<<

The Pennsylvania State University

The Graduate School

Eberly College of Science

STRATEGIES FOR THE

OF INORGANIC

A Dissertation in

Chemistry

by

Matthew Robert Buck

Submitted in Partial Fulfillment of the Requirements for the Degree of

Doctor of Philosophy

August 2013

The dissertation of Matthew R. Buck was reviewed and approved* by the following:

Raymond E. Schaak Professor of Dissertation Advisor Chair of Committee

Thomas E. Mallouk Evan Pugh Professor of Materials Chemistry and Physics Associate Head of the Department of Chemistry

Christine D. Keating Professor of Chemistry

Michael Hickner Assistant Professor of and Engineering

Barbara J. Garrison Shapiro Professor of Chemistry Head of the Department of Chemistry

*Signatures are on file in the Graduate School

iii ABSTRACT

Colloidal nanochemistry is a maturing branch of chemistry that is concerned with the liquid-phase synthesis, manipulation, and properties of nanometer-scale inorganic building blocks that have prescribed sizes, shapes, compositions, and surfaces. Efforts to construct functional materials from such building blocks represent a promising alternative approach to the conventional miniaturization techniques developed for modern microelectronics. Hence, the individual inorganic building blocks can be considered as tailorable fundamental units, from which new materials with designer properties can be constructed. Colloidal nanochemistry is advantageous in this endeavor, because it has the potential to enable exquisite morphological control over individual nanocrystallites, and to permit inexpensive, high-throughput, liquid-phase processing into hierarchical nanostructures.

To address the demand for complex, next-generation with rigorously controlled architectures and interfaces using colloidal nanochemistry, synthetic innovations must be developed. In this dissertation, the total synthesis framework used by organic to prepare complex has been used as a source of inspiration for approaching the colloidal synthesis of such multi-component inorganic nanostructures. Within this framework, we emphasize the importance and utility of single-component crystalline , which can be treated as synthons that can be chemically transformed into more complex derivatives. Furthermore, those transformations can be categorized into classes of reactions that are related to those used in the synthesis of organic molecules. We also demonstrate that concepts used in

iv molecular synthesis – including site-selectivity, regio- and chemoselectivity, orthogonal reactivity, coupling reactions, and protection-deprotection strategies – can be applied to colloidal nanochemistry. Collectively, this synthetic framework represents an emerging paradigm in the synthesis of complex inorganic nanostructures: applying the guiding principles that underpin the multi-step “total synthesis” framework used to produce complex organic molecules and natural products to the stepwise construction of complex multi-component nanostructures.

v TABLE OF CONTENTS

LIST OF FIGURES ...... viii

ACKNOWLEDGEMENTS ...... xix

Chapter 1 Emerging Strategies for the Total Synthesis of Inorganic Nanostructures ...... 1

1.1 Introduction ...... 1 1.2 Colloidal synthesis of single-component crystalline nanoparticles ...... 7 1.3 The “working principles” of molecular total synthesis ...... 11 1.4 reaction libraries ...... 15 1.5 Reaction toolkit for multi-component inorganic nanostructures ...... 25 1.5.1 Orthogonal reactivity and site-selective transformations ...... 26 1.5.2 Merging of multi-domain fragments: “coupling” reactions ...... 46 1.5.3 Protection-deprotection strategies ...... 50 1.5.4 Stepwise construction of complex multi-component inorganic nanostructures ...... 53 1.5.5 Purification, separation, and yield ...... 57 1.6 Research ...... 60 1.7 References ...... 64

Chapter 2 Liquid-Phase Synthesis of Uniform Cube-Shaped GeTe Microcrystals ..... 80

2.1 Introduction ...... 80 2.2 Experimental Details ...... 82 2.2.1 Materials ...... 82 2.2.2 Synthesis of GeTe microcrystallites ...... 83 2.2.3 Conversion of vapor-deposited a-GeTe thin films to GeTe ...... 83 2.2.4 Characterization ...... 84 2.3 Results and Discussion ...... 84 2.4 Conclusions ...... 91 2.5 References ...... 92

Chapter 3 Polymer-Assisted Synthesis of Colloidal Germanium Telluride Nano- Octahedra, Nanospheres, and Nanosheets ...... 96

3.1 Introduction ...... 96 3.2 Experimental Details ...... 99 3.2.1 Materials ...... 99 3.2.2 Synthesis of faceted GeTe nanoparticles ...... 99

vi

3.2.3 Synthesis of amorphous GexTe1-x nanospheres and single-crystal GeTe nanosheets ...... 100 3.2.4 Characterization ...... 101 3.3 Results and Discussion ...... 102 3.3.1 Faceted GeTe nanoparticles ...... 102 3.3.2 Amorphous GexTe1-x alloy nanospheres ...... 111 3.3.3 Single-crystal GeTe nanosheets ...... 115 3.4 Conclusions ...... 121 3.5 References ...... 122

Chapter 4 Investigating the Thermal Decomposition of Co(II) Oleate for Synthesis of Colloidal Wurtzite CoO Nanocrystals ...... 129

4.1 Introduction ...... 129 4.2 Experimental Details ...... 134 4.2.1 Materials ...... 134 4.2.2 Preparation of crude Co(OL)2 ...... 134 4.2.3 Purification of crude Co(OL)2 ...... 135 4.2.4 Synthesis of pencil-shaped wz-CoO nanocrystals ...... 135 4.2.5 Synthesis of cone- and golf tee-shaped wz-CoO nanocrystals ...... 136 4.2.6 Synthesis of bullet-shaped wz-CoO nanocrystals ...... 136 4.2.7 Characterization ...... 137 4.3 Results and Discussion ...... 137 4.3.1 Preparation, purification and FTIR analysis of Co(OL)2 precursors .... 137 4.3.2 Pencil-shaped wz-CoO nanocrystals ...... 141 4.3.3 Cone- and golf tee-shaped wz-CoO nanocrystals ...... 142 4.3.4 Bullet-shaped wz-CoO nanocrystals ...... 147 4.4 Conclusions ...... 152 4.5 References ...... 153

Chapter 5 Engineering Solid-State Complexity into Segmented Metal Using Site-Selective Solution Chemistry ...... 158

5.1 Introduction ...... 158 5.2 Experimental Details ...... 160 5.2.1 electrodeposition and on-wire conversion ...... 160 5.2.2 Characterization ...... 161 5.3 Results and Discussion ...... 161 5.4 Conclusions ...... 165 5.5 References ...... 166

Chapter 6 A Total Synthesis Framework for the Construction of High-Order Colloidal Hybrid Nanoparticles ...... 170

6.1 Introduction ...... 170

vii 6.2 Experimental Details ...... 171 6.2.1 Materials ...... 171 6.2.2 Synthesis of Pt nanoparticles ...... 172 6.2.3 Synthesis of Pt-Fe3O4 heterodimers ...... 173 6.2.4 Synthesis of Au-Pt-Fe3O4 heterotrimers ...... 173 6.2.5 Synthesis of Ag-Pt-Fe3O4 heterotrimers ...... 175 6.2.6 Synthesis of Ni-Pt-Fe3O4 heterotrimers ...... 175 6.2.7 Synthesis of Pd-Pt-Fe3O4 heterotrimers ...... 176 6.2.8 Synthesis of Cu9S5-Au-Pt-Fe3O4 heterotetramers ...... 176 6.2.9 Synthesis of PbS-Au-Pt-Fe3O4 heterotetramers ...... 177 6.2.10 Synthesis of higher-order Au(Pt-Fe3O4)n oligomers ...... 177 6.2.11 Characterization ...... 178 6.3 Results and Discussion ...... 179 6.3.1 M-Pt-Fe3O4 heterotrimers (M = Au, Ag, Ni, Pd) ...... 179 6.3.2 Insights into M-Pt-Fe3O4 heterotrimer formation ...... 185 6.3.3 MxSy-Au-Pt-Fe3O4 heterotetramers (M = Cu, Pb) ...... 189 6.3.4 Higher-order Au(Pt-Fe3O4)n oligomers via coupling ...... 195 6.4 Conclusions ...... 197 6.5 References ...... 197

Chapter 7 Summary and Outlook ...... 202

viii LIST OF FIGURES

Figure 1-1: A multi-step synthetic concept for molecules (top) and nanoparticles (bottom), along with chemical structures and graphical examples ...... 12

Figure 1-2: Classes of compounds that are commonly accessible by solution- phase chemical transformation reactions of metal nanoparticle synthons ...... 15

Figure 1-3: Categories of solution-phase chemical transformation reactions that are commonly used to convert one type of nanoparticle into another ...... 16

Figure 1-4: Representative examples of nanoscale cation exchange reactions: TEM images of a) CdS nanorod precursors and b) CdS-Ag2S superlattice nanorods after partial exchange of the CdS nanorods with Ag+, and HRTEM images of c) wurtzite-type CdSe nanorods, d) wurtzite-type Cu2Se nanorods after ion exchange of the CdSe nanorods in (c) with Cu+, and e) wurtzite-type ZnSe nanorods after ion exchange of the Cu2Se nanorods in (d) with Zn2+. The conserved anion lattice, which leads to retention of the wurtzite structure type, is shown in (f). Panels (a,b) reprinted from Ref. [132] with permission (Copyright 2007, American Association for the Advancement of Science) and panels (c–f) reprinted from Ref. [137] with permission (Copyright 2011, American Chemical Society) ...... 19

Figure 1-5: Representative examples of nanoscale anion-exchange reactions: a) Powder XRD data characterizing the transformation of wurtzite-type ZnO into wurtzite-type ZnS, b) crystal structures of the ZnO precursor and ZnS product, highlighting the retained cation sublattice, and TEM images of c) ZnO nanoparticles and d) hollow single-crystal ZnS nanoparticles synthesized by treating the ZnO nanoparticles in (c) with hexamethyldisilathiane in trioctylphosphine. Panels (a,b) reprinted from Ref. [140] with permission (Copyright 2009, American Chemical Society) and panels (c,d) reprinted from Ref. [141] with permission (Copyright 2009, American Chemical Society) ...... 21

Figure 1-6: Representative examples of nanoscale addition reactions: TEM images of a) Au, b) Ag, c) Pt, and d) AuAg nanoparticle seeds, and e) Au- Fe3O4, f) Ag-Fe3O4, g) Pt-Fe3O4, and h) AuAg-Fe3O4 hybrid nanoparticle products formed by heterogeneous seeded growth of Fe3O4. Reprinted from Ref. [59] with permission (Copyright 2010, American Chemical Society) ...... 22

Figure 1-7: Top: Schematic representation highlighting orthogonal reactivity in striped metal nanowire systems: Reagent X reacts with A to form AX, but does not react with B; reagent Y reacts with B to form BY, but does not react with AX. SEM images of a) Au-Ag-Au striped metal nanowire and b) Au-

ix

Ag2S-Au nanowires formed after reacting with sulfur. SEM images of c) an Ag-Pt striped metal nanowire, d) an Ag-PtPbx striped nanowire after reacting the Ag-Pt nanowire with Pb (element maps show incorporation of Pb into the Pt segment, but not the Ag segment), and e) an Ag2S-PtPbx striped nanowire after treating the Ag-PtPbx nanowire with sulfur. Scale bars: 1 µm. Panels (a– e) reprinted from Ref. [127] with permission (Copyright 2009, American Chemical Society) ...... 28

Figure 1-8: a) TEM image (scale bar: 50 nm), diffraction pattern, and b) Au and c) Ag elemental mapping data for a colloidal Ag-Au-Ag striped nanorod. d) TEM image and elemental mapping data of striped Ag2S-Au- Ag2S nanorods for e) Au, f) S, and g) Au + S, The nanorods were formed after treating the Ag-Au-Ag nanorods with aqueous Na2S. TEM images (scale bars: 200 nm) of h) Bi2Te3 nanowires with Te tips and i) Bi2Te3 nanowires with PbTe tips after treating the Bi2Te3-Te nanowires with Pb. Panels (a–g) reprinted from Ref. [156] with permission (Copyright 2008, American Chemical Society) and panels (h,i) reprinted from Ref. [160] with permission (Copyright 2012, American Chemical Society) ...... 30

Figure 1-9: Top: Schematic representation showing the stepwise conversion of a CdS nanorod into a PbS-CdS striped nanorod by two sequential ion-exchange reactions and a Cu2S-CdS nanorod intermediate. HRTEM images of a) a + Cu2S-CdS striped nanorod formed by treating CdS nanorods with Cu and b) 2+ a PbS-CdS striped nanorod formed by treating Cu2S-CdS nanorods with Pb . Panel (a) reprinted from Ref. [135] with permission (Copyright 2009, American Chemical Society) and panel (b) reprinted from Ref. [134] with permission (Copyright 2009, American Chemical Society) ...... 32

Figure 1-10: TEM images of CdSe nanorods, selectively capped with PbSe at a) both ends (PbSe-CdSe-PbSe) and b) one end (PbSe-CdSe). Reprinted from Ref. [164] with permission (Copyright 2005, American Chemical Society) ...... 34

Figure 1-11: TEM images showing a) Pt nanoparticles photodeposited selectively at the CdSe core region of CdSe/CdS core–shell nanorods (scale bar: 20 nm), b) Au nanoparticles deposited selectively at the CdSe core region of CdSe/CdS core–shell nanorods (HAADF-STEM image, scale bar: 20 nm), c) oxide deposited selectively at the Au end of Au-tipped CdS nanorods, and d) Au and Ag2S selectively deposited at opposite ends of a CdS nanorod (Au-CdS-Ag2S). Panel (a) reprinted from Ref. [166] with permission (Copyright 2008, Wiley-VCH), panel (b) reprinted from Ref. [167] with permission (Copyright 2008, American Chemical Society), panel (c) reprinted from Ref. [169] with permission (Copyright 2010, American Chemical Society), and panel (d) reprinted from Ref. [170] with permission (Copyright 2010, Wiley-VCH) ...... 37

x Figure 1-12: TEM images of a,b) PbSe nanorods grown selectively off the Au domain of Au-Fe3O4 heterodimers to generate PbSe-Au-Fe3O4 heterotrimers, c) Au-SiO2 heterodimers and d) Ag-Au-SiO2 heterotrimers formed by growing Ag selectively off the Au domains, and e,f) various Au-Pt-Fe3O4 heterotrimers formed by the selective growth of e) Au off the Pt domains of Pt-Fe3O4 heterodimers and f) Fe3O4 off the Au domains of Au-Pt heterodimers. Panels (a,b) reprinted from Ref. [180] with permission (Copyright 2006, Wiley-VCH), panels (c,d) reprinted from Ref. [181] with permission (Copyright 2010, American Chemical Society), and panels (e,f) reprinted from Ref. [183] with permission (Copyright 2008, American Chemical Society) ...... 40

Figure 1-13: Top left: Schematic represention of nanoscale elimination reactions by selective dissolution. a) FESEM image of an Au nanowire with 25, 50, and 100 nm gaps created by selectively dissolving the Ni components of Au- Ni multisegmented nanowires. TEM images of b) Au nanoframes formed by facet-selective overgrowth of Au on decahedral Ag nanoparticles and selective dissolution of Ag (scale bar: 50 nm), c) Ag-tipped Au nanorods formed by selectively etching Ag from the sides of Au@Ag core–shell nanorods, d) dented Fe3O4 nanoparticles formed by selectively dissolving the Au domains of Au- Fe3O4 heterodimers, and e) Fe3O4 nanocontainers formed by selectively dissolving the Au domain of Au-Fe3O4 heterodimers with additional Fe3O4 covering the Au. Panel (a) reprinted from Ref. [184] with permission (Copyright 2005, American Association for the Advancement of Science), panel (b) reprinted from Ref. [188] with permission (Copyright 2011, American Chemical Society), panel (c) reprinted from Ref. [190] with permission (Copyright 2012, American Chemical Society), panel (d) reprinted from Ref. [191] with permission (Copyright 2010, Wiley-VCH), and panel (e) reprinted from Ref. [192] with permission (Copyright 2011, American Chemical Society) ...... 44

Figure 1-14: Top left: Schematic represention of hybrid nanoparticle coupling reactions. TEM images of a) Fe3O4-Au-Fe3O4 heterotrimers formed by heating Au-Fe3O4 heterodimers with trace amounts of sulfur and b–d) linear (Au-CdSe)n chains formed by treating Au-CdSe-Au nanorods with I2. Panel (a) reprinted from Ref. [179] with permission (Copyright 2006, American Chemical Society) and panels (b–d) reprinted from Ref. [196] with permission (Copyright 2009, Wiley-VCH) ...... 47

