated SmCo-O NPs: In a flask, 0.05 g SmCo-O NPs, 0.47g

Ca(acac)2, 1ml OAc,1ml OAm and 20 mL ODE were mixed under magnetic stirring

and heated to 110 °C for 1 h. Then 6 mL of HTMA-OH was added into the solution

dropwise. After 0.5 h, the mixture was heated to 200 °C and kept at 200 °C for 2 h.

Then the mixture was cooled down to room temperature and NPs were collected by the

addition of 45 ml ethanol and subsequent centrifugation (8500 rpm, 8 min) for 2 times.

Synthesis of SmCo5 NPs: Here we take 100 nm SmCo5 NPs synthesis process as

an example. 0.2 g CaO-coated 100 nm SmCo-O NPs was firstly heated in air at 185°C

for 5h to remove surfactant. Then the product was ground with 0.45g Ca together and

transferred into a stainless-steel boat. The mixture was annealed in argon at 850 oC for

30 min. After being cooled down to room temperature, the sample was washed with

water in argon atmosphere and magnetic NPs can be collected by a magnet bar.

SmCo5 NPs alignment and compaction: 50 mg SmCo5 NPs in 10 ml ethanol

dispersion were mixed with 250 mg PEG together. After the ethanol fully evaporation

in N2, the SmCo5-polymer complex was placed in standard PPMS powder sample

holder. The complex was heated to 370 K for 100 seconds in 90 kOe field to fully melt

PEG, and then cooled to 270 K for 60 seconds to solidify PEG and fixed the particle

alignment. To compact the SmCo5 NPs, the PEG in mixture was washed away by

78

ethanol and remnant SmCo5 NPs (90 mg) were compacted in a piston cylinder apparatus

at 1 GPa.

Synthesis of Sm2Co17 NPs: Firstly, the 120 nm SmCo8-O NPs were prepared. A

mixture of 0.36 g Co(acac)3, 0.056 g Sm(acac)3 and 20 ml OAm was heated to 120 oC

under Ar flow for 1 h. Then it was further heated up to 230 °C and kept at this

temperature for another 3 h. After it cooled down, the green NPs can be obtained by 40

ml ethanol washing and centrifugation (8500 rpm, 8 min) for 3 times. The CaO coating

and reduction annealing process are the same as the method to 100 nm SmCo5 synthesis

and the 100 nm Sm2Co17 NPs can be obtained.

Synthesis of Sm2Fe17N3 NPs: The 110 nm SmFe-O NPs were firstly prepared. A

mixture of 0.36 g Fe(acac)3, 0.056 g Sm(acac)3, 20 ml OAm was heated to 120 oC under

Ar flow for 1 h. After the degas process, the mixture was heated up to 230 °C, then 20

ml OAc was injected and aging for 1h, following by a temperature rise until 300 oC.

After heating for 3 h, the system was cooled down and the brownish NPs can be

precipitated by 40 ml ethanol washing and centrifugation (8500 rpm, 8 min) for 3 times.

The CaO coating and reduction annealing process are the same as the method to SmCo5

synthesis. Then, 50 mg Sm2Fe17 NPs and 2 g melamine were annealed together at 600

oC in Ar for 6 h and Sm2Fe17 was nitridated to Sm2Fe17N3.

Characterization: Transmission electron microscopy (TEM) images were

79

obtained by a Philips CM 20 operating at 200kV. High resolution transmission electron

microscopy (HRTEM) images were obtained by a JEOL 2100F. Scanning electron

microscopy (SEM) images were obtained on a LEO 1530 microscope at an accelerating

voltage of 10 kV. Elemental mapping analysis work was done on a Hitachi HD2700C.

Powder XRD patterns of the samples were recorded on a Bruker AXS D8-Advanced

diffractometer with CuKa radiation (λ= 1.5418 Å). The Sm/Co and Sm/Fe composition

were determined by elemental analysis using a JY2000 Ultrace inductively coupled

plasma atomic emission spectrometer. Magnetic properties were measured using a

Physical Property Measurement System (PPMS) under a maximum applied field of 90

kOe.

5.3 Result and Discussion

5.3.1 Synthesis of SmCo5 NPs with size control

The key to the success of the general approach to dispersible REM alloy NPs is a

new way of preparing REM oxide NPs with tighter CaO coating for NP stabilization

during the Ca-initiated reductive annealing process. Oleylamine is a weak organic base

and can decompose metal acetylacetonate (acac) to metal oxide at a relatively low

temperature (200-250oC). At this temperature, Co(acac)2 was decomposed to CoO

without further reduction of CoO to metallic Co, and Sm(acac)3 was decomposed to

Sm2O3. We combined these two decomposition chemistry and reacted Sm(acac)3 and

Co(acac)2 in oleylamine with a fixed Sm:Co = 1:4.7 at 230 oC to form rod-like SmCoO

NPs, growing together to form nanoflowers. The dimensions of the SmCoO NRs were

80

controlled by Co(acac)3 concentration (the molar ratio of Co(acac)3 to OAm) – 0.25

mM gave 60 ± 10 nm SmCoO nanoflowers (Figure 5-1a); 0.5 mM produced 110 ± 20

nm SmCoO nanoflowers (Figure 5-1b); and 1 mM yielded 220 ± 40 nm nanoflowers

(Figure 5-1c). X-ray diffraction (XRD) analysis (Figure 5-1d) shows that the SmCoO

NRs made with different dimensions have a pure crystalline hexagonal structure of CoO

without showing obvious Sm2O3 diffraction peaks. To confirm that Sm-O is present in

the oxide NR structure, we analyzed the 100 nm SmCoO NRs by high-angle annular

dark-field scanning TEM (HAADF-STEM) and elemental mapping (Figure 5-1e).

