Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald ....

122
Copyright by Cynthia Stowell 2005

Transcript of Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald ....

Page 1: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

Copyright

by

Cynthia Stowell

2005

Page 2: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

The Dissertation Committee for Cynthia Ann Stowell certifies that this is the

approved version of the following dissertation:

Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping

Committee:

Brian A. Korgel, Supervisor

Paul F. Barbara

John G. Ekerdt

Miguel Jose-Yacaman

Allan H. MacDonald

Page 3: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping

by

Cynthia Ann Stowell, B.S.

Dissertation

Presented to the Faculty of the Graduate School of

The University of Texas at Austin

in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

The University of Texas at Austin

May 2005

Page 4: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

iv

Acknowledgements

I would like to thank all Korgel Group members, past and present, but especially

Chris, Lindsay, Michael, Tobias, Rob, Fred and Aaron for technical discussions and

assistance. I thank Ali, Felice, April, Doh, Tripp, Dayne, Hsing-Yu, Preeti and Danielle

for friendship, moral support and occasional comic relief.

TEM has been essential in the research that follows, and I thank Dr. J. P. Zhou,

and Dr. John Mendehall for instruction, advice and equipment upkeep. Additionally, I

would like to thank Dr. John Lansdown for ICP-MS assistance, Dr. Mehdi Moini for

GC/MS training, Glen Baum for access to a ring tensiometer, Steven Sorey for ESR

training, and Ronald Dass in the Goodenough lab and Utkir Mirsaidov in the Markert lab

for SQUID assistance.

I’d like to thank the staff in the Department of Chemical Engineering for their

support over the years in scheduling classes, purchasing chemicals and equipment,

creating reactor cells, fixing computers and providing extra cookies after seminars.

I also thank my committee members, Dr. Ekerdt, Dr. Jose-Yacaman, Dr. Barbara,

and Dr. MacDonald for their advice and suggestions. Tremendous thanks to my advisor,

Dr. Korgel, for ideas, instruction, guidance, and a different perspective.

Lastly, I would like to thank my parents and brother for their love and support,

and Jason for keeping me company during this entire process.

Page 5: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

v

Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping

Publication No._____________

Cynthia Ann Stowell, Ph.D

The University of Texas at Austin, 2005

Supervisor: Brian A. Korgel

Colloidal nanocrystals have many advantages over those synthesized by other

means due to the flexibility not only in synthesis conditions but in post-synthesis

assembly. Three aspects of colloidal nanocrystals that demonstrate this versatility were

studied: the self-assembled patterning of nanocrystals into arrays through the use of fluid

dynamics, the catalytic properties of nanocrystals as a function of ligand type and

reaction cycle, and the doping of III-V semiconductor nanocrystals with magnetic atoms

during colloidal synthesis.

While much work has been done on the thermodynamically-driven formation of

monodisperse or bi-modal nanocrystal superlattices, another option exists for nanocrystal

self-assembly: formations driven by fluid dynamics. Hexagonal networks of gold

nanocrystals were observed after drop casting gold nanocrystals in chloroform on

different substrates. The honeycomb-shaped structures were calculated to be created by

surface tension driven (Marangoni) convection. The honeycomb networks have a lattice

parameter of 4.3 µm and their formation is highly dependant not only particle size and

size distribution but concentration of particles within solution.

Page 6: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

vi

A new synthesis for iridium nanocrystals was developed. The iridium particles

could be synthesized with any of five stabilizing molecules. Taking advantage of this

fact, the effect of capping ligands on the catalysis of 1-decene hydrogenation was studied.

Ligands that stabilized the iridium well prevented hydrogenation while “weak” capping

ligands allowed the iridium nanocrystals to reach turnover rates as high as 270 s-1.

Recycling the catalytic particles also affected the activity, as the turnover frequency

increased with each cycle until the particle began to agglomerate and fall out of solution.

MnxIn1-xAs and MnxIn1-xP nanocrystals ranging from 2 to 10 nm in diameter were

synthesized with up to xMn=0.025 for InMnAs and xMn=0.11 for InMnP. Surface

exchange and magnetic measurements confirmed that much of the dopant resides in the

nanocrystal core and modifies the magnetic properties of the host material through

antiferromagnetic superexchange interactions. The effective Bohr magnetons of Mn in

the synthesized InMnAs nanocrystals ranged from 2.2 to 3.7 atom MnBµ , and from 3.4

to 5.1 atom MnBµ for InMnP, values below the theoretical value of 5.9 atom MnBµ .

This result is attributed to antisite defects and interstitial doping.

Page 7: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

vii

Table of Contents

List of Tables ...........................................................................................................x

List of Figures ........................................................................................................ xi

Chapter 1: Introduction ............................................................................................1 1.1 Colloidal Nanocrystal Background...........................................................1 1.2 Self Organization of Nanocrystals ............................................................2 1.3 Catalysis in Nanocrystal Systems .............................................................3 1.4 Dilute Magnetic Semiconductors..............................................................5 1.5 Dissertation Overview ..............................................................................8 1.6 References.................................................................................................9

Chapter 2: Microscopic Patterning of Nanocrystals through Fluid Dynamics ......13 2.1 Introduction.............................................................................................13 2.2 Experimental ...........................................................................................14

2.2.1 Nanocrystal synthesis..................................................................14 2.2.2 Droplet preparation and evaporation ..........................................15 2.2.3 Characterization techniques ........................................................16

2.3 Results and Conclusions .........................................................................17 2.3.1 Quality of gold nanocrystals .......................................................17 2.3.2 Qualitative patterning as a function of concentration .................18 2.3.3 Qualitative patterning as a function of particle size and distribution

.....................................................................................................21 2.3.4 Topography of nanocrystal patterns............................................22 2.3.5 Ring and polygonal patterning mechanism.................................23 2.3.6 A mechanism for hexagonal array formation: Marangoni convection

.....................................................................................................24 2.3.7 Marangoni Convection in Nanocrystal Solutions .......................28 2.3.8 Competition of Long Wavelength and Short Wavelength Marangoni

Convection ..................................................................................32 2.3.9 Surface Tension Effects of Polydispersity..................................33

Page 8: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

viii

2.3.10 Hexagonal Arrays in Other Nanocrystal Systems ....................34 2.4 Conclusions.............................................................................................35 2.5 References...............................................................................................36

Chapter 3: Iridium Nanocrystal Synthesis and Surface Coating-Dependant Catalytic Activity .........................................................................................................38 3.1 Introduction.............................................................................................38 3.2 Experimental ...........................................................................................39

3.2.1 Nanocrystal synthesis..................................................................39 3.2.2 Catalysis......................................................................................41 3.2.3 Characterization techniques ........................................................42

3.3 Results and Discussion ...........................................................................43 3.3.1 Oleic acid and oleylamine capped Ir nanocrystals......................43 3.3.2 Weaker binding capping ligands.................................................46 3.3.2.1 Nanocrystal quality ..................................................................47 3.3. Turnover frequency calculations...................................................49 3.3.2 Catalytic activity as a function of recycling ...............................50

3.4 Conclusions.............................................................................................53 3.5 References...............................................................................................53

Chapter 4: Dilute Magnetic III-V Semiconductor Nanocrystals: Synthesis and Magnetic Properties ......................................................................................56 4.1 Introduction.............................................................................................56 4.2 Experimental ...........................................................................................58

4.2.1 Nanocrystal synthesis..................................................................58 4.2.2 Characterization techniques ........................................................61

4.3 Results and Discussion ...........................................................................63 4.3.1 Nanocrystal quality .....................................................................63 4.3.2 Doping.........................................................................................68 4.3.3. ESR results.................................................................................73 4.3.4 Absorbance and photoluminescence data ...................................75 4.3.5 Magnetic Measurements .............................................................78

4.4 Conclusions.............................................................................................88

Page 9: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

ix

4.5 References...............................................................................................89

Chapter 5: Conclusions and Recommendations ....................................................93 5.1 Conclusions.............................................................................................93

5.1.1 Fluid patterning through fluid dynamics.....................................93 5.1.2 Catalysis of iridium nanocrystals................................................93 5.1.3. III-V Dilute magnetic semiconductors.......................................94

5.2 Recommendations for Future Work........................................................95 5.2.1 Iridium nanocrystals....................................................................95 5.2.2 Dilute magnetic semiconductor nanocrystals .............................95

5.3 References...............................................................................................96

Bibliography ..........................................................................................................97

Vita.......................................................................................................................106

Page 10: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

x

List of Tables

Table 2.1: Parameters Measured at 20 °C for 3.5 and 5 nm Au Nanocrystals

Dispersed in Chloroform: σ, σT, κ, Ma.............................................30

Table 4.1: In, Mn, As composition of TOP-capped InMnAs nanocrystals determined

by ICP-MS for different reaction precursors of identical concentration

and reaction conditions. ....................................................................68

Table 4.2: Composition of InMnAs particle cores determined by ICP-MS of non-

ligand exchanged, size selected nanoparticles from the same batch. The

smaller nanocrystals contain proportionally more Mn. ....................73

Table 4.3: Effective Bohr magneton number p , and the nearest neighbor exchange

integral nnJ for InMnAs found by fitting Equation (4.1) to Figure 4.10

(A). ....................................................................................................82

Table 4.4: Effective Bohr magneton number p , and the nearest neighbor exchange

integral nnJ for InMnP found by fitting Equation (4.1) to figure 4.11

(A). ....................................................................................................85

Page 11: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xi

List of Figures

Figure 1.1: Schematic of a hydrogenation reaction on the surface of a nanocrystal.

Hydrogen is adsorbed to the nanocrystal surface. The Pi bond in an

alkene reacts with the metal core. Hydrogen-carbon bonds are formed

and then molecule desorbs from the surface.......................................4

Figure 1.2: Schematic of a compound dilute magnetic semiconductor, magnetic spins

indicated by arrows. (A) Without carrier mediated interactions, the

magnetic dopants do not align ferromagnetically, and (B) with carrier

mediated exchange, the magnetic dopants align antiferromagnetically

with carriers throughout the crystal, producing a net ferromagnetic

alignment of the spins of the dopant atoms. .......................................6

Figure 2.1: TEM images of (A) 3.5 nm diameter gold nanocrystals. Inset show

crystalline nature of particles, with lattice fringes present. (B) 5 nm

diameter gold nanocrystals. (C) 3 – 12.5 nm polydisperse, gold

nanocrystals.......................................................................................18

Figure 2.2: TEM images of microscopic structures formed by 3.5 nm diameter

sterically stabilized Au nanocrystals drop-cast on a carbon substrate

from a chloroform dispersion. (A) Au nanocrystals at 1.67 g/L deposit

as rings. (B) At 0.277 g/L, the number of rings decreases. (C) At 9.26 ×

10-2, rings no longer form; instead, polygonal networks become the

preferred morphology. (D) At 1.25 × 10-2 g/L, ordered repeating

hexagonal networks resembling honeycombs dominate the structural

morphology. (E) Further dilutions (1.25 × 10-3) collapse the order.20

Page 12: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xii

Figure 2.3: TEM images of (A) monodisperse 3.5 nm , (B) monodisperse 5 nm, and

(C) polydisperse 3 to 12.5 nm Au nanocrystals deposited on a carbon

substrate from chloroform at a concentration of 1.25 × 10-2 g/L......22

Figure 2.4: AFM images of (A) hexagonal networks and (B) dewetting rings of Au

nanocrystals. The average height of the hexagonal borders is 15 nm, and

the dewetting rings vary from 25 to 100 nm.....................................23

Figure 2.5: TEM image of dewetting rings within the honeycomb network.......25

Figure 2.6: Illustration of surface tension driven convection. The substrate

temperature, Ts, is higher than T, the temperature of the free interface,

creating a vertical temperature gradient. When Ma>Mac, the film

becomes unstable due to thermal fluctuations on the surface, and

convection occurs. Surface tension increases with decreasing

temperature, causing the surface to spread at regions of higher

temperature and contract at regions of lower temperature. The resulting

surface flow forces warmer liquid near the substrate to rise and replace

the spreading fluid, which leads to vertical convective flow. Flow is

further enhanced because the rising fluid is warmer than the average gas-

solvent interfacial temperature. Convection reaches steady state as

viscous forces dissipate the surface tension induced energy gradient.27

Figure 2.7: Illustration of gold nanocrystal/ chloroform system. ........................31

Figure 2.8: SEM image of a hexagonal network of 3-7 nm sterically stabilized nickel

nanocrystals.......................................................................................35

Page 13: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xiii

Figure 3.1: Oleic acid/oleylamine-coated Ir nanocrystals. (A) TEM image of ~4 nm

diameter size-selected Ir nanocrystals, (B) XRD Ir nanocrystals, peaks

correspond to fcc Ir (JCPDS file 46-1044), (C) SAXS of two different

sizes of Ir nanocrystals isolated by size-selective precipitation from the

same reaction: (□) Ravg= 1.09 nm, σ= ±0.283 nm (○) Ravg= 1.92 nm, σ=

±0.635 nm; the curves correspond to the best fits of Eqn (1) to the data

that were used to obtain Ravg and σ, (D) HRTEM of a 5 nm oleic

acid/oleylamine coated Ir nanocrystal. .............................................44

Figure 3.2: Conversion of 1-decene to decane at 75oC and 3 psig H2 in the presence

of dispersed Ir nanocrystals: (◊) 2 nm oleic acid/oleyl amine coated; (×)

4 nm oleic acid/oleyl amine coated; (+) TOP-coated; (○) 1.5 nm TOAB-

coated particles, TOF = 5 s-1; (□) 5 nm TOPB-coated, TOF= 270 s-1.46

Figure 3.3: TEM images of Ir nanocrystals synthesized with different capping

ligands: (A) TOAB (1.5~3 nm diameter); (B) TOPB (2~5 nm diameter);

and (C) TOP (10~100 nm diameter). ................................................48

Figure 3.4: TOAB-coated Ir nanocrystals after 1-decene hydrogenation reactions (A)

Hydrogenation of 1-decene to decane as a function of catalyst recycling:

(○) First reaction, TOF = 5 s-1; (□) Second Reaction, TOF = 14 s-1; (◊)

Third Reaction, TOF = 50 s-1; (×) Fourth Reaction, TOF = 124 s-1; (∆)

Fifth Reaction, TOF = 38 s-1; (B) TEM images of TOAB-coated Ir

nanocrystals before catalysis; and (C) after four hydrogenation reactions.

...........................................................................................................51

Page 14: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xiv

Figure 4.1: TEM images of (A) 4.5 nm TOP capped InAs nanocrystals and (C) 4.5

nm TOP capped Mn0.01In0.99As nanocrystals. (B) SAXS data of InAs

nanocrystals dispersed in cyclohexane: (□) dp= 4.4 nm, σ=±0.18; (×)

dp=4.2 nm, σ=±0.14; (○) dp=3.6 nm, σ=±0.20. (D) SAXS data for

Mn0.003In0.997As nanocrystals in cyclohexane: (□) dp=3.8 nm,

σ=±0.15; (×) dp=3.7 nm, σ=±0.17; (○) dp=3.76 nm, σ=±0.11; (+)

dp=3.64 nm, σ=±0.13; (∆) dp=3.46 nm, σ=±0.14. Size histograms

determined by TEM agree with those found from SAXS to within ±13%.

(E) HRSEM image of 4 nm diameter Mn0.014In0.986As nanoparticles

on a glassy carbon substrate..............................................................65

Figure 4.2: TEM images of (A) 5 nm InP, (B) and 5 nm Mn0.02In0.98P, and (C) XRD

of 5 nm Mn0.02In0.98P nanocrystals. The diffraction peaks are indexed to

the zinc blend structure of InP (JCPDS file 00-032-0452), and the

addition of Mn has not noticeably affected the peak location, although

the broadening of the peaks due to the nanocrystal small size may

obscure any dopant effects................................................................67

Figure 4.3: TEM of 5 nm InMnP particles after ligand exchange. The ligand

exchange reduces the size of the particles slightly, but does not affect the

crystallinity of the particles...............................................................69

Figure 4.4: Absorbance measurement for size selected InAs nanoparticles for

different durations of pyridine ligand exchange. 0 hours corresponds to

nanoparticles with the trioctylphosphine ligand. ..............................71

Figure 4.5: Percent composition of manganese in size selected, InMnAs

nanoparticles as a function of pyridine ligand exchange duration....72

Page 15: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xv

Figure 4.6: Electron Spin Resonance (ESR) spectra measured at 115K and 9.42

GHz: (A) dp=5 nm, TOP capped InAs nanocrystals in cyclohexane; (B)

dp=5 nm, MnxIn1-xAs (x=0.01) TOP capped nanocrystals in cyclohexane;

(C) dp=5 nm, pyridine capped MnxIn1-xAs (x=0.024) nanocrystals in

powder...............................................................................................74

Figure 4.7: Additional ESR spectra of dp=5 nm, pyridine capped MnxIn1-xAs

(x=0.024) nanocrystals in powder form, extended to less than 1000G.

...........................................................................................................75

Figure 4.8: Room temperature photoluminescence emission (PL) (λ exc= 790 nm) and

absorbance spectra for nanocrystals dispersed in cyclohexane: (A) InAs

and (B) Mn0.02In0.98As.......................................................................77

Figure 4.9: Room temperature photoluminescence emission (PL, solid line) (λ exc=

550 nm) and absorbance spectra (dashed line) for Mn0.02In0.98P

nanocrystals dispersed in chloroform. ..............................................78

Figure 4.10: Magnetization measurements of MnxIn1-xAs nanocrystals. (A)

Temperature dependent magnetic molar susceptibility measured under a

magnetic field of 5000 Oe, the solid lines are the predicted values as

calculated from equation (4.1) in the text and (B) magnetization versus

applied magnetic field measured at 5K: (×) dp=4 nm, x=0.024 (after

pyridine ligand exchange); (□) dp=4 nm, x=0.048; (+) dp=4.6 nm,

x=0.014 (after pyridine ligand exchange); (○) dp=4 nm, x=0.02; ( )

dp=4 nm InAs measured at 15 K. .....................................................80

Page 16: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

xvi

Figure 4.11: Magnetization measurements of MnxIn1-xP nanocrystals. (A)

Temperature dependent magnetic molar susceptibility measured under a

magnetic field of 5000 Oe, the solid lines are the predicted values as

calculated from equation (4.1) in the text and (B) magnetization versus

applied magnetic field measured at 5K: (□) dp=3 nm, x=0.15; (+) dp=5

nm, x=0.08 (after pyridine ligand exchange); (○) dp=3 nm, x=0.11 (after

pyridine ligand exchange); (◊) dp=4 nm InP.....................................84

Figure 4.12: Magnetization measurements of Mn0.02In0.98P nanocrystals on mica

before and after annealing. (A) Temperature dependent magnetic molar

susceptibility measured under a magnetic field of 5000 Oe, the solid line

is the predicted value as calculated from equation (4.1) and (B)

magnetization versus applied magnetic field measured at 5K: (□) dp=5

nm, x=0.2 before anneal; (○) same particles after 650 °C anneal. ...87

Page 17: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

1

Chapter 1: Introduction

Nanocrystals are just that: crystals with dimensions on the order of a nanometer

(10-9 meters). These materials, generally ranging between 1 and 10 nanometers in size,

straddle the line between bulk crystals and individual molecules, and as a result have

unique and often size tunable optical, electronic and magnetic properties.1

While several methods exist for nanocrystal synthesis including laser ablation2

and chemical vapor deposition,3 colloidal synthesis offers a large amount of flexibility

with respect to synthesis conditions, precursor chemistry and post-synthesis assembly.

Because of this versatility, colloidal nanocrystals pose diverse challenges. Three aspects

of colloidal nanocrystals are covered in this dissertation: the self-assembled patterning of

nanocrystals into arrays directed by fluid dynamics, the catalytic properties of iridium

nanocrystals as a function of ligand type and recycling, and Mn- doped group III-V

semiconductor nanocrystals.

1.1 COLLOIDAL NANOCRYSTAL BACKGROUND

Attractive van der Waals forces between nanocrystal cores drive particle

agglomeration and precipitation from solutions. Two methods to create repulsion

between the particles stabilize the nanocrystals – steric and electrostatic stabilization.

Although the electric double layer created in electrostatic colloids can create significant

interparticle repulsion4- enough to keep particles synthesized in the 19th century still in

solution today5- the work in this dissertation focuses on sterically stabilized nanocrystals.

Sterically stabilized nanocrystals remain in solution because capping ligands

surrounding the particle core prevent particle agglomeration. Besides stabilizing the

nanocrystals, these molecules control the solubility of the particles within solvents, the

Page 18: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

2

interparticle spacing in evaporated thin films of nanocrystals, and the growth of the

particles.6

Many nanocrystal materials have been synthesized through colloidal chemistry,

from metals such as gold, silver7 and platinum8 to semiconductors such as silicon,9

indium arsenide,10 and cadmium sulfide.11 The stabilizing ligand can also be adjusted,

either through different synthesis techniques or a post-synthesis ligand exchange, to make

the nanocrystals soluble in a desired solvent.

1.2 SELF ORGANIZATION OF NANOCRYSTALS

If a system is capable of self-organization, templates and lithographic processes

are not needed to produce elaborate patterns. If nanocrystals can assume patterns through

self-assembly, less processing, and therefore money, is required to produce devices. In

nature, complicated self-organization occurs routinely, for example the stripes and spots

found on animal fur can be formed through the interactions in reaction–diffusion

systems.12 At the time of the publication of the material in Chapter 2, most nanocrystal

ordering studied was thermodynamically-driven.

Monodisperse nanocrystals organize into hexagonal close-packed superlattices

because that formation is thermodynamically favored.13-15 The interparticle spacing in

these systems is tunable by adjusting the stabilizing ligand length.16 Bi-modal size

distributions form more complicated patterns, such as AB13 and AB2 lattices.17-20

Since colloidal nanocrystals are typically transferred to substrates in solution,

fluid dynamics is also an important source of self-organization, and competes with

thermodynamically-favored order. Ohara et al. demonstrated that ring structures formed

from nanocrystals are a result of solvent dewetting.21,22 As a thin film of solvent

evaporates, a hole opens up in the solvent, pulling nanocrystals along the expanding

Page 19: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

3

boundary until particle-substrate interactions become too large and the particles are

locked into a ring formation.

Chapter 2 describes the formation of hexagonal cellular arrays of nanocrystals

through another phenomena in fluid dynamics, surface tension driven (or Marangoni)

convection. Since this work in 2001, others have studied the effect of fluid dynamics on

nanocrystal patterning.