Figure 1-15: a) Schematic representation and b–d) corresponding TEM images showing the formation of CdS-PdSx-CdS hybrid particles by installing a PdSx domain on b) CdS nanoparticles to form c) CdS-PdSx dimers, followed by fusion to form d) the CdS-PdSx-CdS products. e) Schematic representation and corresponding f) TEM and g) HRTEM images of PdSx-Co9S8-PdSx

xi

hybrid particles formed from the fusion of the Co9S8 domains on PdSx-Co9S8 dimer particles. Panels (a–d) reprinted from Ref. [178] with permission (Copyright 2009, The Royal Society of Chemistry) and panels (e–g) reprinted from Ref. [179] with permission (Copyright 2007, Wiley-VCH) ...... 49

Figure 1-16: a–d) TEM images showing a four-step synthesis of colloidal ZnSe- CdS-Pt nanorods: a) ZnSe seed nanoparticles, b) ZnSe nanoparticles coated with a thin layer of CdS, c) ZnSe/CdS nanorods grown off the ZnSe/CdS core–shell nanoparticles, and d) ZnSe-CdS-Pt nanorods after site-selective Pt deposition. e,f) Schematic representations of multistep routes to core–shell nanoparticles with controlled metal, anion, and cation compositions. Panels (a–d) reprinted from Ref. [55] with permission (Copyright 2011, American Chemical Society) and panels (e,f) reprinted from Ref. [208] with permission (copyright 2010, American Association for the Advancement of Science) ...... 54

Figure 1-17: (a) Typical DMCR chromatogram showing the applied magnetic flux density (dashed line) and the absorbance, where peaks correspond to fractions that are eluted. Peak A and panel (b) correspond to the absorbance and TEM image, respectively, for as-synthesized Au-Fe3O4 heterodimers eluted with no applied magnetic field. Peaks B, C, and D correspond to the peaks eluted at the applied magnetic flux densities as indicated, and the TEM images of each fraction are shown in panels (c), (d), and (e), respectively. Reprinted with permission from ref. 84 (copyright 2011, Wiley-VCH) ...... 58

Figure 2-1: Experimental and simulated XRD patterns for dried powders and drop-cast films of rhombohedral GeTe. Drop-casting the microcrystallites produces preferred orientation along <002>, indicating a faceted, non- spherical morphology ...... 85

Figure 2-2: (a) SEM micrograph of the GeTe product, which consists of highly faceted, micron-sized crystallites with a cube-shaped morphology. The particle shape is consistent with the preferred orientation observed by XRD. (b) Size distribution histogram for the GeTe crystallites, which have an average edge length of 1.0 ± 0.2 µm ...... 86

Figure 2-3: (a) EDS spectrum and (b) SEM image of the GeTe microcrystals, along with the associated EDS element mapping data showing uniform distribution of (c) Ge and (d) Te throughout each particle ...... 87

Figure 2-4: Two views of the rhombohedral GeTe crystal structure, with the (002) [top] and (111) [bottom] surfaces highlighted. In contrast to the {002} faces, which consist of an alternating arrangement of Ge and Te , the {111} faces are composed of hexagonally close-packed Te (or Ge) atoms ...... 88

xii Figure 2-5: TEM images of (a) a collection of some of the smaller GeTe crystallites, showing generally truncated and faceted cube-shaped morphologies and (b) a region of smaller crystallites produced during imaging via electron beam induced fragmentation. The SAED pattern in (c) corresponds to the smaller fragmented nanocrystals from (b) and indicates highly crystalline rhombohedral GeTe. The SAED pattern in (d) is from a region of crystalline GeTe particles similar in morphology and size to those shown in (a) that began to become amorphous after continued exposure to the electron beam. (e) DSC trace for the GeTe microcrystallites, showing an endotherm at 717 °C corresponding to the melting of GeTe ...... 89

Figure 2-6: (a) SEM image of a 50 nm thick a-Ge film on (111)-Si, after treatment with TOP-Te solution at 250 ºC. EDS-generated elemental mapping confirms the presence of (c) Ge and (d) Te, which are uniformly distributed throughout the converted film, and clearly supported by (b) the Si substrate ...... 91

Figure 3-1: Summary of primary chemical variables and corresponding products ..... 103

Figure 3-2: (a) Simulated and (b) experimental powder XRD patterns corresponding to rhombohedral GeTe, prepared by reacting (Ge[N(SiMe3)2]2) and TOP-Te in ODE at 230 °C. Under otherwise identical conditions, when a 2.5 wt. % solution of PVP-t in ODE is used as the solvent, amorphous solids are obtained even after (c) 1 h and (d) 12 h of heating, suggesting that the polymer prevents crystallization of GeTe. (e) When 2.5 wt. % PVP-t in OLAC is used, the XRD pattern shows only crystalline GeTe, indicating that the reaction is affected significantly by the solvent ...... 104

Figure 3-3: TEM image of the amorphous solids obtained by reacting Ge[N(SiMe3)2]2 and TOP-Te in ODE, in the presence of PVP-t. The product consists primarily of small particulates with average sizes that mostly fall within the range of 3-6 nm, although a few are larger. Scale bar = 20 nm ...... 105

Figure 3-4: (a) Representative TEM image of faceted GeTe nanoparticles with (b) a higher-magnification image highlighting the octahedral morphology and (c) the corresponding electron diffraction pattern ...... 107

Figure 3-5: (a) HRTEM image of a representative GeTe nanoparticle, with an enlarged region in (b), corresponding to region enclosed by the red dashed box in (a), showing lattice fringes matching those of GeTe {022}. (c) The crystal structure of rhombohedral GeTe (purple = Ge, yellow = Te) viewed along its [111] axis, highlighting the atomic arrangement and interatomic distances that correspond well to the image shown in (b) ...... 108

Figure 3-6: (a) Representative SEM image of faceted GeTe nanoparticles and (b) the corresponding EDS spectrum ...... 109

xiii Figure 3-7: TEM image of faceted GeTe nanoparticles, synthesized in the presence of PVP in NONAC at 250°C and at a PVP:Ge2+ ratio of 2 ...... 110

Figure 3-8: (a) Powder XRD data for GeTe nanoparticles synthesized in the presence of PVP in NONAC at 250°C and at a PVP:Ge2+ ratio of 10. The peak positions correspond well to the simulated pattern, but the diffraction intensities from the {111}, {111 }, {222}, and {222} planes are enhanced, suggesting preferred orientation. (b) TEM image of the corresponding GeTe nanoparticles, showing that many have grown into plate-like, 2D shapes, which is likely the cause of the preferred orientation in (a) ...... 110

Figure 3-9: (a) Representative TEM image of GexTe1-x alloy nanospheres with corresponding SAED pattern (inset), showing that the particles are amorphous, and (b) HRTEM image (with an enlargement in the inset, corresponding to the region enclosed by the dashed red box) highlighting the lack of atomic ordering for a representative nanosphere. (c) STEM/EDS line scan spectrum across a GexTe1-x nanosphere (right) as indicated on the HAADF image (left), showing a homogeneous distribution of Ge and Te where x ≈ 0.3 ...... 113

Figure 3-10: (a) TEM image of the GeTe product formed in the presence of excess NaOct. In addition to GexTe1-x nanospheres, the product mixture contains a large number of 2D nanosheets. (b-h) Representative TEM images of the nanosheets, which showcase the diverse 2D morphologies that emerge. The product includes (b) triangles, (c) quadrilaterals, (d,e) hexagons, (f) other polyhedral shapes, and (g-h) folded, multi-layered structures. Unlabeled scale bars are 100 nm...... 116

Figure 3-11: (a) XRD pattern corresponding to the product shown in Figure 6a, confirming the presence of rhombohedral GeTe, with preferred orientation in the [111] direction. (b) SAED pattern of the single triangular nanosheet shown in the inset of (a). The {022} diffraction spots indicate a [111] axis perpendicular to the TEM image, which is consistent with the preferred orientation observed in (a) ...... 116

Figure 3-12: STEM/EDS line scan and corresponding HAADF images for a hexagonal (top) and triangular (bottom) GeTe nanosheet. Ge and Te are uniformly distributed throughout each nanosheet in an approximate 1:1 ratio .... 117

Figure 3-13: Representative AFM data characterizing the thicknesses of colloidal GeTe nanosheets. The line profiles labeled [1], [2], and [3] in (b) correspond to the lines indicated in the AFM image in (a). The vertical position markers in (b) are represented by color-coded dots in (a), which highlight the height differential between the nanosheets and the substrate. The thicknesses of the nanosheets [1], [2], and [3] are 6.7 nm, 13.1 nm, and 7.9 nm, respectively.

xiv

The image in panel (a) includes groups of GexTe1-x alloy nanospheres, which appear white because their height is taller than that of the maximum z-axis scale. (c) High-magnification AFM image and corresponding height profile of a multi-layered GeTe nanosheet with a maximum height of 5.6 nm ...... 118

Figure 3-14: (a,b) High-magnification TEM images of GeTe nanosheets taken at early stages of reaction, which suggest that growth may occur by laterally- oriented fusion of smaller particles. (c) HRTEM image of a 2D GeTe quadrilateral (inset), showing well-defined edges and vertices, but an atomically disordered face. The nanosheets consistently exhibited such disorder, despite the single-crystallinity revealed by electron diffraction. (d) TEM image of several 2D GeTe triangles that have higher-contrast, triangular domains in their centers. This observation suggests that, in rare cases, 2D GeTe nanostructures may grow laterally from a dense GeTe seed ...... 119

Figure 3-15: (a) TEM image and (b) SAED pattern for a representative GeTe nanosheet with a dense, triangular domain at its center, which implies seed- mediated growth. In addition to the {02-2} reflections expected for the GeTe nanosheets, additional diffraction spots are present in (b), which are likely due to the central domain. All the reflections in (b) correspond well to the crystal structure of rhombohedral GeTe. The scale bar in (a) corresponds to 100 nm ...... 121

Figure 4-1: FTIR absorbance spectra of (a) OLAC, (b) Na(OL), (c) purified Co(OL)2, and (d) an impurity that was isolated by centrifugation during preparation of Co(OL)2 ...... 139

Figure 4-2: (a) Experimental XRD patterns for colloidal wz-CoO nanocrystals, along with a simulated pattern for reference. (b) Large-area TEM image of the nanocrystal product, which consists of uniform nanorods with pencil- point tips. A higher magnification image is shown in the inset of (a). These results were obtained by thermally decomposing Co(OL)2 samples that contained the [Co(OL)2-x(OH)x] impurity. The scale bar in the inset of (a) is 20 nm, and the scale bar in (b) is 50 nm ...... 142

Figure 4-3: (a) Low-magnification and (b) higher-magnification TEM images of stout, cone-shaped nanocrystals that were formed by thermolysis of Co(OL)2 in the presence of TMAOH. It appears that the nanocones assembled into pairs, with the bases in contact with one another. (c) SAED of the nanocrystals in (a) and (b), which shows diffraction rings that can be assigned to wz-CoO. (d) Representation of the wz-CoO crystal structure depicting the non-equivalent (001) and (001) planes, and the polar attractions that may have caused the nanocones to assemble pairwise. Oxygen atoms are colored red and cobalt atoms are colored blue. The scale bars in (a) and (b) are 50 nm and 10 nm, respectively ...... 143

xv Figure 4-4: (a) Experimental XRD and SAED (inset) patterns for the wz-CoO nanocrystals shown in (b), with a simulated XRD pattern for reference. (b) TEM image of the nanocrystals, which have rod-shaped domains protruding from a conical base. The nanocrystals in (b) were formed by thermolysis of Co(OL)2 in the presence of Na(OL), at Co(OL)2:Na(OL) molar ratio of 4. When Co(OL)2:Na(OL) of 2 was used, the rod-shaped domains were longer, which can be seen in (c). All scale bars are 50 nm ...... 145

Figure 4-5: (a) Large-area and (b) higher-magnification TEM images of wz-CoO nanocrystals with a bullet-like shape, along with a corresponding SAED pattern in (c). The scale bar in (a) is 100 nm and the scale bar in (b) is 20 nm ... 148

Figure 4-6: FTIR absorbance spectra of [Co(CH3COO)2ž4H2O] and Co(Ol)2, - - which highlight the νsym(COO ) and νasym(COO ) stretches in each compound .... 149

Figure 4-7: FTIR absorbance spectrum of a distillate (top), collected from Co(OL)2 solution in ODE over a heating period of 2 hours at 240 °C. The FTIR spectrum of ODE solvent (bottom) is included as a reference. Peaks marked with an asterisk (*), in the distillate spectrum, do not appear in the spectrum for ODE ...... 150

Figure 4-8: FTIR absorbance spectra of Co(OL)2, before (bottom) and after (top) heating in the solid-state for 2 hours at 240 °C. The spectra highlight the disappearance of a peak 1705 cm-1, and the appearance of a new peak at 1725 cm-1. The spectral changes suggest that free is consumed, and a ketone is generated, as a result of heating ...... 151

Figure 5-1: Metal nanowires in AAO membrane reacted either before (a-c) or after (d-f) Ag backing removed. Nanowires undergo site-specific modifications (b,e) and are then released from the membrane (c,f) ...... 162

Figure 5-2: (a) SEM micrograph with EDS element maps for (b) Pt, (c) Pb, and (d) SEM image with Pt and Pb overlay. (e) XRD data for as-grown Pt nanowires in the AAO membrane, and released PtPb/Pt/PtPb nanowires after conversion. Inset: EDS spectrum for PtPb/Pt/PtPb nanowires ...... 163

Figure 5-3: EDS-generated element maps for (a) PtPb/Pt/PtPb nanowires with short tips, and (b) fully-converted PtPb nanowires, which were both prepared by conversion of template-confined Pt nanowires. c) Anisotropic PtPb/Pt nanowires were made in-membrane, with the Ag backing intact. All scale bars are 2 µm ...... 164

Figure 5-4: A series of SEM images at increasing magnifications for PtPb / Pt / PtPb nanowires that were prepared by reacting template-confined Pt nanowires in (a) 30 mM Pb2+ at 300 ºC, and (b,c) 10 mM Pb2+ at 220 ºC.

xvi These large field images demonstrate the sample-wide uniformity of nanowires that were transformed in-membrane. SEM images of fully converted PtPb nanowires reacted either (d) in the membrane, or (e) after release. The smooth morphology of the wire in (d) demonstrates the effectiveness of the membrane for protecting the sidewalls of the nanowires, which ensures that the shape and morphology of the nanowire are conserved during the chemical conversion process ...... 165

Figure 6-1: Stepwise construction of M-Pt-Fe3O4 heterotrimers (M = Ag, Au, Ni, Pd). A schematic showing the multi-step synthesis of M-Pt-Fe3O4 heterotrimers, along with the most significant possible products and their observed frequencies (expressed as the percentage of observed heterotrimers, not total yield), is shown at the top. Representative TEM images show (a) Pt nanoparticle seeds, (b) Pt-Fe3O4 heterodimers, and (c) Au-Pt-Fe3O4, (d) Ag- Pt-Fe3O4, (e) Ni-Pt-Fe3O4, and (f) Pd-Pt-Fe3O4 heterotrimers. Panel (g) shows photographs of (left) a vial containing Au-Pt-Fe3O4 heterotrimers in hexane, which responds to an external Nd-Fe-B , (middle) the same vial with Au-Pt-Fe3O4 heterotrimers in a larger volume of hexanes, and (right) the same vial after precipitation of the heterotrimers with ethanol. The precipitated heterotrimers collect next to the external magnet. All scale bars are 25 nm ...... 180

Figure 6-2: HRTEM image of a single Au-Pt-Fe3O4 heterotrimer, highlighting the coherent (111) interfaces. Scale bar is 5 nm ...... 182

Figure 6-3: A STEM image of a single Au-Pt-Fe3O4 heterotrimer is shown in (a), along with EDS-generated elemental maps corresponding to (b) Fe, (c), Pt, (d), Au, and (e) superimposed Fe (red), Pt (green), and Au (blue). Scale bar is 5 nm ...... 183

