SYNTHESIS AND CHARACTERIZATION OF...

50
SYNTHESIS AND CHARACTERIZATION OF COBALT NANOPARTICLES PREPARED BY LIQUID-PHASE REDUCTION MARY DONNABELLE L. BALELA UNIVERSITI SAINS MALAYSIA 2008

Transcript of SYNTHESIS AND CHARACTERIZATION OF...

SYNTHESIS AND CHARACTERIZATION OF COBALT NANOPARTICLES PREPARED BY

LIQUID-PHASE REDUCTION

MARY DONNABELLE L. BALELA

UNIVERSITI SAINS MALAYSIA 2008

SYNTHESIS AND CHARACTERIZATION OF COBALT NANOPARTICLES PREPARED BY LIQUID-PHASE REDUCTION

by

MARY DONNABELLE L. BALELA

Thesis submitted in fulfillment of the

requirements for the Degree of Master of Science

July 2008

ii

ACKNOWLEDGEMENTS

I would like to extend my gratitude to the following individuals and

institutions for their significant contributions in making the current research work

possible:

• Dr. Zainovia Lockman, for her constant probing, questioning, and

correcting, for her unbelievable patience, for teaching me the value of

research and what it takes to produce a good one.

• Dr. Azizan Aziz, for the conversations, debates, and arguments, for the

commendable patience as well.

• Prof. Eiichiro Matsubara and Dr. Alberto Amorsolo Jr., for their

valuable suggestions and advices.

• AUNSEED-Net/JICA, for the scholarship and financial assistance.

• Pn. Fong, Pn. Haslina, Zaini, Zul, En. Suhaimi, En. Azrul, En. Azam,

En. Rashid, for the company and assistance during experiments. Laboratory

work was never dull because of these people.

• Mr. Muthu and the staff of the Electron Microscope Laboratory, School

of Biology, USM, for the encouragements and support especially during

those long months of frustrations and failures.

iii

• Mr. Karuna of the Crystallography Laboratory, School of Physics,

USM, for the assistance in the XRD analysis.

• Dr. Mohd and En. Elias of SIRIM-AMREC, for the help on the magnetic

measurements.

• Kak Na, Kak Mila, and the rest of the administrative staff of the School

of Materials and Mineral Resources Engg, USM, for their assistance on

purchases and testing.

• Ng, Chia, and Hwee Yeo, for the help during TEM and XRD testing and for

the interesting discussions in the labs.

• Laili and Dr. Azizan Aziz, for the abstract translation.

• Vilay, Erfan, Tedi, Mr. Toguh, Luan, Viet, Dika, Kim, Nad, Parisha,

and to all my USM friends, for putting up with me these past two years.

• To my family, for their encouragement and support.

• Eugene Clef – for the blind faith, for being there 24/7, for keeping me sane

everyday, for making my life fun, interesting, and worth living.

Most of all, to God Almighty for all the blessings and opportunities He bestowed on

me; for the good health and for each day of being alive.

iv

TABLE OF CONTENTS PAGE

ACKNOWLEDGEMENTS ii

TABLE OF CONTENTS iv

LIST OF TABLES x

LIST OF FIGURES xiii

LIST OF ABBREVIATIONS xxii

LIST OF SYMBOLS xxiii

ABSTRAK xxiv

ABSTRACT xxv

CHAPTER 1: INTRODUCTION

1.1 Background of the Study and Problem Statement 1

1.2 Objectives of the Study 4

1.3 Project Overview 4

CHAPTER 2: LITERATURE REVIEW

2.1 Introduction 6

2.2 Cobalt Allotropes and their Allotropic Phase Transformations:

for Bulk and Nanosized Particles

8

2.3 Formation of Monosized and Uniformly Dispersed Cobalt

and Other Metal Nanoparticles

13

2.3.1 Separation of Nucleation and Growth 14

2.3.2 Control of the Growth Process 18

v

2.3.3 Inhibition of Random Agglomeration of Nanoparticles 21

2.3.4 Reserve of Monomers 24

2.4 Liquid-Phase Synthesis of Cobalt and Other Magnetic Nanoparticles 24

2.4.1 Metal Salt Reduction 26

2.4.2 Reactions in Reverse Micelles 29

2.4.3 Organometallic Syntheses 30

2.4.3.1 High Temperature Reduction 31

2.4.3.2 Thermal Decomposition of Organometallic

Compounds

32

2.5 Tailoring the Particle Size of Cobalt and Other Nanoparticles 32

2.5.1 Influence of Temperature on the Particle Size 33

2.5.2 Influence of Reactant Concentration on Particle Size 35

2.5.3 Influence of Reducing Agent on Particle Size 36

2.5.4 Influence of Protective Agents on Particle Size 37

2.5.5 Influences by Other Factors on the Particle Size 37

2.6 Magnetic Properties of Bulk and Nanosized Cobalt Particles 38

2.7 Oxidation of Bulk and Nanosized Cobalt Particles 46

2.8 Applications of Bulk and Nanosized Cobalt Particles 49

CHAPTER 3: MATERIALS AND METHODS

3.1 Introduction 51

3.2 Raw Materials 52

3.3 Method A: Metal Salt Reduction Using Hydrazine 53

3.3.1 Preparation of Solutions 54

3.3.2 Mixing of Solutions 55

vi

3.3.3 Effects of Process Temperature on the Reduction of Cobalt

Ions and on the Particle Size of Cobalt Nanoparticles

56

3.3.4 Effect of Stirring Speed on the Particle Size of Cobalt

Nanoparticles

56

3.3.5 Effects of Amount of Hydrazine on the Reduction of Cobalt

Ions and on the Particle Size of Cobalt Nanoparticles

57

3.3.6 Effect of Precursor Concentration (CoCl2·6H2O) on the

Reduction of Cobalt Ions and on the Particle Size of Cobalt

Nanoparticles

57

3.3.7 Precipitation of Colloidal Cobalt Nanoparticles 58

3.4 Method B: Metal Salt Reduction by the Polyol Method 58

3.4.1 Solution Preparation and Synthesis of Cobalt Particles 60

3.4.2 Effect of Precursor Concentration ((C2H3O2)2Co·4H2O)

on Cobalt Particle Size

60

3.4.3 Effect of Addition of Nucleating Agent on Cobalt

Particle Size

60

3.4.4 Effect of the Amount of Nucleating Agent on Cobalt

Particle Size

61

3.4.5 Effects of Addition of PVP k15 and its Amount on the

Particle Size and Agglomeration of the Cobalt Particles

62

3.5 Washing, Drying, and Storage of Cobalt Nanoparticles 64

3.6 Characterization of Cobalt Nanoparticles 64

3.6.1 Morphological and Structural Analyses 65

3.6.1.1 Transmission Electron Microscope (TEM) 65

vii

3.6.1.2 Field-Emission Scanning Electron Microscope

(FE-SEM)