Moriarty et al. have studied cellular networks formed by froths created by

particles in solution.23 The influence of surfactants on evaporating droplets and the

resulting deposition patterns produced were studied by Truskett et al.,24 while Narayanan

et al. looked at the self-assembly of nanocrystals during droplet evaporation in situ.25

Rabani et al. modeled the self-assembly of nanocrystals as a function of thermodynamics

and the fluid dynamics of the evaporating droplet, predicting different formations than

those caused from just thermodynamic interactions alone.26 Finally, inverse opal

nanocrystal superlattice films formed as a result of condensed water droplets templating

the nanocrystals around them.27 This technique was perfected so that long range order of

the inverse opal film was possible.28

1.3 CATALYSIS IN NANOCRYSTAL SYSTEMS

Nanocrystal catalysts have many advantages over larger scale catalysts due to an

increased surface-to-volume ratios,29-33 but in some cases take on completely different

catalytic activity at reduced sizes.34 For example, gold nanocrystals are catalytically

active as 3 nm diameter nanocrystals while bulk gold is unreactive.34 The reason for this

change in activity may be due to electron density changes that accompany diminished

size, support effects, or shape effects, as there are more edge and corner atoms present in

nanocrystals.35

Page 20: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

4

Figure 1.1 is a schematic of a hydrogenation reaction occurring at the surface of a

nanocrystal. While in this figure the adsorbtion of molecules to the particle surface is

straightforward, new concerns arise with the addition of stabilizing ligands to the surface

of the nanocrystals.

Figure 1.1: Schematic of a hydrogenation reaction on the surface of a nanocrystal. Hydrogen is adsorbed to the nanocrystal surface. The Pi bond in an alkene reacts with the metal core. Hydrogen-carbon bonds are formed and then molecule desorbs from the surface.

The presence of a ligand bound to the nanocrystal core will affect reaction rates

due to the decreased diffusion of reactants to the particle surface and a change in surface

state of the catalyst caused by the ligand bonds themselves.35 The effect of stabilizing

ligands on the reactivity of colloidal nanocrystals has not been extensively studied with

two exceptions. Li et al. noted that good capping ligand stabilization appears to diminish

catalytic nanocrystal activity in some cases,36 and Narayanan et al. explored the effects of

Ostwalt ripening and precipitation on catalytic activity.37

In Chapter 3, a new synthesis for iridium nanocrystals was developed and the

catalytic activity of a hydrogenation reaction was measured as a function of the type of

stabilizing ligand. The activity was found to be dependant of the number of times the

particles were recycled. This change in activity over time increases the challenge of

HH-H

C=C H H

H H

HHC-C

HHH H

C=CHH

H H

H H H C-C

H H H H

Page 21: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

5

finding suitable catalyst-ligand systems for specific reactions and puts a twist on the goal

of using specific capping ligands as a method to tune the selectivity of catalyst.35

1.4 DILUTE MAGNETIC SEMICONDUCTORS

Dilute magnetic semiconductors (DMS) are a relatively new class of materials

that have garnered attention and captured vision only within the past four decades,38 but

already over a thousand papers have been published on the topic.

Simply put, a DMS is a semiconductor in which a small amount of magnetic

impurity, typically a transition metal, is a substitional dopant within the crystal lattice.

This presence leads to the exchange interaction of electrons in the metal’s d orbital and

the sp band electrons of the host lattice. The addition of this magnetic impurity has

several effects on the material: it may affect the lattice constant and band parameters of

the host crystal; in II-VI semiconductors it may improve the efficiency of

electroluminescence; magnetic doping causes large Zeeman splitting which leads to

optical effects like a giant Faraday rotation; and finally, the dopant atoms may interact

through a carrier mediated superexchange, leading to ferromagnetism.39-41 This last

characteristic has interested many researchers because of the potential use of

ferromagnetic DMS’s for spintronic applications,42 devices which take advantage of both

the charge and spin of an electron, a spin light–emitting diode being an example.43

Figure 1.2 is a schematic of the interactions crucial to the ferromagnetic behavior

of a DMS. A transition metal is substituionally doped into a host lattice. Having the

magnetic dopant distributed throughout the host crystal is not enough alone to produce

ferromagnetism; without a carrier to mediate interactions, the localized magnetic

moments interact antiferromagnetically.44 If enough of an appropriate carrier type is

present, the spin of the magnetic ions interacts antiferromagnetically with the carrier,

eventually aligning the magnetic moment of all ions. The carrier type that is needed for

Page 22: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

6

ferromagnetism is dependent on electronic structure of the host lattice, and can be

predicted along with doping concentration required.45

Figure 1.2: Schematic of a compound dilute magnetic semiconductor, magnetic spins indicated by arrows. (A) Without carrier mediated interactions, the magnetic dopants do not align ferromagnetically, and (B) with carrier mediated exchange, the magnetic dopants align antiferromagnetically with carriers throughout the crystal, producing a net ferromagnetic alignment of the spins of the dopant atoms.

A frequently used magnetic dopant is manganese because it has a large magnetic

moment due to its five unpaired d electrons. The first ferromagnetic DMS systems

synthesized were based on III-V semiconductors because with the substitution of Mn2+

for the cation, a hole was produced that could mediate the ferromagnetic interactions.40

The dopant must be homogeneously distributed throughout the crystal in

ferromagnetic systems, since phase separation often leads to ferromagnetic clusters.46

Only fairly recently have thin films generated by molecular beam epitaxy been able to

produce III-V DMS where sufficient dopant is present to produce ferromagnetism

without phase separation.40 Since this development, Ohno et al. have shown that in

InMnAs heterostructures the critical temperature at which a material transitions from

No Carrier Interactions Carrier Super-Exchange

= Semiconductor Lattice Atomsand = Dopant Atom = Carrier

A B

Page 23: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

7

ferromagnetic to paramagnetic may be tuned through an externally applied electric

field.47

Another recent breakthrough is the discovery that ferromagnetism can occur in

systems besides III-V semiconductors and in some cases have critical temperature above

room temperature. ZnO,48 ZnTe,49 half-metallic oxides,50 Heusler alloys,51 and even

germanium52 have all been doped with magnetic impurities to produce ferromagnetic

DMS’s.

Nanocrystals of DMS systems have more challenges to overcome than in the thin

film case. If each nanocrystal is to contain a few dopant atoms, a doping concentration

larger than the thin film case is required, and this may cause phase segregation or the

migration of impurity elements to the nanocrystal surface.53 Another concern specific to

magnetic nanocrystals is superparamagnetism. When the dimension of a nanocrystal

becomes smaller than the magnetic domain size, thermal energy perturbs the alignment of

the nanocrystal’s magnetic moment.35 This results in the existence of a blocking

temperature, above which the magnetization decreases, even though the material may still

be below the critical temperature. Due to these factors, no ferromagnetic III-V DMS

nanocrystalline systems have been synthesized, while ferromagnetic II-VI DMS

nanocrystals have been the result of extensive amounts of co-doping or annealing into

thin films.54-56

The specific instance of the Mn- doping of III-V semiconductor nanocrystals is

discussed in Chapter 4, but the results may be applied in a more general fashion to other

doping challenges in nanotechnology.

Page 24: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

8

1.5 DISSERTATION OVERVIEW

This dissertation focuses on three aspects of colloidal nanocrystals: nanocrystal

self-assembly driven by fluid dynamics, the catalytic properties of nanocrystals in

solution, and doping of semiconductor nanocrystals during synthesis.

Chapter 2 focuses on nanocrystal self-assembly driven by means other than

thermodynamics. Honeycomb structures with a lattice parameter of 4.3 µm were

produced by drop casting gold nanocrystal/chloroform solutions on different substrates.

The concentration of nanoparticles and the particle size distributions were varied to

determine what factors perturb the formation of the hexagonal arrays, the eventual

conclusion being that this pattern is formed through surface tension driven (Marangoni)

convection.

Chapter 3 discusses a new synthesis for iridium nanocrystals that is suitable for

use with any of five different capping ligand groups. The iridium nanocrystal diameter

and size distribution varied with capping ligand. The catalytic activity of the

nanocrystals was tested as a function of capping ligand present. The effects of catalyst

recycling on catalytic activity was also studied. Results show that loosely bound

stabilizing ligands promoted hydrogenation, while tightly bound ligands opposed it, and

that catalytic activity is extremely sensitive to recycling.

Chapter 4 deals with III-V dilute magnetic semiconductors nanocrystals. InMnP

and InMnAs were synthesized with Mn concentrations within the semiconductor core as

large as 2.5% for InMnAs and 5.7% for InMnP. In addition to standard characterization

techniques, the particles were subjected to magnetic measurements to determine if the

Mn in the particles aligns ferromagnetically. In both cases, the Mn interacted

antiferromagnetically, due to antisite defects or interstitial instead of substitutional

doping.

Page 25: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

9

Chapter five contains a summary of the results and lists recommendations for

future study in these areas.

1.6 REFERENCES

(1) Brus, L. E. J. Chem. Phys. 1983, 79, 5566.

(2) Geohegan, D. B.; A.A., P.; Duscher, G.; Pennycook, S. J. Appl. Phys. Lett. 1998, 72, 2987-2989.

(3) Kruis, F. E.; Fissan, H.; Peled, A. J. AEROSOL SCIENCE 1998, 29, 511-535.

(4) Israelachvili, J. Intermolecular & Surface Forces; 2nd ed.; Acadmeic Press: San Diego, 1992.

(5) Faraday, M. Phil. Trans. Roy. Soc. 1857, 147, 145.

(6) Battaglia, D.; Peng, X. G. Nano Lett. 2002, 2, 1027-1030.

(7) Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D. J.; Whyman, R. J. J. Chem. Soc. Chem. Commun. 1994, 801-802.

(8) Ahmadi, T. S.; Wang, Z. L.; Henglein, A.; ElSayed, M. A. Chem. Mater. 1996, 8, 1161.

(9) Heinrich, J. L.; Curtis, C. L.; Credo, G. M.; Kavanagh, K. L.; Sailor, M. J. Science 1992, 255, 66-68.

(10) Guzelian, A. A.; Banin, U.; Kadavanich, A. V.; Peng, X.; Alivisatos, A. P. Appl. Phys. Lett. 1996, 69, 1432-1434.

(11) Sondi, I.; Siiman, O.; Koester, S.; Matijevic, E. Langmuir 2000, 16, 3107-3118.

(12) Stewart, I. Life's other secret - The new mathematics of the living world; John Wiley & Sons, Inc: New York, 1998.

(13) Saunders, J. V. Philos. Mag. A 1980, 42, 705.

(14) Alder, B. J.; Wainwright, T. E. Phys. Rev. 1962, 127, 459-361.

(15) Pusey, P. N.; VanMegen, W. Nature 1986, 320, 340-342.

(16) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Annu. Rev. Mater. Sci. 2000, 30, 545-610.

Page 26: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

10

(17) Saunders, J. V.; Murray, M. J. Nature 1978, 275, 201-203.

(18) Bartlett, P.; Ottweill, R. H.; Pusey, P. N. Phys. Rev. Lett. 1992, 68, 3801-3804.

(19) Kiely, C. J.; Fink, J.; Brust, M.; Bethell, D.; Schiffrin, D. J. Nature 1998, 396, 444-446.

(20) Kiely, C. J.; Fink, J.; Zheng, J. G.; Brust, M.; Bethell, D.; Schiffrin, D. J. Adv. Mater. 2000, 12, 640.

(21) Ohara, P. C.; Heath, J. R.; Gelbart, W. M. Angew. Chem. Int. Ed. Engl. 1997, 36, 1077.

(22) Ohara, P. C.; Gelbart, W. M. Langmuir 1998, 14, 3418-3424.

(23) Moriarty, P.; Taylor, M. D. R.; Brust, M. Phys. Rev. Lett. 2002, 89, 248303-248301 - 248303-248304.

(24) Truskett, V. H.; Stebe, K. Langmuir 2003, 19, 8271-8279.

(25) Narayanan, S.; Wang, J.; Lin, X. M. Phys. Rev. Lett. 2004, 93, 135503.

(26) Rabani, E.; Reichman, D. R.; Geissler, P. L.; Brus, L. E. Nature 2003, 426, 271-274.

(27) Shah, P. S.; Sigman, M. B.; Stowell, C. A.; Lim, K. T.; Johnston, K. P.; Korgel, B. A. Adv. Mater. 2003, 15, 971-974.

(28) Saunders, A. E.; Shah, P. S.; Sigman, M. B.; Hanrath, T.; Hwang, H. S.; Lim, K. T.; Johnston, K. P.; Korgel, B. A. Nano Lett. 2004, 4, 1943-1948.

(29) Kralik, M.; Biffis, A. J. Mol. Catal. A: Chem. 2001, 177, 113-138.

(30) Lewis, L. N. Chem. Rev. 1993, 93, 2693-2730.

(31) Aiken_III, J. D.; Finke, R. G. J. Mol. Catal. A: Chem. 1999, 145, 1-44.

(32) Bonnemann, H.; Brijoux, W.; Brinkmann, R.; Dinjus, E.; Joussen, T.; Korall, B. Angew. Chem. Int. Ed. Engl. 1991, 30, 1312-1314.

(33) Roucoux, A.; Schulz, J.; Patin, H. Chem. Rev. 2002, 102, 3757-3778.

(34) Valden, M.; Lai, X.; Goodman, D. W. Science 1998, 281, 1647-1650.

(35) Klabunde, K., Ed.; John Wiley and Sons, Inc.: New York, 2001.

Page 27: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

11

(36) Li, Y.; El-Sayed, M. A. J. Phys. Chem. B 2001, 105, 8938-8943.

(37) Narayanan, R.; El-Sayed, M. A. J. Am. Chem. Soc. 2003, 125, 8340-8347.

(38) Galazka, R. R. In 14th International Conference on the Physics of Semiconductors; Wilson, B. L. H., Ed.; Inst. of Phys. Conf. Series: Edinburgh, 1978; Vol. 43, p 133.

(39) Furdyna, J. K. J. Appl. Phys. 1988, 64, R29-R64.

(40) Ohno, H.; Munekata, H.; Penney, T.; von Molnar, S.; Chang, L. L. Phys. Rev. Lett. 1992, 68, 2664-2667.

(41) Sugano, S.; Kojima, N., Eds. Magneto-optics; Springer: New York, 2000.

(42) Chambers, S. A.; Yoo, Y. K. MRS Bulletin 2003, 28, 706-707.

(43) Jonker, B. T.; Erwin, S. C.; Petrouw, A.; Petukhov, A. G. MRS Bulletin 2003, 28, 740 - 748.

(44) Dietl, T.; Ohno, H. MRS Bulletin 2003, 28, 714 - 719.

(45) Sato, K.; Katayama-Yoshida, H. Phys. Stat. Sol. B 2002, 229, 673 - 680.

(46) Munekata, H.; von Molnar, S.; Segmuller, A.; Chang, L. L.; Esaki, L. Phys. Rev. Lett. 1989, 63, 1849.

(47) Ohno, H.; Chiba, D.; Matsukura, F.; Omiya, T.; Abe, E.; Dietl, T.; Ohno, Y.; Ohtani, K. Nature 2000, 408, 944-946.

(48) Sharma, P.; Gupta, A.; Rao, K. V.; Owens, F. J.; Sharma, R.; Ahuja, R.; Osorio, J. M.; Johansson, B.; Gehring, G. A. Nat. Mater. 2003, 2, 673-677.

(49) Ferrand, D.; Cibert, J.; Wasiela, A.; Bourgognon, C.; Tatrenko, S.; Fishman, G.; Andrearczak, T.; Jaroszynski, S.; Kolesnik; Dietl, T. Phys. Rev. B 2001, 63, 085201:085201-085213.

(50) Coey, J. M. D.; Chien, C. L. MRS Bulletin 2003, 28, 720 - 724.

(51) Palstrøm, C. MRS Bulletin 2003, 28, 725-728.

(52) Park, Y. D.; Hanbicki, A. T.; Erwin, S. C.; Hellberg, C. S.; Sullivan, J. M.; Mattson, J. E.; Ambrose, T. F.; Wilson, A.; Spanos, G.; Jonker, S. T. Science 2002, 295, 651-654.

Page 28: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

12

(53) Mikulec, F. V.; Kuno, M.; Bennati, M.; Hall, D. A.; Griffin, R. G.; Bawendi, M. G. J. Am. Chem. Soc. 2000, 122, 2532-2540.

(54) Norberg, N. S.; Kittilstved, K. R.; Amonette, J. E.; Kukkadapu, R. K.; Schwartz, D. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 126, 9387-9398.

(55) Bryan, J. D.; Heald, S. M.; Chambers, S. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 127, 11640-11647.

(56) Norton, D. P.; Pearton, S. J.; Hebard, A. F.; Theodoropoulou, N. A.; Boatner, L. A.; Wilson, R. G. Appl. Phys. Lett. 2003, 82, 239 - 241.

Page 29: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

13

Chapter 2: Microscopic Patterning of Nanocrystals through Fluid Dynamics*

2.1 INTRODUCTION

In many instances, thermodynamics controls the structural organization of

mesoscopic colloid-based materials. For example, size-monodisperse hard sphere

colloids pack into face-centered cubic (fcc) superlattices;1-3 charge-stabilized colloids and

sterically-stabilized nanocrystals can organize into fcc or bcc (body-centered cubic)

lattices, depending on the range of repulsive interaction between particles;4-8 and size-

matched colloids with bimodal size distributions can form ordered phases, such as AB13

and AB2 lattices.9-12 The dynamics of the colloidal organization process, however, can

also determine the structure of colloidal assemblies. Illustrative examples include the

convective self-assembly of microscopically ordered colloidal arrays with macroscopic

planar, spherical, ellipsoidal, and even doughnut shape.13,14 Constraints imposed on the

mobility of hard-sphere nanocrystals on a surface during monolayer deposition can

frustrate equilibrium phase behavior: For example, disordered monolayers occur for zero

surface diffusion (a process known as random sequential adsorption), hexatic phases form

with limited surface diffusion, and the equilibrium phase—hexagonal arrays—form when

the surface mobility is high.15,16 Ohara, Heath and Gelbart have demonstrated the

formation of organized rings of nanocrystals due to solvent dewetting effects during

solvent evaporation.17,18 More recently, Pileni and coworkers demonstrated the formation

of a variety of microstructures, including fingering patterns and polygonal networks of

nanocrystals and rings, which were attributed to thermocapillary flow during

evaporation.19,20 In this chapter is described observations of self-assembled hexagonally

* The contents of this chapter appeared in Nano Lett. 2001, 1, 595-600.

Page 30: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

14

organized networks of gold nanocrystals (See Figures 2.2 (D) and 2.3 (A)). The

hexagonal networks consist of ribbons of nanocrystals approximately 15 nm thick and

500 nm wide; the honeycomb network has a lattice parameter L of ~4.3 µm.

2.2 EXPERIMENTAL

2.2.1 Nanocrystal synthesis

All chemical were bought and used as-is from Aldrich Chemical Co, and all water

was doubly distilled and deionized.

2.2.1.1 Gold nanocrystal synthesis

Samples of relatively size-monodisperse 3.5 nm, Figure 2.1 (A), and 5 nm, Figure

2.1 (B), diameter nanocrystals, and a polydisperse sample with sizes ranging from 3 to

12.5 nm diameters, Figure 2.1 (C), were studied.

3.5 nm diameter, dodecanethiol-stabilized (C12H25SH) gold nanocrystals were

synthesized at room temperature using a modified version of the procedure developed by

Brust, et. al.21 18 mL of an aqueous solution of 30 mM aqueous solution of hydrogen

tetrachloroaurate(III) hydrate (HAuCl4xH2O) was added to 12.5 mL of 0.2 M

tetraoctylammonium bromide ([CH3(CH2)7]4NBr) and stirred for one hour. The organic

phase containing the gold ions was separated from the aqueous phase, and 120 µL (1

mmol) of dodecanethiol was added. After stirring for 5 minutes, the gold ions were

reduced by 15 mL of a 0.44 M aqueous solution of sodium borohydride (NaBH4). This

solution was stirred for 12 hours, and the organic phase containing gold nanocrystals was

collected. The nanocrystals were washed in ethanol and centrifuged at 10,000 RPM for

three minutes to precipitate the particles and remove excess reactants. The nanocrystals

were redispersed in chloroform and centrifuged again to remove poorly capped particles.

Page 31: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

15

Two more ethanol and chloroform cleaning cycles were completed to insure the purity

and quality of the nanocrystals.

5 nm diameter gold particles were synthesized using the same reaction procedure,

except that the reducing agent was added to the reaction flask before the addition of

dodecanethiol.

3 – 12.5 nm polydisperse gold nanocrystals were synthesized by taking a solution

of 5×10-2 g/L of 3.5 nm diameter gold particles in toluene and refluxing for 4 hours at 60

°C.

2.2.1.2 Nickel nanocrystal synthesis

5 nm nickel nanocrystals were synthesized in an argon environment by combining

0.211 g bis(cyclopentadienyl) nickel (C5H6NiC5H6), 480 µL of dodecanethiol, and 25 mL

of toluene that was preheated to 75 °C. The solution was vigorously stirred at 75 °C for

two hours. The resulting solution was alternately centrifuged in chloroform and ethanol

three times to remove excess thiol, uncapped particles, and reaction byproducts.

2.2.2 Droplet preparation and evaporation

Each size sample of gold nanocrystals was examined under different

concentration conditions. Dispersions were prepared with the following concentrations:

1.67 g of gold nanoparticles/L of chloroform, Lg 108.2 1−× , Lg 1026.9 2−× ,

Lg1025.1 2−× and Lg 1025.1 3−× . 3 µL of each nanocrystal dispersion was drop-cast

on a carbon-coated 200 mesh copper Transmission Electron Microscopy (TEM) grid.

Solvent was evaporated from the TEM grid while holding the substrate horizontal with

anti-capillary action tweezers at 20°C in still air.

Page 32: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

16

2.2.3 Characterization techniques

2.2.3.1 Transmission Electron Microscopy (TEM)

The nanocrystal microstructures were imaged using a Phillips EM208 TEM

operated at an acceleration voltage of 120 kV.

2.2.3.2 Scanning Electron Microscopy (SEM)

Scanning Electron Microscopy (SEM) was used to further characterize

microscopic patterning. Nanoparticles in chloroform were drop cast on to a glassy

carbon substrate and examined using a LEO 1530 HRSEM at 7 kV.