Figure 6-4: Characterization data for M-Pt-Fe3O4 heterotrimers (M = Au, Ag, Ni, Pd). (a) Powder XRD data for all of the M-Pt-Fe3O4 heterotrimers and the Pt-Fe3O4 heterodimer for comparison. (b) EDS spectra for all of the M-Pt- Fe3O4 heterotrimers. For the Au-Pt-Fe3O4 and Ag-Pt-Fe3O4 samples, a Ni TEM grid was used, and the Ni and Cu signals originate from this grid. For the Ni-Pt-Fe3O4 and Pd-Pt-Fe3O4 samples, a Cu TEM grid was used, and the Cu signal originates from this grid. (c) SAED patterns for all of the M-Pt- Fe3O4 heterotrimers show distinct diffraction spots or rings for each of the components. (d) UV-visible absorption spectra for the Au-Pt-Fe3O4 and Ag- Pt-Fe3O4 heterotrimers, also showing reference spectra for Au and Ag nanoparticles and the Pt-Fe3O4 ...... 184

Figure 6-5: Au nanoparticles (8.3 ± 0.9 nm) prepared by heating 50 mg HAuCl4 dissolved in 20 mL OLAM for 15 minutes at 90 °C. The UV-Visible absorbance spectrum corresponding to this sample is shown in Figure 6-4(d) .... 185

xvii

Figure 6-6: Chemoselective nucleation in the Ag-Pt-Fe3O4 heterotrimer system. (a-d) TEM images and corresponding schematics showing a series of control experiments, all carried out under the same reaction conditions and designed to probe the chemoselectivity of heterogeneous nucleation in the prototype Ag-Pt-Fe3O4 system: (a) nucleation of Ag on Fe3O4 nanoparticles, (b) nucleation of Ag on Pt nanoparticles, (c) nucleation of Ag on both Fe3O4 and Pt nanoparticles that are present as a physical mixture, and (d) nucleation of Ag exclusively on the Pt domain of Pt-Fe3O4 nanoparticles that are present as interconnected dimers with a solid-state interface. The insets in (b) and (c) show enlarged views of the regions indicated by boxes, highlighting the lower-contrast Ag regions that nucleate on the higher-contrast Pt particles. The scale bars in the main panels correspond to 20 nm and those in the insets correspond to 5 nm ...... 187

Figure 6-7: Pt4f XPS spectra for Pt-Fe3O4 heterodimers and the corresponding Pt seeds, indicating a decreased binding energy of 0.4 eV in the heterodimers ...... 187

Figure 6-8: Stepwise construction of MxSy-Au-Pt-Fe3O4 heterotetramers (M = Cu, Pb). A schematic showing the stepwise construction of MxSy-Au-Pt-Fe3O4 heterotetramers, along with the most significant possible products and their observed frequencies (expressed as the percentage of observed heterotetramers, not total yield), is shown at the top. Representative TEM images of Cu9S5-Au-Pt-Fe3O4 heterotetramers are shown in (a) and (b), with enlarged regions in (b) showing lattice fringes that correspond to the Fe3O4 (311), Cu9S5 (200), Cu9S5 (111), Au (111), and Pt (111) planes. Powder XRD data for the Cu9S5-Au-Pt-Fe3O4 and PbS-Au-Pt-Fe3O4 heterotetramers are shown in (c) and (d), respectively, along with a pattern for Cu9S5-Au heterodimers in (c) for comparison. Panel (e) shows a representative TEM image of the PbS-Au-Pt-Fe3O4 heterotetramers. The composite EDS element map for a single PbS-Au-Pt-Fe3O4 heterotetramer particle in panel (f) includes an overlay of Fe (red), Pt (green), Au (blue), and Pb (orange). A STEM image of a single Cu9S5-Au-Pt-Fe3O4 heterotetramer is shown in (g), along with the corresponding EDS element maps for (h) Fe, (i) Pt, (j) Au, (k) Cu, and (l) superimposed Fe, Pt, Au, and Cu. The scale bars correspond to (a) 25 nm, (b,g) 5 nm, and (e) 10 nm ...... 190

Figure 6-9: (a) Powder XRD pattern corresponding to CuxS-Au-Pt-Fe3O4 tetramers prepared using Cu(II) oleate, indicating the presence of (1) Fe3O4 (2) Au, and (3) Pt. The diffraction pattern is nominally the same as for Au- Pt-Fe3O4 trimers [Fig 6-4(a)], which suggests that the CuxS component is amorphous. (b) EDS spectra for a blank formvar-coated Ni TEM grid (bottom), showing the presence of Cu and Ni, and CuxS-Au-Pt-Fe3O4 tetramers (top) dropcast onto the same grid. The presence of a Cu-containing

xviii phase is inferred by the significantly higher Cu signal in the heterotetramer sample ...... 192

Figure 6-10: The UV-visible absorption spectrum corresponding to the CuxS-Au- Pt-Fe3O4 heterotetramers (blue), along with that of Au-Pt-Fe3O4 heterotrimers (red) for comparison ...... 192

Figure 6-11: TEM images corresponding to control experiments design to probe the chemoselectivity of Cu9S5 nucleation on the Au surface of Au-Pt-Fe3O4 seed particles. Cu9S5-Au heterostructures (which include some heterodimers), shown in (a), clearly indicate that heterogeneous nucleation of Cu9S5 on the Au seeds has occurred. When an identical reaction is carried out using Pt- Fe3O4 heterodimers as seeds, shown in (b), a mixture of irregularly shaped CuxSy nanoparticles and unreacted Pt-Fe3O4 seeds is obtained, clearly indicating that the CuxSy and Pt-Fe3O4 components remain separate from one another. The scale bars correspond to 20 nm in (a) and 50 nm in (b) ...... 194

Figure 6-12: Higher order hetero-oligomers based on the Au-Pt-Fe3O4 heterotrimer building block. Top: A schematic showing the sulfur-mediated coupling of Au-Pt-Fe3O4 heterotrimers to form higher-order oligomers. Bottom: Representative TEM images of coupled products, including linear and bent Fe3O4-Pt-Au-Au-Pt-Fe3O4 hexameric oligomers and higher-order branched nanostructures. Cartoon representations of each hetero-oligomer are also shown. Scale bars correspond to 10 nm ...... 196

xix ACKNOWLEDGEMENTS

Most importantly, I want to thank my family. My Mother and Father are the best parents that anyone could want. I have an instinct to know myself well. For me, science has been about earnestly testing the boundaries of my intellectual and spiritual capabilities, and about realizing my potential. This instinct is what brought me to graduate school, and is certainly the reason that I have been successful here. I am sure that I inherited this instinct from my Mom and Dad. Thank you for your unconditional love, advice, encouragement, sacrifice and support. I realize that it has been a long ten more years, and I promise that I won’t go to school anymore now. Probably. To Adam,

Marianne, Angela, Ande, Kathy, Bob, Jenny, Jessica, Auntie B., Adam Bock and all of my extended family - thank you for understanding what I have been trying to do, for your interest, and for being a great support team. I hope that my writing this means we can finally start spending more time together.

At Penn State I met the greatest love of my life, Farah Dawood. Nobody knows better than Farah that graduate school has brought out the very best in me, and the very worst in me. There are no words to express how grateful I am to have shared the laboratory with you, and to now be sharing a home with you. Thank you for riding all of the highs and lows along with me, and for all of your love, understanding, encouragement, and intelligence. You bring a balance and comfort to me that I could not have done without (boogie). Also, thank you for introducing me to my good friend,

Abby.

xx I chose wisely when I decided to do chemistry with Dr. Raymond Schaak. Ray had dedicated himself to being the greatest teacher, mentor, advisor, and friend that he could be to me - every single day. I feel privileged to have learned a tremendous amount from him about science, communicating, and life, and I am grateful for having had the chance to work in such a creative laboratory. I also owe my (expanded) graduate committee - Dr. Thomas Mallouk, Dr. Christine Keating, Dr. Mary Elizabeth Williams, and Dr. Michael Hickner - many thanks for their fresh perspectives, knowledge, advice, and professional support. Also, I’ll never forget where I come from - Dr. Daniel

Freedman, Dr. Preeti Dhar, Dr. Pamela St. John, Dr. Diego Dominici and others at SUNY

New Paltz - you demanded the deepest commitment to understanding fundamental chemistry from me. My experience at New Paltz changed my life, and I am grateful to you all.

My co-workers in the Schaak lab, past and present, deserve huge thanks for being a pleasure to think with, work with and learn from. This is especially true for those who helped bring my chemistry from the workbench to the journal, including important contributions from Adam Biacchi, Eric Popczun, Dr. Karl Oyler, Dr. Beth Anderson, Dr.

James Bondi, and Dr. Ian Sines. Thank you for elevating my research to the next level.

A special note of thanks goes to an excellent friend and collaborator, Dr. Jacob

Beveridge, with whom I have had countless engaging conversations about chemistry and nanoscience. I can only hope that one day we’ll be bouncing our ideas back and forth, and solving the world’s problems once again, Jake. Also, thank you (begrudgingly) to a fellow Penn State , Dr. Seth Morton, who is one of the most fiercely loyal friends

xxi I have ever had. There is almost no way that the last few months of my graduate career would have worked without the help of Alex Wiltrout. I owe you big time, Troutman.

I spent a great deal of time using the Electron Microscopy Facilities at the Huck

Institute of the Life Sciences, and would like to thank Missy Hazen for all of her hard work and assistance with the TEM. Josh Stapleton, Julie Anderson, Melisa Yashinski, and Trevor Clark, of the Materials Characterization Lab at the Penn State Materials

Research Institute, also went out of their way to help me collect key pieces of data using scanning electron microscopy, , and thermogravimetric . Of course, I am thankful for the funding that kept the wheels turning. This research was supported primarily by the US National Science Foundation (NSF) (grants

DMR-0748943, CHE-0845258, and CHE-1213475), but with additional partial support through the Penn State Materials Research Science and Engineering Center (DMR-

0820404). The work pertaining to GeTe nanosheets was supported by the US

Department of Energy, Office of Basic Energy Sciences, Division of Materials Sciences and Engineering, under award #DE-FG02-08ER46483.

Chapter 1

Emerging Strategies for the Total Synthesis of Inorganic Nanostructures

1.1 Introduction

Sophisticated architectures of growing complexity are becoming desired for inorganic nanostructures, as the applications of these materials continue to expand.

Increasingly stringent boundaries are being placed on dimensionality, uniformity, and functionality, requiring meticulous engineering at the nanometer scale. Among the simplest nanostructures are those that contain single inorganic components, and increasingly rigorous synthetic control is needed to meet design criteria for such systems across a wide range of diverse research areas. For example, metal,1-6 multi-metal,7,8 and metal oxide nanocrystals9-12 bound by high-index crystal facets are desirable catalysts because they have surfaces with large densities of reactive step, ledge and kink sites.13,14

This requires the preparation of nanocrystals with non-equilibrium geometries,15-17 which is challenging because high-index surfaces are inherently energetic. Likewise, single- component magnetic nanoparticles are promising for applications that include in vivo theranostics, but the surface of each particle must be carefully engineered to support physiological stability and targeting motifs, all while maintaining a high magnetic energy density for clear imaging.18-20 Also, nanoparticles with exquisite uniformity can self- assemble into highly periodic crystalline superlattices, and this has given rise to “artificial

Adapted with permission from Angew.Chem Int. Ed. DOI: 10.1002/anie.201207240, “Emerging Strategies for the Total Synthesis of Inorganic Nanostructures” Copyright 2013 WILEY-VCH Verlag GmbH & Co. 2 solids” that are compelling materials for electronics and energy-conversion devices.21-25

These and many other examples emphasize the degree to which complexity and stringency have become central themes in nanomaterials synthesis, even for systems with only a single inorganic component.

As the trend toward complexity of nanoscale materials continues to expand, there is a natural progression from single-component systems to nanostructures that have multiple inorganic components. Of particular interest are multi-component structures that are characterized by solid-state interfaces, which facilitate electronic and magnetic communication between the inorganic domains.26-34 Interfacial phenomena of this type, such as charge transfer and spin-exchange, give rise to useful properties not observed in physical mixtures of the components. Next-generation electronic devices, for example, may contain many nanoscale heterojunctions, including metal-, semiconductor-semiconductor, and metal-metal oxide contacts. Device components such as p-n junctions, Schottky , and logic gates are derived from these points of contact.35 New and unexpected properties can also result from creating intimate contact between complementary solids. Au nanoparticles, for example, become highly active catalysts for CO oxidation when dispersed on a variety of metal oxide supports, despite being chemically inert when isolated or supported on .36,37 Recently, tremendous effort has gone toward developing nanostructured materials that efficiently harvest, store, and utilize solar energy.38-41 A spatially controlled arrangement of inorganic domains, resulting in a prescribed sequence of heterojunctions, is the defining characteristic of many new photoactive nanomaterials.42-44 For example, programmed arrangements of metal and semiconducting domains are designed to energetically direct the flow of

3

45 separated charges, as in the recently proposed ternary Pt-TiO2-IrO2 heterostructure, but such complex targets are still quite a synthetic challenge.

Multi-component inorganic nanostructures are not so unlike molecules, wherein solid-state interfaces are analogous to chemical bonds. This has been noted previously by a number of other researchers.46,47 Electrochemical potential results in charge transfer between two solids in contact, and the magnitude of that potential determines the amount of charge transferred and its reversibility. This is akin to the electronegativity difference between two atoms that share , which characterizes the polarity (or ionicity) of a bond. Even the simplest collection of atoms, a diatomic like H2 or O2, has properties that are vastly different from those of the isolated atoms, just as multi- component nanostructures exhibit modulated properties relative to the constituent inorganic domains.

A complex molecule, such as an organic natural product, can be thought of as a construction of simpler molecular fragments and functional groups, with local geometries, charge distributions, and connectivities that concertedly impart its global function. Certain parts of a molecule may act as a scaffold, providing structural and geometrical support for the more active functional groups. One functional group may be more active than another under certain conditions, or more directly responsible for the chemical and physical properties of the molecule. The presence of a certain element or group may also significantly influence other parts of a molecule, either through steric or inductive effects. Multi-component inorganic nanostructures, created by piecing together simple solid-state building blocks into more complex functional architectures, can be thought of as multi-domain constructs that are conceptually analogous to organic

4 molecules. Nanoparticles that are connected by a solid-state interface often behave very differently from their isolated individual domains, with distinct properties, surface chemistries, reactivities, and affinities to transformation. Many are considered multi- functional, particularly if the different domains can play distinct roles in accomplishing multiple tasks simultaneously. Importantly, adjacent domains can significantly impact the properties of one another via a solid-state heterojunction. One such class of nanostructures is colloidal hybrid nanoparticles,26-30 where the heterojunctions can serve as a platform for studying synergistic processes such as light-induced charge separation48-

51 and carrier dynamics,52,53 ,54-56 metal-support interactions,57-59 and exchange-biased magnetism.60-62 One-dimensional segmented nanowires prepared by sequential, electrodeposition within a porous template are another type of nanostructure that has different materials segments linked together in a linear array.31,63,64 These segmented nanowires have been used extensively as platforms for a wide variety of physical phenomena and applications, including nanoscale locomotion,65-67 self- assembly,68-70 photoconductivity,71 biological sensing,72-74 and .75-77

While there are clearly conceptual analogies between molecules and multi- component inorganic nanostructures, there are important differences as well. In a purified sample of a given molecule, each is precise and identical. The size and shape of a molecule, along with its arrangement of constituent atoms, are defined by the energy- minimized hybridization and directional interaction of orbitals, which gives rise to predictable bond angles and molecular geometries. Atoms and functional groups are spatially arranged, in part, as a result of the synthesis pathway. In contrast to molecules, each nanoparticle in an as-made sample has, at best, variance in its diameter of one or

5 more atomic layers. Such variance at the “artificial ” level guarantees that there will be size variance in any “artificial molecule,” and this is especially problematic for nanoparticles that exhibit size-dependent properties. Spherical nanoparticles have surfaces that are largely equivalent regardless of spatial orientation, so attachment of a second nanoparticle to its surface typically occurs at an arbitrary site and with random

“bond angles,” unless the nanoparticle has clearly defined facets, each with a different surface energy and propensity for attachment or growth of a subsequent domain.30,78,79

Molecules with multiple functional groups that are synthesized using multi-step reactions are rarely accessed in near-100% yield, but chromatographic strategies for separation and purification are both readily available and routinely practiced. Despite the irregularities that are inherent in multi-component inorganic nanostructures, strategies for separation and purification are just beginning to emerge.80-84 Such techniques are still largely at the research and development stage, and are therefore of limited availability and utility. Any variance in the morphological and compositional characteristics of the hybrid nanoparticle population is typically retained in the final product.