From the images, we can see the Sm (red), Co (blue) and O (green) elements distribute

evenly across the NR structure. Inductively coupled plasma-atomic emission

spectroscopy (ICP-AES) analysis further confirms that in the SmCoO NRs, the ratio of

Sm/Co is at 1:4.5. These studies indicate that Sm2O3 is indeed present in the SmCoO

NR structure, but it is in an amorphous state, forming a mixture, not an oxide alloy

structure, with the crystalline CoO.

81

Figure 5-1. TEM images of as-synthesized (a) 60 nm (b) 110 nm (c) 220 nm SmCoO

flower-liked NPs. (d) XRD patterns of SmCoO NPs with different sizes and standard

CoO pattern (JPCDS No. 80-0075). (e) HADDF-STEM image and elemental mapping

of Sm (red), Co (blue) and O (green).

To prevent particles aggregation in the high temperature annealing process, the as-

prepared SmCoO NPs must be protected with an inorganic coating. In the experiment,

a mixture of 50 mg SmCoO NPs and 470 mg Ca(acac)2 was dispersed in 1-octadecene

solution. After the solution was heated to 200 C, 6 ml HTMA-OH was injected and

the solution was kept at 200 C for 1 h to decompose Ca(acac)2 to CaO.28 Figure 5-2

shows a representative TEM image of the SmCoO NPs embedded in the CaO matrix.

Different from the previously reported, we improved the CaO coating followed by an

air annealing step of SmCoO/CaO composite at 185 oC for 5 h. The air annealing can

82

remove the organic surfactant on the surface of particles to avoid CoCx formation and

make the matrix more robust. After surfactant removal, the size and morphology of

SmCoO NPs kept unchanged. Then the SmCoO/CaO was mixed with metallic Ca and

transferred to a stainless-steel boat for reduction annealing. In the annealing process,

we monitored the particles morphology change at different stages by TEM. Here we

take 110 nm SmCoO NPs as an example. After 15 min annealing at 850 oC, the branches

of nano-flower broke and started to diffuse into the central node to form around 100 nm

particles (Figure 5-3). After 30 min annealing, 100 nm polyhedral SmCo5 NPs formed

without obvious branches. Figure. 4-4 a-c show the TEM images of 50 ± 10 nm, 100 ±

20 nm and 200 ± 30 nm SmCo5 NPs, respectively. The HRTEM image of a 100 nm

SmCo5 NP with a 5 nm oxidation shell is shown in Figure 5-4d. The distance of lattice

is 2.16 A, which matches perfect lattice fringes of (200) planes. HAADF-STEM image

and elemental mapping of a typical 100 nm particle show Sm and Co distribute evenly

in a particle (Figure 5-4e). The XRD patterns of 50, 100 and 200 nm as-prepared

SmCo5 NPs match well to standard hexagonal D2d SmCo5 structure, and the intensity

of peaks fits with the standard pattern, indicating that all the SmCo5 NPs are randomly

oriented (Figure 5-4f). The Sm/Co ratio was 1/5 as analyzed by ICP-AES. Magnetic

properties of the SmCo5 NPs were characterized by a physical property measurement

system (PPMS) with fields up to 90 kOe at room temperature. The hysteresis loops

show the as-synthesized SmCo5 NPs are strong ferromagnetic at room temperature

(Figure 5-4 g-i). These NPs exhibited clear size-dependent magnetic properties. For the

50 nm SmCo5 NPs, the coercivity is 19.2 kOe. As the size increased to 100 nm, the

83

coercivity increases to 30.8 kOe correspondingly. And the 200 nm SmCo5 NPs shows

a coercivity of 36.1 kOe. The values are among the largest coercivity of SmCo5 reported.

Besides, the saturation moment (Ms) of the 50, 100 and 200 nm SmCo5 NPs is 58.3,

64.5, 69.2 emu/g, respectively. The values are smaller than the perfect bulk SmCo5 (99

emu/g) due to nanoscale effect and inevitable surface oxidation. The trend of Ms

increase indicates the ratio of oxidation layer to the particle become smaller with

increasing size of SmCo5 NPs.

Figure 5-2. TEM image of 100 nm SmCoO NPs in CaO matrix coating.

Figure 5-3. TEM image of 100 nm SmCoO NPs after 15 min annealing at 850 °C.

84

Figure 5-4. TEM images of annealed (a) 50 nm (b) 100 nm (c) 200 nm polyhedral

SmCo5 NPs. (d) HRTEM of a part of a 100 nm SmCo5 particle. (e) HADDF-STEM

image of a 100 nm SmCo5 particle and elemental mapping of Sm (red) and Co (blue),

showing uniform elemental distribution. (f) XRD patterns of SmCo5 NPs and standard

SmCo5 pattern (JPCDS No. 65-8981). Non-aligned hysteresis loops of (g) 50 nm (h)

100 nm and (i) 200 nm SmCo5 NPs at 300 K.

5.3.2 Alignment of SmCo5 in Polymer Matrix and Compaction of SmCo5 to Pellet

The SmCo5 NPs are polyhedral with a short axis to long axis ratio of 0.85 on

average. The intrinsic hexagonal crystalline structure and shape anisotropy can lead to

anisotropic magnetism. To explore their anisotropic magnetic behavior, SmCo5 NPs

were aligned with the help of polyethylene glycol (PEG) polymer fixing in 90 kOe field.