2.2.3.3 Atomic Force Microscopy (AFM)

The thickness of the ring structures and honeycomb networks was determined

using atomic force microscopy (AFM; Thermomicroscopes Autoprobe CP Research, with

Ultralever cantilevers), using a 10 µm scanning head. Both glassy carbon and mounted

TEM grids were used as substrates for AFM work. When TEM grids were examined, the

AFM was operated in non-contact mode. For glassy carbon substrates, contact mode was

used.

2.2.3.4 Surface tension measurements

Surface tension measurements of each chloroform solution concentration were

made by using a ring tensiometer (Central Scientific Co., Inc.). Each solution was

measured fifteen times to insure repeatability, and each solution was measured at 15.5 °C

and 20 °C to determine the change of surface tension with temperature.

Page 33: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

17

2.3 RESULTS AND CONCLUSIONS

2.3.1 Quality of gold nanocrystals

Figure 2.1 (A), (B) and (C) show the typical particles produced following the

synthesis described above. The TEM in Figure 2.1 (A) shows 3.5 nm diameter gold

nanocrystals with and the inset is a high-resolution TEM image of a nanocrystal with

lattice fringes present, confirming particle crystallinity.

Figure 2.1 (B) is a TEM image of 5 nm gold nanocrystals, and the size

distribution is noticeably larger than those of the 3.5 nm sample. Figure 2.1 (C) is the

most polydisperse sample, consisting of particles between 3 and 12.5 nm in diameter.

Page 34: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

18

Figure 2.1: TEM images of (A) 3.5 nm diameter gold nanocrystals. Inset show crystalline nature of particles, with lattice fringes present. (B) 5 nm diameter gold nanocrystals. (C) 3 – 12.5 nm polydisperse, gold nanocrystals.

2.3.2 Qualitative patterning as a function of concentration

Figure 2.2 shows the progression of patterns formed with decreasing gold

nanocrystal concentration for 3.5 nm diameter particles. At the highest concentration

studied (1.67 g/L), the nanocrystals coalesce into ring structures with an average diameter

of 1.7 µm. As the concentration is lowered ( Lg 1026.9 2−× ), the pattern transitions from

Page 35: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

19

isolated dewetting rings to polygon networks. As the nanocrystal concentration decreases

further ( Lg1025.1 2−× ), the particles organize into networks of uniform hexagons as

shown in Figure 2.2 (D), with a characteristic lattice parameter L=4.3 µm. The

characteristic size of the hexagonal networks formed using 3.5 nm diameter nanocrystals

did not vary with nanocrystal preparation, the grid used, or the area on the grid examined.

However the topography of the TEM grid did affect the honeycomb formation, as

hexagonal arrays limited themselves to “valleys” in the carbon film. The arrays typically

stretched over 300 µm2. At concentrations below Lg1025.1 2−× , the networks no

longer appear, as in Figure 2.2 (E) at a concentration of Lg1025.1 3−× .

Page 36: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

20

Figure 2.2: TEM images of microscopic structures formed by 3.5 nm diameter sterically stabilized Au nanocrystals drop-cast on a carbon substrate from a chloroform dispersion. (A) Au nanocrystals at 1.67 g/L deposit as rings. (B) At 0.277 g/L, the number of rings decreases. (C) At 9.26 × 10-2, rings no longer form; instead, polygonal networks become the preferred morphology. (D) At 1.25 × 10-2 g/L, ordered repeating hexagonal networks resembling honeycombs dominate the structural morphology. (E) Further dilutions (1.25 × 10-3) collapse the order.

The 5 nm and 3 – 12.5 nm diameter samples did not show as radical changes in

morphology through the concentration spectrum examined. The 5 nm sample formed

rings and at some concentrations uneven polygonal structures. 3 – 12.5 nm diameters

only formed ring structures, with the ring concentration decreasing with decreasing

particle concentration.

Page 37: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

21

2.3.3 Qualitative patterning as a function of particle size and distribution

The nanocrystal size and size distribution, affect the morphology of the self-

assembled structures. Figure 2.3 shows representative TEM images of nanocrystal

dispersions deposited from Lg1025.1 2−× dispersions as a function of polydispersity

and particle size. The larger 5 nm nanocrystals, Figure 2.3 (B), and the polydisperse

nanocrystals, Figure 2.3 (C), with sizes ranging from 3 to 12.5 nm, do not organize into

hexagonal networks at this concentration, or at any of the other concentrations tested.

Page 38: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

22

Figure 2.3: TEM images of (A) monodisperse 3.5 nm , (B) monodisperse 5 nm, and (C) polydisperse 3 to 12.5 nm Au nanocrystals deposited on a carbon substrate from chloroform at a concentration of 1.25 × 10-2 g/L.

2.3.4 Topography of nanocrystal patterns

The thickness of the ring structures and the honeycomb networks was determined

using AFM. Figure 2.4 shows an AFM image of the border of one of the hexagonal

convection cells and an image of nanocrystals that have formed a dewetting ring. The

ring structures were significantly thicker than the honeycomb networks of particles, with

Page 39: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

23

heights of greater than 25 nm, compared to ~15 nm for the hexagonal borders. This

result complements the TEM images where the rings appear darker than hexagonal

borders. The measured thicknesses correspond to 4 layers of 3.5 nm particles for

hexagonal cell borders and more than 7 layers for dewetting rings.

Figure 2.4: AFM images of (A) hexagonal networks and (B) dewetting rings of Au nanocrystals. The average height of the hexagonal borders is 15 nm, and the dewetting rings vary from 25 to 100 nm.

2.3.5 Ring and polygonal patterning mechanism

Ohara, Heath, and Gelbart proposed that rings of nanocrystals form as a result of

hole nucleation and growth in thin wetting solvents during evaporation.17,18 In the

absence of thermocapillary effects (i.e., convective motions in the fluid), the film

Page 40: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

24

thickness decreases homogeneously until reaching a critical thickness when the film

becomes unstable and dewets the substrate through a hole nucleation mechanism. The

critical thickness depends on the disjoining pressure of the solvent film on the substrate.

The holes grow as solvent continues to evaporate. As a hole widens in the film,

nanocrystals collect at the air-solvent interface until becoming “pinned” on the substrate.

The solvent continues to evaporate, leaving the nanocrystals fixed on the substrate in the

form of a ring. The separation between the hole nucleation event—which occurs at a

critical fluid film height—and hole growth, leads to relatively size-monodisperse ring

diameters as shown in Figure 2.2 (A). Dewetting rings similar to these have also been

observed for thin polymer films.22-24 As the nanocrystal concentration decreases from

1.67 g/L, the evaporating hole spreads further before nanocrystal pinning occurs. At Lg 1026.9 2−× , the holes grow large enough to intersect in some areas on the TEM grid.

At the intersection of the rings, the nanocrystals restructure to minimize their interfacial

free energy, resulting in the formation of polygonal networks. This dewetting mechanism

has also been well-documented for thin polymer films.22-24

2.3.6 A mechanism for hexagonal array formation: Marangoni convection

Further decreases in the nanocrystal concentration ( Lg1025.1 2−× ) lead to the

formation of organized hexagonal networks. It is highly unlikely that the hexagonal

networks form as a result of intersecting dewetting holes. The thermodynamic packing

constraints and force-field interactions between particles are not sufficiently complex or

long-range to control microstructural organization into honeycomb rings. In essence,

there is no driving force to spatially correlate neighboring dewetting holes into a

hexagonal lattice. In fact, organized hexagonal networks have not been observed in any

dewetting thin polymer films—only disorganized polygonal networks have been

formed.22-24 Additionally, as seen in Figure 2.5, rings were observed within hexagonal

Page 41: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

25

cells, indicating that the rings and hexagons most likely do not both result from dewetting

phenomena, as dewetting does not produce concentric rings. Also, the simultaneous

occurrence of these patterns within the same cells indicate that they form at different

stages of the evaporation process and that two different organizing mechanisms could be

active under these deposition conditions.

Figure 2.5: TEM image of dewetting rings within the honeycomb network.

Pileni and coworkers recently stated that Marangoni convection, or

thermocapillary flow, can direct nanocrystal organization during the evaporation

process.19,20 In relatively thin fluid films, on the order of micrometers thick, solvent

surface tension fluctuations at the air–solvent interface caused by a vertical temperature

gradient across the film can give rise to two-dimensional hexagonally organized flow

patterns.25 As solvent evaporates from a drop-cast nanocrystal dispersion, a vertical

temperature gradient results as evaporation cools the free surface relative to the

isothermal substrate. Marangoni convection, originally observed by Bénard in 1900,26

and later explained by Pearson in the 1958,27 can give rise to hexagonal honeycomb

networks of convective flow cells.25 For a flat liquid film with negligible thickness

Page 42: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

26

compared to its length and fixed temperatures at the top and bottom surface, the

Marangoni number Ma, is

ρνκσ Td

Ma T ∆=

; (2.1)

where, σ is the liquid surface tension, ∆T is the temperature differential across the

fluid film of thickness d, dTd

Tσσ =

, ρ is density, ν is kinematic viscosity, and κ is the

thermal diffusivity. The Marangoni number is a ratio of the force due to the surface

tension gradient to the viscous forces. For a system characterized by Ma less than the

critical Marangoni number 80≅cMa , heat transfer from the underlying hot surface to

the cool air-liquid interface occurs by diffusion. However, when 80=> cMaMa ,

surface tension gradients become large relative to the viscous forces and convective

motion sets in.27,28 This instability in the fluid film can lead to a steady-state flow pattern

of hexagonally organized convective “cells,” depicted in the cross-sectional schematic of

Figure 2.6.

Page 43: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

27

Figure 2.6: Illustration of surface tension driven convection. The substrate temperature, Ts, is higher than T, the temperature of the free interface, creating a vertical temperature gradient. When Ma>Mac, the film becomes unstable due to thermal fluctuations on the surface, and convection occurs. Surface tension increases with decreasing temperature, causing the surface to spread at regions of higher temperature and contract at regions of lower temperature. The resulting surface flow forces warmer liquid near the substrate to rise and replace the spreading fluid, which leads to vertical convective flow. Flow is further enhanced because the rising fluid is warmer than the average gas-solvent interfacial temperature. Convection reaches steady state as viscous forces dissipate the surface tension induced energy gradient.

Steady-state flow occurs as the viscous forces dissipate the thermally-derived

surface tension energy gradients in the film. This behavior has been well-studied in a

variety of micrometer thick fluid films.28 In these systems, the dimensionless wave

number, Ldac π2= ; (2.2)

is on the order of 2, and relates the characteristic lattice spacing of the hexagonal network

L, to the fluid film thickness d.

T - δT T - δTT + δT

d

Ts

T

L

Liquid

Substrate

Air

Page 44: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

28

2.3.7 Marangoni Convection in Nanocrystal Solutions

2.3.7.1 Steady State Approximation

Chloroform, the solvent in this system, is an extremely volatile liquid, and the free

surface is decreasing in temperature rapidly until the droplet completely evaporates. This

system experiences an ever increasing ∆T, while d becomes smaller and smaller.

Complete evaporation can occur in as small a period of time as 50 seconds. The subject

of the behavior and onset of Marangoni convection in evaporating films has been tackled

in at least three papers.27,29,30 Pearson and Tan et al. assume that the heat lost due to

evaporation is steady state, meaning that the heat lost during evaporation is replenished to

the free surface from the fluid below, and the free surface temperature does not change

over time.27,30 Ha and Lai make the free surface temperature a function of time, but they

use a quasi-steady state by assuming that no mass is lost due to evaporation, which is

inaccurate in this present system.29 Solution of the quasi-steady state approximation is

not trivial, and removing the mass condition would require extremely rigorous

calculations. A rough approximation of the heat lost in this solution uses the heat of

vaporization of chloroform. The droplets used were 3 µl in volume, decreased in

temperature by 4.5 °C and took approximately 50 seconds to evaporate completely.

2.7.3.2 Numerical Values of Ma

Few measurements of the surface tension, density, and viscosity of nanocrystal

dispersions exist. Therefore, in order to calculate Ma for the dispersions, ρ, ν, and σ were

measured as a function of nanocrystal concentration, size and size distribution. Table 2.1

lists the parameters measured at 20°C for 3.5 and 5 nm diameter gold dodecanethiol-

capped nanocrystals at different concentrations in chloroform. There was no noticeable

concentration dependence on ν. The same was true for ρ—as expected for these dilute

Page 45: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

29

colloidal dispersions. However, σ (30.7 dyne/cm) for 1.66 g/L 3.5 nm and (29.94

dyn/cm) for 5 nm gold nanocrystals in chloroform at 20ºC exceeded the pure chloroform

value of 27.4 dyn/cm. The surface tension increases with increasing particle

concentration for both particle sizes. The increase in surface tension with increased

concentration presumably results from the much lower volatility of the nanocrystals

relative to the solvent. Values of σT also change significantly with particle concentration,

with larger σT at lower concentrations. The increased solvent exposure at the air-liquid

interface in more dilute nanocrystal dispersions leads to greater Tσ . The thermal

conductivity also changes with nanocrystal concentration. Changes in the thermal

conductivity manifest themselves in Ma as variations in κ.

Page 46: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

30

Table 2.1: Parameters Measured at 20 °C for 3.5 and 5 nm Au Nanocrystals Dispersed in Chloroform: σ, σT, κ, Ma†

concentration (g/l) diameter (nm) σ (dynes/cm) σT (dynes/cm) κ (cm2/s) Ma

1.66 3.5 30.75 -0.5520 9.161×10-4 63.6

2.77×10-1 3.5 30.7 -0.5522 7.553×10-4 77.1

9.26×10-2 3.5 30.63 -0.5531 7.339×10-4 79.5

1.25×10-2 3.5 30.38 -0.5606 7.246×10-4 81.6

1.25×10-3 3.5 30.3 -0.6833 7.233×10-4 99.7

1.66 5 29.94 -0.5421 9.161×10-4 62.4

2.77×10-1 5 29.85 -0.5447 7.553×10-4 76.1

9.26×10-2 5 29.8 -0.5496 7.339×10-4 79

1.25×10-2 5 29.77 -0.5520 7.246×10-4 80.4

1.25×10-3 5 29.67 -0.7392 7.233×10-4 107.8

Ma was calculated using the measured parameters for each dispersion and

estimating ∆T and d (Table 2.1). The cooling of the free surface due to evaporation of

chloroform was estimated as ∆T =4.5°C. This number is produced from the change in

temperature of a 3 µL droplet of chloroform on an omega thermocouple. d= 1.3 µm was

calculated using the relation in Equation 2.2, where dimensionless ac = 2 at the onset of

Marangoni convection and L = 4.3 µm.

In all of the 3.5 nm nanocrystal dispersions, Ma is close to 80, the critical

Marangoni number required for convective flow, and increases with decreasing † For all solutions, ρ = 1.471 (g/ml), υ = 0.58 cP, α = 1.24 × 10-3 (1/K), k = 2.46 × 10-4 (cal/s⋅cm2)), kg = 5.624 × 10-5 (cal/s⋅cm2⋅(C/cm)).

Page 47: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

31

concentration. At 1.25×10-2 g/L—the concentration that produces hexagons—Ma > 80

and Marangoni convection is expected. At the onset of thermocapillary flow, the fluid

sweeps the nanocrystals into the boundary regions between neighboring flow cells. The

solvent velocity at the cell boundaries must reach a boundary condition of zero.

Furthermore, the temperature of the solvent between cells is lowest, which gives the

lowest nanocrystal solubility in the solvent film, causing nanoparticles to collect on the

substrate at the cellular interface, as shown in Figure 2.7. With sufficient concentration

and slow diffusion compared to the evaporation rate, the nanocrystals remain fixed

between flow cells as the solvent evaporates below the film thickness required to host

short wavelength instabilities. The hexagonal networks form using a narrow range of

nanocrystal concentrations because Marangoni convection becomes more favorable with

decreasing concentration while simultaneously there must be enough particles to fill the

gaps between convection cells.

Figure 2.7: Illustration of gold nanocrystal/ chloroform system.

2.3.7.3 Particle Diffusion After Marangoni Convection

It is essential to confirm that diffusion of nanocrystals across the substrate after

Marangoni convection is not large enough so that the ordering is lost. Towards that goal,

T - δT T - δTT + δT

d

Ts

T

Evaporation of Solvent

L

Evaporation of Solvent

Page 48: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

32

the diffusion coefficient of 3.5 nm nanocrystals was estimated using the Stokes-Einstein

equation,

ABAB R

kTDπµ 6

= ; (2.3)

where k is Boltzmann’s constant, µB is the solvent viscosity, RA is the nanocrystal radius,

and T is the solvent temperature. Plugging in the appropriate parameters for these

nanocrystals, DAB = 1.85×10-10 m2/s. A lower limit characteristic time for a 1.3 µm film

to evaporate can be estimated by assuming no convection occurs in the gas phase:

cD

xl

AB∆∆

=ρτ ; (2.4)

ρ is the density of chloroform, l is the stagnant diffusion film thickness (~0.2 µm), and ∆c

is the concentration gradient in the gas phase (9.4×10-4 g/cm3). Considering a fluid film

thickness ∆x 1.3 µm, τ = 2.05×10-5 sec. Using this value of τ, the mean square

displacement traveled by the nanocrystals during solvent evaporation can be estimated

from ( ) 22.04 2/1 == τπ ABDx µm. Since L=4.3 µm in the honeycomb arrays, the

nanocrystals remain relatively localized during the final solvent evaporation, and the

pattern is not erased due to diffusion.

2.3.8 Competition of Long Wavelength and Short Wavelength Marangoni Convection

Films sufficiently thin will tend towards long wavelength Marangoni Convection,

which causes a hole to open up in the film, rather than short wavelength convection.31,32

In order for Marangoni convection to cause the hexagonal arrays seen above, a

nanocrystal/chloroform thin film of d = 1.3 µm must be able to support short range

convection. While this film thickness is short enough to produce long wavelength

convection, there is no proof that it will be the dominant convective force.

Page 49: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

33

Models that determine the onset thickness for long wavelength convection apply

for the steady state condition and how the effects of a non steady state system would

impact the prediction models is unknown.31,32 However, for a range of liquid depths both

long wavelength and short wavelength instabilities can coexist.31 While liquid is being

drawn from one area to another, the rest of the surface can contain short range

instabilities. This may explain why hexagonal patterns did not appear over the entire

TEM grid.

Another possibility is that long wavelength convection did not occur for several

reasons. One factor that discourages long wavelength instabilities is the time scale of

formation, which in some cases is on the order of hours, and would not be a factor in an

evaporation process that is finished within a minute. Also, existing hexagonal patterns

have been known to suppress long wavelength motion,32 so it is possible that short

wavelength Marangoni convection occurring when the droplet was thicker prevented long

wavelength convection from developing.

2.3.9 Surface Tension Effects of Polydispersity

Ordered honeycomb networks were not observed for polydisperse nanocrystals or

larger 5 nm diameter nanocrystals. The 5 nm diameter nanocrystal solutions do not

exhibit values of σT as high as the smaller particles, leading to slightly lower Ma.

Nonetheless, Ma> 80 for the two most dilute solutions, the 1.25×10-2 g/L sample is

shown in Figure 2.3 (B), yet hexagons were not observed for either case. In the most

dilute solution there are certainly not enough particles to leave a residue of the hexagons

on the grid as in the case for the 3.5 nm solution of the same concentration. Since the

Marangoni number is barely above 80 for the 1.25×10-2 g/L solution (assuming d = 1.3

µm), the drive towards convective behavior is not as strong as the 3.5 nm nanocrystal

case, requiring a larger d for Marangoni convection to occur; thus the nanocrystals have

Page 50: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

34

more time to diffuse away from the interface as the solvent evaporates completely.

Indeed, the deposited nanocrystal patterns for the 5 nm particles appear as though they

are approaching the polygonal networks that exist for the smaller nanocrystal samples.

Another factor to consider is the slightly larger polydispersity of the 5 nm sample which

is discussed below for the 3 – 12.5 nm sample.

The polydisperse sample, as shown in Figure 3.2 (C), did not exhibit order at any

of the concentrations tested. It is possible that these samples could not achieve organized

convective flow patterns due to the presence of renegade larger particles that disturbed

the surface tension “field” within their vicinity at the interface, thus preventing spatially

correlated flow patterns to develop. In this respect, it appears that in addition to the value

of Ma, the local uniformity in particles size and interaction on the grid are also important.

2.3.10 Hexagonal Arrays in Other Nanocrystal Systems

Other nanocrystals, in addition to gold, were also observed to form hexagonal

networks. For example, nickel particles with an average diameter of 4 nm produced

honeycomb networks, as shown in the high resolution SEM image of Figure 2.8. In the

sample, particles sizes ranged between 3 and 7 nm, yet networked hexagons formed.

However, the ordering of this sample is less extensive than the gold nanoparticles, which

can perhaps be attributed to the polydispersity of the nickel nanocrystals.

Page 51: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

35

Figure 2.8: SEM image of a hexagonal network of 3-7 nm sterically stabilized nickel nanocrystals

2.4 CONCLUSIONS

In conclusion, the experimental observations presented here demonstrate that

within a narrow concentration range, short wavelength Marangoni instability can occur

during solvent evaporation to produce honeycomb networks of monodisperse sterically–

stabilized 3.5 nm diameter nanocrystals with a lattice parameter of 4.3 µm. Organized

networks were not observed at higher nanocrystal concentrations, giving rise instead to

1.7 µm diameter rings as a result of the dewetting characteristics of the solvent. Rings

have been observed to occur within the hexagonal networks, indicating that ring and

hexagonal network formation occur at two different stages in the evaporation process.

A necessary, but not sufficient condition for honeycomb network formation is that

Ma>80, which depends on the values of several parameters of the nanocrystal solution.

The value of σT changes significantly as a function of particle concentration, with the

most dilute solutions giving the largest changes. The combined changes in thermal

conductivity and σT make the most dilute solutions more prone to Marangoni convection.

Page 52: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

36

However, the most dilute solutions do not form hexagonal networks because the particle

concentrations are too low to leave significant deposits on the TEM grid. The particle

size also affects σT and 5 nm nanocrystals did not form hexagonal networks at identical

concentrations as the 3.5 nm particles. Furthermore, nanocrystal polydispersity frustrates

hexagonal network formation.

2.5 REFERENCES

(1) Saunders, J. V. Philos. Mag. A 1980, 42, 705.

(2) Alder, B. J.; Wainwright, T. E. Phys. Rev. 1962, 127, 459-361.

(3) Pusey, P. N.; VanMegen, W. Nature 1986, 320, 340-342.

(4) Luck, V. W.; Kleir, M.; Wesslau, H. Ber. Bunsen-Ges. Phys. Chem. 1963, 67, 75.