Finally, to construct complex molecules, chemists can draw upon an extensive library of chemical reactions that permits rational modification and linking together of smaller molecular building blocks. Multi-step reactions are performed in a predictive, stepwise manner, using a synthetic framework that is guided by both mechanistic understanding and empirical insight. This contrasts sharply with the limited reaction toolbox and mechanistic understanding available to chemists constructing multi- component inorganic nanostructures, and sequential multi-step reactions are rare. A total synthesis framework that is conceptually analogous to the approach taken by chemists to

6 synthesize complex molecules and natural products would greatly expand our capabilities for the deterministic and by-design synthesis of complex multi-component inorganic nanostructures with pre-defined morphological features and, accordingly, designer properties and functions. Such a synthetic framework would require both an expanded reaction library and mechanistic insights.

Solution-phase routes are attractive strategies for synthesizing single-component inorganic nanostructures, because they have the potential to impart morphology control on the individual crystallites (nanoparticles) and to permit inexpensive, high-throughput, liquid-based processing of films and patterned structures. A wide variety of multi- component nanostructures can be prepared by simple extensions of those strategies, so we begin the sections that follow with a general description of the colloidal, liquid-phase approach for synthesizing single-component, crystalline nanoparticles. We then highlight some of the most important concepts that underpin the of organic molecules. A discussion of the currently available reaction toolbox for nanoparticles follows, emphasizing a classification scheme for these nanoparticle transformation reactions that explicitly links them to molecular reactions. Indeed, if one thinks about individual nanoparticles as conceptually analogous to molecular functional groups, then reactions that modify nanoparticles can be thought of as conceptually analogous to textbook organic reactions. We then move to the application of such reactions to multi- component inorganic nanostructures, which parallels those transformations that target a specific functional group or moiety on a larger molecule that contains multiple reactive sites. Moving beyond reactions, we then detail some of the key synthetic strategies that underpin the construction of both complex molecules and complex multi-component

7 inorganic nanostructures, such as site-specific reactivity, orthogonal reactivity, and protection-deprotection. With this framework, we then discuss how a total synthesis approach could be used to construct high-order hybrid inorganic nanostructures using sequential multi-step reactions. Finally, we discuss the need for purification, separation, and detailed reporting about yield. These procedures and the information that they provide are ubiquitous in organic synthesis but are not mainstream for inorganic nanostructures, despite the capabilities and applications that emerge as a direct result of such post-synthesis work-up and analysis.

1.2 Colloidal synthesis of single-component crystalline nanoparticles

Over the past two decades, physical science, engineering, biology and medicine have become saturated with ideas of next-generation technologies based on materials with nanometer-sized components. Practitioners of synthetic “nanochemistry” are primarily concerned with exerting exquisite control over the composition, size, shape, and uniformity of those components. Single-domain, crystalline nanoparticles

(nanocrystals) are one such building block, which are inorganic particles with at least one dimension measuring between 1 and 100 nanometers (normally tens to a few hundred atoms, in that dimension). Although nanocrystals typically adopt the same crystal structure and composition as the extended (bulk) solid, they exhibit unique physical properties that are modulated as a result of reduced size. For example, the melting temperature of a 2-3 nm particle can be suppressed by as much as half that of the corresponding bulk solid.85 The reduced dimensions also affect remarkably on light-

8 matter interactions, because the allowable energies of electrons are restricted as a result of spatial confinement to within the nanocrystal.86 Semiconductor nanocrystals, as a result of such “quantum” confinement, exhibit size-tunable absorption and emission wavelengths, which is being researched for diverse applications in solid-state lighting, solar energy conversion, and biological detection.87 Similarly, the collective oscillations of free electrons in metal nanocrystals are spatially confined, which gives rise to strong light absorption at wavelengths that can be varied as a function of nanocrystal size and shape.88 Colloidal dispersions of some metal nanocrystals (Ag, Au, Cu) are therefore brightly colored. These examples and many others illustrate the tunable dependence of nanocrystal properties on size and shape, which has earned colloidal nanocrystals the nickname “artificial atoms”.

Colloidal nanocrystals are so-named because they are grown in solution, and they consist of an inorganic core surrounded by a shell of organic molecules. Synthesis usually involves the chemical transformation of soluble inorganic precursors in a high- boiling solvent (250-300 °C), and typical transformations include thermal decomposition, reduction, or oxidation of the precursors. Those reactions transform the precursors into reactive, transient species that are broadly termed monomers, which are subunits of the resulting nanocrystals. As the transformation proceeds, the concentration of monomers increases, and eventually, the solution reaches a state of supersaturation. The formation of nanocrystals from monomer, or nucleation, relieves the instability associated with supersaturation. If the concentration of monomers falls below a critical threshold during nucleation, and the rate of monomer generation does not exceed monomer consumption, the remaining monomer will contribute only to growth of the existing nanocrystals and

9 no new nuclei will be formed.89-92 In this way the nucleation event and the growth stage are temporally separated, which is a (generally agreed upon) condition that is necessary to synthesize nanocrystals that are uniform in size and shape. Despite the extensive development of this approach, and its applicability to a wide variety of materials, there is surprisingly little known about the identity of monomer species. Transient monomers are generated in situ, in solvents that are not volatile, and are therefore difficult to isolate and analyze. nanocrystals, for example, are thought to crystallize from monomers that resemble multinuclear iron-oxo clusters, but no structural information is yet available.93 Although it is possible to gather information about the reaction pathway by studying the isolable crystalline material at various stages of nucleation and growth, monomer generation and consumption processes are a ‘black box’ (with very few exceptions), which limits true mechanistic understanding of nanocrystal formation.

In order to prepare colloidal nanocrystals, the reaction mixture must contain surface-stabilizing . Stabilizing ligands mediate the growth and prevent aggregation of nanocrystals by binding, coating, and protecting their surfaces. They are typically organic molecules that consist of a Lewis-basic polar head group, such as carboxylate, , phosphine, phosphonate, or thiolate, and a sterically bulky tail group.94 Lewis acid-base coordination between head group and surface metal atoms adhere the stabilizers to the nanocrystals, and the tail groups provide steric protection from particle-particle interactions in the form of an organic shell. The tail group of the stabilizer also interacts with the solvent, which imparts solubility to nanocrystals in the surrounding liquid. As a result, nanocrystals can be suspended, processed, and stored in

10 liquids that are compatible with the stabilizing agent. In some cases, polymers are used as stabilizing agents.

Surface-stabilizing ligands are often present in a nanocrystal synthesis during all stages of reaction, from monomer generation, to nucleation, and during growth, and are used as size- and shape-directing agents. Surface ligands can react with, bind to, form complexes with, or otherwise interact with either nanocrystal precursors or the transient monomers, which affects the kinetics of nucleation and growth.95-100 The rate at which monomers are consumed during nucleation, for example, is strongly influenced by the presence of ligands, particularly if monomer- complexes are formed in situ. In general, when monomer is consumed quickly, a large number of smaller-sized nuclei are formed, which leaves a low concentration of monomer left in solution for growth. When monomer is consumed more slowly, a smaller number of larger nuclei are formed, and the concentration of remaining monomers is relatively higher. In the latter case, a large monomer reservoir causes fast, kinetically-controlled growth, which can result in nanocrystals that have unconventional shapes, such as rods, cubes, disks, sheets, stars, or wires.28 The crystal habit can also be altered via preferential adsorption of surface ligands onto particular crystal facets. Nanocrystal growth is arrested in the crystal direction perpendicular to the planes that are most efficiently passivated by adsorbed ligands, allowing for more rapid growth in other crystal directions.95,101 Ligand adsorption during the early stages of nucleation can also prevent atoms from ordering into their equilibrium positions, which leads to breaking of the crystal symmetry in the form of defects.102-104

Crystal defects give rise to anisotropic growth of nanocrystal shapes that are not consistent with the crystal symmetry.

11 Tailoring the size, shape, and composition of single-domain nanoparticles for assembly into hierarchical multi-domain structures, such as “artificial molecules”46 and

“artificial solids”,21-25 can be achieved using colloidal synthesis. The processing of those building blocks into films, patterned structures, and device architectures is enabled by their solubility in liquid media. A wide variety of solution-mediated, post-synthetic treatments, including nanocrystal transformation reactions (section 1.4) and surface chemistry modifications, are being developed continually. The chemical tools for manipulating the dimensions, composition and surfaces of single-component nanostructures define our capability to construct more complex, multi-component nanostructures with designer properties.

1.3 The “working principles” of molecular total synthesis

Many important details go into successfully implementing the total synthesis strategies that are used to construct large, complex molecules, and these are covered extensively in numerous textbooks and articles.105-107 However, stepping back from the details – which are clearly different from those used in the synthesis and chemical transformation of nanoparticles – there are a number of concepts, or “working principles,” that underpin such efforts, and that are important to consider for possible applicability to nanoparticle systems (Figure 1-1). First, it is usually desirable to start with reasonably simple molecular fragments that can be easily made or acquired and that structurally mimic a part of the target molecule [Figure 1-1(a)].108,109 Such molecular building blocks will typically contain reactive functional groups at specific locations,

12 which permit desired chemical modifications to be carried out in a systematic and stepwise manner. The types of chemical reactions used to modify and build outward from these molecular precursors are both extensive and diverse, and new reactions with expanded capabilities continue to be frequently reported [Figure 1-1(b)].110 Many of these reactions are named for the chemists who first reported them and are often considered as

“textbook reactions” because of their widespread use. A non-exhaustive list of classes of reactions includes oxidation, reduction, addition, substitution, elimination, condensation, cleavage, metathesis, coupling, cyclization, cycloaddition,

Figure 1-1: A multi-step synthetic concept for molecules (top) and nanoparticles (bottom), along with chemical structures and graphical examples.

group transfer, and ring opening.

Guiding principles exist for carrying out transformations on molecules where multiple reactive functional groups may be present, or where multiple sites may facilitate a particular reaction [Figure 1-1(c)]. An understanding of the principles governing selectivity allows chemists to direct reactions to occur at specific locations on a molecule,

13 leading to a targeted product.111-113 Chemoselective reactions, for example, target one functional group when others are also present – as exemplified by the borohydride- mediated reduction of ketones and aldehydes to alcohols. This reduction reaction can be carried out in the presence of a carboxylic acid or an ester, which remain unreacted by the borohydride ion. Another working principle is regioselectivity, which preferentially generates products that have a particular connectivity when other isomers may be possible, as is the case for electrophilic aromatic substitution reactions. Here, electron donating and withdrawing groups have a powerful influence over whether substitution occurs at the ortho, meta or para positions of a benzene ring. When selectivity is required but not possible for a particular set of functional groups and available reagents, undesired reaction pathways can sometimes be shut down by installing protecting groups on one or more reactive moieties, while carrying out subsequent reactions only on those that remain exposed.114 After such a reaction is carried out, the protecting group can be removed to re-expose the other functional group(s). Protection-deprotection strategies are powerful for targeting desired reaction sites and achieving selectivity in complex or multi-functional systems. Collectively, such strategies serve to facilitate orthogonal reactivity, where one functional group on a multi-functional molecule reacts in a desired way with a particular reagent under a given set of conditions while other functional groups on the same molecule do not.

Total synthesis consists of the purposeful, stepwise execution of available chemical reactions, implemented in a way that permits precise targeting of the desired functional groups at specific locations on a molecule [Figure 1-1(d)].105,107 Indeed, the power of total synthesis is not simply the reactions and their mechanisms (although this is

14 clearly important), but rather the ability to integrate multiple reactions in a predictable and sequential manner that leads to the rational design and deterministic synthesis of an arbitrary molecule with pre-defined geometries, linkages, functional groups, and spatial organization. Another important practical consequence of this framework is the availability of multiple reactions, each with distinct capabilities and unique sets of advantages and disadvantages, which can lead to the same product. For example, there are a variety of approaches that can result in the installation of an –OH group on a molecule. Considerations such as the reaction conditions, the stability and reactivity of the candidate reagents, the possible side-reactions and conceivable products associated with each approach, and the types and locations of other (possibly competing) functional groups, along with empirical guidelines, dictate when one of the available reactions is preferable over another. Finally, another practical consequence of multi-step pathways to complex molecules is decreasing yield with each additional reaction. In one sense, this is not a problem, because separation and purification tools are both widely available and routinely utilized [Figure 1-1(e)]. However, at the same time, yields are maximized when reactions go as far toward completion as possible and when reactions are carried out as selectively and as efficiently as possible, so that byproducts and undesired isomers are avoided, and the number of steps in the synthesis is minimized.115-117 Indeed, one of the key metrics of a successful multi-step synthesis is the product yield, and data to demonstrate purity of the final product – NMR, GC-MS, IR, UV-Vis, melting point, etc.

– are both readily achievable and demanded by the scientific community. These

“working principles” of molecular total synthesis are also applicable to the construction of multi-component nanostructures, although the details of how the concepts are

15 implemented are very different. The sections that follow will emphasize both the similarities and differences.

1.4 Nanoparticle reaction libraries

When planning the synthesis of a complex molecular target, chemists identify simple building blocks that can be adjoined, split apart, or modified, by relying on libraries of well-understood chemical reactions. These transformations are often classified according to the type of synthon, such as the knowledge that acyl chlorides react with alcohols to produce esters, undergo hydrolysis to make carboxylic acids, react with to generate amides, and react with carboxylic acids to make acid anhydrides.

Figure 1-2: Classes of compounds that are commonly accessible by solution-phase chemical transformation reactions of metal nanoparticle synthons.

16 Alternatively, generalized types of reactions can be grouped together, including nucleophilic substitution, elimination, addition, electrophilic aromatic substitution, and condensation reactions. During the past decade, a large body of research has been dedicated to generating libraries of chemical transformations for nanoparticles, which are the building blocks of multi-component nanostructures. This approach demonstrates that a pre-formed nanoparticle can be utilized as a synthon to generate a more complex derivative, in a manner that is similar to functional group transformation in organic synthesis.118 Nanoparticle transformations, like molecular reactions, can be classified either by synthon type (metallic nanoparticles, ionic nanocrystals, etc.), as highlighted in

Figure 1-2, or by reaction type (e.g. oxidation, coupling, etc.), highlighted in Figure 1-3.

Figure 1-3: Categories of solution-phase chemical transformation reactions that are commonly used to convert one type of nanoparticle into another.

17

Below, we will outline the most important types of nanoparticle reactions, including oxidative diffusion, reductive diffusion, metal diffusion, cation exchange, anion exchange, heterogeneous seeded-growth (addition), dissolution (elimination), galvanic replacement and coupling reactions. Detailed reviews of nanoparticle transformation reactions have been published,119,120 and are not duplicated here; important and more recent highlights are included below.

Single-element metal nanoparticles are straightforward to prepare for many systems, and can be reacted subsequently in a wide variety of ways. Oxidative diffusion reactions can be performed on metal nanoparticles to synthesize the corresponding metal oxides, chalcogenides, phosphides and nitrides. Cobalt nanoparticles, for example, oxidize spontaneously in the presence of oxygen to form CoO,121 react readily with

molecular sources of sulfur and selenium to form a variety of CoxSy and CoxSey

121,122 phases, and transform into CoP or Co2P in the presence of thermally decomposed trioctylphosphine (TOP).123,124 Similarly, when cationic metal precursors are reduced in the presence of metal nanoparticles, metallurgical diffusion reactions can transform metal nanoparticles into alloy and intermetallic phases.125 The rate of diffusion is significantly enhanced when the elemental product of metal salt reduction has a relatively low melting point, such as Pb, Sn, Bi, or In. Moreover, the rate-limiting step of solid-solid diffusion in bulk metallurgical processes is bypassed in reactions involving small metal nanoparticles, because the reduced dimensions result in short diffusion distances.