85

After alignment, their hysteresis loops show decent squareness and improved magnetic

properties (Figure 5-5a). For example, the hysteresis loop of 200 nm SmCo5 shows a

high remanence ratio (Mr/Ms) of 0.92. Compared with the randomly oriented NPs, the

Hc increase to 49.2 kOe (24% increase), Mr increases to 75.5 emu/g (50% enhancement)

and Ms increases to 81.8 emu/g (18% enhancement). Figure 5-6 shows a B-H hysteresis

loop of aligned 200 nm SmCo5 with a calculated (BH)max of 16.8 MGOe. Both

coercivity and (BH)max are the highest values of reported SmCo5 obtained by chemical

methods (Table 4-1). The anisotropic behavior also be observed in the aligned 50 nm

and 100 nm SmCo5 NPs. The aligned 50 nm SmCo5 NPs show a coercivity of 25.3 kOe

and aligned 100 nm SmCo5 NPs show a coercivity of 44.5 kOe. To achieve fabrication

of SmCo5 magnetic device, the NPs should be compacted to a dense bulk material

without polymer. The aligned 200 nm SmCo5 NPs was rinsed by ethanol to remove the

PEG and compressed to a 3 mm x 2 mm bulk squat cylinder under a pressure of 1 GPa

(Figure 5-5b). After compaction, no obvious aggregation occurs to the SmCo5 NPs

according to scanning electron microscope (SEM) characterization (Figure 5-5c). The

hysteresis loops of the compressed SmCo5 were measured in a 90 kOe field. We can

see the anisotropic magnetic behavior was kept after compaction with a (BH)max of 14.2

MGOe (Figure 5-5d), which is still about 35% enhancement compared to the sintered

or warm-compacted SmCo5 bulk magnets reported (10.2 MGOe).29-30

86

Figure 5-5. (a) Hysteresis loops of 50 nm, 100 nm and 200 nm SmCo5 NPs after

external field alignment with PEG at 300 K. (b) A picture of compacted SmCo5

nanomagnet. (c) SEM of the SmCo5 nanomagnet after compaction. (d) Hysteresis loops

of compacted 200 nm SmCo5 nanomagnet at 300 K.

Figure 5-6. B-H hysteresis loops of aligned 200 nm SmCo5 NPs at 300 K.

87

Table 4-1. A list of SmCo5 made by chemical method. The theoretical calculated

(BH)max for perfect SmCo5 is 28.6 MGOe.

Ms (emu/g) Hc (kOe) (BH)max (MGOe) Reference

70 12 6 11

52 17.7 2.75 13

82 20.7 10 18

78 19.3 14.4 22

83 25.7 15.8 23

81.8 49.2 16.8 This work

5.3.3 Synthesis of Sm2Co17 NPs and Sm2Fe17N3 NPs

The method of SmCo5 NPs synthesis can be extended to Sm2Co17 NPs by adjusting

the initial ratio of Sm/Co in SmCoO precursor. With a lower molar ratio of

Sm(acac)3/Co(acac)3 (1/8) in the reaction, the 120 nm SmCo8.5O NPs can be

synthesized (Figure 5-7a). With the same CaO coating process and annealing, the 120

± 20 nm Sm2Co17 NPs were obtained (Figure 5-7b). The XRD shows its structure

matches well to hexagonal Sm2Co17 phase (Figure 5-7c). We should emphasis Sm2Co17

can display two kinds of structure, rhombohedral and hexagonal lattice. Here we only

obtained pure hexagonal Sm2Co17 phase. The hysteresis loop, as shown in Figure 5-7d,

shows a large Hc of 21.2 kOe. After alignment, the Hc reaches to 26.3 kOe, which is

among the highest value of Sm2Co17 reported.31 Compared with 100 nm SmCo5 NPs,

100 nm Sm2Co17 NPs shows a higher Ms of 98.3 emu/g, which is typical for the SmCo

88

alloy with higher Co concentration.

Figure 5-7. (a) TEM image of 120 nm SmCo8.5O NPs. (b) TEM image of 100 nm

Sm2Co17 NPs. (c) XRD patterns of Sm2Co17 NPs and standard hexagonal Sm2Co17

pattern (JPCDS No. 03-065-7762). (d) Hysteresis loop of unaligned (black) and aligned

(red) as-synthesized Sm2Co17 NPs at 300 K.

This method, also can be applied to synthesize Sm2Fe17 NPs by reducing SmFeO

NPs. 110 nm SmFeO nanocubes can be synthesized in a similar condition by thermal

decomposition of Sm(acac)3 and Fe(acac)3 (Figure 5-8a). After an annealing, 100 nm

Sm2Fe17 NPs with a rhombohedral crystalline structure were obtained (Figure 5-8b).

The Sm2Fe17 NPs are also ferromagnetic with a small coercivity of 0.23 kOe (Figure

5-8c). Afterward, the Sm2Fe17 was annealed in the presence of melamine at 600 oC. The

89

melamine will decompose and release ammonia, which can nitride the Sm2Fe17 NPs to

100 ± 20 nm Sm2Fe17N3 NPs (Figure 5-8d). The XRD shows the product owns a

rhombohedral structure, matching well to standard Sm2Fe17N3 (Figure 5-8e). Compared

with the XRD spectrum of Sm2Fe17, the peaks show a small left shift (0.99o), indicating

an interstitial N-doped structure. We should note the 600 oC is the maximum

temperature in nitridation process, above which the Sm2Fe17N3 NPs are not stable and

decompose to SmN + Fe (Figure 5-9). After nitridation, the hysteresis loop of

Sm2Fe17N3 shows a strong ferromagnetic property with Hc to 13.1 kOe and Ms to 121.3

emu/g (Figure 5-8f), much larger than the Sm2Fe17N3 obtained from physical methods.

This is the first time to chemically synthesize nano-sized Sm2Fe17N3 particles.

Figure 5-8. (a) TEM image of a 120 nm SmFeO nanocubes (b) XRD of as-prepared

Sm2Fe17 NPs (black curve) and the standard pattern of rhombohedral structure Sm2Fe17

(red lines, JPCDS No. 01-074-7186). (c) Hysteresis loops of as-prepared 100 nm

Sm2Fe17 NPs at 300 K. (d) TEM of 100 nm Sm2Fe17N3 NPs. (e) XRD of as-prepared

Sm2Fe17N3 NPs (black curve) and the standard pattern of rhombohedral structure

90

Sm2Fe17N3 (red lines, JPCDS No. 00-048-1790). (f) Hysteresis loops of unaligned

(black) and aligned (red) nitrogenized Sm2Fe17N3 NPs at 300 K.

Figure 5-9. XRD of Sm2Fe17 NPs annealed with melamine at 650 oC for 2h (black

curve). The product matches well to standard SmN (red lines) and standard bcc-Fe (blue

lines).