(5) McConnell, G. A.; Gast, A. P.; Huang, J. S.; Smith, S. D. Phys. Rev. Lett. 1993, 71, 2102-2105.

(6) Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Science 1995, 270, 1335-1338.

(7) Whetten, R. L.; Shafigullin, M. N.; Khoury, J. T.; Schaaff, T. G.; Vexmar, I.; Alvarex, M. M.; Wilkinson, A. Acc. Chem. Res. 1999, 32, 397-406.

(8) Korgel, B. A.; Fitzmaurice, D. Phys. Rev. B. 1999, 59, 14191-14201.

(9) Saunders, J. V.; Murray, M. J. Nature 1978, 275, 201-203.

(10) Bartlett, P.; Ottweill, R. H.; Pusey, P. N. Phys. Rev. Lett. 1992, 68, 3801-3804.

(11) Kiely, C. J.; Fink, J.; Brust, M.; Bethell, D.; Schiffrin, D. J. Nature 1998, 396, 444-446.

(12) Kiely, C. J.; Fink, J.; Zheng, J. G.; Brust, M.; Bethell, D.; Schiffrin, D. J. Adv. Mater. 2000, 12, 640.

(13) Denkov, N. D.; Velev, O. D.; Kralchevsky, P. A.; Ivanov, I. B.; Yoshimura, H.; Nagayama, K. Langmuir 1992, 8, 3183-3190.

(14) Velev, O. D.; Lenhoff, A. M.; Kaler, E. W. Science 2000, 287, 2240-2243.

Page 53: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

37

(15) Gray, J. J.; Klein, D. H.; Bonnecaze, R. T.; Korgel, B. A. Phys. Rev. Lett. 2000, 85, 4430-4433.

(16) Gray, J. J.; Klein, D. H.; Korgel, B. A.; Bonnecaze, R. T. Langmuir 2001, 17, 2317-2328.

(17) Ohara, P. C.; Heath, J. R.; Gelbart, W. M. Angew. Chem. Int. Ed. Engl. 1997, 36, 1077.

(18) Ohara, P. C.; Gelbart, W. M. Langmuir 1998, 14, 3418-3424.

(19) Maillard, M.; Motte, L.; Ngo, A. T.; Pileni, M. P. J. Phys. Chem. B. 2000, 104, 11871-11877.

(20) Maillard, M.; Motte, L.; Pileni, M. P. Adv. Mater. 2001, 13, 200-204.

(21) Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D. J.; Whyman, R. J. J. Chem. Soc. Chem. Commun. 1994, 801-802.

(22) Reiter, G. Phys. Rev. Lett. 1991, 68, 75-78.

(23) Sharma, A.; Reiter, G. J. Colloid Interface Sci. 1996, 178, 383-399.

(24) Thiele, U.; Mertig, M.; Pompe, W. Phys. Rev. Lett. 1998, 80, 2869-2872.

(25) Probstein, R. F. Physiochemical Hydrodynamics: an Introduction; Butterworths: Boston, 1989.

(26) Bénard, H. Rev. Gen. Sci. Pures Appl. Bull. 1900, 11, 1261-1271.

(27) Pearson, J. R. A. J. Fluid Mech. 1958, 4, 489-500.

(28) Schatz, M. F.; VanHook, S. J.; McCormick, W. D.; Swift, J. B.; Swinney, H. L. Phys. Rev. Lett. 1995, 75, 1938-1941.

(29) Ha, V. M.; Lai, C. L. Proc. R. Soc. Lond. A 2001, 457, 885-909.

(30) Tan, K. K.; Thorpe, R. B. Chem. Eng. Sci. 1999, 54, 775-783.

(31) VanHook, S. J.; Schatz, M. F.; McCormick, W. D.; Swift, J. B.; Swinney, H. L. Phys. Rev. Lett. 1995, 75, 4397-4400.

(32) VanHook, S. J.; Schatz, M. F.; Swift, J. B.; McCormick, W. D.; Swinney, H. L. J. Fluid Mech. 1997, 345, 45-78.

Page 54: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

38

Chapter 3: Iridium Nanocrystal Synthesis and Surface Coating-Dependant Catalytic Activity‡

3.1 INTRODUCTION

The catalytic properties of solid-supported metal nanocrystals have been studied

for decades due to their high surface area-to-volume ratios and ability to catalyze many

industrially important reactions;1-5 however, some recent reports have found that

nanoscale materials can in some cases give rise to new unique catalytic activity.6 These

findings—enabled to a large extent by significant recent advances in metal nanocrystal

synthetic control—7-10 have renewed interest in developing a better fundamental

understanding of metal nanocrystal catalytic properties by studying metal nanocrystals as

“model” materials.10

Iridium (Ir) is a particularly interesting transition metal catalyst; it is well-known

for promoting alkene, aldehyde and ketone hydrogenation,11-18 with more reactions

possible when used in a bimetallic system;19,20 furthermore, there have been reports of

unusually high catalytic activity from molecular Ir clusters.21 A few reports of Ir

nanocrystal synthesis exist in the literature;12,15,16,18,22 however, existing synthetic

methods for Ir nanocrystals are not well developed and better quality particles are needed

for catalytic studies. Here we demonstrate the synthesis of high quality Ir nanocrystals

by reducing (methylcyclopentadienyl)(1,5-cyclooctadiene)Ir or ((MeCp)Ir(COD)), an

organometallic compound developed as a chemical vapor deposition (CVD) precursor for

Ir films, in the presence of organic capping ligands to control the particle size.23 The

nanocrystal synthesis was carried out with four different capping ligands: oleic

acid/oleylamine, trioctylphosphine (TOP), tetraoctylammonium bromide (TOAB) and

‡ The contents of this chapter have been submitted to Nano Letters.

Page 55: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

39

tetraoctylphosphonium bromide (TOPB). The ligand chemistry greatly influences the

nanocrystal size and size distribution and the catalytic activity of the nanocrystals.

3.2 EXPERIMENTAL

3.2.1 Nanocrystal synthesis

3.2.1.1 Chemicals

All chemicals were of the highest purity available and used without further

purification. Oleic acid (C18H34O2H, 65%), oleylamine (C18H37N, 70 %), dioctylether

([CH3(CH2)7]2O, 97%), trioctylphosphine or TOP (C24H51P, 90 %) were purchased from

Fluka. Tetraoctylammonium bromide or TOAB (C32H68N Br, 98 %),

tetraoctylphosphonium bromide or TOPB (C32H68P Br, 97 %), decane (C10H22, 99 %), 1-

decene (C10H20, 94 %) and 1,2 hexadecanedediol (C16H34O2, 90%) were purchased from

Aldrich. (Methylcyclopentadienyl) (1,5-cyclooctadiene) iridium ( (C6H7)(C8H12)Ir, 99

%) was purchased from Strem. All solvents used were of analytical grade and purchased

from Aldrich.

3.2.1.2 Oleic acid and oleylamine capped iridium

Iridium nanocrystals capped with oleic acid and oleylamine were produced in a

method similar to Sun et al.’s for FePt.24 In a 25 ml- three neck, round bottom flask, 7.5

ml of dioctylether was used to dissolve 0.195 g of hexadecanediol. 0.08 µl of oleic acid

and 0.085 µl of oleylamine were added. Septa were placed on two of the necks of the

flask, while a condenser and a stopcock valve connected to the third. The flask was

connected to a Schlenk line, and three freeze-pump-thaw cycles were performed on the

contents, to remove oxygen. This flask was heated up to 290 °C under nitrogen flow.

0.19 g of (Methylcyclopentadienyl) (1,5-cyclooctadiene) iridium was measured into

another 25 ml- three neck, round bottom flask, and 1 ml of dioctylether was added. The

Page 56: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

40

flask was prepared using the same method as the first to remove oxygen, but kept at room

temperature. The contents of this flask were then injected into the 290 °C flask, and the

reaction was allowed to proceed for 2 hours. The product was a dark brown liquid that,

after cleaning by centrifuging the product in alternating solutions of a hexanol and

ethanol mixture and chloroform, resulted in brown precipitate. The nanocrystals

redisperse in various non-polar organics solvents including chloroform, cyclohexane and

toluene. The size distribution could be narrowed by size selective precipitation using

ethanol as an antisolvent.

3.2.1.3 Trioctylphosphine capped iridium

Iridium particles were synthesized with TOP as the capping ligand by using a

similar injection method, with flasks prepared as in the previous method. 0.39 grams of

hexadecanediol were dissolved into 10 ml of TOP in a 25 ml- three neck, round bottom

flask which was deoxygenated and brought to 290 °C. 0.38 g of

(methylcyclopentadienyl) (1,5-cyclooctadiene) iridium was added to a second 25 ml-

three neck, round bottom flask with 2 ml of TOP, deoxygenated and then injected into the

290 °C flask. After heating for 1.5 hours, the solution cooled and a brown product was

extracted and cleaned by centrifuging in alternating hexanol and chloroform solutions.

3.2.1.4 Pyridine capped iridium

A ligand exchange was performed on TOP capped particles, replacing the ligand

with pyridine. This was accomplished by dispersing TOP capped particles in hexane,

adding excess pyridine, and stirring the solution for 24 hours at room temperature.

Afterwards, the particles were cleaned by precipitation in hexane. For more description

of this procedure, see section 4.2.

Page 57: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

41

3.2.1.5 Iridium annealed on silicon substrate

4 nm Oleic Acid and Oleylamine Ir particles were drop cast on to a silicon

substrate and then annealed on SiO2 at 625 °C for one hour, in an attempt to completely

remove the stabilizing ligands from the Ir core.

3.2.1.6 Tetraoctylammonium bromide and tetraoctylphosphonium bromide capped iridium

Iridium particles with TOAB and TOPB ligands were synthesized using only one

25 ml- three neck, round bottom flask. 0.2 g of hexadecanediol, 7 ml of dioctylether, 0.2

g of (methylcyclopentadienyl) (1,5-cyclooctadiene) iridium, and 0.76 g of either TOAB

or TOPB were measured into the flask. The solution was freeze-pump-thawed for three

cycles, and heated up to 270 °C for 30 minutes. The product was a black liquid and

particles were isolated using one rinse in ethanol. Rinsing the solution multiple times in

ethanol resulted in the removal of capping ligands from the iridium surface, so exposure

to ethanol and other antisolvents was kept at a minimum.

3.2.2 Catalysis

1-decene hydrogenation reactions were carried out in a 100 ml round bottom flask

that acted as a batch reactor. Three psig of hydrogen gas were bubbled into 1-decene at

75 °C. Iridium nanoparticles in 1-decene were injected into the flask as to create a

1000:1 decene to iridium mass ratio. Initially the reaction temperature dropped to 70 °C

but stabilized back at 75 °C after 1 minute. Aliquots of the solution were removed every

minute for the first 20 minutes of the reaction, then every 5 minutes until one hour, and at

every 10 minutes after that. Aliquots were placed in screw cap gas chromatography vials.

Particles were later recovered by vacuum evaporation of the liquid away from the

particles.

Page 58: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

42

3.2.3 Characterization techniques

3.2.3.1 Transmission Electron Microscopy

High resolution transmission microscopy (HRTEM) images were obtained using a

JEOL 2010F microscope operating at 200 kV. Low resolution transmission microscopy

(LRTEM) images were obtained using a Phillips EM208 microscope operating at 120

kV. The samples were prepared by drop casting dilute solutions of iridium particles in

chloroform on 200 mesh carbon-coated copper grids.

3.2.3.2 Powder X-Ray Diffraction

Powder X-Ray diffraction (XRD) was performed on a Burker-Nonius Powder

Diffractometer using Cu Kα radiation (λ = 1.54 Å). Iridium samples in chloroform were

drop cast on a quartz substrate.

3.2.3.3 Small Angle X-Ray Scattering

Small angle X-ray scattering (SAXS) measurements were performed using as

rotating copper-anode generator by Bruker Nonius, Molecular Metrology, Inc. operated

at 3.0 kW on iridium nanocrystals dispersed in cyclohexane. X-rays accessed the sample

though the Kapton windows of the sample holder. The scattering angle was calibrated

using a silver behenate (CH3(CH2)20)COOAg) standard. The experimental SAXS data

was corrected for background scattering and sample absorption.

3.2.3.4 Gas Chromatography/ Mass Spectroscopy

The catalytic properties of oleic acid and oleylamine, TOP, TOAB, and TOPB

particles were determined by analyzing the hydrogenation reaction aliquots using a

Hewlett Packard 5890 series II gas chromatography mass spectrophotomer (GCMS). The

GCMS was calibrated to determine the relative amounts of 1-decene and decane in each

aliquot, and in this way the hydrogenation reaction progress and rate were monitored.

Page 59: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

43

3.3 RESULTS AND DISCUSSION

3.3.1 Oleic acid and oleylamine capped Ir nanocrystals

3.3.1.1 Quality of nanocrystals

The oleic acid/oleylamine coated Ir nanocrystals obtained after purification

ranged from 1.5 nm to 5 nm in diameter. Size selective precipitation using

chloroform/ethanol as the solvent/antisolvent pair could be used to narrow the size

distribution if desired. Figure 3.1 shows transmission electron microscopy (TEM), X-ray

diffraction (XRD) and small angle X-ray scattering (SAXS) data of the Ir nanocrystals.

TEM and XRD confirm that the nanocrystals are crystalline face-centered cubic (fcc) Ir.

The lattice spacings observed by TEM match fcc Ir. Analysis of the XRD peak

broadening using the Scherrer equation gives an estimate of the particle diameter of 5.3

nm, which corresponds well with the average diameter of 5 nm determined from TEM

images.

Page 60: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

44

0.01

0.1

1

10

100

0.4 0.8 1.2 1.6 2 2.4 2.8 3.2 3.6

Nor

mal

ized

Inte

nsity

q (nm-1)

C

Figure 3.1: Oleic acid/oleylamine-coated Ir nanocrystals. (A) TEM image of ~4 nm diameter size-selected Ir nanocrystals, (B) XRD Ir nanocrystals, peaks correspond to fcc Ir (JCPDS file 46-1044), (C) SAXS of two different sizes of Ir nanocrystals isolated by size-selective precipitation from the same reaction: (□) Ravg= 1.09 nm, σ= ±0.283 nm (○) Ravg= 1.92 nm, σ= ±0.635 nm; the curves correspond to the best fits of Eqn (1) to the data that were used to obtain Ravg and σ, (D) HRTEM of a 5 nm oleic acid/oleylamine coated Ir nanocrystal.

Page 61: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

45

Figure 3.1 (C) shows SAXS data for oleic acid/oleylamine-coated Ir nanocrystals

dispersed in hexane. Unlike TEM, which can be used to probe the size of only a very

small number of nanocrystals in the sample, SAXS provides an accurate measure of the

average nanocrystal size and size distribution by probing >1010 nanocrystals.25-28 In

SAXS, the scattered X-ray intensity )(qI , is measured as a function of the wave vector,

λπθ /4)sin(=q , which depends on the scattering angle θ2 , and the X-ray wavelength

( λ =0.154 nm). In a nanocrystal dispersion, the particles are generally sufficiently dilute

and nonaggregating such that scattering results from “isolated” nanocrystals of radius R,

and )(qI is only a function of the size distribution )(RN , and the shape factor for a

sphere, ( ) ( ) ( ) ( )( ) ( )[ ] 23cossin3 qRqRqRqRqRP −= :25-28

∫∝ dRRqRPRNqI 6)()()(. (3.1)

Assuming that the nanocrystal size distribution is Gaussian, ]2/)(exp[)2/(1)( 22 σπσ avgRRRN −−= , Eqn (3.1) can be fit to the data in Figure 3.1

(C), to obtain the average radius avgR , and the standard deviation, σ . The Ir

nanocrystals measured in Figure 3.1 (C) by SAXS have diameters of Ravg= 1.09 nm (σ=

±0.283 nm) and Ravg= 1.92 nm (σ= ±0.635 nm).

3.3.1.2 Catalytic properties

Figure 3.2 shows the 1-decene/decane conversion in the presence of 2 nm and 4

nm oleic acid/oleylamine coated Ir nanocrystals. It is clear that the oleic acid/oleylamine

coated Ir nanocrystals did not catalyze 1-decene hydrogenation. 1.5 nm and 5 nm

diameter oleic acid/oleylamine particles were also tested and did not convert a

measurable amount of decene after 5 hours. 4 nm annealed Ir particles on silicon also did

catalyze the hydrogenation reaction after 5 hours. Perhaps reduced surface area of the

iridium on a substrate or a smaller iridium to reactant ratio is responsible.

Page 62: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

46

0

20

40

60

80

100

0 50 100 150 200 250

Dec

ene

Con

cent

ratio

n (%

)

Reaction Time (Min.)

Figure 3.2: Conversion of 1-decene to decane at 75oC and 3 psig H2 in the presence of dispersed Ir nanocrystals: (◊) 2 nm oleic acid/oleyl amine coated; (×) 4 nm oleic acid/oleyl amine coated; (+) TOP-coated; (○) 1.5 nm TOAB-coated particles, TOF = 5 s-1; (□) 5 nm TOPB-coated, TOF= 270 s-1.

3.3.2 Weaker binding capping ligands

The oleic acid/oleylamine coated Ir nanocrystals were of relatively high quality in

terms of crystallinity, narrow size and shape distribution, and good dispersibility;

however, they exhibited no ability to catalyze the hydrogenation of 1-decene to decane.

Since Ir is a well-known hydrogrenation catalyst and recent reports have demonstrated

the catalytic activity of Ir nanocrystals in ionic liquids,16-18 the absence of catalytic

activity of oleic acid/oleylamine coated Ir nanocrystals prompted the exploration of

weaker-binding capping ligands to determine if the capping ligands were in fact

preventing catalysis on the nanocrystal surface. Recently, Li and El-Sayed noted that

very good capping ligand stabilization appears to diminish catalytic nanocrystal activity

in some cases,29 and Narayanan and El-Sayed showed that nanocrystals can also ripen

and precipitate under the catalytic reaction conditions due to poor capping, leading to

diminished reactivity over time due to decreased surface area-to-volume ratio.30

Page 63: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

47

There are several capping ligands, such as trioctylphosphine (TOP),

tetraoctylammonium bromide (TOAB) and tetraoctylphosphonium bromide (TOPB), that

are known to be “weak” binders to metal and semiconductor nanocrystal surfaces.31,32

Since there is little predictive knowledge about the influence of capping ligand chemistry

on metal nanocrystal catalysis, Ir nanocrystals were synthesized using all three of these

ligands and then studied for their catalytic activity.

3.3.2.1 Nanocrystal quality

All four capping ligands stabilized Ir nanocrystals; however, the quality of the

nanocrystals varied widely. Figure 3.3 shows TEM images of TOP, TOAB and TOPB-

coated Ir nanocrystals. The TOP-capped particles were the poorest quality—extremely

polydisperse, with diameters ranging from 10 to 100 nm and irregular shapes. TOAB-

capped Ir nanocrystals were crystalline and relatively size-monodisperse, with diameter

ranging from 1.5 to 3 nm. TOPB-capped Ir nanocrystals were also crystalline, but were

larger and slightly more polydisperse than the TOAB-capped nanocrystals, averaging 4

nm in diameter.

Page 64: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

48

Figure 3.3: TEM images of Ir nanocrystals synthesized with different capping ligands: (A) TOAB (1.5~3 nm diameter); (B) TOPB (2~5 nm diameter); and (C) TOP (10~100 nm diameter).

Page 65: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

49

3.3.2.2 Catalytic activity

The capping ligand chemistry significantly affects the Ir nanocrystal catalytic

activity. Figure 3.2 shows the conversion of 1-decene to decane as a function of reaction

time in the presence of TOP-, TOAB-, and TOPB-capped Ir nanocrystals. TOP-capped

and pyridine capped Ir did not promote 1-decene hydrogenation, but both TOAB and

TOPB-coated Ir nanocrystals were catalytically active. ~1.5 nm diameter TOAB-capped

nanocrystals converted 42% of the 1-decene after 230 minutes and 5 nm diameter TOPB-

coated Ir nanocrystals converted 100% of the decene after 60 minutes.

3.3. Turnover frequency calculations

Turnover frequencies (TOF), defined as the number of molecules n, reacted at

each available catalytic site per time t,

=

dtdn

NTOF

act

1 , (3.2)

were calculated assuming that all surface atoms are active with

( )nAAN UCPact = , (3.3)

where PA is the surface area of the average particle size, UCA is the surface area of an Ir

unit cell face, and n is the number of Ir atoms in a unit cell face. Although only a small

portion of surface Ir atoms can actually serve as catalytic active sites—many will be

bonded to capping ligands and unavailable for catalysis—it is common practice to take

the total number of surface atoms as the number of active catalytic sites when the value is

not known.33 Using Eqn (3.2) and the average nanocrystal diameter measured using

TEM, TOF=4 s-1 for ~1.5 nm TOAB-coated nanocrystals and TOF=270 s-1 for ~5 nm

TOPB-coated nanocrystals.

Page 66: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

50

3.3.2 Catalytic activity as a function of recycling

Recycling the nanocrystals through several hydrogenation reactions changed their

catalytic activity. Figure 3.4 shows 1-decene conversion to decane in the presence of 1.5

nm diameter TOAB-coated Ir nanocrystals used in sequential reactions at 75oC for 230

min. The nanocrystal catalytic activity increased with subsequent reactions until the fifth

reaction cycle, when the catalytic activity dropped significantly. During the first reaction

cycle, the nanocrystals achieve only a 42% conversion of 1-decene and a TOF of 4 s-1. A

second hydrogenation yielded 100% conversion in the same 230 minute period with

TOF=13 s-1. The third and fourth cycles led to 100% conversion in shorter times and

significantly higher TOFs: 50 s-1 and 124 s-1, respectively. Additionally, the lag time in

the initial stages of the reaction shortened with each subsequent reaction.

Page 67: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

51

0

20

40

60

80

100

0 50 100 150 200 250

Dec

ene

Con

cent

ratio

n (%

)

Reaction Time (Min.)

A

Figure 3.4: TOAB-coated Ir nanocrystals after 1-decene hydrogenation reactions (A) Hydrogenation of 1-decene to decane as a function of catalyst recycling: (○) First reaction, TOF = 5 s-1; (□) Second Reaction, TOF = 14 s-1; (◊) Third Reaction, TOF = 50 s-1; (×) Fourth Reaction, TOF = 124 s-1; (∆) Fifth Reaction, TOF = 38 s-1; (B) TEM images of TOAB-coated Ir nanocrystals before catalysis; and (C) after four hydrogenation reactions.