Intermetallic and alloy nanoparticles of FePt, FePt3, CoPt, PtPb, PtSn, PtBi, PtSb, Cu3Pt

18 and other derivatives, for instance, can all be prepared in solution using Pt nanoparticles as a reactive precursor.125-127

Diffusion reactions involving the oxidation of metal nanoparticles are far more common than transformations involving the reduction of ionic nanocrystals, but there are a few important examples. Reduction of metal oxide particles is a common method to prepare supported metal nanoparticle catalysts, and this is usually accomplished by thermal treatment in hydrogen or by a chemical reagent such as formaldehyde or hydrazine.128 Also, it was shown recently that TOP, when heated in the presence of binary metal chalcogenide nanoparticles, induces partial reduction to selectively extract sulfur or selenium.129 The as-made particles were typically chalcogenide-rich phases, such as SnSe2, FeS2, NiSe2, and CoSe2, which conveniently convert to the most metal- rich phases (SnSe, FeS, Ni3Se2, and Co9Se8) by heating in TOP at temperatures ranging from 65° to 270°C.

Ionic nanocrystals, including the so-called quantum dots, are susceptible to useful cation exchange reactions (Figure 1-4), which are typically driven by a higher solubility of the synthon nanoparticles in the surrounding medium than the target compound.130-139

For example, CdS and CdSe nanoparticles react rapidly at room temperature with either

+ + Cu or Ag in methanol/toluene mixtures to form Ag2S, Ag2Se, Cu2S, or

130.132,134 Cu2Se. Partial cation exchange has also been realized, giving rise to segmented

132 135 CdS-Ag2S and CdS-Cu2S nanorods, starting from CdS nanorods [Figure 1-4(a,b)].

Significantly, many of these transformations proceed with morphological retention of the substrate particle due to conservation of the anion sublattice during exchange,136 and are reversible based on the concentration and solvation energies of incoming and outgoing

19

Figure 1-4: Representative examples of nanoscale cation exchange reactions: TEM images of a) CdS nanorod precursors and b) CdS-Ag2S superlattice nanorods after partial ion exchange of the CdS nanorods with Ag+, and HRTEM images of c) wurtzite-type CdSe nanorods, d) wurtzite- type Cu2Se nanorods after ion exchange of the CdSe nanorods in (c) with Cu+, and e) wurtzite- 2+ type ZnSe nanorods after ion exchange of the Cu2Se nanorods in (d) with Zn . The conserved anion lattice, which leads to retention of the wurtzite structure type, is shown in (f). Panels (a,b) reprinted from Ref. [132] with permission (Copyright 2007, American Association for the

20 Advancement of Science) and panels (c–f) reprinted from Ref. [137] with permission (Copyright 2011, American Chemical Society). ions. Recently, multiple cation exchange reactions have been carried out sequentially using CdSe nanocrystals as a starting material, which were first reacted with Cu+ to make

2+ 137 Cu2-xSe (0≤x<1), and then reacted with Zn to generate ZnSe [Figure 1-4(c-e)]. In this case, the transformation from the more stable Cu2-xSe to lower lattice energy ZnSe was made favorable by high temperatures (250°C) and strong interactions between the solution-phase ligands and Cu+ that was extracted from the lattice. Another important consequence of anion sublattice conservation is the transfer of crystal structure information to the product nanocrystals, which has enabled access to non-equilibrium solids that are otherwise difficult to prepare. In the previous example, hexagonal (hcp) wurtzite-type CdSe nanocrystals converted to a metastable hcp phase of Cu2-xSe (the stable hcp phase is 1:1 CuSe), and again to metastable hcp wurtzite–type ZnSe [Figure 1-

4(f)]. [ZnSe typically crystallizes as the cubic (fcc) sphalerite phase.]

Anion exchange is also a powerful technique for reacting ionic nanocrystals

(Figure 1-5). Wurtzite-type ZnO nanoparticles, for example, are an intermediate in the reaction between ZnCl2 and thio- or selenourea in polyol solvent containing hydroxide ions.140 The transient ZnO particles acted as a structural template for the formation of metastable wurtzite-type ZnS and ZnSe [Figure 1-5(b)], again highlighting the conserved sublattice crystal structure during the transformation [Figure 1-5(a)]. Building upon this observation, single-crystal wurtzite-ZnS shells with a hollow interior were generated by reacting colloidal ZnO nanocrystals with hexamethyldisilathiane [Figure 1-5(c,d)].141 A void was formed in the core of each ZnS particle as a result of the nanoscale Kirkendall

21 effect, where outward O2- diffusion was faster than inward S2- diffusion, and the hexagonal pyramidal shape of the ZnO nanocrystals was conserved in the process. More recently, wurtzite-type CdS was reacted with tri-n-octylphosphine telluride (TOP=Te) to

142 form zinc blende CdTe nanocrystals. If this reaction was allowed to go only partially to

Figure 1-5: Representative examples of nanoscale anion-exchange reactions: a) Powder XRD data characterizing the transformation of wurtzite-type ZnO into wurtzite-type ZnS, b) crystal structures of the ZnO precursor and ZnS product, highlighting the retained cation sublattice, and TEM images of c) ZnO nanoparticles and d) hollow single-crystal ZnS nanoparticles synthesized by treating the ZnO nanoparticles in (c) with hexamethyldisilathiane in trioctylphosphine. Panels (a,b) reprinted from Ref. [140] with permission (Copyright 2009, American Chemical Society) and panels (c,d) reprinted from Ref. [141] with permission (Copyright 2009, American Chemical Society).

22 completion, anisotropically phase-segregated CdS-CdTe heterodimers were isolated, due to strain induced by the different crystal structures at the CdS-CdTe interface.

Heterogeneous seeded-growth reactions have been used to generate libraries of multi-component colloidal hybrid nanostructures, including an extensive variety of core- shell and a smaller (but growing) number of non-concentric heterooligomer-type geometries (Figure 1-6). This synthetic innovation is relatively new, and is a result of two primary motivations: 1) The ability to introduce additional functionality into an already functional unit, giving rise to multi-functional nanostructures with useful property combinations that may be able to accomplish multiple tasks simultaneously and

2) the desire to study how inorganic domains synergistically communicate when adjoined

Figure 1-6: Representative examples of nanoscale addition reactions: TEM images of a) Au, b) Ag, c) Pt, and d) AuAg nanoparticle seeds, and e) Au-Fe3O4, f) Ag-Fe3O4, g) Pt-Fe3O4, and h) AuAg-Fe3O4 hybrid nanoparticle products formed by heterogeneous seeded growth of Fe3O4. Reprinted from Ref. [59] with permission (Copyright 2010, American Chemical Society).

by a solid-state interface. Conceptually, the introduction of a new functionality to a pre- existing unit is analogous to an addition reaction in organic synthesis, such as the

23 epoxidation of olefins. The study of heterogeneous seeded growth has already given rise to numerous metal-metal, metal-metal oxide, metal-semiconductor, metal oxide- semiconductor, and semiconductor-semiconductor hybrid nanostructures with a variety of two-component architectures, and the library is growing rapidly. These results are summarized in a number of recent review articles.26-30

As the name implies, seeded-growth “additions” require a second inorganic domain to grow directly from the surface of a pre-formed nanoparticle seed, which is usually accomplished by selecting reaction conditions that suppress homogenous nucleation of isolated nanoparticles. Conditions that promote heterogeneous nucleation include much lower concentrations of molecular precursors and lower temperatures than are normally required to nucleate single-component nanoparticles. The final morphology of colloidal hybrid nanostructures (core-shell, anisotropic, multi-domain, etc.) however, is influenced by many variables. In many cases, these variables are elucidated largely based on empirical observations and they can be system-dependent, including lattice matching, solvent polarity, heating rate, seed-to-precursor ratio, surfactant concentration and the availability of reactive crystal facets on the seeds. Despite the lack of materials- general observations regarding what governs the morphology of colloidal hybrid nanostructures, sophisticated hypotheses are continually being developed, substantiated by in-depth characterization of solid-solid interfaces and mechanistic insights.143-145 For example, Cao and co-workers carried out extensive high-resolution transmission electron microscope (HRTEM) studies of FePt-In2O3 heterodimer nanoparticles and revealed that

FePt (110) and In2O3 (222) junctions and FePt (200) and In2O3 (440) junctions were both present in their samples, with the population heavily weighted toward the former. This

24 was surprising, because the lattice pair FePt (200) / In2O3 (440) was more closely matched than the FePt (110) / In2O3 (222) pair, which suggested that the coordination environment of atoms at the FePt-In2O3 interface and bond strengths play a more important role than minimization of lattice mismatch.144

An analogue of organic elimination reactions, rather than addition, is also useful in nanostructure synthesis, and can be considered as the simple dissolution or etching of an inorganic solid. There have been many examples of both in situ and post-synthetic modification of metal nanoparticles (i.e. Au, Ag, Pt, Pd) using crystallographically- selective wet etchants, which have resulted in unconventional, non-equilibrium morphologies.146-149 Metal nanoparticles can also be sacrificially oxidized by metal cations having a higher reduction potential, and thereby galvanically replaced to create new metal and alloyed nanostructures that often have a hollow interior.150-153 For example, Ag nanocrystals with a wide variety of shapes can be prepared readily, and

– subsequently reacted with AuCl4 ions, to make hollow Au and AuAg alloy nanostructures.

Pre-formed nanoparticles can also be controllably positioned or grouped into stoichiometrically-controlled clusters using “coupling” processes, which has been reviewed.46,47,154 Coupling reactions allow nanoparticles to be spatially and geometrically arranged in close proximity to one another, to examine how the properties of the ensemble depend on its dimensionality, and differ from those of the isolated particles.

This is particularly interesting for coupling plasmonically-active (i.e. Au, Ag, Cu) nanoparticles, because hybridization of surface plasmon resonance is strongly dependent on interparticle distance, which has important implications for monitoring biological

25 processes using microscopy and spectroscopy.154 Coupling of inorganic nanoparticles is often facilitated by molecular interactions with a high degree of chemical recognition such as DNA base-pairing, antibody-receptor binding, predictable Lewis acid-base pairing (such as Au-thiol) and surface-ligand . In some cases, coupling is achieved by oriented attachment, where two or more nanoparticles become crystallographically aligned, and form a solid-solid interface at a particular facet. This process eliminates a significant fraction of undercoordinated surface atoms, and is therefore driven by minimization of surface energy.155 Alternatively, groupings of nanoparticles can also be encapsulated within inorganic glass and/or polymer matrices, or for larger colloids, can form through controlled de-wetting or liquid-phase encapsulation techniques. A variety of architectures have been achieved via controlled coupling, including dimers, trimers, tetramers, linear chains, branched chains, and other network- type geometries. Another strategy for nanoparticle coupling involves the thermally- driven interdiffusion of two or more pre-existing domains. When applied to multi- domain nanostructures, which will be discussed in another section, high-order oligomers and complex solid-state networks can be formed without the need for molecular of biological linkers.

1.5 Reaction toolkit for multi-component inorganic nanostructures

The previous section described many conversion chemistry reactions that have been used to transform single-domain nanoparticles, along with their similarity to organic transformations. Far fewer reports exist, however, where analogous reactions are used to

26 modify inorganic nanostructures that have multiple domains. When the substrate to be modified possesses more than one component, a number of interesting questions arise: Is the given reaction selective for one of the components, or will multiple domains react under these conditions? Can reaction conditions be found that promote preferential conversion of one component? Does the component to be modified react differently in the heterostructure than it does when isolated? How is the mechanical and structural integrity of the heterojunction affected by the reaction – is it robust or weak? In many ways, these questions and the associated challenges are similar to those that organic chemists encounter when planning the synthesis of a complex molecule. Below, we highlight the synthetic concepts that are unique to multi-component structures – including orthogonal reactivity, site-selective transformations, coupling reactions, and protection- deprotection strategies – along with the reactions that have been demonstrated to facilitate this in nanoparticle systems.

1.5.1 Orthogonal reactivity and site-selective transformations

Some of the most powerful tools for transforming multi-component nanostructures are those reactions that proceed with site-selectivity, which enables the controlled positioning of a specific functionality within a multi-functional object. Simply put, site-selective reactions either fully or partially transform one domain in a multi- domain nanostructure while leaving others unmodified. Site-selective transformations are enabled when the reactivities of different components within a heterostructure are said to be orthogonal to one another, that is, when one (or part of one) component is

27 susceptible to transformation under a particular set of conditions while the other components do not react. Site-selectivity is in some cases predictable – over a wide range of relevant reaction conditions, for example, standard reduction potentials (E°,

Mn+/M) dictate that certain components within a heterostructure will react readily with an oxidant (i.e. Fe, Co, Ni, Cu) while others will not (i.e. Au, Pt, Fe2O3). Orthogonal reactivity is not always intuitive, however, and those transformations that proceed site- selectively are fertile ground for in-depth study and development.

Oxidative, reductive, and metallurgical diffusion-mediated reactions are useful for modifying single-component nanoparticles, but application of these types of transformations to multi-domain nanostructures is comparatively rare. In many of these cases, however, diffusion reactions can site-selectively transform a targeted component without significantly altering the heterostructured morphology. To demonstrate this principle, Leonard et al. investigated some metallurgical diffusion reactions of physical mixtures of metal nanoparticles (Ni, Cu, Rh, Pd, Ag, Pt, Au, and Sn) with metal salts under reducing conditions at several temperatures, and observed a number of orthogonal relationships.127 Although different nanoparticles reacted with the same metal salt solution, they did so at significantly different temperatures, providing a basis for reacting a targeted component in the presence of several reactive species. For example, many metal nanoparticles reacted with Sn to form binary intermetallics, but AuSn formed near room temperature, Ni3Sn4 formed at 100 °C, Ag4Sn formed at 175 °C, and PdSn2 and

PtSn formed at 200 °C. These observations, along with the knowledge that some combinations of metal nanoparticles and metal salts did not react, allowed one-pot two- component orthogonal reactivity to be demonstrated in several systems. A mixture of Ag

28

Figure 1-7: Top: Schematic representation highlighting orthogonal reactivity in striped metal nanowire systems: Reagent X reacts with A to form AX, but does not react with B; reagent Y reacts with B to form BY, but does not react with AX. SEM images of a) Au-Ag-Au striped metal nanowire and b) Au-Ag2S-Au nanowires formed after reacting with sulfur. SEM images of c) an Ag-Pt striped metal nanowire, d) an Ag-PtPbx striped nanowire after reacting the Ag-Pt nanowire with Pb (element maps show incorporation of Pb into the Pt segment, but not the Ag segment), and e) an Ag2S-PtPbx striped nanowire after treating the Ag-PtPbx nanowire with sulfur. Scale bars: 1 µm. Panels (a–e) reprinted from Ref. [127] with permission (Copyright 2009, American Chemical Society).

and Ni nanoparticles, for example, were reacted with SbCl3 at 100 °C to form intermetallic NiSb, leaving the Ag unreacted, and the mixture of Ag and NiSb

29 nanoparticles was then reacted with SnCl2 at 175 °C to form a mixture of Ag4Sn and unreacted NiSb nanoparticles.

These and other derived orthogonal relationships were then applied to the controlled modification of electrochemically-deposited segmented nanowires, which resulted in the synthesis of various linearly-patterned, multi-element, multi-domain nanowires (Figure 1-7). For example, three-domain Au-Ag-Au striped nanowires reacted with sulfur to generate Au-Ag2S-Au wires, wherein sulfur selectively reacted with the Ag domains and not with Au [Figure 1-7(b,c)].127 Similarly, Seo and co-workers have shown that colloidally-grown Ag-Au-Ag nanorods dispersed homogeneously in water [Figure 1-

156 8(a-c)] can be converted to Ag2S-Au-Ag2S by reacting with Na2S [Figure 1-8(d-g)].

Electrodeposited nanowires with Ag and Pt segments reacted under reducing conditions with Pb(CH3COO)2 to form Ag-PtPbx nanowires (PtPbx is a mixture of PtPb and Pt4Pb phases), which were reacted subsequently with sulfur to generate Ag2S-PtPbx segmented nanowires [Figure 1-7(c-e)].127 This two-step synthesis relied on the fact that Pb0 reacts readily with Pt over a wide range of temperatures, but not with Ag, and that PtPbx is resistant to sulfidation, allowing sequential, site-selective conversion of each domain.

Solution-phase diffusion reactions represent an alternative to the ostensibly tedious, system-specific optimization required for electrochemical co-deposition of such multi- element and non-conducting segments. Other spatially organized, multi-domain nanostructures, such as lithographically patterned surfaces and wires, are also viable substrates for stepwise modification via diffusion reactions.127

Unlike template-grown striped nanowires and patterned surfaces, colloidal hybrid nanoparticles can be dispersed homogeneously in a variety of solutions, and would seem

30 to be ideal substrates for chemical modification using site-selective diffusion reactions.