66

3.6.2 Phase Analysis and Crystallite Size Determination 67

3.6.3 Chemical Analysis 68

3.6.4 Room-Temperature Magnetic Measurements 68

3.7 Chemical Reactivity (Thermal Oxidation Properties) of the Cobalt

Particles Prepared by the Polyol Method

69

CHAPTER 4: RESULTS AND DISCUSSION

4.1 Introduction 70

4.2 Method A: Metal Salt Reduction Using Hydrazine 70

4.2.1 Effects of Process Temperature on the Reduction of

Cobalt Ions and on the Particle Size of Cobalt Nanoparticles

71

4.2.2 Effect of Stirring Speed on Particle Size 81

4.2.3 Effect of Amount of Hydrazine on the Reduction of

Cobalt Ions and on the Particle Size of Cobalt Nanoparticles

85

4.2.4 Effect of Cobalt Chloride Concentration on the Reduction

of Cobalt Ions and on the Particle Size of

Cobalt Nanoparticles

93

4.2.5 Summary of Results 100

4.3 Method B: Metal Salt Reduction by the Polyol Method 101

4.3.1 Effect of Cobalt Acetate Concentration on the Particle Size 108

4.3.2 Effect of Addition of Nucleating Agent on the Cobalt

Particle Size

114

viii

4.3.3 Effect of the Amount of Nucleating Agent on the Cobalt

Particle Size

122

4.3.3.1 Effect of the Amount of AgNO3 on the Cobalt

Particle Size

122

4.3.3.2 Effect of the Amount of PdCl2 on the Cobalt

Particle Size

127

4.3.3.3 Effect of the Amount of H2PtCl6 on the Cobalt

Particle Size

131

4.3.4 Effect of the Addition of PVP k15 on the Particle Size

and Agglomeration of Cobalt Particles

136

4.3.5 Summary of Results 144

4.4 Room-Temperature Magnetic Properties of the Cobalt Particles

Prepared by the Polyol Method

145

4.4.1 Size-Dependence of the Coercivity (Hc) of the

Cobalt Particles

148

4.4.2 Size-Dependence of the Saturation Magnetization (Ms)

and Remanence (Mr/Ms) of the Cobalt Particles

150

4.4.3 Summary of Results 152

4.5 Chemical Reactivity (Thermal Oxidation Properties)

of the Cobalt Particles Prepared by the Polyol Method

152

4.5.1 Chemical Reactivity (Thermal Oxidation Properties)

of Pure Cobalt Particles

154

4.5.2 Chemical Reactivity (Thermal Oxidation Properties)

of the Cobalt Nanoparticles with Pd Nucleus

157

ix

4.5.3 Chemical Reactivity (Thermal Oxidation Properties)

of the Cobalt Nanoparticles with Pt Nucleus

161

4.5.4 Summary of Results 164

CHAPTER 5: CONCLUSIONS AND RECOMMENDATIONS

5.1 Introduction 165

5.2 Conclusions 165

5.2.1 Method A: Metal Salt Reduction Using Hydrazine 165

5.2.2 Method B: Metal Salt Reduction by the Polyol Method 166

5.2.3 Room-Temperature Magnetic Properties of the Cobalt

Particles Prepared by the Polyol Method

166

5.2.4 Chemical Reactivity (Thermal Oxidation Properties)

of the Cobalt Particles Prepared by the Polyol Method

166

5.3 Recommendations for Future Research 167

REFERENCES 168

APPENDIX: REPRODUCIBILITY

LIST OF PUBLICATIONS

LIST OF PRESENTATIONS

x

LIST OF TABLES

PAGE

Table 2.1 Lattice Parameters, Lattice Surface Area (A), Lattice Volume

(V), and Calculated Densities (ρ) for HCP-Co, FCC-Co, and

ε-Co (ICDD Card No. 5-0727 and 15-0806; Dinega and

Bawendi, 1991; Sun and Murray, 1999; Ram, 2001)

9

Table 2.2 Comparison of Diffraction Peaks and Relative Intensities of

11 nm Co Nanocrystals and β-Mn (Sun and Murray, 1999).

12

Table 2.3 Magnetic Moment per Atom and Magnetocrystalline

Anisotropy of HCP-Co, FCC-Co, and ε-Co (Cullity, 1972;

Shukla et al., 2006).

13

Table 2.4 Some Protective Agents Used in the Synthesis of Co

Nanoparticles.

23

Table 2.5 Standard Electrode Potential, E˚ and the Corresponding

Reduction Reactions Taken at STP and in Aqueous Solutions

at Room Temperature (Cushing et al., 2005).

27

Table 2.6 Saturation Magnetization at 20 ˚C (Ms) and 0 K (Mo) and the

Curie Temperatures (Tc) for Fe, Co, and Ni (Cullity, 1972)

41

Table 3.1 Raw Materials

53

Table 3.2 Reactant Concentrations and Reaction Temperatures

Investigated

56

Table 3.3 List of Stirring Speeds Studied

57

Table 3.4 Mole Ratios of N2H4 to Co2+ (R) Experimented 57

xi

Table 3.5 Initial Co2+ Concentrations Employed in the Study

58

Table 3.6 Amounts of (C2H3O2)2Co·4H2O Studied

60

Table 3.7 Concentrations of Reactants Used

61

Table 3.8 Mole Ratios of Nucleating Agent (NA) to Co2+ Studied

62

Table 3.9 Experimental Conditions Employed to Determine the Effect

of PVP k15

63

Table 3.10 ICDD Cards Employed in the Analysis of the XRD Patterns

67

Table 4.1 Synthesis of Co Nanoparticles at Different Temperatures

72

Table 4.2 Calculated Particle and Crystallite Sizes of Co Nanoparticles

Produced at Different Stirring Speeds

84

Table 4.3 Calculated Particle and Crystallite Sizes of Co Nanoparticles

Produced at Different Mole Ratios of N2H4 to Co2+ (R)

92

Table 4.4 Calculated Particle and Crystallite Sizes of Co Nanoparticles

Produced at Increasing Precursor Concentration.

99

Table 4.5 Values of the Mean Particle Size, Calculated Crystallite Size,

and Elemental Composition by EDX Analysis

113

Table 4.6 Values of the Relative Standard Deviation (%) and Crystallite

Size (nm) Obtained from the TEM and XRD Analyses of Co

Particles Nucleated with Ag

125

xii

Table 4.7 Values of the Relative Standard Deviation (%) and Crystallite

Size (nm) Obtained from the TEM and XRD Analyses of Co

Particles Nucleated with Pd

129

Table 4.8 Values of the Mean Particle Size (nm), Relative Standard

Deviation (%), and Crystallite Size (nm) Obtained from the

TEM and XRD Analyses of Co Particles Seeded with Pt.

132

xiii

LIST OF FIGURES

PAGE

Fig. 2.1 Representations of the (a) HCP and (b) FCC crystal structures

of Co (Callister; 2000)

9

Fig. 2.2 Unit cell of ε-cobalt: (a) unit cell cube filled with eight atoms

of Type I (dark) and twelve atoms of Type II (light); (b) the

111-projection of the same cube showing three-fold symmetry

along its main diagonal (Dinega and Bawendi, 1999).

11

Fig. 2.3 La Mer model for the formation of monodispersed systems by

controlled homogenous nucleation and growth: I.

prenucleation period; II. nucleation step; III. growth step

(Redrawn from Viau et al., 1996).

16

Fig. 2.4 (a) TEM image of Co50Ni50 nanoparticles produced by

heterogeneous nucleation with Pt in polyol (dm = 6 nm) and its

corresponding (b) XRD pattern with increasing mole ratios of

Pt to (Co+Ni) showing only peaks of Co and Ni: a. dm = 44

nm; b. dm = 15 nm; c. dm = 5 nm (Viau et al., 2001).

18

Fig. 2.5 Plot illustrating the radius difference (δr/r) among the

nanoparticles as a function of (a) particle radius (r) and (b)

growth period (t) for diffusion-controlled growth (I),

polynuclear growth (II), and mononuclear growth (III)

mechanisms (Modified from Cao, 2004).

20

Fig. 2.6 Scheme depicting the growth mechanism by nucleation and

agglomeration (Capek, 2006)

21

Fig. 2.7 Preparation of metal nanoparticles in homogeneous solution by

liquid-phase synthesis (Toshima, 2003).

25

xiv

Fig. 2.8 Hot injection technique for organometallic synthesis of

magnetic nanoparticles; TC= thermocouple; Compositions of

Solution A and B are detailed in the text (Talapin et al., 2004)

31

Fig. 2.9 Summary of the processes involved during the nucleation and

growth of the nanoparticles in solution.

34

Fig. 2.10 A typical hysteresis loop for a ferromagnetic material (Hibst

and Schwab, 1994).

39

Fig. 2.11 Particle size versus coercivity. The largest coercivity is

achieved at the critical size dc corresponding to the single

domain limit of a material; SPM = Superparamagnetic (Talapin

et al., 2004).

42

Fig. 2.12 Schematic of the energy barrier, ∆E, needed to rotate the

magnetization of a uniaxial magnetic nanoparticle (dotted lines

show the easy axes of magnetization) (Talapin et al., 2004).

44

Fig. 3.1 Schematic diagram of the synthesis of cobalt nanoparticles by

metal salt reduction using hydrazine

54

Fig. 3.2 Flowchart of the synthesis of Co nanoparticles via metal salt

reduction by the polyol method

59

Fig. 4.1 (a) Fibrous materials formed from CoCl2, EG, and N2H4 at 60

˚C, (b) very fine Co nanoparticles produced during reaction at

70 ˚C, (c) large Co nanoparticles prepared at 80 ˚C and (d) its

corresponding particle size distribution. Mole Co2+ = 1.0 mmol

(30 mM); Mole ratio of OH¯ to Co2+ = 2; Mole ratio of N2H4 to

Co2+ = 30.

73

xv

Fig. 4.2 X-ray diffraction pattern of T80. F= FCC; H = HCP. Inset

shows the SEM micrograph of the precipitated nanoparticles of

T80.

78

Fig. 4.3 Typical TEM images of (a) S350 and (b) S420. Insets are the

images obtained at a lower magnification. [Co2+] = 40 mM;

Mole ratio of OH¯ to Co2+ = 2; Mole ratio of N2H4 to Co2+ (R)

= 30.

82

Fig. 4.4 Particle size distributions obtained for (a) S350 and (b) S420.

84

Fig. 4.5 X-ray diffractograms of samples (a) T80/S250 (dm = 12.43

nm), (b) S350 (dm = 5.54 nm), and (c) S420 (dm = 4.10 nm).

H= HCP; F= FCC.

85

Fig. 4.6 TEM images of (a) R10, (b) R12, and (c) R15. [Co2+] = 40

mM; Mole ratio of OH¯ to Co2+ = 2; Stirring rate = 420 rpm.