5.4 Conclusion

In summary, we present a general method to synthesize Sm-based permanent

nanomagnets (SmCo5, Sm2Co17 and Sm2Fe17N3) by Ca reduction of the SmCoO NPs or

SmFeO NPs. The size of SmCoO NPs can be rationally controlled by tuning the

concentration of precursors in the reaction. With the protection of CaO matrix, the

particle aggregation in annealing process is successfully prevented and uniform sizes

(50 nm, 100 nm and 200 nm) SmCo5 NPs with ultra large coercivity can be obtained.

Among them, 200 nm SmCo5 NPs owns the largest Hc of 49 kOe. After external

magnetic alignment, the 200 nm SmCo5 NPs show an anisotropic magnetic behavior

with (BH)max reaching to 16.8 MGOe. After compaction, the anisotropic magnetic

behavior can be kept. The synthesis method is not limited to SmCo5 NPs, but also can

91

be extended to Sm2Co17 NPs by modifying Sm/Co ratio and to Sm2Fe17 NPs by

replacing Co(acac)3 with Fe(acac)3, which are further nitridated to Sm2Fe17N3 NPs. The

synthesis of Sm based permanent nanomagnets provides a viable way to make

permanent magnetic device with ultra large coercivity and energy product. Work on

controlled synthesis of exchange-coupled SmCo-M (M = Co, Fe, or FeCo)

nanocomposites is in process to realize the superior magnetic performance.

92

References:

1 N. Jones, Nature, 2011, 472, 22.

2 R. Skomski and J. M. D. Coey, Phys. Rev. B: Condens. Matter Mater. Phys., 1993,

48, 15812.

3 H. A. Davies, J. Magn. Magn. Mater., 1996, 157, 11.

4 E. F. Kneller and R. Hawig, IEEE Trans. Magn., 1991, 27, 3588.

5 J. Bauer, M. Seeger, A. Zern and H. Kronmuller, J. Appl. Phys., 1996, 80, 1667.

6 G. C. Hadjipanayis, J. Magn. Magn. Mater., 1999, 200, 373.

7 Z. C. Wang, S. Z. Zhou, M. C. Zhang and Y. Qiao, J. Appl. Phys., 2000, 88, 591.

8 T. Schrefl, H. Kronmuller and J. Fidler, J. Magn. Magn. Mater., 1993, 127, L273.

9 C. B. Rong, Y. Zhang, N. Poudyal, X. Y. Xiong, M. J. Kramer and J. P. Liu, Appl.

Phys. Lett., 2010, 96, 102513.

10 H. Zeng, J. Li, J. P. Liu, Z. L. Wang and S. H. Sun, Nature, 2002, 420, 395.

11 H. L. Li, X. H. Li, D. F. Guo, L. Lou, W. Li and X. Y. Zhang, Nano Lett., 2016, 16,

5631.

12 I. A. Alomari and D. J. Sellmyer, Phys. Rev. B: Condens. Matter Mater. Phys., 1995,

52, 3441.

13 F. Liu, J. H. Zhu, W. L. Yang, Y. H. Dong, Y. L. Hou, C. Z. Zhang, H. Yin and S.

H. Sun, Angew. Chem., Int. Ed., 2014, 53, 2176.

14 T. L. Zhang, H. Y. Liu and C. B. Jiang, Appl. Phys. Lett., 2015, 106, 162403.

15 B. Balasubramanian, R. Skomski, X. Z. Li, S. R. Valloppilly, J. E. Shield, G. C.

Hadjipanayis and D. J. Sellmyer, Nano Lett., 2011, 11, 1747.

93

16 T. L. Zhang, H. Y. Liu, Z. H. Ma and C. B. Jiang, J. Alloys Compd., 2015, 637, 253.

17 B. Shen, A. Mendoza-Garcia, S. E. Baker, S. K. McCall, C. Yu, L. H. Wu and S. H.

Sun, Nano Lett., 2017, 17, 5695.

18 X. H. Li, L. Lou, W. P. Song, Q. Zhang, G. W. Huang, Y. X. Hua, H. T. Zhang, J.

W. Xiao, B. Wen and X. Y. Zhang, Nano Lett., 2017, 17, 2985.

19 X. H. Li, L. Lou, W. P. Song, G. W. Huang, F. C. Hou, Q. Zhang, H. T. Zhang, J.

W. Xiao, B. Wen and X. Y. Zhang, Adv. Mater., 2017, 29, 1606430.

20 N. Liakakos, T. Blon, C. Achkar, V. Vilar, B. Cormary, R. P. Tan, O. Benamara, G.

Chaboussant, F. Ott, B. WarotFonrose, E. Snoeck, B. Chaudret, K. Soulantica and M.

Respaud, Nano Lett., 2014, 14, 3481.

21 M. Yue, X. Y. Zhang and J. P. Liu, Nanoscale, 2017, 9, 3674.

22 K. Gandha, K. Elkins, N. Poudyal, X. B. Liu and J. P. Liu, Sci. Rep., 2014, 4, 5345.

23 L. Y. Zheng, B. Z. Cui and G. C. Hadjipanayis, Acta Mater., 2011, 59, 6685.

24 W. L. Zuo, X. Zhao, T. Y. Zhao, F. X. Hu, J. R. Sun and B. G. Shen, Sci. Rep., 2016,

6, 25805

25 Y. P. Wang, Y. Li, C. B. Rong and J. P. Liu, Nanotechnology, 2007, 18, 4.

26 H. W. Gu, B. Xu, J. C. Rao, R. K. Zheng, X. X. Zhang, K. K. Fung and C. Y. C.

Wong, J. Appl. Phys., 2003, 97, 7589.

27 K. Ono, Y. Kakefuda, R. Okuda, Y. Ishii, S. Kamimura, A. Kitamura and M. Oshima,

J. Appl. Phys., 2002, 91, 8480.

28 Y. L. Hou, Z. C. Xu, S. Peng, C. B. Rong, J. P. Liu and S. H. Sun, Adv. Mater., 2007,

19, 3349.