Page 68: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

52

The increase in TOF indicates that the nanocrystal surfaces become increasingly

catalytically active with each reaction, most likely as a result of ligand desorption and

reduced surface capping. It is well known that capping ligands on nanocrystal surfaces

are in dynamic equilibrium with free solvated ligands, and ligand desorption is always a

concern when considering nanocrystal dispersion stability and particle aggregation. The

ligand desorption kinetics depend on the binding strength between the ligand functional

group and the inorganic nanocrystal surface. Due to their relatively weak binding, TOAB

and TOPB can be stripped from the nanocrystal surfaces by excessive precipitation with

antisolvent (5 to 10 precipitations) preventing redispersibility; whereas, oleic

acid/oleylamine are bound relatively strongly to the Ir nanocrystal surfaces and could be

reprecipitated many times and still be redispersed. The higher TOF of TOPB-coated

nanocrystals relative to TOAB-coated nanocrystals appears to relate to the difference in

ligand binding strength: Ir-N has a shorter equilibrium bond length than Ir-P, indicating

that Ir-N has a larger binding energy and stronger attachment to the surface.34,35 This is

consistent with the observed smaller diameter of TOAB-capped nanocrystals relative to

TOPB-capped nanocrystals and the higher TOF of the TOPB-coated nanocrystals.

Basically, more Ir surface atoms are exposed for catalysis with TOPB-coating.

In the fifth reaction, the catalytic activity decreased significantly, reaching 100%

conversion after 150 min with a TOF=38 s-1. Comparison of TEM images of

nanocrystals before and after catalysis (Figures 3.4 (B) and 3.4 (C)) reveal that the

nanocrystals exhibit an increasing amount of particle aggregation. TEM also showed that

the Ir particles increase slightly in size (due to Ostwald ripening) during each reaction,

which was accounted for in the TOF calculations in each reaction. The reduced TOF

after the fifth cycle (the average Ir particle size is 3 nm) is consistent with decreased

available surface area per particle for catalysis due to aggregation. For comparison, the

Page 69: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

53

TOF calculated by considering the available surface area-to-volume ratio of particles as

“aggregates” with average “diameter” of 10 nm gives a value of TOF=127 s-1, which is

the same as measured in the fourth reaction cycle. The nanocrystals appear to become

better catalysts as the capping ligand surface coverage decreases, until the dispersion

becomes unstable and the particles aggregate.

3.4 CONCLUSIONS

Ir nanocrystal catalysis of 1-decene hydrogenation to decane is extremely

sensitive to capping ligand chemistry and ligand coverage. “Good” capping ligands—

i.e., those that stabilize robust nanocrystals with very narrow size distributions—appear

to be poor choices for catalytic applications. However, the nanocrystals must be of

reasonably high quality with good dispersibility in multiple reaction cycles, so the ligands

must be strong enough binders to stabilize nanocrystals, but “weak” enough to provide

reactant access to the metal surface. These studies reveal that capping ligands play a

central role in the catalytic activity of transition metal nanocrystals and cannot be

ignored.

3.5 REFERENCES

(1) Kralik, M.; Biffis, A. J. Mol. Catal. A: Chem. 2001, 177, 113-138.

(2) Lewis, L. N. Chem. Rev. 1993, 93, 2693-2730.

(3) Aiken_III, J. D.; Finke, R. G. J. Mol. Catal. A: Chem. 1999, 145, 1-44.

(4) Bonnemann, H.; Brijoux, W.; Brinkmann, R.; Dinjus, E.; Joussen, T.; Korall, B. Angew. Chem. Int. Ed. Engl. 1991, 30, 1312-1314.

(5) Roucoux, A.; Schulz, J.; Patin, H. Chem. Rev. 2002, 102, 3757-3778.

(6) Valden, M.; Lai, X.; Goodman, D. W. Science 1998, 281, 1647-1650.

(7) Narayanan, R.; El-Sayed, M. A. Nano Lett. 2004, 4, 1343-1348.

(8) Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2004, 108, 5726-5733.

Page 70: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

54

(9) Kim, S. W.; Park, J.; Jang, Y.; Chung, Y.; Hwang, S.; Hyeon, T.; Kim, Y. W. Nano Lett. 2003, 3, 1289-1291.

(10) Son, S. U.; Jang, Y.; Yoon, K. Y.; Kang, E.; Hyeon, T. Nano Lett. 2004, 4, 1147-1151.

(11) Lin, Y.; Finke, R. G. Inorg. Chem. 1994, 33, 4891-4910.

(12) Yee, C. K.; Jordan, R.; Ulman, A.; White, H.; King, A.; Rafailovich, M.; Sokolov, J. Langmuir 1999, 15, 3486-3491.

(13) Hayek, K.; Goller, H.; Penner, S.; Rupprechter, G.; Zimmermann, C. Catal. Lett. 2004, 92, 1-9.

(14) Lin, Y.; Nomiya, K.; Finke, R. G. Inorg. Chem. 1993, 32, 6040-6045.

(15) Mevellec, V.; Roucoux, A.; Ramirez, E.; Phillippot, K.; Chaudret, B. Adv. Synth. Catal. 2004, 346, 72-76.

(16) Fonseca, G. S.; Scholten, J. D.; Dupont, J. Synnlett 2004, 1525-1528.

(17) Dupont, J.; Fonseca, G. S.; Umpierre, A. P.; Fichtner, P. F. P.; Teixeira, S. R. J. Am. Chem. Soc. 2002, 124, 4228-4229.

(18) Fonseca, G. S.; Umpierre, A. P.; Fichtner, P. F. P.; Teixeira, S. R.; Dupont, J. Chem. Eur. J. 2003, 9, 3263-3269.

(19) Okada, M.; Nakamura, M.; Moritani, K.; Kasai, T. Surf. Sci. 2003, 523, 218-230.

(20) Shiblisma, N. Int. J. Hydrogen Energy 1993, 18, 141-147.

(21) Anderson, J. R.; Howe, R. F. Nature 1977, 268, 129-130.

(22) Shah, P. S.; Husain, S.; Johnston, K. P.; Korgel, B. A. J. Phys. Chem. B 2001, 105, 9433-9440.

(23) Hoke, J. B.; Stern, E. W.; Murray, H. H. J. Mater. Chem. 1991, 1, 551-554.

(24) Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Science 2000, 287, 1989-1992.

(25) Glatter, O.; Kratky, O., Eds. Small Angle X-ray Scattering; Academic Press Inc.: New York, 1982.

Page 71: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

55

(26) Guinier, A.; Fournet, G. Small-Angle Scattering of X-Rays; Wiley: New York, 1955.

(27) Lifshin, E., Ed. X-ray Characterization of Materials; Wiley-VCH: New York, 1999.

(28) Korgel, B. A.; Fitzmaurice, D. Phys. Rev. B 1999, 59, 14191-14201.

(29) Li, Y.; El-Sayed, M. A. J. Phys. Chem. B 2001, 105, 8938-8943.

(30) Narayanan, R.; El-Sayed, M. A. J. Am. Chem. Soc. 2003, 125, 8340-8347.

(31) Mattoussi, H.; Cumming, A. W.; Murray, C. B.; Bawendi, M. G.; Ober, R. J. Chem. Phys. 1996, 105, 9890-9896.

(32) Saunders, A. E.; Sigman, M. B.; Korgel, B. A. J. Phys. Chem. B 2004, 108, 193-199.

(33) Dumonteil, C.; Lacroix, M.; Geantet, C.; Jobic, H.; Breysse, M. J. Catal. 1999, 187, 464-473.

(34) Dahlenburg, L.; Gotz, R. Eur. J. Inorg. Chem 2004, 888-905.

(35) Housecroft, C. E.; O'Neill, M. E.; Wade, K.; Smith, B. C. J. Organomet. Chem. 1981, 213, 35-43.

Page 72: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

56

Chapter 4: Dilute Magnetic III-V Semiconductor Nanocrystals: Synthesis and Magnetic Properties§

4.1 INTRODUCTION

The addition of impurity elements to a host lattice is a common approach for

tuning electronic, optical, mechanical and magnetic properties of materials. In the case of

nanostructured materials, their small size and large surface area to volume ratios present

a significant challenge to impurity doping. Extremely high doping concentrations must

be achieved to incorporate a meaningful number of atoms into the structure. For

example, 12 Mn impurity atoms in a 4 nm diameter InAs semiconductor nanocrystal

corresponds to a doping concentration greater than 1020 cm-3 (compare this to typical n-

and p-dopant levels in silicon 1016 to 1018 cm-3!). Doping levels of this magnitude often

exceed the solubility limits of the impurity element in the host lattice. Furthermore, high

dopant solubility does not guarantee effective nanostructure doping. The very high

surface area to volume ratios in nanocrystals have been found to enhance impurity

segregation to the particle surface through a “self-annealing” process in the core that

intensifies with increasing synthesis temperature.1 Since reasonably high synthesis

temperatures are typically required to achieve core crystallinity and high quality optical

properties,2 this presents a challenge to nanocrystal doping.1,3 Nonetheless, other studies

have shown that surface segregation does not necessarily occur in all materials3 and can

be eliminated in some cases by manipulating the reaction chemistry.1,4-7

Group III-V dilute magnetic semiconductor (DMS) nanocrystals, such as the Mn-

doped InAs and InP nanocrystals studied here, provide an interesting testbed for a study

of impurity doping of nanostructures. The Mn solubility is low—in the range of 1016 to

§ The contents of this chapter, in part, appear in Nano Lett. 2003, 3, 1441-1447

Page 73: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

57

1017 cm-3—8 yet Mn doping levels (~1020 cm-3) far exceeding the solubility limit in GaAs

and InAs hosts have been achieved by kinetically-limited molecular beam epitaxy

(MBE).9 The ability to achieve high dopant levels of magnetic impurity elements in

these materials subsequently led to the discovery of unique properties, such as

ferromagnetism and electric field dependent magnetic susceptibilities.10 These properties

result from coupling between charge carriers and the Mn d-electrons. Only recently has

manifested itself in this manner in the more thoroughly studied Group II-VI DMS

materials with very high Mn solubility due to the scarcity of charge carriers within those

systems.11-17

Nanocrystals of Group III-V DMS materials could exhibit magnetic and optical

properties modified by nanoscale dimensions. For example, enhanced electron-hole

exchange interactions in quantum dots have been speculated to lead to unique magnetic

properties.18 In this context, ferromagnetic DMS thin films have been fabricated and

studied, however, there has not been an example of group II-V DMS nanocrystals

exhibiting ferromagnetism at any temperature.19-22 Due to the high Mn solubility in II-VI

hosts, the synthetic nanocrystal research in the literature to date has focused almost

entirely on Group II-VI DMS materials, including transition metal doped CdSe,1,4

ZnS,23,24 ZnSe3, CdS25, TiO226 and ZnO.6,27 Although the magnetic impurity in these

produces properties such as magnetic circular dichroism and large Zeeman splitting, this

class of materials has provided ferromagnetism only in the cases of extensive co-doping26

and annealed nanocrystal thin films.24,27 Initially, the only example of Group III-V DMS

quantum dots had been synthesized by MBE, with a broad size and shape distribution.20,21

Colloidal nanocrystal synthesis could potentially yield nanocrystals with controlled size

and shape, and the process could be scaled up to produce large quantities of material;

Page 74: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

58

however, it must be demonstrated that kinetic trapping of a magnetic impurity element in

a crystalline semiconductor host is possible by arrested precipitation.

In this chapter is presented the colloidal synthesis of MnxIn1-xAs and MnxIn1-xP

nanocrystals ranging from 2.2 to 10 nm in diameter, along with compositional, optical,

and magnetic characterization. The MnxIn1-xAs nanocrystals were synthesized with Mn

compositions as high as 1.25 atomic % (xMn=0.025) within the semiconductor core. The

MnxIn1-xP nanocrystals were synthesized with Mn compositions as high as 5.7 atomic %

(xMn=0.11) encorporated within the core. The nanocrystals exhibit size-dependent

photoluminescence emission (PL) red-shifted with respect to InAs and InP particles of

equivalent size, and antiferromagnetic superexchange between Mn impurity atoms. In

contrast to their Group II-VI counterparts, such as Mn-doped ZnSe and CdSe, these

dopant levels are at least two orders of magnitude above the bulk solubility limit. Doping

levels of this magnitude were accessed through the kinetic trapping during nanocrystal

synthesis that occurs in the dehalosilylation synthesis reaction.

4.2 EXPERIMENTAL

4.2.1 Nanocrystal synthesis

4.2.1.1 InMnAs nanocrystal synthesis

MnxIn1-xAs nanocrystals were synthesized using a modification of the high

temperature dehalosilylation colloidal synthesis for InAs developed by Guzelian, et al.28

Depending on the amount of nanocrystals desired for characterization, the synthesis was

carried out in either a 1 mL stainless steel batch reactor or using standard airless

procedures on a Schlenk line. The stainless steel reactors were loaded and sealed inside a

nitrogen glove box with 0.9 mL of a stock solution of 0.25 g InCl3 (Strem), 0.165 g

tristrimethylsilylarsine (((CH3)3Si)3As, prepared according to the procedure described by

Page 75: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

59

Wells et al.29) and 6.8 mL trioctylphosphine, (C8H17)3P (TOP, Aldrich). The reactor was

placed in a preheated brass heating block at the synthesis temperature, which ranged

between 250°C and 280°C (below the TOP boiling temperature of 290°C). After reacting

for one to three hours, the particles were extracted from the cell with chloroform. Higher

reaction temperature improved the crystallinity, while longer reaction times led to larger

nanocrystals. Reaction byproducts were separated from the nanocrystals through

alternating rinses with ethanol and chloroform. The nanocrystals were size selectively

precipitated using chloroform and ethanol as the solvent-antisolvent pair. The procedure

on the Schlenk line was slightly modified from the small batch reactions. In a three-

necked flask under nitrogen, 20 ml TOP and 0.73 g InCl3 were stirred and heated to the

reaction temperature (between 250°C and 280°C). Upon reaching the desired

temperature, 0.71 g ((CH3)3Si)3As was injected by syringe. Under reflux, the reaction

proceeded for one to three hours. Mn doping was explored using a variety of precursors,

including MnCl2, MnBr2, dimethylmanganese (MnMe2), and manganese (II)

phthalocyanine (C32H16MnN8). Each precursor was added to the reaction mixture at a

0.16 M Mn concentration.

4.2.1.2 InMnP nanocrystal synthesis

MnxIn1-xP nanocrystals were produced using a synthesis based on the InP

nanocrystal work by Battaglia et al.30 Standard airless procedures were followed on a

Schlenk line and all chemicals were stored in a nitrogen filled glove box. In a three-

necked flask under nitrogen, 25 ml of 1-octadecene (ODE, C18H36, Aldrich), 0.09 g

Indium acetate (In(C2H3O2)3, Strem), 0.08 g of manganese acetate (Mn(C2H4O2)2, City

Chemical, LLC) and 0.27 g of myristic acid (C14H22O2, Aldrich) were stirred and heated

to the reaction temperature 280 °C. Upon reaching the desired temperature, 0.05 g

tristrimethylsilylphoshine (((CH3)3Si)3P, (Strem)) in 2 ml of ODE was injected by

Page 76: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

60

syringe. The reaction progressed at this temperature for 1.5 hours, and the products were

extracted with chloroform. The particles were precipitated by centrifuging in a mixture

of ethanol and butanol as an antisolvent. Subsequent cleaning was accomplished by

alternating centrifugation with chloroform and ethanol.

Size selective precipitation was not easily accomplished with these particles. The

sizes produced from this reaction were more monodisperse than InMnAs, so size

selective precipitation did not yield noticeably more monodisperse particles. As a result,

it was difficult to produce particles of different sizes with the same Mn concentration.

Fortunately, the average sizes of the particles could be tuned by adjusting the reaction

conditions. Larger particles were formed at longer reaction times. Smaller particles were

formed with shorter reaction times and higher reaction temperatures.

4.2.1.3 Ligand exchange

To replace either TOP or myristic acid with pyridine, a ligand exchange was

performed per the procedure developed by Mikulec et al.1 MnInAs or InMnP

nanocrystals were dissolved in approximately 1 ml of toluene and then stirred for 24

hours in 15 ml of pyridine. The MnInAs or InMnP nanocrystals were then precipitated

using hexane as an antisolvent.

4.2.1.3 Nanocrystal annealed thin films

InMnP nanocrystals were annealed into a thin film to eliminate any

superparamagnetic effects associated with individual nanocrystals. A concentrated

solution of ligand exchanged particles in pyridine was drop cast on to a 0.5 cm × 0.5 cm

slice of mica (muscovite, Alfa Aesar) substrate resting on a warm hot plate. After a mass

comparable to that of the substrate had been deposited on the mica, the sample was

annealed in a 650 °C furnace under nitrogen for 6 hours.

Page 77: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

61

4.2.2 Characterization techniques

4.2.2.1 High Resolution Transmission Electron Microscopy (HRTEM)

HRTEM was performed using a JEOL 2010F TEM at an operating voltage of

200kV with a 200 mesh carbon coated copper TEM grid as the substrate. The Energy

Dispersive X-Ray Spectroscopy (EDS) feature of this tool was used to determine the In:P

ratio for InMnP particles. This technique was not used for other elemental analysis

because the technique listed in section 4.2.2.6 was more sensitive and had higher

precision for all elements assayed in this work except for phosphorous.

4.2.2.2 High Resolution Scanning Electron Microscopy (HRSEM)

HRSEM images were acquired using a LEO 1530 SEM from samples drop cast

on a glassy carbon substrate.

4.2.2.3 Powder X-Ray Diffraction (XRD)

Powder X-Ray diffraction (XRD) was performed on a Burker-Nonius Powder

Diffractometer using Cu Kα radiation (λ = 1.54 Å). InMnP samples in chloroform were

drop cast on a quartz substrate.

4.2.2.4 Small Angle X-Ray Scattering (SAXS)

SAXS measurements were performed using a rotating copper-anode generator

(Bruker Nonius, Molecular Metrology, Inc.) operated at 3.0 kW on InAs and InMnAs

nanocrystals dispersed in cyclohexane. Scattered photons were collected on a multiwire

gas-filled detector calibrated using silver behenate (CH3(CH2)20COOAg). All

experimental data were corrected for background scattering and sample absorption. q is

the magnitude of the scattering vector, q = 4π/λ sin(θ), where λ is the wavelength (0.154

nm) and 2θ is the scattering angle. The angle-dependent scattered intensity I(q), from a

dilute dispersion of non-interacting nanocrystals of radius R depends on the normalized

Page 78: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

62

size distribution N(R), and the shape factor P(qR), ∫∝ RRqRPRNqI d)()()( 6 .31-36 The

average diameter Ravg and size distribution were determined using the shape function for

spherical particles, 2

3)()cos()sin(3)(

−=

qRqRqRqRqRP , and a Gaussian size distribution,

−−= 2

2

2

)(exp

21)(

σπσavgRR

RN , where σ is the standard deviation.

4.2.2.5 Absorbance and Photoluminescence (PL)

UV-vis absorbance measurements were carried out using a Cary 500 Scan, Varian

spectrophotometer. PL measurements were performed with a Fluorolog F-3 equipped

with a Jorbin-Yvon/ SPEX InGaAs detector for the IR measurements of InMnAs

measurements and with a PMT detector for the visible frequencies of InMnP.

4.2.2.6 Inductively Coupled Plasma Mass Spectroscopy (ICP-MS)

Elemental analysis was obtained using a Micromass Platform inductively coupled

plasma mass spectrometer (ICP-MS). Samples were prepared by drop casting the

nanocrystals into a plastic centrifuge tube. Concentrated nitric acid was added to dissolve

the particles and then diluted with deionized water to provide manganese, arsenic, and

indium concentrations within the required ranges based upon calibration curves created

from elemental standards. Phosphorous was not measured on this equipment because the

background signal at and around the mass of phosphorous is too great to measure

concentration accurately.

Page 79: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

63

4.2.2.7 Electron Spin Resonance (ESR)

Electron spin resonance (ESR, IBM Instruments, Inc. ER 300) spectroscopy

provides another measure of Mn doping and the nature of the dopant. ESR was used to

analyze InMnAs only, as the tool was decommisioned prior to InMnP work.

4.2.2.8 Superconducting Quantum Interference Device Magnetomety (SQUID)

Temperature-dependent and field-dependent magnetization measurements of the

InMnAs and InMnP nanocrystals were conducted using a superconducting quantum

interference device magnetometer (SQUID, Quantum Design). Thin films of pyridine,

myristic acid or TOP nanocrystals were drop cast onto a glass substrate and then scraped

into a Lilly #4 gelatin sample holder and secured with cotton for the measurements.

4.3 RESULTS AND DISCUSSION

4.3.1 Nanocrystal quality

4.3.1.1. InMnAs

Figure 4.1 (A) and (C) show TEM images of InAs and Mn0.01In0.99As

nanocrystals, respectively, after size selective precipitation. Prior to size selective

precipitation, the InAs and MnxIn1-xAs nanocrystal size distributions were broad, with

particles ranging from 2 to 10 nm in diameter. Size selective precipitation narrowed the

size distribution to standard deviations about the average particle diameter (σ) less than

±20% (as determined by both TEM and SAXS (Figure 1.4 (B) and (D),31-36 with σ as low

as ±11% for the best samples. HRSEM, such as in Figure 1.4 (E), provided another

confirmation of the overall quality of the MnxIn1-xAs nanocrystal synthesis. There was

no evidence of a pure MnAs phase in any of the MnxIn1-xAs samples.

Page 80: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

64

InMnAs particles synthesized in this fashion also may form macroporous thin

films when deposited with chloroform in humid air as a result of the condensation of

water as documented by Shah et al.37

Page 81: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

65

10-4

10-3

10-2

10-1

100

101

102

0 0.5 1 1.5 2 2.5 3 3.5

Nor

mal

ized

Int

ensi

ty

q (nm-1)

B

10-6

10-5

10-4

10-3

10-2

10-1

100

101

102

0 0.5 1 1.5 2 2.5 3 3.5N

orm

aliz

ed I

nten

sity

q (nm-1)

D

Figure 4.1: TEM images of (A) 4.5 nm TOP capped InAs nanocrystals and (C) 4.5 nm TOP capped Mn0.01In0.99As nanocrystals. (B) SAXS data of InAs nanocrystals dispersed in cyclohexane: (□) dp= 4.4 nm, σ=±0.18; (×) dp=4.2 nm, σ=±0.14; (○) dp=3.6 nm, σ=±0.20. (D) SAXS data for -Mn0.003In0.997As nanocrystals in cyclohexane: (□) dp=3.8 nm, σ=±0.15; (×) dp=3.7 nm, σ=±0.17; (○) dp=3.76 nm, σ=±0.11; (+) dp=3.64 nm, σ=±0.13; (∆) dp=3.46 nm, σ=±0.14. Size histograms determined by TEM agree with those found from SAXS to within ±13%. (E) HRSEM image of 4 nm diameter Mn0.014In0.986As nanoparticles on a glassy carbon substrate.