88

Fig. 4.7 TEM images of (a) R20, (b) R40, and (c) R50. [Co2+] = 40

mM; Mole ratio of OH¯ to Co2+ = 2; Stirring rate = 420 rpm.

89

Fig. 4.8 Plot of average particle size (nm) against mole ratio of N2H4 to

Co2+ (R).

90

Fig. 4.9 Particle size distributions obtained for (a) R10, (b) R 12, and

(c) R15.

90

Fig. 4.10 Particle size distributions obtained for (a) R20, (b) R 40, and

(c) R50.

91

Fig. 4.11 X-ray diffraction patterns of (a) R10 (dm = 7.09 nm), (b) R20

(dm = 4.05 nm), (c) R30 (dm = 4.10 nm), and (d) R50 (dm =

3.59 nm). H = HCP; F = FCC.

93

xvi

Fig. 4.12 Plot of mean particle size (nm) against cobalt chloride

concentration (mM)

95

Fig. 4.13 Typical TEM image of Co nanoparticles obtained at 30 mM

(P30). Mole ratio of N2H4 to Co2+ (R) = 50; Mole ratio of OH¯

to Co2+ = 2; Stirring rate = 420 rpm.

95

Fig. 4.14 Typical TEM images of Co nanoparticles obtained at (a) 50

mM (P50), (b) 70 mM (P70), and (c) 90 mM (P90). Insets are

images taken at lower magnification. Mole ratio of N2H4 to

Co2+ (R) = 50; Mole ratio of OH¯ to Co2+ = 2; Stirring rate =

420 rpm.

96

Fig. 4.15 Particle size distributions obtained for (a) P30 and (b) P50.

97

Fig. 4.16 Particle size distributions obtained for (a) P70 and (b) P90.

98

Fig. 4.17 X-ray diffraction patterns of (a) P90 (dm = 6.79 nm), (b) P70

(dm = 5.94 nm), (c) R50/P40 (dm = 3.59 nm), and (d) P30 (dm =

3.21 nm). H = HCP; F = FCC.

99

Fig. 4.18 Typical (a) TEM and (b) SEM images of C5; Co acetate = 5.0

mmol; PG = 50 ml; SEM magnification = 10,000x

105

Fig. 4.19 The corresponding particle size distribution for C5.

106

Fig. 4.20 X-ray diffraction pattern of C5. H = HCP; F = FCC

107

Fig. 4.21 Plot of mean particle size (nm) against the amount of Co

acetate (mmol)

109

xvii

Fig. 4.22 Typical TEM images of (a) C3 (dm = 388 nm), (b) C4 (dm =

364 nm), and (c) C6 (dm = 152 nm). Vol of PG = 50 ml; Mole

ratio OH¯ to Co2+ = 2.5; dm = mean diameter

110

Fig. 4.23 Corresponding SEM images of (a) C3 (dm = 388 nm), (b) C4

(dm = 364 nm), and (c) C6 (dm = 152 nm) taken at 10,000x

magnification. Vol of PG = 50 ml; Mole ratio OH¯ to Co2+ =

2.5; dm = mean diameter

111

Fig. 4.24 Particle size distributions for (a) C3 (dm = 388 nm), (b) C4 (dm

= 364 nm), and (c) C6 (dm = 152 nm) obtained by analyzing

300 particles for each sample

112

Fig. 4.25 X-ray diffractograms of as-prepared Co nanoparticles from (a)

C3 (dm = 388 nm), (b) C4 (dm = 364 nm), (c) C5 (dm = 298

nm). H = HCP; F = FCC

114

Fig. 4.26 TEM images of (a) Ag2.5 (dm = 106 nm), (b) Pd2.5 (dm = 121

nm), and (c) Pt2.5 (dm = 61 nm). Inset is a TEM image of Pt2.5

at higher magnification. Arrows point to the impurity layer

surrounding the nanoparticles. Nucleation ratio = 2.5 x 10-2;

Amount of Co2+ = 5.0 mmol

117

Fig. 4.27 The corresponding SEM images of (a) Ag2.5 (dm = 106 nm),

(b) Pd2.5 (dm = 121 nm), and (c) Pt2.5 (dm = 61 nm).

Nucleation ratio = 2.5 x 10-2; Amount of Co2+ = 5.0 mmol

118

Fig. 4.28 Particle size distributions of (a) Ag2.5, (b) Pd2.5, and (c) Pt2.5.

120

Fig. 4.29 XRD diffraction patterns of (a) C5, (b) Pd2.5, (c) Ag2.5, and

(d) Pt2.5. H = HCP; F = FCC; Ag = Ag peaks

121

xviii

Fig. 4.30 Plot of mean particle size (nm) of the Co particles obtained

using varying AgNO3 nucleation ratio (10-2). Amount of Co

acetate = 5.0 mmol

123

Fig. 4.31 TEM images of (a) Ag0.5 (dm = 369 nm), (b) Ag1.5 (dm = 357

nm), and (c) Ag2.0 (dm = 291 nm). Amount of Co acetate = 5.0

mmol.

124

Fig. 4.32 Corresponding particle size distributions of (a) Ag0.5, (b)

Ag1.5, and (c) Ag2.0.

125

Fig. 4.33 X-ray diffraction patterns of (a) Ag0.5, (b) Ag1.5, (c) Ag2.0,

and (d) Ag2.5. F = FCC-Co; H = HCP-Co; Ag = Ag peaks.

126

Fig. 4.34 Plot of mean particle size (nm) of the Co particles obtained

using varying PdCl2 nucleation ratio (10-2). Amount of Co

acetate = 5.0 mmol.

127

Fig. 4.35 TEM images of (a) Pd3.5 (dm = 112 nm), (b) Pd4.0 (dm = 92

nm), (c) Pd4.5 (dm = 88 nm), and (d) Pd5.0 (dm = 57 nm).

Amount of Co acetate = 5.0 mmol.

128

Fig. 4.36 Corresponding particle size distributions of (a) Pd3.5, (b)

Pd4.0, (c) Pd4.5, and (d) Pd5.0.

130

Fig. 4.37 X-ray diffraction patterns of (a) Pd2.5, (b) Pd3.5, (c) Pd4.0,

and (d) Pd4.5. F = FCC-Co; H = HCP-Co; Pd = Pd peaks.

131

Fig. 4.38 TEM images of (a) Pt0.5(dm = 63 nm), (b) Pt1.5 (dm = 57 nm),

(c) Pt3.5 (dm = 59 nm), (d) Pt4.0 (dm = 52 nm), (e) Pt4.5 (dm =

49 nm), and (f) Pt5.0 (dm = 59 nm). Amount of Co acetate =

5.0 mmol.

133

xix

Fig. 4.39 Corresponding particle size distributions of (a) Pt0.5, (b) Pt1.5,

(c) Pt3.5, and (d) Pt4.0.

134

Fig. 4.40 Corresponding particle size distributions of (a) Pt4.5, and (b)

Pt5.0.

135

Fig. 4.41 X-ray diffraction patterns of (a) Pt1.5, (b) Pt2.5, (c) Pt3.5, (d)

Pt4.5, and (e) Pt5.0. F = FCC-Co; H = HCP-Co.

136

Fig. 4.42 Typical TEM micrographs of samples (a) Co-0.075 (dm = 181

nm), (b) CoPd-0.075 (dm = 62 nm), (c) CoAg-0.20 (dm = 166

nm), and (d) CoPt-0.20 (dm = 78 nm). Amount of Co acetate =

5.0 mmol; Nucleation ratio = 2.5 x 10-2.

138

Fig. 4.43 Particle size distributions for samples (a) Co-0.075, (b) CoPd-

0.075, (c) CoAg-0.20, and (d) CoPt-0.20.

141

Fig. 4.44 X-ray diffraction patterns of (a) pure Co particles produced by

homogeneous nucleation and (b) Co particles nucleated with

PdCl2, with and without the addition of PVP k15. Amount of

Co acetate = 5.0 mmol; Nucleation ratio = 2.5 x 10-2; Weight

ratio of Co(OH)2 to PVP k15 = 0.075; H= HCP; F = FCC.

142

Fig. 4.45 X-ray diffraction patterns of Co particles produced by

heterogeneous nucleation with (a) AgNO3 and (b) H2PtCl6,

with and without the addition of PVP k15. Amount of Co

acetate = 5.0 mmol; Nucleation ratio = 2.5 x 10-2; Weight ratio

of Co(OH)2 to PVP k15 = 0.20; H= HCP; F = FCC.

143

Fig. 4.46 Magnetization (emu/g) vs. applied magnetic field (Tesla) for

the Co particles having mean particle size of (a) 364 nm (C4),

(b) 298 nm (C5), and (c) 152 nm (C6).