There are, however, surprisingly few examples. Heterodimers of Pt-Fe3O4 are typically

Figure 1-8: a) TEM image (scale bar: 50 nm), electron diffraction pattern, and b) Au and c) Ag elemental mapping data for a colloidal Ag-Au-Ag striped nanorod. d) TEM image and elemental mapping data of striped Ag2S-Au-Ag2S nanorods for e) Au, f) S, and g) Au + S, The nanorods were formed after treating the Ag-Au-Ag nanorods with aqueous Na2S. TEM images (scale bars: 200 nm) of h) Bi2Te3 nanowires with Te tips and i) Bi2Te3 nanowires with PbTe tips after treating the Bi2Te3-Te nanowires with Pb. Panels (a–g) reprinted from Ref. [156] with permission (Copyright 2008, American Chemical Society) and panels (h,i) reprinted from Ref. [160] with permission (Copyright 2012, American Chemical Society).

31 prepared by thermally decomposing Fe(CO)5 in the presence of Pt nanoparticle seeds, which forms intermediate Pt-Fe heterodimers that require the Fe domain to be subsequently reacted with O2. Similarly, a conformal shell of Fe has been deposited around Au seeds, and subsequently oxidized by bubbling air through the product mixture

157 to obtain Au/Fe3O4 core/hollow-shell (also called yolk/shell) nanoparticles. Other yolk/shell structures, which are interesting platforms for biological imaging and drug

121 158 159 delivery, include Pt/CoO, FePt/Fe3O4, and FePt/CoS2, which were prepared by reacting the corresponding Pt/Co, FePt/Fe, and FePt/Co core/shell nanoparticles with either O2 or sulfur. The Kirkendall effect was responsible for the formation of voids between the core and the shell. Very recently, Te-Bi2Te3 hybrid nanoparticles with a barbell-like morphology have been synthesized for thermoelectric studies, where Bi2Te3 rods are flanked by plate-like tips of Te [Figure 1-8(h)]. The Te termini were then

2+ reacted selectively with Pb under reducing conditions to form PbTe-Bi2Te3 hybrid particles with the same morphology [Figure 1-8(i)], and the authors speculate that Pb could possibly be replaced with a variety of other metals that react with Te, such as Ag,

Sn, and Sb.160

Orthogonal reactivity in cation exchange reactions has been observed for several ionic nanocrystal systems and utilized to synthesize multi-domain nanostructures.

Information regarding the synthesis of dimensionally-controlled cadmium chalcogenide nanocrystals is abundant, so that CdX (X = S, Se, Te) quantum dots, rods, and branched structures have been ideal starting points for studying cation exchange reactions.

Although Cd2+ cations are extracted and replaced readily with soft metal cations (Ag+,

Cu+, Pd2+, Pt2+) in polar solvent mixtures (methanol/toluene, aqueous/organic),130,133,139

32 conditions that allow other metal cations (Pb2+, Zn2+) to exchange with Cd2+ are more difficult to find. To circumvent this challenge, Luther et al devised a two-step synthesis of PbS nanocrystals that relied on the formation of trialkylphosphino-Cu(I) and -Ag(I) complexes to efficiently extract either Cu+ or Ag+ from their respective lattices in the presence of Pb2+ ions.134 In the previous section we described a similar, sequential cation exchange-enabled synthesis of ZnSe nanocrystals, using CdSe as a template and

137 proceeding through a Cu2Se intermediate phase. In other words, the reactivities of CdX

2+ 2+ and Cu2X (or Ag2X), with cations like Pb and Zn , are orthogonal over a wide range of

Figure 1-9: Top: Schematic representation showing the stepwise conversion of a CdS nanorod into a PbS-CdS striped nanorod by two sequential ion-exchange reactions and a Cu2S-CdS nanorod intermediate. HRTEM images of a) a Cu2S-CdS striped nanorod formed by treating CdS + nanorods with Cu and b) a PbS-CdS striped nanorod formed by treating Cu2S-CdS nanorods with Pb2+. Panel (a) reprinted from Ref. [135] with permission (Copyright 2009, American Chemical Society) and panel (b) reprinted from Ref. [134] with permission (Copyright 2009, American Chemical Society).

33 reaction conditions. Sadtler et al thoroughly investigated partial cation exchange

+ between CdS nanorods and Cu , and found that the energy of CdS-Cu2S interface formation was minimized when Cu2S nucleation occurred at the nanorod ends, followed by inward growth.135 Furthermore, the reaction was kinetically favored at the Cd-rich terminus of the nanorods, so that CdS nanorods with either one or two Cu2S segments could be synthesized [Figure 1-9(a)]. The orthogonal reactivity of CdS and Cu2S enabled

2+ site-selective modification of the Cu2S domains via cation exchange with Pb , leaving the CdS domains unreacted, and templating formation of a variety CdS-PbS nanorod

134 heterostructures [Figure 1-9(b)]. Nanorods with alternating CdS and Ag2S segments were also reacted with Pb2+, leading to a topologically controlled, site-selective cation exchange with Ag2S, and a multi-domain nanostructure with numerous PbS segments embedded in CdS nanorods.

Heterogeneous seeded-growth reactions are site-selective when deposition is favored at a particular location on a seed particle, which may have several sites that could accommodate the addition of a new domain. For example, nanoparticle addition to highly faceted or anisotropically-shaped seeds can favor growth from a particular area, such as the terminus of a nanorod. Nanocrystals having the hcp wurtzite structure (i.e.

CdX, ZnO, ZnS) are routinely grown into highly anisotropic shapes, including nanorods and branched multipods, due to incomplete passivation of high-energy (001) surfaces, resulting in preferential growth along the c-axis.161,162 In the seminal paper by Mokari et al, Au domains were grown site-selectively from either one or both ends of CdSe nanorods, which was also explained by the high surface energy and weak ligand adsorption to the (001)-terminated tips.163 The selectivity was assigned to a mechanism

34 wherein soluble Au complexes were preferentially adsorbed to the CdSe nanorod tips, followed by reduction. Another interesting feature of the wurtzite structure is the lack of inversion symmetry perpendicular to its c-axis, which renders (001) and (001) facets compositionally and energetically non-equivalent, and implies a driving force for terminus-selective heterogeneous growth from wurtzite-type nanorods. Kudera et al utilized this non-equivalence to controllably deposit PbSe domains on either one or both tips of CdS and CdSe nanorods (Figure 1-10).164 In addition to crystal structure effects, the surface energetics associated with site-selective addition to anisotropic seeds are

Figure 1-10: TEM images of CdSe nanorods, selectively capped with PbSe at a) both ends (PbSe-CdSe-PbSe) and b) one end (PbSe-CdSe). Reprinted from Ref. [164] with permission (Copyright 2005, American Chemical Society).

complicated by many variables, including ligand dynamics, surface curvature, tapering, and epitaxy. This subject has been thoroughly reviewed by Carbone and Cozzoli.26 Our focus will be on heterogenous seeded-growth reactions where a new domain was added

35 to a multi-domain nanostructure, with an emphasis on site-selectivity that was promoted by the heterostructured seed.

The eccentric CdSe/CdS core/rod shell hybrid nanocrystals developed by Manna et al are an interesting platform for studying site-selective seeded-growth from a multi- domain nanostructure.165 Organometallic precursors are decomposed in the presence of

CdSe/CdS seeds to make metal-semiconductor nanostructures for charge separation and light harvesting studies. Dukovic et al have shown, for example, that photodeposition of

Pt onto single-component CdS nanorods gave random decoration of the lateral facets with multiple Pt domains. When the CdSe/CdS core/rod shell structures were used as seeds instead, only one large Pt domain was grown, at a location corresponding to the embedded CdSe core [Figure 1-11(a)].166 The authors speculated that photoexcited electrons and holes experienced confinement due to a potential minimum in the CdSe domain, which may have influenced the site-selective growth. Similarly, Menagen et al reported growth of Au domains localized to the buried CdSe core [Figure 1-11(b)].167

Initially, they observed random decoration of the nanorod sidewalls with small Au clusters. The tendency of conduction electrons to flow from the CdS domain to the lower-potential CdSe seed caused oxidation of the Au clusters, which increased the likelihood that larger Au domains would grow and persist at the seed position. In a subsequent report by Carbone et al, the tendency for growth to occur near the core region of the seeds was surmounted by assisting Au(III) reduction with UV photoexcitation, where a large, plasmonically-active Au domain could be spatially directed toward one end of the CdSe/CdS nanorods.168 More recently, UV excitation of Au-tipped CdSe/CdS nano-matchsticks was used to selectively photodeposit Pd and Fe metal onto the Au

36 domains.169 Heterogenous deposition and subsequent oxidation of Fe gave rise to termination of each nanorod with Au/FexOy yolk-shell domains [Figure 1-11(c)], which is similar to what was observed by Schevchenko et al during thermal decomposition of

157 Fe(CO)5 over Au seeds. Instead of heterostructured termini, UV photodeposition of Pd resulted in random alloying between the Au domains and Pd.169

Thermally-induced decomposition of Co2(CO)10 over the dot-in-rod CdSe/CdS seeds generated matchstick-like hybrid nanoparticles, with one Co domain linked selectively to the terminus farthest from the CdSe core.62 In contrast, replacing the core/rod shell seeds with single-domain CdS nanorods resulted in random decoration of the lateral sidewalls with Co. Site-selective Co growth was rationalized based on a difference in permanent dipole moment (arising from asymmetry perpendicular to the wurzite c-axis) between the core/shell CdSe/CdS seeds and pure CdS nanorods, which potentially originated from lattice strain around the CdSe-CdS interface. Also, single- domain Co nanocrystals synthesized under identical control conditions had the cubic ε-

Co crystal structure, but Co-CdSe/CdS matchsticks contained hcp-Co, presumably from adjoining to the hcp-CdS apex. Synergistic interaction at the heterojunction between Co and CdS also gave rise to unexpected at room temperature.

Chakrabortty studied the reduction of Au(III) salts in the presence of CdSe/CdS core/rod shell seeds, and reported site-selective deposition of Au domains based on systematic variations in precursor concentration.170 At low Au(III) concentrations, Au-

CdSe/CdS with a matchstick hybrid structure were obtained, with the Au domain located at the nanorod tip farthest from the CdSe core, similar to the deposition of Co. The formation of the substrate core/shell nanorods results from heterogeneous nucleation of

37 CdS over CdSe seeds, followed by kinetically-favored growth of the (001) facets, which are terminated primarily by sulfur atoms [The (001) face is Cd-rich]. Au-CdSe/CdS nanomatchsticks likely resulted from rapid attachment of Au atoms to the S-rich (001) face, in the limit of low precursor concentrations. Systematically increasing the Au(III) concentration gave rise to symmetrically-tipped Au-CdSe/CdS dumbells, and above a threshold concentration, to random decoration. Notably, many randomly decorated nanorods had one larger Au domain co-located with the CdSe core, consistent with

Figure 1-11: TEM images showing a) Pt nanoparticles photodeposited selectively at the CdSe core region of CdSe/CdS core–shell nanorods (scale bar: 20 nm), b) Au nanoparticles deposited selectively at the CdSe core region of CdSe/CdS core–shell nanorods (HAADF-STEM image, scale bar: 20 nm), c) iron oxide deposited selectively at the Au end of Au-tipped CdS nanorods, and d) Au and Ag2S selectively deposited at opposite ends of a CdS nanorod (Au-CdS-Ag2S). Panel (a) reprinted from Ref. [166] with permission (Copyright 2008, Wiley-VCH), panel (b) reprinted from Ref. [167] with permission (Copyright 2008, American Chemical Society), panel (c) reprinted from Ref. [169] with permission (Copyright 2010, American Chemical Society), and panel (d) reprinted from Ref. [170] with permission (Copyright 2010, Wiley-VCH).

38 aforementioned reports. Replacing Au(III) with similarly-low concentrations of Ag(I) induced site-selective cation exchange at one end of the nanorod, forming Ag2S-

CdSe/CdS matchsticks. The high degree of selectivity in Au and Ag2S growth enabled

CdSe/CdS nanorods to be reacted sequentially with Au(III) and then Ag(I), leading to asymmetrically-tipped Ag2S-CdSe/CdS-Au dumbbell-like hybrid nanoparticles with four distinct inorganic domains [Figure 1-11(d)].

To reduce the lattice mismatch-induced strain that can lead to interfacial defects and charge-carrier trap states in core-shell nanocrystals, seeded-growth reactions have been carried out sequentially to grow multiple, concentric shells around a nanocrystal core. Even for core-shell systems with a modest lattice mismatch, such as CdSe/CdS (3.9

%), growth of more than a few (>2) monolayers of shell material introduces significant strain.171 In core-multishell “onion-like” nanostructures, multiple heterojunctions extend radially from the center, which are characterized by a gradient of gradually increasing lattice mismatch and band gap offset, allowing the growth of thicker shells that more adequately confine excitons within the core. Examples of core-multishell nanostructures prepared by stepwise seeded-growth include InAs/CdSe/ZnSe,172 CdSe/CdS/ZnS,173-175 and the quantum dot-well-dot system CdS/HgS/CdS.176,177 Similarly, Xu et al deposited

178 multiple metallic shells around a magnetic Fe3O4 core in a sequential manner. Growth of thicker Au shells from Fe3O4/Au core/shell seeds resulted in red-shifting of the visible absorbance due to surface plasmon resonance [SPR], while growth of a Ag shell formed

Fe3O4/Au/Ag core/multiple shell structures that caused blue-shifting in the SPR band.

When compared to situations in which core/shell-type seeds are used, heterogeneous seeded-growth reactions become more complex when the substrate seeds

39 possess multiple, chemically distinct surfaces that are exposed to the surrounding growth solution. In a situation where two-component heterodimers are utilized as seeds, for example, from which surface does the third component grow, and why? Although complex mixtures of multi-component products might be expected in such reactions, heterogeneous nucleation is often favored at one particular surface. One of the first examples involved the growth of lead chalcogenide domains from Au-Fe3O4 heterodimer seeds, which formed ternary colloidal hybrid nanoparticles with unprecedented semiconducting-metallic-magnetic functionality [Figure 1-12(a,b)].179,180 Heterogeneous nucleation of lead chalcogenides always occurred site-selectively at the Au surface, and never Fe3O4, giving rise exclusively to multi-domain nanoparticles with PbX-Au-Fe3O4

(X = S, Se) connectivity. The authors proposed that site-selective addition of PbS domains was driven by rapid adsorption of S atoms onto the solvent-exposed Au surfaces, and the subsequent reaction of those S atoms with Pb-oleate precursor complex.

Clear mechanistic differences between the reaction of Pb-oleate complex with the sulfur and selenium [trialkylphosphine-selenium complex, TOP-Se] precursors suggested that a different driving force was responsible for heterogeneous nucleation of PbSe.180 Several experiments suggested that Au-Fe3O4 seeds simply provide a low-energy Au surface onto which small PbSe nuclei attach, coalesce and grow. The PbSe morphology was controllable by varying the Au-Fe3O4 to Pb-oleate [seed-to-precursor] ratio, where higher seed-to-precursor ratios gave rise to spherical PbSe domains, and lower ratios generated rod-shaped PbSe domains. When single-component Au nanoparticles were used as seeds over a broad range of seed-to-precursor ratios, however, Au/PbSe hybrid particles with a core/shell morphology formed, and rod-shaped PbSe domains were not observed.

40

Figure 1-12: TEM images of a,b) PbSe nanorods grown selectively off the Au domain of Au- Fe3O4 heterodimers to generate PbSe-Au-Fe3O4 heterotrimers, c) Au-SiO2 heterodimers and d) Ag-Au-SiO2 heterotrimers formed by growing Ag selectively off the Au domains, and e,f) various Au-Pt-Fe3O4 heterotrimers formed by the selective growth of e) Au off the Pt domains of Pt-Fe3O4 heterodimers and f) Fe3O4 off the Au domains of Au-Pt heterodimers. Panels (a,b) reprinted from Ref. [180] with permission (Copyright 2006, Wiley-VCH), panels (c,d) reprinted from Ref. [181] with permission (Copyright 2010, American Chemical Society), and panels (e,f) reprinted from Ref. [183] with permission (Copyright 2008, American Chemical Society).