94

29 C. N. Chinnasamy, J. Y. Huang, L. H. Lewis, B. Latha, C. Vittoria and V. G. Harris,

Appl. Phys. Lett., 2008, 93, 3.

30 C. N. Chinnasamy, J. Y. Huang, L. H. Lewis, C. Vittoria and V. G. Harris, Appl.

Phys. Lett., 2010, 97, 22–23.

31 V. Pop, O. Isnard, I. Chicinas, D. Givord and J. M. Le Breton, J. Optoelectron. Adv.

Mater., 2006, 8, 494–500.

95

Chapter 6

Synthesis and self-assembly of non-rare earth permanent

nanomagnets

Reprinted with permission from Nanoscale. 2015, 7, 16165. Copyright © The Royal

Society of Chemistry 2015.

96

6.1 Introduction

Iron oxide based nanoparticles (NPs) with adjustable size and magnetic properties

have attracted tremendous research and development interests because of the important

application potential for future biochemical research,1–8 high performance permanent

magnets,9–11 and high density magnetic tape recording.12–13 Generally, two common

types of magnetic iron oxides are well developed: the cubic structured spinel-type

ferrites with a formula of MFe2O4 (M = Mn, Fe, Co, Ni, etc.) and hexagonal barium

ferrite, or BaFe, with a general formula BaFe12O19. The spinel ferrites show weak

ferrimagnetic property and its magnetically isotropic. In contrast, the hexagonal BaFe

shows a great magnetic anisotropy along their magnetic easy axis (crystallographic c-

axis), the value of which reaches to 5 × 105 J m−3.14,15 This magnetic character makes

BaFe used as permanent magnetic materials with great chemical stability. Recently, the

hard magnetic BaFe were prepared in nanostructured plates by physical method like

ball milling and tested as a novel medium for magnetic tape recording.16–20 To maximize

the magnetic recording density and minimization of device volume in tape recording

media and the magnetic energy storage ability in ferrite magnets, monodispersed BaFe

NPs with adjustable magnetic properties need to be obtained and assembled in either

two dimensional (2D) films (for magnetic recording) or densely packed 3D device (for

permanent magnets).

BaFe is conventionally prepared by physical methods or synthesized by solid-state

reactions of BaCO3 and Fe(OH)3 at above 1000 °C. This is because forming hexagonal

structures from the normal cubic oxide precursors needs more energy.21–23 Previous

97

research also include the direct solid reaction between Ba and Fe hydroxides24,25 or

oxides,26–28 hydrothermal reaction, solvothermal reaction,29–31 and organic phase

preparation of Fe3O4/BaCO3 core/shell NPs followed by a high temperature annealing

in O2.32 However, these solid state reactions and high temperature annealing often result

in incomplete alloy formation between the Ba- and Fe-oxides and multiple Ba-Fe-O

phases existence. The high temperature annealing also causes uncontrolled aggregation

of BaFe, making it extremely difficult to control the BaFe size and magnetic properties.

To develop a better approach to BaFe NPs and their assemblies, we tested the organic-

phase decomposition of both Ba- and Fe-precursors together at temperatures above

200 °C. We observed our Ba-doped iron oxide NPs, referred as Ba–Fe–O NPs, could

be synthesized by thermal decomposition of Fe(acac)3 (acac = acetylacetonate) and

Ba(stearate)2 at 320 °C in 1-octadecene with oleic acid and oleylamine as surfactants.

The composition of Ba to Fe can be precisely controlled by the precursors. The as-

synthesized Ba–Fe–O NPs were well-dispersed in hexane and could be transfer to a

substrate and further assembled into 2D arrays. After thermal annealing, these ba doped

iron oxide NPs were converted to hexagonal crystalline BaFe NPs, showing much

enhanced magnetic properties. Here we highlight this new synthesis and self-assembly

of Ba–Fe–O NPs to hexagonal BaFe magnetic arrays. Moreover, the synthesis could

provide a method to prepare other doped iron oxide NPs, such as strontium-doped iron

oxide (Sr–Fe–O) NPs and hexagonal SrFe NPs.

6.2 Experimental Details

98

Chemicals: Iron(III) acetylacetonate (99%), barium stearate (technical grade) were

purchased from Strem Chemicals. Strontium stearate (98.5%) was purchased from

VWR International. Oleic acid (90%), oleylamine (70%) and 1-octadecene (90%) were

all purchased from Sigma-Aldrich. All chemicals were used as received without further

purification.

Synthesis of Ba-Fe-O NPs: In a typical synthesis of Ba0.082-Fe-O NPs, barium

stearate (60 mg, 0.085 mmol), iron(III) acetylacetonate (300 mg, 0.85 mmol), oleic acid

(1 mL), oleylamine (8 mL) and 1-octadecene (3 mL) were mixed and magnetically

stirred at room temperature under a gentle flow of Ar gas for 20 min. Then the mixture

was heated directly to 320 oC at 10 oC/min. The reaction was kept at this temperature

for 1.5 h. Then the mixture was cooled to room temperature by removing the heating

mantle. The NPs were precipitated by 2-propanol (30 mL) and collected by

centrifugation (8500 rpm, 8 min). The product was re-dispersed in hexane and separated

again by adding ethanol followed by centrifugation (8500 rpm, 8 min). The final

product was dispersed in hexane for further characterization.

Assembly of Ba-Fe-O NPs: Monolayer assembly of the Ba0.082-Fe-O NPs was

prepared using the water-air interfacial self-assembly approach reported previously.

Briefly, the NPs were dispersed in the mixture of hexane and toluene (v:v = 1:1) at the

concentration of 0.5 mg/mL. 160 L of the dispersion was drop-cast on the water

surface in the Teflon column (diameter: 3.8 cm). Upon complete evaporation of the

organic solvent, the formed monolayer assembly floating on the water surface was

transferred to TEM Cu grids or Si substrates for further characterization. Multilayer

99

assembly of the NPs was prepared by drop-casting 16 L of the dispersion on a Si

substrate (0.7 cm × 0.7 cm).