146

xx

Fig. 4.47 Magnetization (emu/g) vs. applied magnetic field (Tesla) for

the Co particles having mean particle size of (a) 112 nm

(Pd3.5), (b) 61 nm (Pt2.5), (c) 49 nm (Pt4.5).

147

Fig. 4.48 Plot showing the effect of the particle size (nm) on the

coercivity (Hc, Oe) of the Co particles having average

diameters from 49 to 364 nm at room temperature. Inset is the

plot of coercivity (Hc, Oe) versus the inverse of the particle

size (d-1, nm-1) of the Co particles with diameters > 61 nm.

149

Fig. 4.49 Plots of saturation magnetization (Ms, emu/g) against (a) mean

particle size (nm) and (b) inverse of the mean particle size (d-1,

nm-1) of the Co particles with mean diameters of 49 to 364 nm.

151

Fig. 4.50 TGA curve and the corresponding % weight gain of pure Co

particles (298 nm) after thermal oxidation from room

temperature to 700 ˚C and 3 hour isothermal treatment at about

700 ˚C.

155

Fig. 4.51 X-ray diffraction pattern of Co particles oxidized to Co3O4.

156

Fig. 4.52 TGA curves and the corresponding % weight gains of (a) pure

Co particles (298 nm); Pd seeded Co particles with nucleation

ratio of (b) 2.5 x 10-2 (121 nm), (c) 4.0 x 10-2 (92 nm), and (d)

5.0 x 10-2 (57 nm).

158

Fig. 4.53 XRD patterns of the oxides from (a) pure Co particles (298

nm) and (b) Co particles with Pd (Nucleation ratio = 2.5 x 10-2,

121 nm). # = Co3O4; PdO = PdO peaks; Pd = Pd peaks.

159

xxi

Fig.4.54 XRD patterns of Pd seeded Co particles (Nucleation ratio = 2.5

x 10-2, 121 nm) oxidized for 3 hours at (a) 400 ˚C, (b) 200 ˚C,

(c) 100 ˚C, and (d) room temperature; # = Co3O4 peaks; Co =

Co peaks.

160

Fig. 4.55 TGA curves and the corresponding % weight gains of (a) pure

Co particles (298 nm); Pt seeded Co nanoparticles with

nucleation ratio of (b) 2.5 x 10-2 (61 nm), (c) 3.5 x 10-2 (59

nm), and (d) 4.0 x 10-2 (52 nm).

163

Fig. 4.56 XRD patterns of the oxides from (a) pure Co particles (298

nm) and (b) Co nanoparticles with Pt (nucleation ratio = 2.5 x

10-2, 61 nm). # = Co3O4

163

xxii

LIST OF ABBREVIATIONS AFM : Antiferromagnetic AOT : Bis(2-ethylhexyl) S-sodium Sulfosuccinate CTAB : Centrimethyl Ammonium Bromide DDAB : Didodecyldimethyl Ammonium Bromide DMF : N,N dimethylformamide EDTA : Ethylenediamine-N,N,N’,N’-tetracetic EDX : Energy Dispersive X-ray Spectroscopy EG : Ethylene Glycol FCC : Face-Centered Cubic FE-SEM : Field Emission Scanning Electron Microscope HCP : Hexagonal Close-Packed LA : Lauric Acid NTA : Nitrilotriacetic Acid OA : Oleic Acid PG : Propylene Glycol (ph)3P : Triphenylphosphine PR3 : Trialkylphosphine SDS : Sodium Dodecylsulfosuccinate STP : Standard Temperature and Pressure TEM : Transmission Electron Microscope TGA : Thermogravimetric Analysis TOPO : Trioctylphosphane Oxide Triton X-100 : Polyoxyethylene(10) Isooctylphenyl Ether

VSM : Vibrating Sample Magnetometer XRD : X-ray Diffraction

xxiii

LIST OF SYMBOLS a : Lattice parameter A : Area Ag : Silver Ar : Argon Au : Gold BH4¯ : Borohydride ion β : Beta Co : Cobalt Co2(CO)8 : Dicobalt octacarboynyl E˚ : Standard electrode potential ε : Epsilon Fe : Iron H2 : Hydrogen gas LiBEt3H : Lithium hydride MBH4 : Metal tetraborohydrates Mn : Manganese N2H4 : Hydrazine N2H5

+ : Hydrazinium ion NaBH4 : Sodium borohydride NaOH : Sodium hydroxide Ni : Nickel Pd : Palladium Pt : Platinum PVP : Poly vinyl pyrrolidone ρ : Density µB : Bohr magneton V : Volume

xxiv

SINTESIS DAN PENCIRIAN NANOPARTIKEL KOBALT MELALUI PENURUNAN FASA-CECAIR

ABSTRAK

Nanopartikel Co telah dihasilkan secara penurunan fasa-cecair menggunakan

N2H4 dan propilena glikol tanpa sebarang agen perlindungan. Nanopartikel yang

amat halus dan bulat disediakan dengan penurunan Co2+ menggunakan N2H4 dalam

larutan etilena glikol berbes. Nanopartikel kelihatan diskret dan terserak dengan baik

bila ditaburkan dalam etilena glikol. Walaubagaimanapun nanopartikel bergumpal

ke saiz submikron bila dimendakan dari larutan. Diperhatikan juga bahawa

penurunan Co2+ oleh N2H4 adalah perlahan, walaupun suhu, kepekatan N2H4 dan

Co2+ tindak balas dipertingkatkan. Sebaliknya partikel Co yang lebih besar dan

bersaiz seragam antara 30 nm hingga 400 nm serta berbentuk kuasi-sfera dihasilkan

dalam propilena glikol dan saiz partikel dikawal dengan mengubah kepekatan

prekursor dan menambah agen penukleusan AgNO3, PdCl2 dan H2PtCl6.

Meningkatkan amaun prekursor, termasuk nisbah agen penukleusan AgNO3 atau

PdCl2 menghasilkan partikel Co yang kecil. Dalam kes Pt, saiz partikel kekal dalam

semua nisbah agen penukleusan H2PtCl6 yang dikaji. Saiz partikel Co yang paling

kecil 30 – 80 nm disintesis dengan Pt. Pembelauan Sinar –X menunjukkan kedua

kaedah menghasilkan campuran Co heksagonal susun-padat dan kubik berpusat

muka. Sukatan magnet pada suhu bilik menunjukkan partikel Co yang disediakan

dalam propilena glikol semuanya ferromagnet dengan had domain-tunggal 61 nm.

Pengoksidaan termal sampel menunjukkan partikel-partikel yang dihasilkan tidak

mudah teroksida dibawah 200°C kerana kewujudan lapisan bendasing pasif.

Melepasi suhu ini, lapisan bendasing ini terurai dan partikel Co teroksida kepada

Co3O4.

xxv

SYNTHESIS AND CHARACTERIZATION OF COBALT NANOPARTICLES PREPARED BY LIQUID-PHASE REDUCTION

ABSTRACT

Co nanoparticles were produced by liquid-phase reduction using N2H4 and

propylene glycol without any protective agent. Very small, though polydispersed,

spherical Co nanoparticles were prepared by Co2+ reduction using N2H4 in a basic

ethylene glycol solution. The nanoparticles appeared to be discrete and well-

separated when dispersed in ethylene glycol. However, they agglomerated to

submicron-sized particles when precipitated from the solution. It was also inferred

that the reduction of Co2+ by N2H4 was generally very slow, though the reaction was

enhanced by an elevated temperature, a high N2H4 concentration, and a large amount

of Co2+. On the other hand, larger but monodispersed quasi-spherical Co particles

with sizes between 30 nm to 400 nm were produced in propylene glycol and the

particle size was tailored by varying the precursor concentration and by adding

nucleating agents such as AgNO3, PdCl2, and H2PtCl6. Increasing the amount of

precursor, as well as AgNO3 or PdCl2 nucleation ratios, formed smaller Co particles.

In the case of samples with Pt, the particle size remained almost constant in all

H2PtCl6 nucleation ratios explored. Even so, the smallest Co particles of 30 – 80 nm

in diameters were synthesized with Pt. XRD then showed that both methods

produced mixtures of HCP and FCC-Co. Room-temperature magnetic

measurements revealed that the Co particles prepared in propylene glycol were all

ferromagnetic with a single-domain limit of 61 nm. Thermal oxidation of the

samples showed that they were not susceptible to oxidation below 200 ˚C due to the

passivation of the impurity layer. Beyond this temperature, the impurity layer

decomposed and the Co particles oxidized to Co3O4.

1

CHAPTER 1 INTRODUCTION

1.1 Background of the Study and Problem Statement

Nanoparticles are zero-dimensional nanostructures (Cao, 2004; Gubin et al.,

2005) and are generally classified according to their composition: metal oxides,

noble metals, transition metals, magnetic metals, etc. (Cao, 2004; Sergeev, 2006).