Remarkably, nanorod-shaped PbSe growth was due the presence of the Fe3O4 domain, which limited the surface area of Au that was exposed, and that could accommodate nucleation of PbSe. This ensured that significant quantities of Pb and Se precursor

41 remained available post-nucleation, which is consistent with anisotropic growth and the resultant rod-like domains. Furthermore, even lower concentrations of Au-Fe3O4 seeds resulted in PbSe domains with multiple, rod-like, and branched domains.

In a way that is analogous to organic molecule synthesis, a particular sequence of transformations is often required to construct a multi-domain nanostructure with a targeted morphology or connectivity. The order in which sequential seeded-growth additions are carried out has a profound effect on whether or not site-selective heterogeneous nucleation is observed. Colloidal hybrid Ag-Au-SiO2 nanoparticles, for example, were formed by partial encapsulation of spherical Au nanoparticles in silica, followed by site-selective growth of Ag from the Au surface that remained exposed

[Figure 1-12(c,d)].181 Partial encapsulation was facilitated by two ligands, 4- mercaptophenylacetic acid [4-MPAA] and poly(acrylic acid) [PAA], that competed with each other to replace native citrate ligands on the Au surface. The portion of the Au surface coated in 4-MPAA was amenable to encapsulation with silica, and the PAA- terminated surface was not, resulting in partial encapsulation. [Note that the reactivity of

4-MPAA-, and PAA- functionalized Au surfaces are orthogonal with respect to silica deposition.] The resulting Ag-Au-SiO2 connectivity would have been nearly impossible to obtain if Ag was first deposited onto Au, followed by silica growth. First of all, deposition of Ag onto Au seeds typically results in core-shell structures, due to the similar lattice constants of Au and Ag. Notably, the non-concentric morphology obtained here gave rise to a new SPR band at 674 nm, in addition to the visible SPR bands that Au and Ag display at 525 and 405 nm, which was assigned to longitudinal plasmon coupling between the Au and Ag domains. Secondly, site-selective encapsulation of Au in the

42 presence of Ag is unlikely, because Au and Ag colloids can both accommodate growth of a silica shell.182

Zhang et al studied various sequences of seeded-growth reactions involving Pt,

183 Au, and Fe3O4, and obtained a wide variety of ternary morphologies [Figure 1-12(e,f)].

Multi-domain nanostructures comprised of branched Pt nanorods with either one or multiple tips made of Au were used as seeds for the addition of Fe3O4 domains via thermal decomposition of Fe(CO)5. The final morphology of the Pt-Au-Fe3O4 heterostructures was assigned to variation in the decomposition rate of Fe(CO)5, which was observably dependent on its concentration and the reaction temperature. Provided that the reaction temperature and precursor-to-seed ratio were sufficiently low, growth of

Fe3O4 site-selectively added to either one or multiple Au domains when they were below a certain size. In contrast, encapsulation of the branched Pt domains with Fe3O4 was favored when the Au domains were relatively larger. Above a threshold concentration of

Fe(CO)5, or temperature, the Au-Pt seeds were completely encapsulated by Fe3O4. When pre-formed Pt-Fe3O4 heterodimers were used as seeds instead, Au domains grew site- selectively from the Pt surface, and not from Fe3O4, which demonstrated that the order of reactions plays a significant role in determining heterostructure morphology.

Just as seeded-growth represents a transformational analogue of organic addition reactions for the bottom-up construction of inorganic nanostructures, elimination reactions correspond to the top-down modification of multi-component nanostructures via deconstruction. This is accomplished using oxidizing agents that either partially or completely dissolve one or more components in a site-selective manner. One prominent example, termed “on-wire ”, involved wet-chemical etching of template-

43 grown striped nanowires with either alternating Au-Ag or Au-Ni segments [Figure 1-

13(a)].184 Selective elimination of Ag or Ni was accomplished by treatment with either

HNO3 or a methanolic solution of H2O2 and NH4OH, respectively, which left Au segments intact. The striped nanowires were thereby transformed into arrays of Au disks, separated by size-tunable gaps perpendicular to the longitudinal nanowire axis.

Insulating gaps in an otherwise conductive nanowire are an interesting platform for studying gap functionalization and its effect on electrical transport,185 Surface-enhanced

Raman scattering [SERS],186 and bio-molecular sensing.187 Another demonstration of selective dissolution involved the use of multiply-twinned Ag nanoparticles as morphological templates, and resulted in remarkable, ultra-thin [<2 nm] nanoframes of

188 Au [Figure 1-13(b)]. Under optimized conditions, HAuCl4 was reduced with ascorbic acid, selectively depositing Au onto the edges and corners of decahedral Ag nanoparticles. Hydrogen peroxide solution was then used to dissolve the Ag component, leaving decahedral Au nanoframes behind. Other faceted Ag nanoparticles, such as pentagonal nanorods and icosahedra, were also viable templates for Au nanoframe formation. Notably, the same techniques have also been used to selectively dealloy a target metal from bimetallic alloy nanoparticles, such as the dissolution of Ag from AuAg alloy nanostructures.189

Recently, selective dissolution of Ag from the lateral sidewalls of Au/Ag core/shell nanorods resulted in dumbbell-like nanostructures with Au/Ag core/shell domains remaining only at the tips [Figure 1-13(c)].190 Metallic Ag was oxidized using

Fe3+ ions, and cetyltrimethylammonium bromide (CTAB) surfactant promoted etching by strongly ligating to Ag+. Silver atoms along the lateral facets were oxidized faster than

44

Figure 1-13: Top left: Schematic represention of nanoscale elimination reactions by selective dissolution. a) FESEM image of an Au nanowire with 25, 50, and 100 nm gaps created by selectively dissolving the Ni components of Au-Ni multisegmented nanowires. TEM images of b) Au nanoframes formed by facet-selective overgrowth of Au on decahedral Ag nanoparticles and selective dissolution of Ag (scale bar: 50 nm), c) Ag-tipped Au nanorods formed by selectively etching Ag from the sides of Au@Ag core–shell nanorods, d) dented Fe3O4 nanoparticles formed by selectively dissolving the Au domains of Au- Fe3O4 heterodimers, and e) Fe3O4 nanocontainers formed by selectively dissolving the Au domain of Au-Fe3O4 heterodimers with additional Fe3O4 covering the Au. Panel (a) reprinted from Ref. [184] with permission (Copyright 2005, American Association for the Advancement of Science), panel (b) reprinted from Ref. [188] with permission (Copyright 2011, American Chemical Society), panel (c) reprinted from Ref. [190] with permission (Copyright 2012, American Chemical Society), panel (d) reprinted from Ref. [191] with permission (Copyright 2010, Wiley-VCH), and panel (e) reprinted from Ref. [192] with permission (Copyright 2011, American Chemical Society).

45 those at the nanorod tips, which contradicted previous studies pertaining to wet chemical etching of pure Au nanorods. Adsorption of CTAB to the nanorod sidewalls, which correspond to {110} crystal facets for fcc metals, was expected to impart favored reaction kinetics at the tips. Zeta potential measurements and thermogravimetric analysis indicated that the CTAB populates the surface of core/shell Au/Ag nanorods to a much lower extent than pure Au nanorods. It was therefore concluded that the Ag dissolution rate was proportional to intrinsic surface energies [for fcc, 110 > 100 > 111]. The Ag - tipped Au nanorods exhibit higher catalytic activity toward the room-temperature reduction of p-nitrophenol than either Au/Ag core/shell or pure Au nanorods, which the authors ascribed to an electronic synergetic effect between Au and Ag. Similarly, selective elimination of either Au or Fe3O4 from Au-Fe3O4 heterodimers was used to demonstrate that Au-Fe3O4 particles exhibit a synergetic enhancement in catalyzing the

191 reduction of H2O2. Single-component Au and Fe3O4 nanoparticles were obtained by dissolving either the Au domain using KI/I2 solution, or the Fe3O4 domain with H2SO4

[Figure 1-13(d)]. The authors were then able to compare the catalytic activity of Au,

Fe3O4, and the parent Au-Fe3O4 heterodimers separately and under identical conditions.

Selective etching of Au with I2 was also demonstrated by George et al, which eliminated the Au domain from a series of asymmetric core/shell and Au-Fe3O4

192 heterodimers [Figure 1-13(e)]. Thermal decomposition of Fe(CO)5 in the presence of

Au nanoparticles, at temperatures in the 180-200°C range, results in slow seeded- growth of Fe3O4 shells. Increasing the temperature, however, causes the core and shell to become increasingly non-centrosymmetric, and eventually the Au and Fe3O4 surfaces spatially segregate into the heterodimer morphology. The authors found that I2 molecules

46 selectively leached Au from either the asymmetric core/shell Au/Fe3O4 particles or the

Au-Fe3O4 heterodimers, while the remaining Fe3O4 domain remained morphologically unchanged. The core/shell domains were transformed into porous shells with a large internal void space, or so-called nano-containers, while the heterodimers evolved into nanoparticles with a large concave region. Inorganic structures that mimic containers are desirable for applications that require storage and controlled release of guest species, and the authors showed that the cavity of “dense Fe3O4-shell Fe3O4” hybrid nano-containers could indeed be loaded with the molecular anti-cancer agent cis-platin. After coating the cis-platin loaded dense Fe3O4-shell Fe3O4 nano-containers with a polymer to confer water dispersability, ultrasonication was applied to rupture the porous shell-like domain, releasing the drug into solution. Selective I2 etching of Au was also carried out on AuPt-

Fe3O4 heterostructures, which deconstructed the AuPt domains by completely leaching

Au, leaving small (1-3 nm) Pt domains trapped within the void space or concave region.

The Pt/Fe3O4 nanocontainers were efficient photocatalysts for reduction of an organic dye, and were magnetically recoverable from the product mixture.

1.5.2 Merging of multi-domain nanostructure fragments: “coupling” reactions

In order to construct large molecules, chemists frequently link together smaller molecular fragments using coupling reactions. Similarly, reactions that link nanoparticles into higher-order hybrid structures have the potential to significantly expand our capabilities for designing and synthesizing larger multi-domain nanostructures in a predictable manner. As an example, Gao et al used a nanoscale soldering technique to

47 weld multi-segmented nanowires together, with an emphasis on environmentally responsible “lead-free” compositions.193 The terminal segment(s) of each striped nanowire was made of Sn solder, which aided joint formation between striped nanowires when heated in a liquid under agitation. Various combinations of nanowires with multiple segments of Au, Ni, and Sn were thereby merged into complex multi-segmented networks, which could be an interesting technique for fabricating electronic devices that require robust and electrically conductive interconnects.

Figure 1-14: Top left: Schematic represention of hybrid nanoparticle coupling reactions. TEM images of a) Fe3O4-Au-Fe3O4 heterotrimers formed by heating Au-Fe3O4 heterodimers with trace amounts of sulfur and b–d) linear (Au-CdSe)n chains formed by treating Au-CdSe-Au nanorods with I2. Panel (a) reprinted from Ref. [179] with permission (Copyright 2006, American Chemical Society) and panels (b–d) reprinted from Ref. [196] with permission (Copyright 2009, Wiley- VCH).

Several examples of solid-state coupling for colloidal hybrid nanoparticles involved the solution-mediated merging and thermal sintering of solvent-exposed Au

48 domains. For example, Au-Fe3O4 heterodimers were heated in the presence of trace amounts of sulfur, which caused the Au to coalesce into a single domain, bringing two or

179 three Fe3O4 domains together [Figure 1-14(a)]. The coupling of Au-Fe3O4 heterodimers into higher-order oligomers was only observed if sulfur was added to the reaction mixture. Spontaneous adsorption of sulfur atoms was responsible for displacing the hydrophobic surface ligands from Au, which colloidally destabilized and drove the controlled aggregation of the Au domains in non-polar media. In Chapter 6, we apply this strategy to coupling linear heterotrimers of Au-Pt-Fe3O4, and obtain a variety of oligomer-type hybrid structures with as few as two and as many as six heterotrimers bridged by fused Au domains.194

Metal-semiconductor networks of alternating Au dots and CdSe nanorods were obtained by heating dumbbell-shaped Au-CdSe-Au hybrid nanoparticles in the presence

195,196 of trace molecular I2 [Figure 1-14(b-d)]. The role of I2 is to destabilize the Au surface, leading to controlled aggregation, and is therefore similar to the role of trace sulfur. The chain-like networks were dielectrophoretically assembled into devices that were designed to measure voltage-current response. Thermal annealing of the devices caused (as-made) CdSe nanorods that were randomly decorated with small Au nanoparticles to undergo a rearrangement in morphology, which was monitored in a previous study using in situ TEM.197 Au atoms and small Au clusters diffused along the lateral nanorod facets upon heating and coalesced at the Au-CdSe heterojunctions, forming much larger Au domains. In contrast to the randomly decorated Au-CdSe components, which formed kinetically during low-temperature deposition of Au, the annealed heterostructure possessed high-quality, epitaxially well-defined Au-CdSe

49 interfaces with increased interfacial area, which is important for the application of such metal-semiconductor nanostructures to electronics.

Figure 1-15: a) Schematic representation and b–d) corresponding TEM images showing the formation of CdS-PdSx-CdS hybrid particles by installing a PdSx domain on b) CdS nanoparticles to form c) CdS-PdSx dimers, followed by fusion to form d) the CdS-PdSx-CdS products. e) Schematic representation and corresponding f) TEM and g) HRTEM images of PdSx-Co9S8-PdSx hybrid particles formed from the fusion of the Co9S8 domains on PdSx-Co9S8 dimer particles. Panels (a–d) reprinted from Ref. [178] with permission (Copyright 2009, The Royal Society of Chemistry) and panels (e–g) reprinted from Ref. [179] with permission (Copyright 2007, Wiley- VCH).

Teranishi et al demonstrated that incomplete surface passivation, similar to the destabilization of Au domains by sulfur atoms or I2, leads to the solid-state coupling of hybrid nanoparticles. Seeded-growth of Co9S8-PdSx heterodimers was accomplished by

50 reacting Co(acac)2 with 1-octadecanethiol in the presence of PdSx seeds [Figure 1-15(a- d)]. At long reaction times, however, the dominant product consisted of two PdSx

198 domains bridged by Co9S8, resulting in a three-domain, peanut-shaped morphology.

Since 1-octadecanethiol behaved as a surface-stabilizing ligand, but was also consumed in the reaction with Co(acac)2, the authors ascribed the formation of PdSx-Co9S8-PdSx heterostructures to the diffusion-mediated coalescence of Co9S8 domains, driven by insufficient surface passivation. Increasing the concentration of alkanethiol in the reaction mixture suppressed coupling, and the acorn-shaped Co9S8-PdSx heterodimers were then the dominant morphological product. Similarly, rod-like CdS domains were bridged through the thermally-induced coupling of CdS-PdSx heterodimers, forming dumbbell- and flower-shaped CdS-PdSx heterostructures through the fusion of PdSx domains [Figure 1-15(e-g)].199

1.5.3 Protection-deprotection strategies

Protective groups enable site-selective modifications to be carried out on molecules that possess multiple reactive centers. Temporary shielding of targeted functional groups creates orthogonally reactive sites within a molecule, which facilitates transformation of the unprotected groups. Although protection-deprotection sequences are ubiquitous in organic synthesis, the deliberate application of such strategies to building inorganic nanostructures remains largely unexplored. One simple example was discussed in a previous section, where fabrication of Au nanodisk arrays with size- tunable gaps was achieved by selectively eliminating Ag and Ni segments that were

51 susceptible to dissolution.184 Before the wet-etching step, the striped nanowires were dispersed onto a glass substrate, and a corrosion-resistant coating of either SiO2 or Au/Ti bilayer was applied. The nanowire sidewalls that were facing up became coated, and the sidewalls that were in contact with the substrate were physically protected. This protection scheme allowed the half-exposed Ag and Ni segments to be etched away in the next step, while the corrosion-resistant layer provided structural support to the Au nanodisk array.