Characterization: Transmission electron microscopy (TEM) and high-resolution

TEM (HRTEM) images were collected using a Philips CM20 and JEOL 2010 with an

accelerating voltage of 200 kV, respectively. Scanning electron microscopy (SEM)

images of the assemblies were acquired on a LEO 1530 microscope at an operating

voltage of 10 kV. Energy dispersive X-ray (EDX) spectrum was obtained by Oxford

energy-disperse X-ray spectroscopy equipped in the SEM at an operating voltage of 20

kV. X-ray diffraction (XRD) patterns of the samples were collected on a Bruker AXS

D8-Advanced diffractometer with Cu Kα radiation (λ = 1.5418 Å). The Ba/Fe

composition was determined by elemental analysis using a JY2000 Ultrace ICP Atomic

Emission Spectrometer. Magnetic properties were measured on a Lakeshore 7404 high

sensitivity vibrating sample magnetometer (VSM) with fields up to 14.5 kOe at room

temperature (~298K).

6.3 Result and Discussion

6.3.1 Synthesis of Ba doped Iron Oxide with Composition Control

Under the synthetic conditions with the amount of Fe(acac)3 (300 mg) and oleic

acid (1 mL) fixed, the composition of Ba in Ba–Fe–O NPs was controlled by the

amount/concentration of Ba(stearate)2 or oleylamine (Table 1). For example, adding 20

mg of Ba(stearate)2 to the reaction mixture produced Ba–Fe–O NPs with a Ba/Fe atomic

ratio of 0.04, denoted as Ba0.04–Fe–O. 40 mg (or 60 mg) of Ba(stearate)2 gave Ba0.055–

100

Fe–O (or Ba0.065–Fe–O) NPs. In the synthesis, we also noticed that oleylamine played

two roles in NP stabilization and in promoting metal precursor decomposition.33,34

Reacting 300 mg of Fe(acac)3 with 60 mg of Ba-stearate in 8 mL oleylamine and 7 mL

1-octadecene yielded Ba0.075–Fe–O NPs. By reducing the volume of 1-octadecene to 3

mL, Ba0.082–Fe–O NPs were obtained. These NPs have the Ba composition close to the

ideal Ba/Fe ratio of 0.083 in the pure BaFe phase. If only oleylamine was used as the

solvent, then Ba0.095–Fe–O NPs were synthesized. The Ba/Fe ratio of 0.095 is very close

to the initial precursor ratio of Ba(stearate)2/Fe(acac)3 (0.099) used in the reaction,

indicating almost complete precursor decomposition and metal ratio carry-over to the

final Ba–Fe–O product.

Figure 6-1 shows a typical transmission electron microscopy (TEM) image of

the 15 ± 0.5 nm Ba0.04–Fe–O NPs. The high resolution TEM image of a representative

NP is shown in Figure 6-1B. The distance of the lattice fringe was measured to be ∼2.6

Å, corresponding to the lattice spacing of (311) planes in the spinel Fe3O4. Crystal

defects are also visible in the NP as marked by the dashed circular lines, which can be

ascribed to the lattice mismatch caused by the Ba doping. Figure 1C shows the TEM

image of the monodisperse Ba0.082–Fe–O NPs of 13 ± 0.5 nm. The Ba–Fe–O NPs were

further characterized by energy dispersive X-ray (EDX) spectroscopy, confirming the

existence of Ba and Fe at the atomic ratio of 0.093, which is close to that obtained from

the ICP-AES analysis.

101

Figure 6-1. (A) TEM image of the as-synthesized Ba0.04–Fe–O NPs. (B) HR-TEM

image of a representative Ba0.04–Fe–O NP. (C) TEM image of the as-synthesized

Ba0.082–Fe–O NPs.

The X-ray diffraction (XRD) pattern of the as-synthesized Ba0.04–Fe–O NPs is

shown in Figure 6-2A. The pattern matches with the spinel Fe3O4 structure, but the

broad diffraction peaks infer the presence of small crystalline domains, which supports

what is observed from Figure 6-1B. The magnetic hysteresis loop of the as-synthesized

Ba0.04–Fe–O NPs (Figure 6-2B) indicates that these NPs are magnetically soft. Due to

the existence of Ba and the induced crystal defects, the saturation moment (Ms) of the

NPs is relatively small (31 emu g−1) compared to the pure single crystalline Fe3O4 NPs

at a similar size (∼65 emu g−1).33 In order to convert the as-synthesized Ba–Fe–O NPs

into BaFe NPs, we first tested different annealing conditions to ensure that the Fe3O4

structure could be oxidized to Fe2O3,35 followed by the formation of the BaFe phase via

102

the diffusion of Ba2+ into the Fe2O3 lattice.36 In O2 at 600 °C for 1 h, the Ba0.04–Fe–O

NPs show no obvious structure change and the diffraction peaks become sharper,

indicating that the annealing enlarges the crystal domain within each NP, which is

further supported by their soft magnetic properties with higher Ms than the as-

synthesized NPs (Figure 6-2B). Ba0.082–Fe–O NPs behave similarly once annealed the

same way. When annealed at 700 °C in O2 for 1 h, the as-synthesized Ba0.04–Fe–O NPs

are converted into hexagonal BaFe, which is ferromagnetic with the coercivity (Hc) of

2800 Oe and Ms of 40 emu g−1 (Figure 6-2B). Both XRD and magnetic data (the loop

shows a two-phase behavior) indicate that in the annealed Ba0.04–Fe–O NPs, the Fe2O3

phase co-exists with the BaFe phase due to the non-stoichiometric composition of Ba

in the NP structure. The increase of Ba composition from 0.055 to 0.065 reduces the

amount of Fe2O3 present in the annealed Ba–Fe–O NPs (Figure 6-2C). When the Ba

composition is above 0.075, no obvious diffraction peaks of Fe2O3 are observed for the

annealed Ba–Fe–O NPs. The diffraction peaks of the annealed Ba0.082–Fe–O NPs match

well with those of the hexagonal BaFe, indicating the formation of a pure BaFe phase.