Like all nanostructures, the properties of nanoparticles are dependent on their size

and shape. The variation of their properties at the nanoscale is not a result of a

scaling factor, but stems from different causes in different materials (Burda et al.,

2005). In semiconductor, it is due to the further confinement of the electronic

motion to a length scale that is comparable to or less than the length scale of the

electronic motion in bulk semiconductors (Moriarty, 2001). In noble metals, it

results from the strong absorption of radiation within the visible region leading to

the collective oscillation of the electrons in the conduction band, called surface

plasmon resonance, from the surface of one particle to another (Burda et al., 2005).

In transition metals, it arises from the large surface to volume ratio resulting to high

chemical activities (Toshima, 2003). And in magnetic metals, it is due to finite-size

and surface effects, which become increasingly important as the particle size of the

magnetic material is reduced (Batlle and Labarta, 2002).

Co is a well-known ferromagnetic material which is commonly used as an

alloying element in permanent magnets (Goldman, 1999). It exists in two forms:

HCP (hexagonal close-packed) and FCC (face-centered cubic). HCP is the stable

phase at room temperature, whereas FCC is stable at temperatures above 450 ˚C

(Dinega and Bawendi, 1999).

2

In nanosized, Co particles display a wide range of interesting size-dependent

structural, electrical, magnetic, and catalytic properties (Puntes et al., 2001). In

particular, because of their large surface area, Co nanoparticles showed high

chemical reactivity, which makes them suitable for catalysis (Hyeon, 2002).

Furthermore, below a critical size of 20 nm, they behave as singe-domain particles

displaying quantum size effects, superparamagnetism, large magnetic anisotropies,

and a maximum coercivity (Ram, 2001). Moreover, at this size range, a new

metastable phase can also be formed, called ε-Co that has a complex β-Mn structure

and has properties in between the HCP and FCC phases. ε-Co can be prepared by

solution chemistry, particularly via organometallic route in the presence of tight

binding ligands (Dinega and Bawendi, 1999; Sun and Murray, 1999). Thus, these

previous studies prove that Co as a nanomaterial is an excellent and interesting

system to study.

Future applications of Co nanoparticles in the fields of separation

technology, information storage systems, catalysis, and biomedicine (Kodama,

1999) require the nanoparticles to be discrete, identical in size and shape, and

uniform in composition and crystal structure (Cao, 2004). However, formation of

nanoparticles satisfying these requirements prove to be difficult due to their high

surface energy, their intrinsic magnetic properties, and the inherent limitations of the

available processes. In addition, there are also concerns regarding the

reproducibility of the existing methods.

By far, liquid-phase synthetic routes are the most successful in the

preparation of monodispersed Co nanoparticles. Examples of liquid-phase processes

3

are metal salt reduction, reverse micelles, and thermal decomposition of

organometallic precursor. Among these methods, direct metal salt reduction in

solution is the simplest, fastest, and the least expensive, which are desirable for

future attempts of large-scale production (Guo et al., 2002).

An important component of metal salt reduction processes is the protective

agent. Protective agents such as surfactants that form a layer of molecular

membrane around the nanoparticles and polymers that provide steric hindrance

between nanoparticles are added during the reaction to inhibit particle agglomeration

and to control the particle growth. However, besides posing as an additional cost to

the total cost of the synthesis, they may complicate the cleaning process of the

nanoparticles since they are not easy to remove by simple washing and might require

secondary cleaning process, such as annealing, to decompose the protective agents

covering the nanoparticles (Wu and Chen, 2003). Moreover, selection of the

suitable protective agent for a particular reaction is a tedious task of trial-and-error,

since the surface chemistry of the nanoparticles is complex due to their high surface

energy and their tendency to accumulate surface charges. Thus, the interaction of the

protective agent with the solvent and with the nanoparticles plus its properties, such

as the solubility in the solvent, molecular weight, and the amount needed for

stabilization, must be taken into account when choosing the right protective agent.

Thus, in some cases, agglomeration of nanoparticles still occurred even in the

presence of protective agents (Chou and Huang, 2001; Lee and Chen, 2006). One

way to solve the problem is by the choice of a solvent that can also act as protective

agent. In this way, not only the nanoparticles can be protected, but the process will

be cheaper and more straightforward.

4

In this study, preparation of Co nanoparticles was performed using Co salt

reduction technique, without adding protective agents. Polyalcohols such as

ethylene glycol and propylene glycol were used as solvents and their ability to act as

the dispersing agent and growth inhibitors were also evaluated based from the

average particle size and size distribution of the Co nanoparticles. Therefore, the

development of “protective-agent free” methods will greatly simplify the formation

of cobalt nanoparticles by liquid-phase reduction.

1.2 Objectives of the Study

As mentioned, the formation of Co nanoparticles is accomplished by Co salt

reduction technique without any protective agent. The objectives of this work are:

(a) to prepare uniformly sized, monodispersed, and size-tunable Co

nanoparticles using a simplified Co salt reduction process.

(b) to investigate the morphology, crystal structure, magnetic properties, and

oxidation behavior of the Co nanoparticles.

1.3 Project Overview

In achieving the objectives, three main experiments were conducted. The

first experiment involves the reduction of Co ions in a basic ethylene glycol solution

using hydrazine, N2H4, as reductant. The size of the Co nanoparticles were tuned by

varying the reaction temperature, stirring speed, precursor concentration, and

amount of N2H4. However, in all samples, the particle size distribution was broad

with a relative standard deviation greater than 20%. Furthermore, severe

agglomeration of the nanoparticles was observed after precipitation.

5

To obtain Co nanoparticles with narrower size distribution, a second

experiment using the polyol method was explored, which was accomplished by

heating an alkaline solution of propylene glycol containing Co acetate up to the

boiling temperature for about two hours with reflux. Control of particle size was

achieved by varying the precursor concentration and by the addition of nucleating

agents. A set of experiments was also performed to study the effect of PVP k15 as

protective agent on the morphology and agglomeration of the nanoparticles.

Furthermore, this was also done to demonstrate how addition of protective agent can

complicate the synthesis and the cleaning procedure of the nanoparticles.

Standard characterization techniques such as TEM, XRD, FE-SEM, EDX,

and VSM were then performed on the synthesized Co nanoparticles.

The third experiment was conducted to study the oxidation behavior of

monodispersed Co nanoparticles obtained from the second experiment. Firstly, this

is important to compare the oxidation behavior of bulk Co or Co in a bar form to that

of the nanoparticles and to investigate the effects of the nucleating agent and the

impurity layer on the surface of the nanoparticles on their oxidation properties.

Thermal oxidation of Co nanoparticles was achieved by exposing the nanoparticles

in flowing oxygen gas and increasing temperature in TGA. Oxidized nanoparticles

were also sent for XRD analysis to identify the phases present.

6

CHAPTER 2 LITERATURE REVIEW

2.1 Introduction

Nanotechnology deals with the ability to fully control the properties of

nanostructured materials through their size, shape, and composition, as well as with

the fabrication of reproducible complex structures from simpler systems (Eggeman

et al., 2006). The development of nanotechnology is expected to proceed at several

levels: materials, devices, and systems (Salata et al., 2004). Currently, the most

advanced in terms of scientific knowledge and potential industrial application is the

materials level (Hyeon, 2002; Salata et al., 2004).

Nanostructured materials refer to those materials whose structural elements

(crystallites, clusters, or molecules) have sizes less than 100 nm at least in one

dimension (Moriarty, 2001). These include nanoparticles, nanorods, nanowires,

nanotubes, and thin films among others (Cao, 2004). Interest on these materials lies

on their novel electrical, optical, chemical and magnetic properties that have great

technological and fundamental scientific importance due to their remarkable

difference from the properties of their bulk counterparts. Furthermore, by

controlling their size and interactions, the fundamental properties of these materials

maybe tailored (Cushing et al., 2004). Thus, it is clear that utilizing these materials

fully and effectively depends on the understanding of their preparation, their general

properties, and the factors that influence these properties.

As mentioned, nanoparticles are zero-dimensional nanostructured materials

in which all characteristics linear dimensions are of the same order of magnitude

7

(Cao, 2004; Gubin et al., 2005). A special class of nanoparticles, i.e. magnetic

nanoparticles, exhibit unique phenomena such as superparamagnetism, high

coercivity, high saturation field, and exchanged anisotropy due to a thin layer of

antiferromagnetic oxide formed around the ferromagnetic core of the nanoparticles

because of their subsequent exposure to air (Tartaj et al., 2003). These phenomena

are attributed to size and surface effects that dominate the behavior of magnetic

particles in nanosized range (Batlle and Labarta, 2002). Typical examples are Co

nanoparticles.