Page 82: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

66

4.3.1.2 InMnP

Figure 4.2 (A) and (B) show TEM images of InP and Mn0.02In0.98P nanocrystals,

respectively, synthesized with a capping ligand of myristic acid. Since size selective

precipitation was not effective with these product, and the particles shown here did not

receive that procedure. The InP nanocrystals are crystalline, and average 5 nm in

diameter, as determined by TEM. The InMnP nanocrystals also have easily recognizable

lattice fringes, and the sample shown also averages 5 nm in diameter. The addition of

Mn doesn’t appear to impact the nanocrystals superficially, and both InP and InMnP

capped with myristic acid tend to clump together.

Figure 4.2 (C) shows XRD performed to verify the crystal structure of the

particles and to determine if phase separation into MnP occurred. As evidenced by the

diffraction peaks, the only phase that exists is the zinc blende structure of InP. It is

interesting to not that the addition of small amount of Mn within the crystal core has not

affected the diffraction peaks significantly.

Page 83: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

67

20 30 40 50 60 70 80 90

Inte

nsity

(arb

itrar

y un

its)

2θ (degrees)

(1 1

1)

(2 2

0)

(3 1

1)

(4 2

0)(2 0

0)

(4 2

2)

(5 1

1)

C

Figure 4.2: TEM images of (A) 5 nm InP, (B) and 5 nm Mn0.02In0.98P, and (C) XRD of 5 nm Mn0.02In0.98P nanocrystals. The diffraction peaks are indexed to the zinc blend structure of InP (JCPDS file 00-032-0452), and the addition of Mn has not noticeably affected the peak location, although the broadening of the peaks due to the nanocrystal small size may obscure any dopant effects.

Page 84: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

68

4.3.2 Doping

4.3.2.1 Precursor chemistry for InMnAs

The Mn precursor chemistry significantly affects the doping, with MnBr2 giving

the highest dopant concentrations. Table 4.1 shows the resulting dopant levels achieved

from each Mn precursor under similar reaction conditions. The addition of MnBr2 to the

synthesis as an Mn source did not significantly affect the average particle diameter

relative to the pure InAs synthesis. Dopant levels in 4 nm diameter TOP-capped MnxIn1-

xAs nanocrystals reached as high as xMn=0.05.

Table 4.1: In, Mn, As composition of TOP-capped InMnAs nanocrystals determined by ICP-MS for different reaction precursors of identical concentration and reaction conditions.

Nanoparticle Composition (%)Precursor

Mn In As

MnBr2 0.74 48.61 50.65

MnCl2 0.47 53.31 46.22

Mn (II) Phthalocyanine 0.01 61.33 38.66MnMe2 No Reaction

4.3.2.2 Ligand exchange: reduction in Mn

To confirm whether the dopant was bound to the nanocrystal surface or located

within the core of the nanocrystals, we performed a chemical surface exchange with

pyridine on many MnxIn1-xAs samples. Mikulec, et al.1 demonstrated that repeated

exposure to pyridine removed surface-bound Mn from CdSe nanocrystals, as pyridine

forms a stronger bond with the nanocrystal surface than TOP and strips the surface

molecules as it replaces them. This hypothesis is tested in section 4.3.2.4. A significant

amount of Mn was found to be associated with the nanocrystal surface. For example,

prior to ligand exchange, InMnP nanocrystals with xMn=0.048 decreased in Mn content to

Page 85: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

69

xMn=0.024, and InMnP nanocrystals with xMn=0.15 decreased to xMn=0.11. One quarter

to one half the Mn dopant was located at the particle surface, as was typical for all of the

samples synthesized by these procedures.

Figure 4.3 is a TEM image of 5 nm InMnP particles after a ligand exchange. The

particles are still crystalline, and as seen in the figure, pyridine-capped particles tend to

clump together a little more than with their original capping ligand.

Figure 4.3: TEM of 5 nm InMnP particles after ligand exchange. The ligand exchange reduces the size of the particles slightly, but does not affect the crystallinity of the particles.

4.3.2.3 Precursor chemistry for InMnP: non isovalent doping

Somaskandan et al. introduced Mn into InP through a 3+ valance precursor instead

of a 2+ valance precursor, like Mn acetate.22 By this technique, they claim that they can

produce more doping than by using a 2+ valance dopant, achieving as much as 1.35 %

Mn because the dopant is isovalent to the In for which it substitutes. Despite the valance

of precursor, the Mn in the core manifests itself as Mn2+.

The procedure described within this chapter used a non isovalent doping source

and the Mn2+ concentration within the semiconductor core after ligand exchange reached

as high as 5.7%. These findings would indicate that the synthesis conditions and

Page 86: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

70

technique is more important than the valance of the dopant. The reaction pathway plays a

prominent role in the success of dopant incorporation.

4.3.2.4 Ligand exchange: effectiveness

Since in this work it is extremely important to verify the location of Mn atoms

within the nanocrystal, the assumption that nanocrystals lose a layer of core atoms when

undergoing ligand exchange was tested.

The absorbance of InAs nanocrystals after different amounts of time in pyridine is

plotted in Figure 4.4. Comparing the absorbance peaks to the work by Guzelian et al.,28

the size of that particles is determined. The particles before ligand exchange are 4.5 nm

in diameter, decreasing to 3.8 nm after 23.5 hours. The unit cell of zinc blende InAs is

6.058 Å, meaning that more than a monolayer of InAs has been stripped from the initial

size. With longer ligand exchange durations, the particle size actually increases slightly,

while the particle size distribution, as evidenced by the width of the absorption peaks

increases. This is most likely due to the reabsorption of solvated atoms to the particle

surface due to Oswald ripening.

Page 87: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

71

1 1.2 1.4 1.6 1.8 2

Nor

mal

ized

Abs

orba

nce

Energy (eV)

0 hours

23.5 hours

67 hours

96.5 hours

Figure 4.4: Absorbance measurement for size selected InAs nanoparticles for different durations of pyridine ligand exchange. 0 hours corresponds to nanoparticles with the trioctylphosphine ligand.

In addition to measuring the size as a function of ligand exchange duration, the

amount of Mn present in InMnAs particles was plotted as shown in Figure 4.5. InMnAs

originally containing 0.15 % Mn with TOP as a capping ligand was placed in a pyridine

bath. After approximately 24 hours of ligand exchange, these particles lost roughly half

of the manganese within the particle, and for periods longer, the amount of manganese

actually increases, consistent with the reabsorption of manganese to the particle surface,

and with the information presented in Figure 4.4.

Page 88: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

72

0.06

0.08

0.1

0.12

0.14

0.16

0 50 100 150

Am

ount

of M

anga

nese

(%)

Ligand Exchange Duration (hours)

Figure 4.5: Percent composition of manganese in size selected, InMnAs nanoparticles as a function of pyridine ligand exchange duration.

4.3.2.5 Mn concentration as a function of particle size

The Mn composition systematically increased with slower heat-up rates of the

reaction vessel. For example, heat up times on the order of a few minutes gave

xMn=0.015, whereas, heat up times of 15 to 45 minutes led to significantly higher dopant

levels of xMn ranging from 0.02 to 0.05. The Mn concentration also appeared to vary

with particle diameter: smaller nanocrystals contained a higher Mn concentration, as

shown in Table 4.2. During ligand exchange, a proportionally larger percentage of Mn

was removed from the smaller particles due to their higher surface area to volume ratios

and higher surface-adsorbed Mn.

Page 89: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

73

Table 4.2: Composition of InMnAs particle cores determined by ICP-MS of non-ligand exchanged, size selected nanoparticles from the same batch. The smaller nanocrystals contain proportionally more Mn.

4.3.3. ESR results

ESR spectra measured at 115K are shown in Figure 4.6. The TOP capped InAs

and MnInAs nanoparticles were measured as concentrated dispersions in cyclohexane,

and the pyridine-treated MnInAs nanocrystals were measured as a powder. The ESR

spectra of MnInAs nanocrystals before and after chemical surface exchange revealed the

characteristic sextet centered at H=3.3 kG (g=2.02) of the d5 configuration of Mn

dopant.38,39 This configuration has been attributed to the localized trapping of an electron

by a substitutional Mn2+ dopant.38 A reference sample of undoped InAs nanocrystals

shows a flat background. The signal shape depends sensitively on the amount of Mn in

the nanocrystals and the location of the ions in the particle. The hyperfine splitting found

from curve B in Figure 4.6 is -14 cm 1074 −× , which is consistent with expectations for

Mn tetrahedrally coordinated in an ionic lattice.39 As the Mn concentration increases

from xMn=0.01 to xMn=0.024, the sextet signal weakens in the ESR spectra, as seen in

curve C in Figure 4.6. Szczytko and coworkers38 have indicated that the absence of the

hyperfine splitting results from increased Mn-Mn interactions at high dopant

concentrations. The Mn dopant level of the sample with xMn=0.012 is consistent with

Nanoparticle Composition (%) Nanoparticle Diameter, nm Mn In As

3.3 0.21 45.88 53.91 3.1 0.25 43.25 56.50 2.7 0.27 42.74 56.98 2.5 0.32 41.16 58.51 2.2 0.38 46.71 52.90

Page 90: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

74

dopant levels that have eliminated the hyperfine splitting in ESR spectra obtained from

thin MnInAs films.40 As a point of note, the ESR signal in curve C in Figure 4.6 could be

observed at room temperature, while the sextet in the ESR spectra of InAs nanocrystals

with lower Mn concentrations did not appear at temperatures exceeding 150 K.

Figure 4.6: Electron Spin Resonance (ESR) spectra measured at 115K and 9.42 GHz: (A) dp=5 nm, TOP capped InAs nanocrystals in cyclohexane; (B) dp=5 nm, MnxIn1-xAs (x=0.01) TOP capped nanocrystals in cyclohexane; (C) dp=5 nm, pyridine capped MnxIn1-xAs (x=0.024) nanocrystals in powder.

Figure 4.7 is an additional ESR spectra taken of pyridine capped InMnAs crystals,

(x=0.024), with the field range increased. If Mn3+ dopants exist within the InAs matrix,

then a signal will result from the d4 configuration, with a signal evident at a field slightly

larger than 1000 G, as calculated by Szczytko et al.38 The absence of Mn3+ signal,

especially when compared to the strength of the Mn2+ signal indicated that the presence

Page 91: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

75

of Mn3+ is negligible, so not only is the Mn successfully incorporated into the core of the

nanocrystals, the valence of the dopant has been preserved.

Figure 4.7: Additional ESR spectra of dp=5 nm, pyridine capped MnxIn1-xAs (x=0.024) nanocrystals in powder form, extended to less than 1000G.

4.3.4 Absorbance and photoluminescence data

4.3.4.1 InMnAs

Figure 4.8 compares the absorbance spectra for size-selected InAs and MnInAs

nanocrystals. All the PL spectra in Figure 4.8 were obtained at room temperature with an

excitation wavelength of 790 nm. The exciton peak energies in the absorbance spectra

and PL maxima are significantly blue-shifted from the bulk InAs band gap energy of 0.36

eV (3433 nm) due to quantum confinement. For InAs and InMnAs nanocrystals of equal

size, the exciton peak energies appear to be similar, independent of Mn dopant. In

contrast, Mn doping decreases the PL peak energy relative to undoped InAs of equal size,

and noticeably broadens the peak. The Mn-induced red shift in the PL peak energy is

consistent with an Mn-related acceptor state, as has been found both experimentally and

theoretically to lie approximately 0.1 eV above the valence band in the bulk.40,41 The

increased PL peak breadth for the nanocrystals doped with Mn cannot be attributed to a

Page 92: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

76

difference in size distribution, as both TEM and SAXS confirmed similar size

distributions for both samples; however, one cannot completely rule out the effect of a

compositional distribution in the sample. Nonetheless, the tail on the PL peak for the

MnInAs nanocrystals is consistent with the additional impurity levels due to the Mn

dopant.40,41

Page 93: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

77

Figure 4.8: Room temperature photoluminescence emission (PL) (λ exc= 790 nm) and absorbance spectra for nanocrystals dispersed in cyclohexane: (A) InAs and (B) Mn0.02In0.98As.

4.3.4.2. InMnP

Figure 4.9 shows the absorbance and photoluminescence spectra as synthesized 4

nm myristic acid capped InMnP nanocrystals. The PL spectra were obtained at room

Norm

alized PL Intensity

Energy (eV)11.22 1.6 0.8

3.0 nm

3.1 nm

3.2 nm

3.5 nm

3.6 nm

3.8 nm

4.0 nm

dpInAs

600 800 1000 1200 1400 1600

Nor

mal

ized

Abs

orba

nce

2.7 nm

2.8 nm

2.9 nm

3.2 nm

3.7 nm

4.2 nm

4.3 nm

InMnAs

Wavelength (nm)

Norm

alized PL Intensity

Energy (eV)11.22 1.6 0.8

3.0 nm

3.1 nm

3.2 nm

3.5 nm

3.6 nm

3.8 nm

4.0 nm

dpInAs

600 800 1000 1200 1400 1600

Nor

mal

ized

Abs

orba

nce

2.7 nm

2.8 nm

2.9 nm

3.2 nm

3.7 nm

4.2 nm

4.3 nm

InMnAs

Wavelength (nm)

Page 94: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

78

temperature with an excitation wavelength of 550 nm. The exciton peak energy in the

absorbance spectra and PL maxima are significantly blue-shifted from the bulk InP band

gap energy of 1.27 eV (976 nm) due to quantum confinement. The peak below was

compared to the absorbance spectra for InP as found in the work by Battaglia et al.30

Undoped InP has an exciton peak generally between 500 and 600 nm, depending on the

length of reaction time. The InMnP exciton peak below of approximately 675 nm is

substantially red shifted and less defined than that of InP. This is possibly due to a larger

size distribution. The emission spectra has a tail similar to that in InMnP, and again is

consistent with an additional impurity level due to the Mn dopant.41

400 500 600 700 800 900PL/

UV

-vis

Inte

nsity

(Arb

itrar

y U

nits

)

Wavelength (nm)

Figure 4.9: Room temperature photoluminescence emission (PL, solid line) (λ exc= 550 nm) and absorbance spectra (dashed line) for Mn0.02In0.98P nanocrystals dispersed in chloroform.

4.3.5 Magnetic Measurements

4.3.5.1 InMnAs

The temperature-dependent magnetization data shown in Figure 4.10 (A) were

obtained at a constant field at 5 kOe with temperatures ranging from 100K to 5K. The

Page 95: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

79

magnetization was also measured as a function of applied field at 5 K, as shown in Figure

4.10 (B). Pure InAs nanocrystals were measured as a baseline and found to be

diamagnetic as expected, with a molar magnetic susceptibility -136 molcm 102.49 −×−=molarχ , very close to the literature value of

-136 molcm 103.55 −×−=molarχ .42 The InMnAs nanocrystals exhibit positive

susceptibilities before and after chemical surface exchange, indicating conclusively that

Mn incorporates into the nanocrystal core. Additionally, constant field temperature scans

of InMnAs samples starting at 5K produced no local maxima susceptibility value,

indicating that there isn’t a blocking temperature associated with superparamagnetic

materials. This also is strong evidence that there is no phase separation into MnAs, as it

is ferromagnetic in the bulk, and would be superparamagnetic on these length scales.

Elemental analysis, ESR spectroscopy and magnetization measurements confirm the most

important result of this study, that kinetic trapping of Mn dopant in an InAs nanocrystal

host lattice can be induced by choosing the appropriate reaction pathway for the

synthesis.

Page 96: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

80

0 20 40 60 80 100

χmol

ar (x

103 c

m3 /m

ol)

Temperature (K)

0

0.5

1

1.5

2

2.5

3

3.5

A

-1.5

-1

-0.5

0

0.5

1

1.5

-60 -40 -20 0 20 40 60

Mag

netiz

atio

n (e

mu/

g)

Field (kOe)

B

Figure 4.10: Magnetization measurements of MnxIn1-xAs nanocrystals. (A) Temperature dependent magnetic molar susceptibility measured under a magnetic field of 5000 Oe, the solid lines are the predicted values as calculated from equation (4.1) in the text and (B) magnetization versus applied magnetic field measured at 5K: (×) dp=4 nm, x=0.024 (after pyridine ligand exchange); (□) dp=4 nm, x=0.048; (+) dp=4.6 nm, x=0.014 (after pyridine ligand exchange); (○) dp=4 nm, x=0.02; ( ) dp=4 nm InAs measured at 15 K.

Page 97: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

81

A closer analysis of the magnetization data provides information about the

nature of the magnetic interactions in the MnInAs nanocrystals for comparison with the

well-studied films fabricated by kinetically-controlled MBE. The molar susceptibilities

are relatively low, ranging from -134 molcm 105 −×≈molarχ at room temperature up to -133 molcm 103 −×≈molarχ at 5K, indicating that the magnetic coupling is weak. An

analysis of the magnetic interactions can be obtained by comparing the magnetic data

with a high temperature expansion of ( )Tmolarχ to second order in inverse powers of

temperature:43

( ) ( ) ( )

++=

TSS

TkgNT

B

BA

θµχ 11

3

2molar

, (4.1)

where

( ) ( )nnMn

BnnnMn

B

Jyk

SSJzyk

SS 123

13

+=

+= ∑θ

. (4.2)

In Equations (4.1) and (4.2), NA is Avogadro’s number, µB is the Bohr magneton,

kB is Boltzmann’s constant, T is temperature, g is the Landé g-factor, S is the spin, Mny

is the Mn mole fraction in the sample, zn is the number of nearest Mn neighbors to each

Mn atom, Jn and Jnn are the distance-dependent exchange integral and the nearest-

neighbor exchange integral, respectively. Equation (4.1) was fit to the measured values

of ( )Tmolarχ shown in Figure 4.10 (A). From the high temperature limit of equation

(4.1), we can determine the effective number of Bohr magnetons p, contributed by each

magnetic impurity atom to the magnetization: ( ) ( ) BSSJLSgp µ1+= , where

g JLS( ) =32

+12

S S +1( )− L L +1( )J J +1( )

. The values of p and nnJ determined from these

curve fits are shown in Table 4.3. Values of p range from 2.2 to 3.7 atomMn Bµ , which

is significantly lower than p=5.9 atomMn Bµ expected for Mn dopant with the d5

electron configuration. This lower value however is consistent with both theoretical and

Page 98: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

82

experimental findings. For example, Jain, et al.41 computed a local magnetic moment of

~4.1 atomMn Bµ as the Mn impurity in the lattice serves as an acceptor—an idea that

appears to be consistent with PL spectra. Experimentally, however, local magnetic

moments as low as 2.4 atomMn Bµ have been measured for MnxGa1-xAs thin films.44,45

Although the low magnetic moments of the Mn impurity in the III-V semiconductor hosts

still remain a subject of debate, it is believed that AsGa antisite defects could contribute to

this behavior.44,45 As seen in Table 4.2, the MnInAs nanocrystals are As rich in the core,

which could indicate a significant number of AsGa antisite defects. A high concentration

of Mn atoms would energetically favor the formation of AsGa defects through acceptor-

donor interactions—which would be consistent with the Mn d5 electron configuration

measured by ESR spectroscopy in Figure 4.6. According to Bergqvist, et al.,44 AsGa

antisite defects induce antiferromagnetic coupling between Mn in the lattice. From Figure

4.10 (A), the nearest-neighbor exchange integrals, Jnn kB , ranged from –2.0 to –4.3 K,

indicating antiferromagnetic superexchange interactions between the Mn impurity atoms

in the InAs host lattice. These values are higher than the measured bulk value of

KkJ Bnn 6.1−= found for n-type Mn-doped InAs,46 yet is lower than Jnn kB typical in

Mn-doped II-VI materials, such as CdMnTe ( KkJ Bnn 8−= ) .11

Table 4.3: Effective Bohr magneton number p , and the nearest neighbor exchange

integral nnJ for InMnAs found by fitting Equation (4.1) to Figure 4.10 (A).

Diameter

(nm) xMn Ligand p (µB/Mn Atom) Jnn/kB

(K) 4 0.020 TOP 2.2 -3.674 0.024 Pyridine 3.66 -4.29

4.6 0.014 Pyridine 2.73 -2.014 0.048 TOP 3.03 -2.12

Page 99: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

83

4.3.5.2 InMnP powder

The temperature-dependent magnetization data shown in Figure 4.11 (A) were

obtained at a constant field at 5 kOe with temperatures ranging from 100K to 5K. The

magnetization was also measured as a function of applied field at 5 K, as shown in Figure

4.11 (B). Pure InP nanocrystals were measured as a baseline and found to be

diamagnetic as expected, with a molar magnetic susceptibility -136 molcm 102.35 −×−=molarχ , very close to the literature value of -136 molcm 106.45 −×−=molarχ .42 Like the InMnAs nanocrystals, the InMnP exhibit

positive susceptibilities before and after chemical surface exchange, indicating

conclusively that Mn incorporates into the nanocrystal core. The InMnP particles have

almost an order of magnitude larger per mass susceptibility than the InMnAs particles

due to the larger amount of Mn incorporated per sample. Also like InMnAs, a constant

field temperature scans of InMnP samples starting at 5K produced no local maxima

susceptibility value, indicating that there isn’t a blocking temperature associated with

superparamagnetic materials. This also is strong evidence that there is no phase

separation into MnP, as it is ferromagnetic in the bulk, and would be superparamagnetic

on these length scales. The molar susceptibility for InMnP in the constant field

temperature scan is larger than that for InMnAs, and is a result of a larger value of

effective Bohr magnetons, which will be calculated below. Again elemental analysis and

magnetization measurements confirm that Mn dopant was kinetically trapped within the

host lattice.

Page 100: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

84

0

1

2

3

4

5

6

7

0 20 40 60 80 100

χmol

ar (x

103 c

m3 /m

ol)

Temperature (K)

A

-10

-5

0

5

10

-60 -40 -20 0 20 40 60

Mag

netiz

atio

n (e

mu/

g)

Field (kOe)

Figure 4.11: Magnetization measurements of MnxIn1-xP nanocrystals. (A) Temperature dependent magnetic molar susceptibility measured under a magnetic field of 5000 Oe, the solid lines are the predicted values as calculated from equation (4.1) in the text and (B) magnetization versus applied magnetic field measured at 5K: (□) dp=3 nm, x=0.15; (+) dp=5 nm, x=0.08 (after pyridine ligand exchange); (○) dp=3 nm, x=0.11 (after pyridine ligand exchange); (◊) dp=4 nm InP.