It is worth noting that our annealing is performed at a lower temperature and shorter

time than previous syntheses,24–29 therefore the NP morphology is better preserved.

Table 6-1. Experimental conditions for synthesizing Ba–Fe–O NPs with different Ba

compositions.

103

Figure 6-2. (A) XRD patterns and (B) room temperature hysteresis loops of the Ba0.04–

Fe–O NPs before and after O2 annealing treatment. (C) XRD patterns and (D) room

temperature hysteresis loops of the Ba–Fe–O NPs with different Ba compositions after

annealing in O2 at 700 °C for 1 h.

Ba/Fe composition dependent magnetic properties of the annealed Ba–Fe–O NPs

were studied and their hysteresis loops are shown in Figure 6-2D. As the Ba

composition increases from 0.055 to 0.082, the Hc of the annealed NPs increases from

3120 Oe to 5260 Oe. Their Ms increases as well from 42 emu g−1 to 54 emu g−1 due to

the increased BaFe phase purity, which is further confirmed by the single-phase

104

hysteresis loop from the annealed Ba0.082–Fe–O NPs. When the atomic ratio of Ba/Fe is

over the optimal value required for the formation of BaFe (0.083) at 0.095, the annealed

NPs show a decreased Hc (5150 Oe) and Ms (50 emu g− 1). The above results

demonstrate that the new synthetic method described in this paper is a facile approach

to Ba–Fe–O and further to hard magnetic BaFe NPs.

6.3.2 Self-assembly of As-synthesized BaFe

The as-synthesized Ba–Fe–O NPs are well dispersed in hexane, allowing easy self-

assembly of these NPs into well-defined NP arrays. Using the water–air interface self-

assembly method, we fabricated a monolayer assembly of the Ba0.082–Fe–O NPs.

Figure 6-3A shows a TEM image of the monolayer assembly transferred onto a carbon

coated Cu grid. The SEM image of the monolayer array transferred onto a Si substrate

is shown in Figure 6-3B. After O2 annealing at 700 °C for 1 h, the morphology of the

monolayer was well maintained, as shown in Figure 6-3C. No obvious NP

sintering/aggregation in the monolayer array was observed. However, the magnetic

signal generated from this monolayer array is too weak to be easily detected. To

increase the magnetic signal from the assembly, we prepared a multilayer array of the

Ba0.082–Fe–O NPs by drop-casting the NP dispersion (hexane, 0.5 mg mL−1) directly

on a Si substrate and by controlling the evaporation of hexane. Figure 6-3D show the

SEM images of the densely packed assemblies of the Ba0.082–Fe–O NPs. Different from

the monolayer assembly that maintains the morphology after O2 annealing treatment,

the multilayer assembly exhibits some NP aggregation/sintering after the same

105

treatment (Figure 6-3E). However, the grain size of the NPs after annealing is still

around 50 nm. Room temperature magnetic properties of the annealed multilayer

assembly were determined with the magnetic field perpendicular (out-of-plane) and

parallel (in-plane) to the assembly plane. Figure 6-3F shows the hysteresis loops of the

annealed assembly. The in-plane loop shows the Hc of 4100 Oe, which is much larger

than that of the out-of-plane one (Hc = 2050 Oe). Moreover, the in-plane loop is squarer

compared to the out-of-plane loop, suggesting that the easy axis of the magnetization

lies in the plane of the film. Such assembly may be especially useful to fabricate 3D

stacks of BaFe NPs as new nanostructured magnets for energy product optimization.

Figure 6-3. (A) TEM image of the monolayer assembly of Ba0.082–Fe–O NPs. (B) SEM

image of the monolayer assembly deposited on a Si substrate. (C) SEM image of the

106

monolayer assembly after annealing in O2 at 700 °C for 1 h. (D) SEM images of the

multilayer assembly of Ba0.082-Fe-O NPs deposited on a Si substrate by the drop-casting

method. (E) SEM image of the multilayer assembly after annealing in O2 at 700 °C for

1 h. (F) Room temperature hysteresis loops of the multilayer assembly after annealing

in O2 at 700 °C for 1 h.

6.4 Conclusion

In conclusion, we have reported a facile organic-phase synthesis of monodisperse

Ba–Fe–O NPs through thermal decomposition of Ba(stearate)2 and Fe(acac)3 in 1-

octadecene with oleic acid and oleylamine as surfactants. The Ba/Fe composition is

tuned from 0.04 to 0.095 by controlling the ratio of Ba(stearate)2/Fe(acac)3 or the

volume of oleylamine and 1-octadecene. The as-synthesized Ba–Fe–O NPs, especially

the Ba0.082–Fe–O NPs, can be easily converted into hexagonal BaFe by annealing under

an O2 atmosphere at 700 °C for 1 h, showing strong ferromagnetic properties with Hc

reaching 5260 Oe and a Ms of 54 emu g−1. More importantly, these monodisperse Ba–

Fe–O NPs are well dispersed in hexane and can be easily assembled into densely packed

2D arrays and further converted into oriented BaFe magnets. Our reported synthetic

method and self-assembly approach provides a unique way of fabricating ferromagnetic

ferrite arrays that may be important for magnetic energy storage and data storage

applications.

107

References

1 J. T. Jang, H. Nah, J. H. Lee, S. H. Moon, M. G. Kim and J. Cheon, Angew. Chem.,

Int. Ed., 2009, 48, 1234.

2 W. K. Moon, J. Kim, H. S. Kim, N. Lee, T. Kim, H. Kim, T. Yu, I. C. Song and T.

Hyeon, Angew. Chem., Int. Ed., 2008, 47, 8438.