Therefore, in this chapter, the structural, magnetic and oxidation properties

of bulk Co are described and these macroscopic properties are then compared to the

reported properties of nanosized Co particles. Since monodispersed and discrete

nanoparticles are required to realize their applications, the general principles for the

formation of uniformly sized nanoparticles are explained and the common liquid-

phase synthetic routes for Co and other magnetic nanoparticles are also described.

Emphasis is given on these methods by virtue of their wide application and success

in the preparation of monodispersed and high quality Co nanoparticles.

Furthermore, the general effects of synthesis parameters such as temperature,

precursor concentration, reductant, and protective agent on the nanoparticle size and

size distribution are also discussed. Lastly, the potential applications of Co

nanoparticles in separation technology, catalysis, high-density magnetic recording,

and biomedicine are reviewed.

8

2.2 Cobalt Allotropes and their Allotropic Phase Transformations: for Bulk and Nanosized Particles

It has been known that bulk cobalt exists in two allotropes, namely: (a)

hexagonal close-packed Co (HCP-Co) and (b) face-centered cubic Co (FCC-Co).

Thermodynamically, HCP-Co is the stable phase below 450 ˚C, while FCC-Co

forms at higher temperatures (Erbudak et al., 1997; Sato et al., 1997; Dinega and

Bawendi, 1999). Accordingly, these phases are in equilibrium at about 422 – 427 ˚C

(Dinega and Bawendi, 1999).

Because of the difference in crystal structure, variations in physical and

magnetic properties between the two polymorphs arise. HCP-Co is slightly denser

than FCC-Co, even though both phases are close-packed structures (Dinega and

Bawendi, 1999). Furthermore, HCP-Co is also magnetically harder than the FCC

phase due to its magnetic anisotropy and high coercivity as compared to the

symmetrical and low coercivity FCC phase (Shukla et al., 2006). Thus, the highly

anisotropic HCP-Co is more desirable for magnetic recording and other permanent

magnet applications. On the other hand, the FCC phase is more suitable for soft

magnetic applications (Dinega and Bawendi, 1999; Sun and Murray, 1999; Shukla et

al., 2006). The HCP and FCC crystal structures of Co are illustrated in Fig. 2.1 (a)

and (b) respectively, while the values of the crystal lattice parameter, lattice surface

area, lattice volume, and calculated densities for bulk HCP and FCC-Co are

summarized in Table 2.1.

The HCP→FCC bulk phase transformation is reversible and weakly first

order. It is also accompanied by a small change in enthalpy (∆H) of about 440

9

kJ/mol and in volume (∆V/V) which is approximately equal to 0.0036 (Erbudak et

al., 1997; Ram, 2001). In addition, the FCC→HCP phase transition is martensitic,

resulting in a substantial thermal hysteresis between the heating and cooling

transformations and a considerable change in the stacking sequence.

Fig. 2.1 Representations of the (a) HCP and (b) FCC crystal structures of Co (Callister; 2000) Table 2.1 Lattice Parameters, Lattice Surface Area (A), Lattice Volume (V), and Calculated Densities (ρ) for HCP-Co, FCC-Co, and ε-Co (ICDD Card No. 5-0727 and 15-0806; Dinega and Bawendi, 1991; Sun and Murray, 1999; Ram, 2001)

Crystal Structure

Lattice Parameters,

nm A, 10-2 nm2 V, 10-3 nm3 ρ, g cm-3

HCP a = 0.2507 c = 0.4070 93.90 66.50 8.836

FCC a = 0.3545 75.40 44.55 8.788

ε a = 0.6097 37.17 22.66 8.635

Though thermodynamics predicts that HCP-Co is the stable phase at room

temperatures, both phases can co-exists at ambient temperature in bulk samples.

Furthermore, studies on Co produced by high-temperature crystallization often

yielded mixed phases in one sample, presumably due to the low activation energy of

formation of stacking faults for both HCP and FCC-Co (Dinega and Bawendi, 1999;

Sun and Murray, 1999).

(a) (b)

10

In the case of fine Co particles, recent works on the preparation of Co

nanoparticles by low temperature solution chemistry have similarly produced

mixtures of HCP and FCC-Co, with FCC as the predominant phase as indicated by

the highest peak in the x-ray diffractograms and supported by the electron

diffraction patterns (Toneguzzo et al., 2000; Guo et al., 2002). Additionally, it was

then found that FCC-Co has become more stable at ambient conditions when the

particle size is reduced. Ram (2001) explained this as the influence of the lower

total surface energy of FCC-Co over the HCP phase. From his works, he established

that the crystal structure of Co nanoparticles has a direct correlation with their size,

and due to the lower surface energy and the consequent lower enthalpy of formation

of FCC-Co than HCP-Co, Co nanoparticles are most likely to assume the FCC

crystal structure under conditions of equilibrium to minimize their internal energy.

Hence, he stipulated that within the size range of 10 – 20 nm, the FCC phase is

stable at temperatures below 700 ˚C.

However, there were also evidences that suggest that the stability of FCC-

Co at low temperatures is not solely due to particle size (Sato et al., 1997; Kodama,

1999). In support of this, Sato et al. (1997) annealed large Co nanoparticles of about

200 nm in size above the HCP→FCC transition temperature. Nonetheless, slow

cooling of the annealed nanoparticles did not convert them to HCP-Co, but resulted

to the formation of pure FCC-Co. Lattice defects such as stacking faults,

dislocations, and surface oxidation were identified as possible factors that affect the

crystal structure of Co nanoparticles (Sato et al., 1997; Ram, 2001; Kodama, 1999).

However, Sato et al. (1997) did not explain in detail on how the effects were like.

Then again, a more intensive in-situ structural study is needed in order to

11

qualitatively understand the HCP→FCC phase transformation in Co nanoparticles

and its difference from the bulk phase transition.

Increasing development in the field of solution-phase chemical synthesis of

Co nanoparticles has led to the discovery of a new phase, albeit metastable, in Co

nanoparticles. This phase, designated by Dinega and Bawendi (1999) as epsilon-Co

(ε-Co), has a complex cubic primitive structure, belonging to the P4132 space group

with a unit cell parameter a = 0.6097 ± 0.0001 nm (Dinega and Bawendi, 1999; Sun

and Murray, 1999). Additionally, ε-Co as shown in Fig. 2.2 has 20 Co atoms which

are grouped into two types: (a) 12 Type I atoms and (b) 8 Type II atoms that differ

only in their local coordination in the lattice.

Fig. 2.2 Unit cell of ε-cobalt: (a) unit cell cube filled with eight atoms of Type I (dark) and twelve atoms of Type II (light); (b) the 111-projection of the same cube showing three-fold symmetry along its main diagonal (Dinega and Bawendi, 1999).

Moreover, the ε-Co crystal structure is related to the β-Manganese (β-Mn)

structure, a high-temperature phase of Mn stable between 727 – 1095 ˚C, that also

has 20 atoms and a lattice parameter of 0.63 nm (Sun and Murray, 1999). Table 2.2

gives the diffraction comparison of the β phase of Mn and the 11 nm Co

nanocrystals produced by Sun and Murray (1999) via the reduction of CoCl2

(a) (b)

12

contained in dioctylether solution using superhydride (LiBEt3H) in the presence of

oleic acid (OA) and trialkylphosphine (PR3).

Table 2.2 Comparison of Diffraction Peaks and Relative Intensities of 11 nm Co Nanocrystals and β-Mn (Sun and Murray, 1999)

dCo, Å Rel. Intenstiy dβ-Mn, Å Rel. Intenstiy hkl 2.73 5 2.823 5 210 2.49 4 2.577 5 211 2.16 8 2.231 7 220 2.04 100 2.104 100 221 1.94 56 1.997 60 310 1.85 31 1.904 25 311 1.63 9 1.6872 9 321 1.44 7 1.4874 6 330 1.37 7 1.4115 5 420 1.20 23 1.2377 25 510 1.13 18 1.1721 20 520

Formation of ε-Co is only possible by solution-phase chemistry, particularly

by organometallic route, and generally, using a combination of 2 or more tight

binding ligands or surfactants like OA, PR3, triphenylphosphine ((ph)3P), and

trioctylphosphane oxide (TOPO) (Dinega and Bawendi, 1999). Organometallic

processes, being not thermodynamically controlled, allowed the preparation of

metastable structures like ε-Co (Dinega and Bawendi, 1999; Puntes et al., 2002;

Shukla et al., 2006; Yang et al., 2006; Cheng and Hight Walker, 2007). On the other

hand, the role of these coordinating ligands in the preparation of ε-Co was explained

by Dinega and Bawendi (1999) as responsible for changing the energetics of the

crystal growth by binding tightly around the growing crystal and around the

dissolved Co atoms. In their study, fabrication of Co nanoparticles by thermal

decomposition of dicobalt octacarboynyl, Co2(CO)8, in the absence of TOPO

produced exclusive FCC phase.