Page 101: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

85

A closer analysis of the magnetization data provides information about the nature

of the magnetic interactions in the InMnP nanocrystals for comparison with the well-

studied films fabricated by kinetically-controlled MBE. The molar susceptibilities are

relatively low, ranging from -134 molcm 105.7 −×≈molarχ at room temperature up to -133 molcm 101.6 −×≈molarχ at 5K, indicating that the magnetic coupling is weak.

Using the technique in the previous section , the effective Bohr magneton number

and nearest neighbor exchange integral were found by fitting Figure 4.11 (A) to Equation

(4.1), and are listed in Table 4.4. Values of p range from 3.4 to 5.1 atomMn Bµ , which

is slightly lower than the theoretical value of p=5.9. atomMn Bµ . The same arguments

as above apply to possible explanations for the antiferromagnetic interactions between

the Mn atoms, except the reasoning about antisite defects. The In:P ratio is typically

around 55:45, and so there is no obvious evidence of antisite defects. From Figure 4.10

(A), the nearest-neighbor exchange integrals, Jnn kB , ranged from –0.22 to –0.27 K,

indicating slight antiferromagnetic superexchange interactions between the Mn impurity

atoms in the InAs host lattice. These values are lower than the measured bulk value of

KkJ Bnn 6.1−= for n-type Mn-doped InAs, again indicating that antisite defects are not

a major factor in the interactions.46 A possible explanation for the lack of

ferromagnetism is that the Mn atoms aren’t able to interact with carriers within the crystal

because the Mn atoms are not substitutionally doped.

Table 4.4: Effective Bohr magneton number p , and the nearest neighbor exchange integral nnJ for InMnP found by fitting Equation (4.1) to Figure 4.11 (A).

Diameter

(nm) xMn Ligand p (µB/Mn

Atom) Jnn/kB

(K) 3 0.112 Pyridine 5.1 -0.273 0.148 Myristic Acid 3.75 -0.295 0.06 Pyridine 3.96 -0.22

Page 102: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

86

4.3.5.3 InMnP annealed film

5 nm In0.98Mn0.02P particles were annealed on a mica substrate. The temperature

dependent magnetization curves are shown in Figure 4.12 (A). The before anneal curve

follows a paramagnetic curve similar to those in Figure 4.11 (A). Fitting Equation (4.1)

to the data produces a p value of 5.8. atomMn Bµ and a Jnn kB value of -0.07 K. After

annealing, the curve is diamagnetic with a temperature dependence. Figure 4.12 (B) is a

plot of field dependent magnetization data. Both before and after annealing curves are

centered around the origin. ICP-MS verified that the concentration of Mn did not change

during the annealing process.

In studies of II-VI DMS nanocrystals, Dang et al. showed that the magnetic

properties of nanocrystals can change through annealing by changing the crystalline

structure of the material.47 An annealing temperature of 650 °C is not high enough to

force a crystal structure change, but only to remove the pyridine ligand and sinter the

nanocrystals together. It is possible that the anneal speeded up Mn diffusion and caused a

phase separation but X-Ray diffraction and more magnetization characterization in the

form of zero field cooling and lower field temperature sweep data is required to move

beyond speculation.

Page 103: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

87

-4

-2

0

2

4

6

8

10

0 50 100 150 200 250 300

χmol

ar (x

103 c

m3 /m

ol)

Temperature (K)

A

-4

-3

-2

-1

0

1

2

3

4

-60 -40 -20 0 20 40 60

Mag

netiz

atio

n (e

mu/

g)

Field (kOe)

B

Figure 4.12: Magnetization measurements of Mn0.02In0.98P nanocrystals on mica before and after annealing. (A) Temperature dependent magnetic molar susceptibility measured under a magnetic field of 5000 Oe, the solid line is the predicted value as calculated from equation (4.1) and (B) magnetization versus applied magnetic field measured at 5K: (□) dp=5 nm, x=0.2 before anneal; (○) same particles after 650 °C anneal.

Page 104: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

88

4.4 CONCLUSIONS

In conclusion, we have demonstrated that kinetic trapping of dopants in a

nanocrystal host lattice is possible, even in high temperature arrested precipitation

reactions. MnxIn1-xAs and MnxIn1-xP nanocrystals ranging from 2 to 10 nm in diameter

were synthesized with up to xMn=0.025 for InMnAs and xMn=0.11 for InMnP. Surface

exchange and magnetic measurements confirmed that much of the dopant resides in the

nanocrystal core and modifies the magnetic properties of the host material through

antiferromagnetic superexchange interactions. Based upon studies of bulk p-doped

Group III-V DMS materials fabricated by MBE, the doping levels achieved here are

sufficiently high to expect ferromagnetism.48 However, at temperatures as low as 5K,

only paramagnetic behavior was observed for the nanocrystals with the appearance of

antiferromagnetic interactions between the Mn dopant atoms at lower temperatures.

Ferromagnetism has generally been absent in bulk n-doped MnInAs films.46 In

the nanocrystal, each Mn atom contributes significantly less than 5.9 atomMn Bµ (i.e.,

2.2 to 3.7 atom MnBµ ) to the total magnetization of the nanocrystals, which may result

from the donor-acceptor interactions between Mn2+ or dangling bonds and AsIn antisite

defects which would be consistent with n-doping. The InMnAs quantum dot PL was

redshifted and broader than the pure InAs PL due to the presence of the magnetic

impurity.

In the InMnP nanocrystals, each Mn atom contributed slightly less the predicted

value for Mn2+, ranging from 3.4 to 5.1 atomMn Bµ , possibly due to interstitial instead

of substitutional doping. The InMnP nanocrystal PL has a tail similar but less extreme

than that of InMnAs, indicating the presence of the Mn impurity.

Page 105: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

89

Potentially, kinetic impurity trapping in nanocrystals could enable the study of a

wide range of doped nanostructures, including other DMS material, such as Mn-doped

GaP which has exhibited room temperature ferromagnetism in the bulk.49,50

4.5 REFERENCES

(1) Mikulec, F. V.; Kuno, M.; Bennati, M.; Hall, D. A.; Griffin, R. G.; Bawendi, M. G. J. Am. Chem. Soc. 2000, 122, 2532-2540.

(2) Murray, C. B.; Norris, D. J.; Bawendi, M. G. J. Am. Chem. Soc. 1993, 115, 8706-8715.

(3) Norris, D. J.; Yao, N.; Charnock, F. T.; Kennedy, T. A. Nano Lett. 2001, 1, 3-7.

(4) Hanif, K. M.; Meulenberg, R. W.; Strouse, G. F. J. Am. Chem. Soc. 2002, 124, 11495-11502.

(5) Raola, O. E.; Strouse, G. F. Nano Lett. 2002, 2, 1443-1447.

(6) Radovanovic, P. V.; Norberg, N. S.; McNally, K. E.; Gamelin, D. R. J. Am. Chem. Soc. 2002, 124, 15192-15193.

(7) Jun, Y.; Jung, Y.; Cheon, J. J. Am. Chem. Soc. 2002, 124, 615-617.

(8) Illegems, M.; Dingle, R.; Rupp, J. L. W. J. Appl. Phys. 1975, 46, 3059-3065.

(9) Munekata, H.; Ohno, H.; Vonmolnar, S.; Harwit, A.; Segmuller, A.; Chang, L. L. J. Vac. Sci. Technol., B 1990, 8, 176-180.

(10) Ohno, H.; Chiba, D.; Matsukura, F.; Omiya, T.; Abe, E.; Dietl, T.; Ohno, Y.; Ohtani, K. Nature 2000, 408, 944-946.

(11) Furdyna, J. K. J. Appl. Phys. 1988, 64, R29-R64.

(12) Ohno, H. Science 1998, 281, 951-956.

(13) Ferrand, D.; Cibert, J.; Wasiela, A.; Bourgognon, C.; Tatrenko, S.; Fishman, G.; Andrearczak, T.; Jaroszynski, S.; Kolesnik; Dietl, T. Phys. Rev. B 2001, 63, 085201:085201-085213.

(14) Cho, Y. M.; Choo, W. K.; Kim, H.; Kim, D.; Ihm, Y. E. Appl. Phys. Lett. 2002, 80, 3358-3360.

Page 106: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

90

(15) Yoon, S. W.; Cho, S. B.; We, S. C.; Yoon, S.; Suh, B. J.; Song, H. K.; Shin, Y. J. J. Appl. Phys. 2003, 93, 7879-7881.

(16) Sharma, P.; Gupta, A.; Rao, K. V.; Owens, F. J.; Sharma, R.; Ahuja, R.; Osorio, J. M.; Johansson, B.; Gehring, G. A. Nat. Mater. 2003, 2, 673-677.

(17) Theodoropoulou, N. A.; Hebard, A. F.; Norton, D. P.; Budai, J. D.; Boatner, L. A.; Lee, J. S.; Khim, Z. G.; Park, Y. D.; Overberg, M. E.; Pearton, S. J.; Wilson, R. G. Solid-State Electron. 2003, 47, 2231-2235.

(18) Hoffman, D. M.; Meyer, B. K.; Ekimov, A. I.; Merkulov, I. A.; Efros, A. L.; Rosen, M.; Couino, G.; Gacoin, T.; Boilot, J. P. Solid State Commun. 2000, 114, 547-550.

(19) Shim, M.; Wang, C.; Norris, D. J.; Guyot-Sionnest, P. MRS Bulletin 2001, 1005.

(20) Jeon, H. C.; Jeong, Y. S.; Kang, T. W.; Kim, T. W.; Chung, K. J.; Jhe, W.; Song, S. A. Adv. Mater. 2002, 14, 1725-1727.

(21) Guo, S. P.; Ohno, H.; Shen, A.; Matsukura, F.; Ohno, Y. Appl. Surf. Sci. 1998, 132, 797-802.

(22) Somaskandan, K.; Tsoi, G. M.; Wenger, L. E.; Brock, S. L. Chem. Mater. 2005, 17, 1190-1198.

(23) Borse, P. H.; Srinivas, D.; Shinde, R. F.; Date, S. K.; Vogel, W.; Kulkarni, S. K. Phys. Rev. B 1999, 60, 8659-8664.

(24) Norton, D. P.; Pearton, S. J.; Hebard, A. F.; Theodoropoulou, N. A.; Boatner, L. A.; Wilson, R. G. Appl. Phys. Lett. 2003, 82, 239 - 241.

(25) Feltin, N.; Levy, L.; Ingert, D.; Pileni, M. P. J. Phys. Chem. B 1999, 103, 4-10.

(26) Bryan, J. D.; Heald, S. M.; Chambers, S. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 127, 11640-11647.

(27) Norberg, N. S.; Kittilstved, K. R.; Amonette, J. E.; Kukkadapu, R. K.; Schwartz, D. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 126, 9387-9398.

(28) Guzelian, A. A.; Banin, U.; Kadavanich, A. V.; Peng, X.; Alivisatos, A. P. Appl. Phys. Lett. 1996, 69, 1432-1434.

(29) Wells, R. L., Self, M. F., Johansen, J. D., Laske, J. A., Aubuchon, S. R., Jones, L. J. Inorganic Synthesis 1997, 31, 150 - 158.

Page 107: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

91

(30) Battaglia, D.; Peng, X. G. Nano Lett. 2002, 2, 1027-1030.

(31) Glatter, O.; Kratky, O., Eds. Small Angle X-ray Scattering; Academic Press Inc.: New York, 1982.

(32) Lifshin, E., Ed. X-ray Characterization of Materials; Wiley-VCH: New York, 1999.

(33) Korgel, B. A.; Fitzmaurice, D. Phys. Rev. B 1999, 59, 14191-14201.

(34) Korgel, B. A.; Fullam, S.; Connolly, S.; Fitzmaurice, D. J. Phys. Chem. B 1998, 102, 8379-8388.

(35) Mattoussi, H.; Cumming, A. W.; Murray, C. B.; Bawendi, M. G.; Ober, R. Phys. Rev. B 1998, 58, 7850-7863.

(36) Guinier, A.; Fournet, G. Small-Angle Scattering of X-Rays; Wiley: New York, 1955.

(37) Shah, P. S.; Sigman, M. B.; Stowell, C. A.; Lim, K. T.; Johnston, K. P.; Korgel, B. A. Adv. Mater. 2003, 15, 971-974.

(38) Szczytko, J.; Twardowski, A.; Palczewska, M.; Jablonski, R.; Furdyna, J.; Munekata, H. Phys. Rev. B 2001, 63, 085315.

(39) Almelch, N.; Goldstein, B. Phys. Rev. 1962, 128, 1568.

(40) Linnarsson, M.; Janzen, E.; Monemar, B.; Kleverman, M.; Thilderkvist, A. Phys. Rev. B 1997, 55, 6938-6944.

(41) Jain, M.; Kronik, L.; Chelikowsky, J. R.; Godlevsky, V. V. Phys. Rev. B 2001, 64, 245205.

(42) Sahu, T. Phys. Rev. B 1991, 43, 2415-2418.

(43) Ashcroft, N. W.; Mermin, N. D. Solid State Physics; Harcourt College Publishers: New York, 1976.

(44) Bergqvist, L.; Korzhavyi, P. A.; Sanyal, B.; Mirbt, S.; Abrikosov, I. A.; Nordstrom, L.; Smirnova, E. A.; Mohn, P.; Svendlindh, P.; Eriksson, O. Phys. Rev. B 2003, 67, 205201.

(45) Korzhavyi, P. A.; Abrikosov, I. A.; Smirnova, E. A.; Bergqvist, L.; Mohn, P.; Mathieu, R.; Svendlindh, P.; Sadowski, J.; Isaev, E. I.; Vekilov, Y. K.; Eriksson, O. Phys. Rev. Lett. 2002, 88, 187202.

Page 108: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

92

(46) Von Molnar, S.; Munekata, H.; Ohno, H.; Chang, L. L. J. Magn. Magn. Mater. 1991, 93, 356-364.

(47) Dang, Q.; Wen, G.; Sorensen, C. M.; Klabunde, K. Israel Journal of Chemistry 2001, 41, 63-68.

(48) Ohno, H.; Munekata, H.; Penney, T.; von Molnar, S.; Chang, L. L. Phys. Rev. Lett. 1992, 68, 2664-2667.

(49) Overberg, M. E.; Gila, B. P.; Thaler, G. T.; Abernathy, C. R.; Pearton, S. J.; Theodoropoulou, N. A.; McCarthy, K. T.; Arnason, S. B.; Hebard, A. F.; Chu, S. N. G.; Wilson, R. G.; Zavada, J. M.; Park, Y. D. J. Vac. Sci. Technol., B 2002, 20, 969-979.

(50) Pearton, S. J.; Overberg, M. E.; Thaler, G. T.; Abernathy, C. R.; Kim, J.; Ren, F.; Theodoropoulou, N. A.; Hebard, A. F.; Park, Y. D. Phys. Status Solidi A 2003, 195, 222-227.

Page 109: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

93

Chapter 5: Conclusions and Recommendations

5.1 CONCLUSIONS

5.1.1 Fluid patterning through fluid dynamics

Experimental observations discussed in Chapter 2 demonstrated that within a

narrow concentration range, short wavelength Marangoni instabilities can occur during

the solvent evaporation of 3.5 nm gold nanocrystal solutions. After complete solvent

evaporation cellular honeycomb networks are left on the substrate, the interfaces of

which consist of the monodisperse sterically–stabilized 3.5 nm diameter gold

nanocrystals. The lattice parameter of these systems is 4.3 µm. At higher concentrations

of nanocrystals dispersed in chloroform, organized networks were not observed, instead

1.7 µm diameter rings form as a result of the dewetting characteristics of the solvent.

The honeycomb array formation is governed by the Marangoni number, a

dimensionless number in which surface tension plays a key role. The surface tension,

and change in surface tension with respect to temperature are strongly dependant on

nanocrystal concentration. These values impact the Marangoni number, which must be

Ma>80 for surface tension driven convection to occur, making surface tension driven

convection concentration dependant. It was also noted that nanocrystal systems with

polydisperse particle sizes did not form hexagonal arrays due to irregular surface tension

fields.

5.1.2 Catalysis of iridium nanocrystals

Ir nanocrystal catalysis of 1-decene hydrogenation to decane is extremely

sensitive to capping ligand chemistry and ligand coverage. “Good” capping ligands—

i.e., those that stabilize robust nanocrystals with very narrow size distributions—appear

Page 110: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

94

to be poor choices for catalytic applications. However, the nanocrystals must be of

reasonably high quality with good dispersibility in multiple reaction cycles, so the ligands

must be strong enough binders to stabilize nanocrystals, but “weak” enough to provide

reactant access to the metal surface. The work discussed in Chapter 3 revealed that

stabilizing ligands play a central role in the catalytic activity of transition metal

nanocrystals and cannot be ignored.

5.1.3. III-V Dilute magnetic semiconductors

MnxIn1-xAs and MnxIn1-xP nanocrystals ranging from 2 to 10 nm in diameter were

synthesized with up to xMn=0.025 for InMnAs and xMn=0.11 for InMnP. Surface

exchange and magnetic measurements confirmed that much of the dopant resides in the

nanocrystal core and modifies the magnetic properties of the host material through

antiferromagnetic superexchange interactions. At temperatures as low as 5K, only

paramagnetic behavior was observed for the nanocrystals with the appearance of

antiferromagnetic interactions between the Mn dopant atoms at lower temperatures.

In the InMnAs nanocrystals, each Mn atom contributes significantly less than 5.9

atomMn Bµ (i.e., 2.2 to 3.7 atom MnBµ ) to the total magnetization of the

nanocrystals, which may result from the donor-acceptor interactions between Mn2+ or

dangling bonds and AsGa antisite defects which would be consistent with n-doping.

In the InMnP nanocrystals, each Mn atom contributed slightly less the predicted

value for Mn2+, ranging from 3.4 to 5.1 atomMn Bµ , possibly due to interstitial instead

of substitutional doping.

Page 111: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

95

5.2 RECOMMENDATIONS FOR FUTURE WORK

5.2.1 Iridium nanocrystals

Since the new synthesis technique described in Chapter 3 is robust and suitable

for several stabilizing ligands, the nanocrystals produced would be ideal seeds for core-

shell nanocrystal synthesis. A particularly interesting system to study would be the

iridium-gold system, which has already proven unique in terms of catalytic selectivity in

thin film systems.1 There are many synthetic routes available for the formation of such

structures, as other catalytic core-shell particles have already been synthesized, like

gold/palladium, gold/platinum, and platinum/ruthenium.2-4

5.2.2 Dilute magnetic semiconductor nanocrystals

5.2.2.1 Other III-V systems

The work in Chapter 4 focused on InMnAs and InMnP because InAs and InP

nanocrystals had the most repeatable existing synthesis procedures. While doped GaAs,

GaN, and GaP would have higher ferromagnetic transition temperatures,5 the reliable

colloidal synthetic routes to those materials weren’t available at the time this work was

completed. Within the past year, some progress has been made on other reactions to

produce III-V semiconductors,6,7 and they may be repeatable and suitable for doping

experiments in different III-V systems.

5.2.2.2 III-V DMS wires

While synthesis routes towards certain III-V semiconductor nanocrystals may not

exist, there are synthetic routes to other geometries, like wires and whiskers that are

available.8,9 While most DMS studies have focused on either nanocrystals or thin films,

DMS wires have been ignored.

Page 112: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

96

5.2.2.3 Biomedical applications

Recently Tanaka et al. suggested CdS:Mn/ZnS as a probe for biological

applications, relying on optical and magnetic signals.10 However, a major concern for

this system is the extreme toxicity of cadmium. InMnP would also make a suitable probe

system if outfitted with a water soluble capping ligand, but may not suffer from the same

toxicity as cadmium.11

5.3 REFERENCES

(1) Okada, M.; Nakamura, M.; Moritani, K.; Kasai, T. Surf. Sci. 2003, 523, 218-230.

(2) Lou, Y.; Maye, M.; Han, L.; Luo, J.; Zhong, C. Chem. Commun. 2001, 5.

(3) Nashner, M. S.; Frenkel, A. I.; Somerville, D.; Hills, C. W.; Shapley, J. R.; Nuzzo, R. G. J. Am. Chem. Soc. 1998, 120, 8093-8101.

(4) Harpeness, R.; Gedanken, A. Langmuir 2004, 20.

(5) Dietl, T.; Ohno, H.; Matsukura, F.; Cibert, J.; Ferrand, D. Science 2000, 287, 1019-1022.

(6) Liu, F. S.; Liu, Q. L.; Liang, J. K.; Song, G. B.; Yang, L. T.; Luo, J.; Zhou, Y. Q.; Dong, H. W.; Rao, G. H. J. MATER. RES. 2004, 19, 3484-3489.

(7) Gao, S. M.; Zhu, L. Y.; Xie, Y.; Qian, X. B. EUR. J. INORG. CHEM. 2003, 3, 557-561.

(8) Duan, X.; Huang, Y.; Cui, Y.; Wang, J.; Lieber, C. M. Nature 2001, 409, 66.

(9) Hiruma, K.; Yazawa, M.; Katsuyama, T.; Ogawa, K.; Haraguchi, K.; Koguchi, M.; Kakibayashi, H. J. Appl. Phys. 1995, 77, 447-462.

(10) Santra, S.; Yang, H.; Holloway, P. H.; Stanley, J. T.; Mericle, R. A. J. Am. Chem. Soc. 2005, 127, 1656-1657.

(11) Tanaka, A.; Hirata, M.; Omura, M.; Inoue, N.; Ueno, T.; Homma, T.; Sekizawa, K. J. Occup. Health 2002, 44.

Page 113: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

97

Bibliography

Ahmadi, T. S.; Wang, Z. L.; Henglein, A.; ElSayed, M. A. Chem. Mater. 1996, 8, 1161.

Aiken_III, J. D.; Finke, R. G. J. Mol. Catal. A: Chem. 1999, 145, 1-44.

Alder, B. J.; Wainwright, T. E. Phys. Rev. 1962, 127, 459-361.

Almelch, N.; Goldstein, B. Phys. Rev. 1962, 128, 1568.

Anderson, J. R.; Howe, R. F. Nature 1977, 268, 129-130.

Ashcroft, N. W.; Mermin, N. D. Solid State Physics; Harcourt College Publishers: New York, 1976.