3 K. Cheng, S. Peng, C. J. Xu and S. H. Sun, J. Am. Chem. Soc., 2009, 131, 10637.

4 T. Hyeon, J. E. Lee, N. Lee, H. Kim, J. Kim, S. H. Choi, J. H. Kim, T. Kim, I. C.

Song, S. P. Park and W. K. Moon, J. Am. Chem. Soc., 2010, 132, 552.

5 J. H. Lee, Y. M. Huh, Y. Jun, J. Seo, J. Jang, H. T. Song, S. Kim, E. J. Cho, H. G.

Yoon, J. S. Suh and J. Cheon, Nat. Med., 2007, 13, 95.

6 J. H. Lee, J. T. Jang, J. S. Choi, S. H. Moon, S. H. Noh, J. W. Kim, J. G. Kim, I. S.

Kim, K. I. Park and J. Cheon, Nat. Nanotechnol., 2011, 6, 418.

7 C. Xu, J. Xie, D. Ho, C. Wang, N. Kohler, E. G. Walsh, J. R. Morgan, Y. E. Chin and

S. Sun, Angew. Chem., Int. Ed., 2008, 47, 173.

8 C. J. Xu, B. D. Wang and S. H. Sun, J. Am. Chem. Soc., 2009, 131, 4216.

9 N. Poudyal and J. P. Liu, J. Phys. D: Appl. Phys., 2013, 46.

10 J. Park, Y. K. Hong, S. G. Kim, S. Kim, L. S. I. Liyanage, J. Lee, W. Lee, G. S. Abo,

K. H. Hur and S. Y. An, J. Magn. Magn. Mater., 2014, 355, 1.

11 R. C. Pullar, Prog. Mater. Sci., 2012, 57, 1191.

12 Q. Dai, D. Berman, K. Virwani, J. Frommer, P. O. Jubert, M. Lam, T. Topuria, W.

Imaino and A. Nelson, Nano Lett., 2010, 10, 3216.

13 L. Wu, P. O. Jubert, D. Berman, W. Imaino, A. Nelson, H. Zhu, S. Zhang and S.

108

Sun, Nano Lett., 2014, 14, 3395.

14 Z. V. Golubenko, L. P. Ol’khovik, Y. A. Popkov, Z. I. Sizova and A. S. Kamzin,

Phys. Solid State, 1998, 40, 1718.

15 J. Wang, F. Zhao, W. Wu and G. M. Zhao, J. Appl. Phys., 2011, 110, 096107.

16 D. Berman, R. Biskeborn, N. Bui, E. Childers, R. D. Cideciyan, W. Dyer, E.

Eleftheriou, D. Hellman, R. Hutchins, W. Imaino, G. Jaquette, J. Jelitto, P. O. Jubert,

C. Lo, G. McClelland, S. Narayan, S. Oelcer, T. Topuria, T. Harasawa, A. Hashimoto,

T. Nagata, H. Ohtsu and S. Saito, IEEE Trans. Magn., 2007, 43, 3502.

17 T. Harasawa, R. Suzuki, O. Shimizu, S. Olcer and E. Eleftheriou, IEEE Trans. Magn.,

2010, 46, 1894.

18 P. O. Jubert, B. Biskeborn, D. N. Qiu, A. Matsumoto, H. Noguchi and O. Shimizu,

IEEE Trans. Magn., 2011, 47, 386.

19 G. Cherubini, R. D. Cideciyan, L. Dellmann, E. Eleftheriou, W. Haeberle, J. Jelitto,

V. Kartik, M. A. Lantz, S. Olcer, A. Pantazi, H. E. Rothuizen, D. Berman, W. Imaino,

P. O. Jubert, G. McClelland, P. V. Koeppe, K. Tsuruta, T. Harasawa, Y. Murata, A.

Musha, H. Noguchi, H. Ohtsu, O. Shimizu and R. Suzuki, IEEE Trans. Magn., 2011,

47, 137.

20 S. Furrer, M. A. Lantz, J. B. C. Engelen, A. Pantazi, H. E. Rothuizen, R. D. Cideciyan,

G. Cherubini, W. Haeberle, J. Jelitto, E. Eleftheriou, M. Oyanagi, Y. Kurihashi, T.

Ishioroshi, T. Kaneko, H. Suzuki, T. Harasawa, O. Shimizu, H. Ohtsu and H. Noguchi,

IEEE Trans. Magn., 2015, 51, 3100207.

21 W. Zhong, W. P. Ding, Y. M. Jiang, N. Zhang, J. R. Zhang, Y. W. Du and Q. J. Yan,

109

J. Am. Ceram. Soc., 1997, 80, 3258.

22 L. Affleck, M. D. Aguas, Q. A. Pankhurst, I. P. Parkin and W. A. Steer, Adv. Mater.,

2000, 12, 1359.

23 J. Jin, S. Ohkoshi and K. Hashimoto, Adv. Mater., 2004, 16, 48.

24 L. C. Li, K. Y. Chen, H. Liu, G. X. Tong, H. S. Qian and B. Hao, J. Alloys Compd.,

2013, 557, 11.

25 A. Yourdkhani, D. Caruntu, A. K. Perez and G. Caruntu, J. Phys. Chem. C, 2014,

118, 1774.

26 P. Xu, X. J. Han and M. J. Wang, J. Phys. Chem. C, 2007, 111, 5866.

27 J. Zhang, J. Fu, F. Li, E. Xie, D. Xue, N. J. Mellors and Y. Peng, ACS Nano, 2012,

6, 2273.

28 M. Manikandan and C. Venkateswaran, J. Magn. Magn. Mater., 2014, 358, 82.

29 D. Lisjak and S. Ovtar, J. Phys. Chem. B, 2013, 117, 1644.

30 D. Primc and D. Makovec, Nanoscale, 2015, 7, 2688.

31 D. Primc, D. Makovec, D. Lisjak and M. Drofenik, Nanotechnology, 2009, 20,

315605.