13

ε-Co has properties in between the HCP and FCC phases as seen in Table

2.1. Additionally, it is also considered as a soft magnetic material similar to FCC-

Co (Dinega and Bawendi, 1999; Sun and Murray, 1999). The magnetic moment per

atom and magnetocrystalline anisotropy of ε-Co are compared to the values for bulk

HCP-Co and FCC-Co in Table 2.3. The reduced magnetic properties of ε-Co offer

several advantages: (a) it facilitates the stabilization of nanoparticles during

synthesis, (b) it assists in the particle size classification, and (c) it favors the

formation of ordered films for future applications in magnetic recording (Sun and

Murray, 1999).

Table 2.3 Magnetic Moment per Atom and Magnetocrystalline Anisotropy of HCP-Co, FCC-Co, and ε-Co (Cullity, 1972; Shukla et al., 2006).

Crystal Structure Magnetic Moment per Atom, µB

Magnetocrystalline Anisotropy, 106 erg cm-3

HCP 1.72 4.2 FCC 1.75 2.7 ε 1.70 1.5

µB = Bohr Mangeton = 0.927 x 10-20 erg/Oe

Phase transition of ε-Co to the HCP or FCC structures can be easily done by

annealing at the proper temperature. Heating the sample at 300 ˚C quantitatively

converts ε-Co to HCP-Co, while annealing to 500 ˚C completely transforms it to the

FCC structure. In addition, subsequent cooling of the heated samples does not revert

them back to the original ε-Co structure (Petit et al., 2007).

2.3 Formation of Monosized and Uniformly Dispersed Cobalt and Other Metal Nanoparticles

Preparation of monosized and uniformly dispersed metal nanoparticles is

very challenging because of the strong tendency of the nanoparticles to agglomerate

14

in order to reduce their high surface energy (Sugimoto, 2001; Cao, 2004). These

attractive forces, a combination of Brownian motion and attractive van der Waals

forces in nanoparticles, are even greater for Co nanoparticles due to the added

influence of their magnetic property (Yondong, 2003). Thus, various strategies have

been proposed to facilitate the formation of well-dispersed and uniformly sized

metal nanoparticles.

Generally, all monodispersed systems, with those having approximately 5%

or less in the relative standard deviation of the particle size distribution, fulfill the

following conditions: (a) the separation of nucleation and growth stages during

synthesis; (b) if possible, the control of the growth mode; (c) the prevention of

random agglomeration of growing particles; and (d) the reserve of monomers

(Sugimoto, 2001; Cao, 2004).

2.3.1 Separation of Nucleation and Growth

This condition is critical in ensuring that all nuclei are formed during the

nucleation stage and that no further nucleation happens during the subsequent

growth, such that all preformed nuclei will grow at the same time leading to the

formation of monosized nanoparticles (Sugimoto, 2001; Cao, 2004; Cushing et al.,

2004; Burda et al., 2005). This requirement was first expressed by La Mer and co-

workers in their study on sulfur sols in the 1950s (Viau et al., 1996; Watzky and

Finke, 1997; Yondong, 2003). Their proposed mechanism on the formation of

monodispersed colloidal particles in homogeneous solutions is the widely cited La

Mer mechanism.

15

According to the La Mer model, formation of colloidal particles is achieved

by nucleation from a supersaturated homogenous solution followed by the diffusive

growth either by molecular addition or by agglomeration to other solid particles

(Watzky and Finke, 1997; Dong et al., 2001; Yondong, 2003; Burda et al., 2005).

The condition of supersaturation is necessary since the energy barrier to nucleation

can only be surmounted in supersaturated solutions, where the probability of

molecular encounters like those needed for the stepwise sequence of molecular

addition is sufficiently high (Watzky and Finke, 1997).

Fig. 2.3 is the La Mer diagram for the formation of monodispersed colloidal

particles by homogenous nucleation and growth. During the prenucleation period

(Stage I), the concentration of the metal (solute) in the solution increases and

reaches the saturation concentration, which corresponds to the equilibrium solubility

of the solute in the particular solvent. However, nucleation does not take place until

the metal concentration attains the critical supersaturation value (Stage II) above the

equilibrium solubility. This critical value represents the activation energy for

spontaneous nucleation. After the initial nucleation, the solute concentration drops

below the critical supersaturation level (Stage III) and nucleation stops. At the same

time, particle growth proceeds until the metal concentration is again equal to the

equilibrium solubility. Therefore, nucleation and growth are separated producing

uniformly sized colloidal particles. Obviously, a very fast nucleation rate is

desirable for a complete separation of nucleation from the growth stage in order to

generate monodispersed metal particles.

16

Fig. 2.3 La Mer model for the formation of monodispersed systems by controlled homogenous nucleation and growth: I. prenucleation period; II. nucleation step; III. growth step (Redrawn from Viau et al., 1996).

In practice, automatic separation of nucleation and growth is difficult to

accomplish and thus, the solute concentration is normally brought abruptly to a very

high supersaturation and quickly decreased below the critical supersaturation level

so that a sharp nucleation is achieved (Cao, 2004). Such techniques are called

supersaturation quenching. Examples of these processes are reduction of pH in the

forced hydrolysis of metal ions, dilution with the solvent, addition of some growth

accelerators, and a quick change of temperature (Sugimoto, 2001).

The La Mer model for the separation of nucleation and growth has been

widely applied in various syntheses of monodispersed particles in homogeneous

solutions. However, inconsistencies were found in some cases, such as in the

formation of gold sols, where kinetics was found to play a part, which does not agree

with the supersaturation theory (Watzky and Finke, 1997). A more recent

mechanism was proposed by Watzky and Finke (1997) to account for the role of

kinetics in the separation of nucleation and growth. Their works on iridium

nanocluster formation using hydrogen gas as reductant demonstrated that (a)

saturation concentration

nucleation concentration

Time

Con

cent

ratio

n of

met

al a

tom

s in

sol

utio

n I II III

17

nucleation in a solution is accomplished via low-level continuous slow process in a

solution that is far from being saturated and (b) a subsequent autocatalytic surface-

growth step achieves the key separation in time necessary for the synthesis

monodispersed nanoclusters. Furthermore, in contrast to the La Mer mechanism, the

autocatalytic surface growth process is generally not diffusion-limited. Instead, the

rate determining step is the metal incorporation onto the particle surface, which in

turn is dependent on the surface area. Nonetheless, even with the existence of this

new mechanism, the short nucleation time model of La Mer is still more commonly

used to explain the precipitation process for many chemical reactions, including the

formation of monosized nanoparticles from solution.

A more drastic measure to separate growth from nucleation is by seeding,

first used by Zsigmondy for Au colloids (Sugimoto, 2001). In this process, seed

crystals are introduced into a monomer solution under a relatively low

supersaturation below the critical supersaturation level. Thus, nucleation process is

performed in a completely separate system. This route, known as the seed mediated-

method, has also been used to synthesize monometallic and bimetallic (core-shell)

nanoparticles (Qian and Yang, 2005).

A variation of this process was employed by Viau et al. (2001) in the

preparation of CoNi nanopowders in polyol. Very fine nanoparticles of Pd, Pt, or

Ag were formed in situ by the introduction of their salts into the metal solution.

Since reduction of noble metals occurs first, Co and Ni atoms preferentially deposit

on the surface of noble metal nanoparticles resulting to uniformly sized CoNi

nanoparticles. Fig 2.4 (a) shows a TEM image of their Co50Ni50 nanoparticles seeded

18

with Pt, while its x-ray diffraction pattern is in Fig. 2.4 (b). Note that no peak

related to Pt was identified from the x-ray diffractogram indicating that all Pt

nanoparticles were covered by Co and Ni. A more recent work on the use of

heterogeneous nucleation employing Ag and Pt as nucleating agents on the

formation of Co80Ni20 nanoparticles was published by Luna et al. (2004).

Fig. 2.4 (a) TEM image of Co50Ni50 nanoparticles produced by heterogeneous nucleation with Pt in polyol (dm = 6 nm) and its corresponding (b) XRD pattern with increasing mole ratios of Pt to (Co+Ni) showing only peaks of Co and Ni: a. dm = 44 nm; b. dm = 15 nm; c. dm = 5 nm (Viau et al., 2001).

2.3.2 Control of the Growth Process

Control of growth mode is important in the synthesis of monodispersed

nanoparticles since the size distribution of the particles is influenced by the

subsequent growth process. Moreover, the final particle size will also be determined

by the growth step. Nanoparticles, in general, grow by molecular addition through

the stepwise addition of monomeric entities of the solute or by secondary growth

process, wherein agglomeration of primary particles occur which form larger

secondary particles. (Viau et al., 1996; Yondong, 2003; Burda et al., 2005).