Bartlett, P.; Ottweill, R. H.; Pusey, P. N. Phys. Rev. Lett. 1992, 68, 3801-3804.

Battaglia, D.; Peng, X. G. Nano Lett. 2002, 2, 1027-1030.

Bergqvist, L.; Korzhavyi, P. A.; Sanyal, B.; Mirbt, S.; Abrikosov, I. A.; Nordstrom, L.; Smirnova, E. A.; Mohn, P.; Svendlindh, P.; Eriksson, O. Phys. Rev. B 2003, 67, 205201.

Bénard, H. Rev. Gen. Sci. Pures Appl. Bull. 1900, 11, 1261-1271.

Bonnemann, H.; Brijoux, W.; Brinkmann, R.; Dinjus, E.; Joussen, T.; Korall, B. Angew. Chem. Int. Ed. Engl. 1991, 30, 1312-1314.

Borse, P. H.; Srinivas, D.; Shinde, R. F.; Date, S. K.; Vogel, W.; Kulkarni, S. K. Phys. Rev. B 1999, 60, 8659-8664.

Brus, L. E. J. Chem. Phys. 1983, 79, 5566.

Brust, M.; Walker, M.; Bethell, D.; Schiffrin, D. J.; Whyman, R. J. J. Chem. Soc. Chem. Commun. 1994, 801-802.

Bryan, J. D.; Heald, S. M.; Chambers, S. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 127, 11640-11647.

Chambers, S. A.; Yoo, Y. K. MRS Bulletin 2003, 28, 706-707.

Page 114: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

98

Cho, Y. M.; Choo, W. K.; Kim, H.; Kim, D.; Ihm, Y. E. Appl. Phys. Lett. 2002, 80, 3358-3360.

Coey, J. M. D.; Chien, C. L. MRS Bulletin 2003, 28, 720 - 724.

Dahlenburg, L.; Gotz, R. Eur. J. Inorg. Chem 2004, 888-905.

Dang, Q.; Wen, G.; Sorensen, C. M.; Klabunde, K. Israel Journal of Chemistry 2001, 41, 63-68.

Denkov, N. D.; Velev, O. D.; Kralchevsky, P. A.; Ivanov, I. B.; Yoshimura, H.; Nagayama, K. Langmuir 1992, 8, 3183-3190.

Dietl, T.; Ohno, H.; Matsukura, F.; Cibert, J.; Ferrand, D. Science 2000, 287, 1019-1022.

Dietl, T.; Ohno, H. MRS Bulletin 2003, 28, 714 - 719.

Duan, X.; Huang, Y.; Cui, Y.; Wang, J.; Lieber, C. M. Nature 2001, 409, 66.

Dumonteil, C.; Lacroix, M.; Geantet, C.; Jobic, H.; Breysse, M. J. Catal. 1999, 187, 464-473.

Dupont, J.; Fonseca, G. S.; Umpierre, A. P.; Fichtner, P. F. P.; Teixeira, S. R. J. Am. Chem. Soc. 2002, 124, 4228-4229.

Faraday, M. Phil. Trans. Roy. Soc. 1857, 147, 145.

Feltin, N.; Levy, L.; Ingert, D.; Vincent, E.; Pileni, M. P. J. Appl. Phys. 2000, 87, 1415-1423.

Ferrand, D.; Cibert, J.; Wasiela, A.; Bourgognon, C.; Tatrenko, S.; Fishman, G.; Andrearczak, T.; Jaroszynski, S.; Kolesnik; Dietl, T. Phys. Rev. B 2001, 63, 085201:085201-085213.

Fonseca, G. S.; Umpierre, A. P.; Fichtner, P. F. P.; Teixeira, S. R.; Dupont, J. Chem. Eur. J. 2003, 9, 3263-3269.

Fonseca, G. S.; Scholten, J. D.; Dupont, J. Synnlett 2004, 1525-1528.

Furdyna, J. K. J. Appl. Phys. 1988, 64, R29-R64.

Galazka, R. R. In 14th International Conference on the Physics of Semiconductors; Wilson, B. L. H., Ed.; Inst. of Phys. Conf. Series: Edinburgh, 1978; Vol. 43, p 133.

Gao, S. M.; Zhu, L. Y.; Xie, Y.; Qian, X. B. EUR. J. INORG. CHEM. 2003, 3, 557-561.

Page 115: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

99

Geohegan, D. B.; A.A., P.; Duscher, G.; Pennycook, S. J. Appl. Phys. Lett. 1998, 72, 2987-2989.

Glatter, O.; Kratky, O., Eds. Small-Angle X-ray Scattering; Academic Press Inc.: New York, 1982.

Gray, J. J.; Klein, D. H.; Bonnecaze, R. T.; Korgel, B. A. Phys. Rev. Lett. 2000, 85, 4430-4433.

Gray, J. J.; Klein, D. H.; Korgel, B. A.; Bonnecaze, R. T. Langmuir 2001, 17, 2317-2328.

Guinier, A.; Fournet, G. Small-Angle Scattering of X-rays; Wiley: New York, 1955.

Guo, S. P.; Ohno, H.; Shen, A.; Matsukura, F.; Ohno, Y. Appl. Surf. Sci. 1998, 132, 797-802.

Guzelian, A. A.; Banin, U.; Kadavanich, A. V.; Peng, X.; Alivisatos, A. P. Appl. Phys. Lett. 1996, 69, 1432-1434.

Ha, V. M.; Lai, C. L. Proc. R. Soc. Lond. A 2001, 457, 885-909.

Hanif, K. M.; Meulenberg, R. W.; Strouse, G. F. J. Am. Chem. Soc. 2002, 124, 11495-11502.

Harpeness, R.; Gedanken, A. Langmuir 2004, 20.

Hayek, K.; Goller, H.; Penner, S.; Rupprechter, G.; Zimmermann, C. Catal. Lett. 2004, 92, 1-9.

Heinrich, J. L.; Curtis, C. L.; Credo, G. M.; Kavanagh, K. L.; Sailor, M. J. Science 1992, 255, 66-68.

Hiruma, K.; Yazawa, M.; Katsuyama, T.; Ogawa, K.; Haraguchi, K.; Koguchi, M.; Kakibayashi, H. J. Appl. Phys. 1995, 77, 447-462.

Hoffman, D. M.; Meyer, B. K.; Ekimov, A. I.; Merkulov, I. A.; Efros, A. L.; Rosen, M.; Couino, G.; Gacoin, T.; Boilot, J. P. Solid State Commun. 2000, 114, 547-550.

Hoke, J. B.; Stern, E. W.; Murray, H. H. J. Mater. Chem. 1991, 1, 551-554.

Housecroft, C. E.; O'Neill, M. E.; Wade, K.; Smith, B. C. J. Organomet. Chem. 1981, 213, 35-43.

Illegems, M.; Dingle, R.; Rupp, J. L. W. J. Appl. Phys. 1975, 46, 3059-3065.

Page 116: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

100

Israelachvili, J. Intermolecular & Surface Forces; 2nd ed.; Acadmeic Press: San Diego, 1992.

Jain, M.; Kronik, L.; Chelikowsky, J. R.; Godlevsky, V. V. Phys. Rev. B 2001, 64, 245205.

Jeon, H. C.; Jeong, Y. S.; Kang, T. W.; Kim, T. W.; Chung, K. J.; Jhe, W.; Song, S. A. Adv. Mater. 2002, 14, 1725-1727.

Jonker, B. T.; Erwin, S. C.; Petrouw, A.; Petukhov, A. G. MRS Bulletin 2003, 28, 740 - 748.

Jun, Y.; Jung, Y.; Cheon, J. J. Am. Chem. Soc. 2002, 124, 615-617.

Kiely, C. J.; Fink, J.; Brust, M.; Bethell, D.; Schiffrin, D. J. Nature 1998, 396, 444-446.

Kiely, C. J.; Fink, J.; Zheng, J. G.; Brust, M.; Bethell, D.; Schiffrin, D. J. Adv. Mater. 2000, 12, 640.

Kim, S. W.; Park, J.; Jang, Y.; Chung, Y.; Hwang, S.; Hyeon, T.; Kim, Y. W. Nano Lett. 2003, 3, 1289-1291.

Klabunde, K., Ed.; John Wiley and Sons, Inc.: New York, 2001.

Korgel, B. A.; Fullam, S.; Connolly, S.; Fitzmaurice, D. J. Phys. Chem. B 1998, 102, 8379-8388.

Korgel, B. A.; Fitzmaurice, D. Phys. Rev. B 1999, 59, 14191-14201.

Korzhavyi, P. A.; Abrikosov, I. A.; Smirnova, E. A.; Bergqvist, L.; Mohn, P.; Mathieu, R.; Svendlindh, P.; Sadowski, J.; Isaev, E. I.; Vekilov, Y. K.; Eriksson, O. Phys. Rev. Lett. 2002, 88, 187202.

Kralik, M.; Biffis, A. J. Mol. Catal. A: Chem. 2001, 177, 113-138.

Kruis, F. E.; Fissan, H.; Peled, A. J. AEROSOL SCIENCE 1998, 29, 511-535.

Lewis, L. N. Chem. Rev. 1993, 93, 2693-2730.

Li, Y.; El-Sayed, M. A. J. Phys. Chem. B 2001, 105, 8938-8943.

Lifshing, E., Ed. X-ray Characterization of Materials; Wiley-VCH: New York, 1999.

Lin, Y.; Nomiya, K.; Finke, R. G. Inorg. Chem. 1993, 32, 6040-6045.

Lin, Y.; Finke, R. G. J. Am. Chem. Soc. 1994, 116, 8335-8353.

Page 117: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

101

Linnarsson, M.; Janzen, E.; Monemar, B.; Kleverman, M.; Thilderkvist, A. Phys. Rev. B 1997, 55, 6938-6944.

Liu, F. S.; Liu, Q. L.; Liang, J. K.; Song, G. B.; Yang, L. T.; Luo, J.; Zhou, Y. Q.; Dong, H. W.; Rao, G. H. J. MATER. RES. 2004, 19, 3484-3489.

Lou, Y.; Maye, M.; Han, L.; Luo, J.; Zhong, C. Chem. Commun. 2001, 5.

Luck, V. W.; Kleir, M.; Wesslau, H. Ber. Bunsen-Ges. Phys. Chem. 1963, 67, 75.

Maillard, M.; Motte, L.; Ngo, A. T.; Pileni, M. P. J. Phys. Chem. B. 2000, 104, 11871-11877.

Maillard, M.; Motte, L.; Pileni, M. P. Adv. Mater. 2001, 13, 200-204.

Mattoussi, H.; Cumming, A. W.; Murray, C. B.; Bawendi, M. G.; Ober, R. J. Chem. Phys. 1996, 105, 9890-9896.

Mattoussi, H.; Cumming, A. W.; Murray, C. B.; Bawendi, M. G.; Ober, R. Phys. Rev. B 1998, 58, 7850-7863.

McConnell, G. A.; Gast, A. P.; Huang, J. S.; Smith, S. D. Phys. Rev. Lett. 1993, 71, 2102-2105.

Mevellec, V.; Roucoux, A.; Ramirez, E.; Phillippot, K.; Chaudret, B. Adv. Synth. Catal. 2004, 346, 72-76.

Mikulec, F. V.; Kuno, M.; Bennati, M.; Hall, D. A.; Griffin, R. G.; Bawendi, M. G. J. Am. Chem. Soc. 2000, 122, 2532-2540.

Moriarty, P.; Taylor, M. D. R.; Brust, M. Phys. Rev. Lett. 2002, 89, 248303-248301 - 248303-248304.

Munekata, H.; Ohno, H.; Vonmolnar, S.; Harwit, A.; Segmuller, A.; Chang, L. L. J. Vac. Sci. Technol. B 1990, 8, 176-180.

Murray, C. B.; Norris, D. J.; Bawendi, M. G. J. Am. Chem. Soc. 1993, 115, 8706-8715.

Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Science 1995, 270, 1335-1338.

Murray, C. B.; Kagan, C. R.; Bawendi, M. G. Annu. Rev. Mater. Sci. 2000, 30, 545-610.

Narayanan, R.; El-Sayed, M. A. J. Am. Chem. Soc. 2003, 125, 8340-8347.

Narayanan, S.; Wang, J.; Lin, X. M. Phys. Rev. Lett. 2004, 93, 135503.

Narayanan, R.; El-Sayed, M. A. Nano Lett. 2004, 4, 1343-1348.

Page 118: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

102

Narayanan, R.; El-Sayed, M. A. J. Phys. Chem. B 2004, 108, 5726-5733.

Nashner, M. S.; Frenkel, A. I.; Somerville, D.; Hills, C. W.; Shapley, J. R.; Nuzzo, R. G. J. Am. Chem. Soc. 1998, 120, 8093-8101.

Norberg, N. S.; Kittilstved, K. R.; Amonette, J. E.; Kukkadapu, R. K.; Schwartz, D. A.; Gamelin, D. R. J. Am. Chem. Soc. 2004, 126, 9387-9398.

Norris, D. J.; Yao, N.; Charnock, F. T.; Kennedy, T. A. Nano Lett. 2001, 1, 3-7.

Norton, D. P.; Pearton, S. J.; Hebard, A. F.; Theodoropoulou, N. A.; Boatner, L. A.; Wilson, R. G. Appl. Phys. Lett. 2003, 82, 239 - 241.

Ohara, P. C.; Heath, J. R.; Gelbart, W. M. Angew. Chem. Int. Ed. Engl. 1997, 36, 1077.

Ohara, P. C.; Gelbart, W. M. Langmuir 1998, 14, 3418-3424.

Ohno, H.; Munekata, H.; Penney, T.; von Molnar, S.; Chang, L. L. Phys. Rev. Lett. 1992, 68, 2664-2667.

Ohno, H. Science 1998, 281, 951-956.

Ohno, H.; Chiba, D.; Matsukura, F.; Omiya, T.; Abe, E.; Dietl, T.; Ohno, Y.; Ohtani, K. Nature 2000, 408, 944-946.

Okada, M.; Nakamura, M.; Moritani, K.; Kasai, T. Surf. Sci. 2003, 523, 218-230.

Overberg, M. E.; Gila, B. P.; Thaler, G. T.; Abernathy, C. R.; Pearton, S. J.; Theodoropoulou, N. A.; McCarthy, K. T.; Arnason, S. B.; Hebard, A. F.; Chu, S. N. G.; Wilson, R. G.; Zavada, J. M.; Park, Y. D. J. Vac. Sci. Technol. B 2002, 20, 969-979.

Palstrøm, C. MRS Bulletin 2003, 28, 725-728.

Park, Y. D.; Hanbicki, A. T.; Erwin, S. C.; Hellberg, C. S.; Sullivan, J. M.; Mattson, J. E.; Ambrose, T. F.; Wilson, A.; Spanos, G.; Jonker, S. T. Science 2002, 295, 651-654.

Pearson, J. R. A. J. Fluid Mech. 1958, 4, 489-500.

Pearton, S. J.; Overberg, M. E.; Thaler, G. T.; Abernathy, C. R.; Kim, J.; Ren, F.; Theodoropoulou, N. A.; Hebard, A. F.; Park, Y. D. Phys. Status Solidi A 2003, 195, 222-227.

Probstein, R. F. Physiochemical Hydrodynamics: an Introduction; Butterworths: Boston, 1989.

Page 119: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

103

Pusey, P. N.; VanMegen, W. Nature 1986, 320, 340-342.

Rabani, E.; Reichman, D. R.; Geissler, P. L.; Brus, L. E. Nature 2003, 426, 271-274.

Radovanovic, P. V.; Norberg, N. S.; McNally, K. E.; Gamelin, D. R. J. Am. Chem. Soc. 2002, 124, 15192-15193.

Raola, O. E.; Strouse, G. F. Nano Lett. 2002, 2, 1443-1447.

Reiter, G. Phys. Rev. Lett. 1991, 68, 75-78.

Roucoux, A.; Schulz, J.; Patin, H. Chem. Rev. 2002, 102, 3757-3778.

Sahu, T. Phys. Rev. B 1991, 43, 2415-2418.

Santra, S.; Yang, H.; Holloway, P. H.; Stanley, J. T.; Mericle, R. A. J. Am. Chem. Soc. 2005, 127, 1656-1657.

Sato, K.; Katayama-Yoshida, H. Phys. Stat. Sol. B 2002, 229, 673 - 680.

Saunders, J. V.; Murray, M. J. Nature 1978, 275, 201-203.

Saunders, J. V. Philos. Mag. A 1980, 42, 705.

Saunders, A. E.; Shah, P. S.; Sigman, M. B.; Hanrath, T.; Hwang, H. S.; Lim, K. T.; Johnston, K. P.; Korgel, B. A. Nano Lett. 2004, 4, 1943-1948.

Saunders, A. E.; Sigman, M. B.; Korgel, B. A. J. Phys. Chem. B 2004, 108, 193-199.

Schatz, M. F.; VanHook, S. J.; McCormick, W. D.; Swift, J. B.; Swinney, H. L. Phys. Rev. Lett. 1995, 75, 1938-1941.

Shah, P. S.; Husain, S.; Johnston, K. P.; Korgel, B. A. J. Phys. Chem. B 2001, 105, 9433-9440.

Shah, P. S.; Sigman, M. B.; Stowell, C. A.; Lim, K. T.; Johnston, K. P.; Korgel, B. A. Adv. Mater. 2003, 15, 971-974.

Sharma, A.; Reiter, G. J. Colloid Interface Sci. 1996, 178, 383-399.

Sharma, P.; Gupta, A.; Rao, K. V.; Owens, F. J.; Sharma, R.; Ahuja, R.; Osorio, J. M.; Johansson, B.; Gehring, G. A. Nat. Mater. 2003, 2, 673-677.

Shiblisma, N. Int. J. Hydrogen Energy 1993, 18, 141-147.

Shim, M.; Wang, C.; Norris, D. J.; Guyot-Sionnest, P. MRS Bulletin 2001, 1005.

Page 120: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

104

Somaskandan, K.; Tsoi, G. M.; Wenger, L. E.; Brock, S. L. Chem. Mater. 2005, 17, 1190-1198.

Son, S. U.; Jang, Y.; Yoon, K. Y.; Kang, E.; Hyeon, T. Nano Lett. 2004, 4, 1147-1151.

Sondi, I.; Siiman, O.; Koester, S.; Matijevic, E. Langmuir 2000, 16, 3107-3118.

Stewart, I. Life's other secret - The new mathematics of the living world; John Wiley & Sons, Inc: New York, 1998.

Sugano, S.; Kojima, N., Eds. Magneto-optics; Springer: New York, 2000.

Sun, S.; Murray, C. B.; Weller, D.; Folks, L.; Moser, A. Science 2000, 287, 1989-1992.

Szczytko, J.; Twardowski, A.; Palczewska, M.; Jablonski, R.; Furdyna, J.; Munekata, H. Phys. Rev. B 2001, 63, 085315.

Tan, K. K.; Thorpe, R. B. Chem. Eng. Sci. 1999, 54, 775-783.

Tanaka, A.; Hirata, M.; Omura, M.; Inoue, N.; Ueno, T.; Homma, T.; Sekizawa, K. J. Occup. Health 2002, 44.

Theodoropoulou, N. A.; Hebard, A. F.; Norton, D. P.; Budai, J. D.; Boatner, L. A.; Lee, J. S.; Khim, Z. G.; Park, Y. D.; Overberg, M. E.; Pearton, S. J.; Wilson, R. G. Solid-State Electron. 2003, 47, 2231-2235.

Thiele, U.; Mertig, M.; Pompe, W. Phys. Rev. Lett. 1998, 80, 2869-2872.

Truskett, V. H.; Stebe, K. Langmuir 2003, 19, 8271-8279.

Valden, M.; Lai, X.; Goodman, D. W. Science 1998, 281, 1647-1650.

VanHook, S. J.; Schatz, M. F.; McCormick, W. D.; Swift, J. B.; Swinney, H. L. Phys. Rev. Lett. 1995, 75, 4397-4400.

VanHook, S. J.; Schatz, M. F.; Swift, J. B.; McCormick, W. D.; Swinney, H. L. J. Fluid Mech. 1997, 345, 45-78.

Velev, O. D.; Lenhoff, A. M.; Kaler, E. W. Science 2000, 287, 2240-2243.

Von Molnar, S.; Munekata, H.; Ohno, H.; Chang, L. L. J. Magn. Magn. Mater. 1991, 93, 356-364.

Wells, R. L., Self, M. F., Johansen, J. D., Laske, J. A., Aubuchon, S. R., Jones, L. J. Inorganic Synthesis 1997, 31, 150 - 158.

Page 121: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

105

Whetten, R. L.; Shafigullin, M. N.; Khoury, J. T.; Schaaff, T. G.; Vexmar, I.; Alvarex, M. M.; Wilkinson, A. Acc. Chem. Res. 1999, 32, 397-406.

Yee, C. K.; Jordan, R.; Ulman, A.; White, H.; King, A.; Rafailovich, M.; Sokolov, J. Langmuir 1999, 15, 3486-3491.

Yoon, S. W.; Cho, S. B.; We, S. C.; Yoon, S.; Suh, B. J.; Song, H. K.; Shin, Y. J. J. Appl. Phys. 2003, 93, 7879-7881.

Page 122: Copyright by Cynthia Stowell 2005 · John G. Ekerdt Miguel Jose-Yacaman Allan H. MacDonald . Aspects of Colloidal Nanocrystals: Patterning, Catalysis and Doping by Cynthia Ann Stowell,

106

Vita

Cynthia Ann Stowell, the eldest of two children, was born in Philadelphia,

Pennsylvania on August 16, 1975 to Carole Nancy Stowell and Frederick William

Stowell. Cindy graduated Council Rock High School in Newtown, Pennsylvania in 1993

and entered Virginia Polytechnic Institute and State University in Blacksburg, Virginia

the following fall. She alternated semesters working a co-op job for Bechtel at the

Savannah River Site and graduated cum laude from Virginia Polytechnic Institute and

State University in December 1997 with a Bachelor of Science degree in chemical

engineering. After graduation, she worked for a year and a half at Lockheed Martin in

Manassas, Virginia as a photolithography engineer. In August of 1999 she entered

graduate school at the University of Texas at Austin in the department of chemical

engineering. While doing research on colloidal nanocrystals for six years under Dr.

Brian Korgel, she studied the Russian language, joined the UT ultimate frisbee team, and

after numerous injuries instead took up writing for the college newspaper, the Daily

Texan.

Permanent address: 32 Williams Way, Downingtown, Pennsylvania 19335

This dissertation was typed by the author.