Particle growth via molecular addition involves four major steps: (a)

generation of growth species, (b) diffusion of growth species from bulk to the

(a) (b)

2θ (˚)

19

growth surface, (c) adsorption of the growth species onto the growth surface, and (d)

surface growth through irreversible incorporation of the growth species onto the

solid surface (Cao, 2004). The first three steps are diffusion-limited processes,

whereas the last step is a surface-controlled process (Cao, 2004). Therefore, the

particle growth mechanism depends on which among these steps is rate limiting,

either a diffusion-controlled or surface-reaction controlled process.

In a diffusion-limited growth, self sharpening of the size distribution is

expected to occur as long as the supersaturation is high, though less than the critical

supersaturation level to avoid renucleation during particle growth (Sugimoto, 2001).

Fig. 2.5 shows that the radius difference among the nanoparticles decreases with the

increase in particle size and with prolonged growth time in a diffusion-controlled

growth, which supports the formation of uniformly sized nanoparticles (Cao, 2004).

To attain a diffusion-dependent growth, several approaches can be done such as by

keeping the amount of the growth species extremely low resulting to very large

diffusion distance, by increasing the viscosity of the solution as in the case of

ethylene glycol when NaOH is added, and by the introduction of a diffusion barrier

such as protective agents like gelatin and PVP (poly vinyl pyrrolidone).

On the other hand, self sharpening of size distribution does not occur in

surface-reaction controlled growth and only the relative width of the size distribution

against the mean size is narrowed (Sugimoto, 2001). Surface-reaction growth can be

further divided into two mechanisms: mononuclear growth and polynuclear growth

(Cao, 2004). In mononuclear growth, the growth process proceeds layer by layer,

such that the growth species are incorporated onto one layer and continues to

20

another layer only after the previous layer is completed. The difference in the

particle size among the nanoparticles increases with the particle size and growth

time as seen in Fig. 2.5. Therefore, mononuclear growth does not promote the

formation of monodispersed nanoparticles.

Fig. 2.5 Plot illustrating the radius difference (δr/r) among the nanoparticles as a function of (a) particle radius (r) and (b) growth period (t) for diffusion-controlled growth (I), polynuclear growth (II), and mononuclear growth (III) mechanisms (Modified from Cao, 2004).

However, in polynuclear growth, the surface process is extremely rapid that

the second growth layer proceeds before the first layer is completed (Viau, 1996;

Cao, 2004). Then, the relative width of the particle size distribution narrows as long

as the particle growth continues. From Fig. 2.5, it can be noted that the radius

difference among the particles is reduced with the increase in size and with longer

growth period. Nonetheless, a predominant diffusion-controlled growth is desirable

for the formation of monosized nanoparticles.

δr/r δr/r

r t

I

II

III

I

II

III

(a) (b)

21

In contrast, secondary growth by agglomeration should be avoided since it

usually results to a broad size distribution of secondary particles of various shapes.

This growth mode also involves diffusion, but in this case, of the primary particles

towards a growing cluster of nanoparticles, formed by absorption of the nearest

primary particles due to the strong van der Waals forces of large clusters (Sugimoto,

2001). Fig. 2.6 illustrates the model for the nucleation and agglomerative growth in

a homogeneous solution (Capek, 2006). It should be noted however that given

sufficient time for growth, uniformly sized particles can be produced by

agglomeration, though the particles are normally in the submicrometer and

micrometer range.

Fig. 2.6 Scheme depicting the growth mechanism by nucleation and agglomeration (Capek, 2006)

2.3.3 Inhibition of Random Agglomeration of Nanoparticles

Besides prevention of secondary growth of nanoparticles, contact between

particles during growth stage should be avoided since these contact points serve as

active sites for the deposition of the solute (Viau et al., 1996; Sugimoto, 2001; Cao,

2004; Cushing et al., 2004; Burda et al., 2005). Furthermore, once the growth

species deposits on preferentially on these contact points, the particles in contact will

be permanently cemented together. Common measures for the inhibition of particle

agglomeration are by electrostatic repulsion and by steric stabilization.

22

Electrostatic repulsion makes use of the electric double layer, wherein

charged particles exert repulsive forces against each other as function of the zeta

potential and Debye length (Sugimoto, 2001; Cao, 2004). These repulsive forces

results from the chemisorption of charged species at the surfaces of the

nanoparticles. This can be accomplished by diluting the solution or by adjusting the

pH. Even so, electrostatic stabilization is limited by the following facts: (a) it is a

kinetic type of stabilization, thus, it is almost impossible to re-disperse agglomerated

nanoparticles; (b) it is only applicable to dilute systems; (c) it cannot be used on

electrolyte sensitive systems; and (d) it is difficult to apply to multiphase systems,

since different solids develop varying surface charge and electric potential in a given

condition (Cao, 2004).

A more typical measure against particle coagulation is by steric stabilization.

It is a widely applied method than electrostatic stabilization due to issues on the

stability of nanoparticles at very high or low pH (Burda et al., 2005) when

electrostatic stabilization is used and is generally accomplished by the addition of

protective agents like polymers, surfactants, and complexing agents, as adsorptive to

particles (Sugimoto, 2001). Also, protective agents not only prevent agglomeration,

but often act as growth inhibitors and shape controllers as well. Moreover, polymer

layer adsorbed on the surface of the nanoparticles also serves as diffusion barrier to

growth species resulting in a diffusion-limited growth (Cao, 2004). Table 2.4 gives

some of the common protective agents that were used in the synthesis of Co

nanoparticles.

23

Table 2.4 Some Protective Agents Used in the Synthesis of Co Nanoparticles Protective Agents References

Trialkylphosphine, PR3 Sun and Murray, 1999; Yang et al., 2006.

Trioctylphosphane oxide, TOPO Dinega and Bawendi, 1999; Puntes et al., 2001; Shukla et al., 2006; Cheng and Hight Walker, 2007.

Oleic acid, OA

Sun and Murray, 1999; Viau et al., 2001; Puntes et al., 2002; Luna et al., 2003; Shukla et al, 2006; Yang et al, 2006.

Lauric acid, LA Luna et al., 2003.

Sodium dodecylsulfosuccinate, SDS Rivas et al, 1999; Gibot et al., 2005. Didodecyldimethyl ammonium bromide, DDAB Chen et al., 1995; Ram, 2001.

Centrimethyl ammonium bromide, CTAB Gibot et al., 2005

1,2 propanediol Viau et al., 1994; Toneguzzo et al., 2000; Luna et al., 2004.

1,2 hexadecanediol Luna et al., 2003. Alkylphosphines molecules such as PR3 and TOPO due to their steric

bulkiness, in combination with the organic molecules of oleic acid, are generally

employed in the preparation of ε-Co by organometallic method. As mentioned,

these molecules were found to be responsible for changing the energetics of crystal

growth during synthesis in favor of the formation of the less dense ε-Co (Dinega and

Bawendi, 1999). In addition, by changing the mole ratio of these protective agents

in the solution, anisotropic Co nanoparticles can be produced (Puntes et al., 2001;

Shukla et al., 2006). On the other hand, cationic (SDS) and anionic (DDAB and

CTAB) surfactants are normally used in microemulsion and metal salt reduction

techniques to fabricate either a single phase FCC-Co or a mixture of both phases

(Chen et al., 1995; Rivas, 1999). Meanwhile, polyalchohols such as 1,2 propanediol

and 1,2 hexadecanediol can act simultaneously as solvent, protective agent, and

reductant during the fabrication of monodispersed Co nanoparticles by the polyol

24

process (Viau et al., 1994; Toneguzzo et al., 2000; Luna et al., 2004). Synthetic

routes for the preparation of Co nanoparticles such as organometallic route,

microemulsion, metal salt reduction, and polyol method will be described in detail in

the succeeding sections.

2.3.4 Reserve of Monomers

In the formation of monodispersed particles, two conflicting requirements

are present during growth stage: moderate supersaturation to avoid renucleation and

an ample concentration of monomers for particle growth (Cao, 2004). To resolve

this, a monomer reservoir, which regulates the release of metal ions, must be built.

This is accomplished by introducing complexing agents like EDTA, NTA, and citric

acid that shield a large concentration of multivalent metal cations, and thus

decreases the supersaturation of free metal ions in the solution (Sugimoto, 2001). In

addition to preventing concurrent nucleation during particle growth, complexing

agents also inhibits agglomeration of particles during this stage.

2.4 Liquid-Phase Synthesis of Cobalt and Other Magnetic Nanoparticles

Various synthetic routes have been exploited in the preparation of uniformly

sized and monodispersed Co nanoparticles. These methods are generally grouped

into two: physical and chemical methods. Physical methods make use of high-

energy treatment in gaseous and solid states to form magnetic nanoparticles (Gubin

et al, 2005). Examples are laser vaporization, arc discharge, chemical vapor

deposition, and low-energy beam cluster deposition, which utilize special

installations or apparatus during synthesis (Cao, 2004; Burda et al, 2005; Gubin et

al., 2005).