Oxidative etching of CTAB-functionalized Au nanorods is a convenient approach for lowering their aspect ratio, because dissolution is favored at the nanorod ends, which enables facile tuning of the longitudinal SPR absorbance.200 The site-selective nature of this transformation implies that transverse oxidation of Au nanorods should be challenging. Bao et al devised a chemical protection scheme to surmount the chemoselective barrier, by selectively depositing protective Ag2O domains onto Au nanorod tips.201 Oxidation was prevented from occurring at these sites, and instead produced dimples and significant roughening along the Au nanorod sidewalls, which are interesting features for plasmon-enhanced spectroscopy. Controllable red-shifting in the longitudinal SPR band of up to 250 nm was achieved by varying the etching time.

Chemical protection was also applied to colloidal nanocrystal transformation, to synthesize quasi-hemispheric metallic particles that feature a deep cavity, or

202 “nanobowls”. Heterodimer nanoparticles Ag-Fe3O4 were reacted with AuCl, which induced oxidation and galvanic replacement of Ag atoms with Au. The authors proposed that the key process was formation of a uniform Au shell around the Ag domain, excluding the portion of the Ag surface that was adjoined to and protected by the Fe3O4.

52

Subsequent detachment of the Ag/Au core/partial shell domain from the Fe3O4 protecting group allowed Ag ions that initially formed bonds with Fe3O4 to be ejected from the nanocrystal, along with concomitant thickening of the Au shell. This process resulted in enlargement of the interior cavity and of the particle diameter. The authors also demonstrated that the nanobowls could carry cargo by depositing small Au nanoparticles inside the cavities.

A final example of chemical protection enabled manipulation of the hierarchical

203 assembly of colloidal Pt-Fe3O4 heterodimers on the surface of WS2 nanotubes. Without the use of protective groups, Sahoo et al showed that Pt-Fe3O4 heterodimers were oriented with Pt domains attached to WS2, and the magnetic domains facing outward.

This chemoselectively-driven attachment was ascribed to favorable soft-soft acid-base interactions, based on Pearson’s hardness, between Pt and the sulfur-rich WS2 surface. A protective layer of sterically bulky, thiol-terminated ligands [O-[2-(3- mercaptopropionylamino)ethyl]-O’-methylpolyethylene glycol 5000; (SH-PEG-OCH3)] was then applied to the Pt domains, which reversed the directionality of assembly, presumably by diminishing the effect of those soft-soft interactions. After replacing the native surface ligands on Pt (oleic acid, oleylamine) with the protective SH-PEG-OCH3, attachment to WS2 occurred almost exclusively through the Fe3O4 domain.

These reports demonstrate the applicability and versatility of protection strategies to nanostructure synthesis, and many more examples are expected to emerge. In Chapter

5, we describe our own physical protection strategy for making site-selective modifications to template-grown metal nanowires. Ultimately, protection-deprotection strategies could help to facilitate the controlled placement of nanoparticles into any

53 desired, arbitrary sequence, leading to hybrid nanoparticle analogues of molecular isomers, each with distinct heterojunctions and materials linkages that impact properties such as the direction of electron transfer, the selectivity of a catalytic reaction, and the coupling of surface plasmons.

1.5.4 Stepwise construction of complex multi-component inorganic nanostructures

When the synthetic tools and concepts outlined in the preceding sections are integrated, a total synthesis framework emerges for the predictable construction of large molecules built from smaller molecular fragments. Such an approach follows naturally from the available functional groups, the large library of chemical reactions, and strategies that facilitate their implementation in a logical manner. The availability of nanoparticle “synthons”, a growing library of chemical reactions that support predictable modifications of nanoparticles, and demonstrations of orthogonal reactivity, site selectivity, protection-deprotection collectively suggest that a similar approach may be viable for constructing higher-order multi-component inorganic nanostructures. Indeed, a total synthesis framework for multi-component nanostructures has the potential to permit chemists to respond to increasingly stringent and sophisticated design criteria for functional nanomaterials. The efficient harvesting of sunlight for the renewable production of solar fuels and electricity, for example, is one of the most significant demand-driven pursuits in modern science,204 and relies critically on our ability to synthesize by-design nanostructures. Acharya et al. explain that by-design semiconductor materials must efficiently absorb solar photons and convert the energy to

54 long-lived charge separated states, but the ground state must also be regenerated at a rate that competes with oxidation and degradation of the light absorber, due to electron vacancies (holes) that remain in the valence band. To this end, the authors designed a colloidal hybrid nanostructure with the ability to efficiently localize holes in one of its domains, at potentials that promoted hole transfer to ligands in the surrounding medium.55

Figure 1-16: a–d) TEM images showing a four-step synthesis of colloidal ZnSe-CdS-Pt nanorods: a) ZnSe seed nanoparticles, b) ZnSe nanoparticles coated with a thin layer of CdS, c) ZnSe/CdS nanorods grown off the ZnSe/CdS core–shell nanoparticles, and d) ZnSe-CdS-Pt nanorods after site-selective Pt deposition. e,f) Schematic representations of multistep routes to core–shell nanoparticles with controlled metal, anion, and cation compositions. Panels (a–d) reprinted from Ref. [55] with permission (Copyright 2011, American Chemical Society) and panels (e,f) reprinted from Ref. [208] with permission (copyright 2010, American Association for the Advancement of Science).

Known seeded-growth reactions were carried out in a stepwise fashion [Figure 1-

16(a-d)] to generate a ZnSe/CdS core/rod shell nanocrystal with a metallic Pt domain at the end farthest from ZnSe. The four-step synthesis required ultrathin CdS shells to be deposited onto pre-formed ZnSe seeds, followed by unidirectional seeded-growth of the

55 rod-shaped CdS domains, and site-selective addition of Pt tips. Collectively, the band structures of the adjoined components form an energy gradient that permits transfer of photogenerated electrons in the CdS nanorod to the Pt domain, and for holes to localize in the ZnSe core. Photo-generated holes transferred to the lower-energy ZnSe domain were consequently reduced by electron-donating ligands, mercaptoundecanoic- or mercaptopropionic acid [MUA, MPA]. Localized hole transfer was evidenced by complete photoluminescence quenching in the intermediate ZnSe/CdS core/rod shell

“seeds”, and after seeded-growth of Pt, the hybrid nanostructures were able to generate

H2 fuel from the catalytic photoreduction of water. Hydrogen production was sustainable for long periods of time, provided that the solution was periodically regenerated with fresh electron-donating ligands. Notably, isomorphological hybrid nanocrystals with a

ZnTe core, where the valence band energy is roughly 0.3 eV above the highest occupied molecular orbital [HOMO] energy of MUA and MPA, did not produce H2.

Seeded-growth of an inorganic shell, specifically for luminescent quantum dots, was an innovation designed to mitigate performance losses associated with under- coordinated atoms and imperfections at the surface of a nanocrystal core.205,206 An engineered core/shell band structure effectively separates exciton wavefunctions in the core from the nanocrystal surface, leading to enhanced quantum emission yields and stability. This approach is limited by epitaxy between the core and shell material, however, because seeded-growth of more than a few atomic layers creates mismatch- induced strain, leading to grain boundaries, dislocations, and other defects. Such imperfections create trap states that are deleterious to performance, and ideally, shell materials should be thick and defect-free.207 An impressive three-step synthesis addressed

56 this demand, and generated a variety of metal/semiconductor core/shell nanocrystals with high-quality, thickness-tunable, single-crystalline shells [Figure 1-16(e)].208 First, a crystalline Ag shell was deposited over a metallic [Au, Pt, FePt, Pd] core via heterogeneous seeded-growth, and subsequently converted by oxidative diffusion to amorphous Ag2X [X = S, Se, Te].

The amorphous Ag2X domain then served as a platform for cation exchange, which generated the single-crystalline MnXy [M = Cd, Zn, Pb] shells, enabling the authors to directly deposit thick, single-crystal shells in core/shell systems with significant lattice mismatch. Importantly, the thermodynamics of cation exchange were mediated by

Pearson’s hard-soft acid base interactions at the nanocrystal surface and in the surrounding medium, which was the key to accessing single crystal shells, independent of the crystallinity or structure of the core. The authors also found reaction conditions that

2+ the allowed the amorphous Ag2S shells to react partially with Cd , which was followed by cation exchange of the remaining Ag+ with Pb2+. The resulting nanostructures were approximately 50/50 CdS/PbS two-domain shells surrounding a spherical Au core, which is an important demonstration of compositional engineering at the nanoscale [Figure 1-

17(e)]. Compositional control of the anionic component within the shell domain was also achieved by reacting the intermediate Ag shells with a predetermined ratio of sulfur and selenium complexes in the oxidative diffusion step. Amorphous shells Ag2S(1-a)Sa with a controllable ratio a were then reacted with Cd2+, resulting in single-crystal shells of a ternary CdS(1-a)Sa alloy [Figure 1-16(f)].

57 1.5.5 Purification, Separation, and Yield

Capabilities for designing and synthesizing complex multi-component nanostructures are rapidly expanding, and the conceptual links between the synthesis and properties of molecular and nanoscale systems are clear. As one constructs complex molecules using multi-step pathways, even sequential reactions that can be carried out in high yields propagate to significantly decrease the overall yield of the final product.

Consider the well-established solid-phase synthesis of polypeptides,209 which has been optimized to occur in extremely high yield. To construct a polypeptide that contains 26 amino acids, the final yield would be 77% if each individual amino acid coupling step could be carried out with a yield of 99%. If the yield of each step were decreased to

95%, then the overall yield of the 12-mer peptide would be dramatically decreased to

25%.210 This underscores two important issues: (a) the larger the number of reactions or steps that are required to synthesize a target product, the larger the number of possible by-products and the lower the yield of the desired product, and (b) separation of the target product from the by-products is absolutely critical for generating pure samples. This is true for both molecules and nanostructures.

The separation and purification of molecules is both routine and required, as is reporting of product yields. The tools to do this are also well established and readily available to most chemists. For nanoscale systems, the picture can be quite different.

Most researchers perform work-up procedures and rudimentary purification procedures that separate particles from soluble by-products, excess ligands, and unreacted precursors, as well as from other particles that differ very significantly in size. Simple

58

Figure 1-17: (a) Typical DMCR chromatogram showing the applied magnetic flux density (dashed line) and the absorbance, where peaks correspond to fractions that are eluted. Peak A and panel (b) correspond to the absorbance and TEM image, respectively, for as-synthesized Au- Fe3O4 heterodimers eluted with no applied magnetic field. Peaks B, C, and D correspond to the peaks eluted at the applied magnetic flux densities as indicated, and the TEM images of each fraction are shown in panels (c), (d), and (e), respectively. Reprinted with permission from ref. 84 (copyright 2011, Wiley-VCH).

59 precipitation and centrifugation procedures help with this, and placing a strong magnetic next to a reaction vial can help to separate particles that are magnetic at room temperature from those that are not. However, it is more difficult and less routine to carry out separations that are capable of discriminating among particles with subtle differences that include size (in a nominally uniform sample with some polydispersity), shape, or composition, as well as nuclearity and spatial arrangement for multi-component nanostructures. Validating that such purification techniques are successful is also a challenge.

Existing techniques for the separation and purification of molecules – including membrane filtration, size exclusion , solvent extraction, density gradient centrifugation, and electrophoresis – can be adapted to nanoscale systems,82 although with some important limitations. Density gradient centrifugation has proven effective for separating particles of different sizes, shapes, materials, and polymorphs.211-214 For hybrid particles that contain multiple inorganic domains, density gradient centrifugation has also permitted the separation of assemblies with different numbers of constituent particles, e.g. monomers vs. dimers vs. trimers vs. tetramers vs. higher-order oligomers.83,215

Differential magnetic catch and release (DMCR), a liquid-phase capillary chromatography technique that separates magnetic particles based on differences in their magnetic moments,80,81 has been successful at separating mixtures of colloidal hybrid nanoparticles into their constituent fractions.84 For example, DMCR was used to identify two distinct sizes of Fe3O4 nanoparticles that form along with colloidal hybrid Au-Fe3O4 particles, ultimately separating them and producing a sample with a significantly higher yield of the target Au-Fe3O4 particles (Figure 1-17). DMCR was also used to separate

60 two distinct populations of FePt-Fe3O4 heterodimers that appeared to be statistically identical by TEM, but that had different magnetic properties due to subtle variations in the Fe content of the FePt domains.84 Such separation tools will be critical both for generating pure samples of multi-component products for which many competing by- products are possible, as well as for interrogating the physical properties inherent to a single populations of particles rather than an ensemble mixture.

1.6 Research

Applications and fundamental scientific studies that require high-quality multi- functional nanostructures continue to expand, and they demand rigorous control over the placement and connectivity of nanoscale components in a manner that facilitates synergistic interactions. Such multi-component systems also require robust, powerful, and highly sophisticated synthetic tools. The total synthesis framework used to construct complex molecules is a source of inspiration for multi-component nanostructures because of the analogies between molecules and hybrid nanoparticles, the expanding capabilities for predictable nanoscale synthesis and modification, and the demonstration of nanoscale analogues of concepts that underpin the multi-step synthesis of complex molecules.

Although the synthesis of artificial atoms – high-quality single-domain nanoparticles with controllable sizes, shapes, and compositions – is becoming its own mature, branch of chemistry, an expanded library of available nanoparticle synthons needs to be developed. Established chemistry for generating high-quality nanocrystals of one material may not be applicable to others, so that each nanocrystal system often

61 requires its own synthetic troubleshooting and optimization. Germanium telluride

(GeTe) nanostructures, for example, are a demonstrated platform for studying the effects of scaling on reversible, amorphous-to-crystalline phase transitions that are important for data storage and computing applications, and for understanding ferroelectric behavior at the nanometer scale. Despite the interest in GeTe, and the well-developed colloidal approaches used to make other IV-VI tellurides, there have been very few reports of solution-grown GeTe nanostructures. A solution-phase route to rhombohedral GeTe crystallites with cube-shaped morphologies is described in Chapter 2.216 The cubic shape enables the microcrystallites to be deposited onto planar substrates to produce highly textured (002) oriented films. During TEM imaging, the GeTe particles undergo electron beam induced fragmentation and, in some cases, partial amorphization.

Previous synthetic reports for GeTe nanostructures, including the work described in Chapter 2, have suggested the presence of a strong, attractive force between crystalline

GeTe particles that is only weakly mediated by molecular surface-passivating ligands.

As a result, approaches to synthesizing GeTe have relied critically on the immediate, rapid cooling and quenching of high-temperature (200-250 °C) reactions, to avoid agglomeration of nanocrystallites into irregularly-shaped, insoluble precipitates. Chapter

3 describes the preparation of colloidal GeTe nanostructures in the presence of surface- stabilizing polymers, which mediate particle-particle interactions and prevent aggregation of GeTe crystallites more effectively than conventional molecular stabilizers.217 As a result, several novel GeTe nanostructures are formed, including faceted octahedral nanoparticles, amorphous GexTe1-x alloy nanospheres and single-crystal 2D GeTe nanosheets. The colloidal stability conferred by the polymer may provide the key

62 experimental degree of freedom necessary to achieve higher-order morphology control for GeTe and related materials.

When the synthetic chemistry of a single-component nanoparticle system is anomalous, or significantly different than its congeners, there is fertile ground for reaction development, mechanistic study, and innovation. Despite the standard colloidal approaches for preparing many other high-quality IV-VI semiconductor (GeS, GeSe,

SnS, SnSe, SnTe, PbS, PbSe, PbTe) nanostructures, for example, GeTe behaves differently in a variety of respects. The colloidal synthesis of binary cobalt(II) and nickel(II) oxide nanostructures via thermolysis of fatty carboxylate salts is another such anomaly. Fatty acid salts of 3d transition metals are known to thermolyze in high-boiling

(typically, 300-330 °C) organic solvents to generate colloidal nanocrystals in high yield.

Gram-scale quantities of uniformly-sized Mn(II), Fe(II,III), Fe(III), and Zn(II) oxide nanocrystals can be produced in a facile, one-step, reaction by thermolysis of their cis-9- octadecanoate (oleate) salts in non-coordinating solvents. The same transformations, surprisingly, do not occur readily for Co(II) or Ni(II) oleates. Reports on this subject are contradictory, and synthetic protocols are often irreproducible. In Chapter 4, an in-depth investigation pertaining to the thermal decomposition of Co(II) oleate is described.

Several conditions wherein Co(II) oleate thermolysis occurs straightforwardly are identified, that can be related to the presence of impurities, the in situ generation of side- products, and the practical handling of atmosphere. These insights have enabled the shape-controlled synthesis of colloidal Co(II) oxide nanocrystals that feature Co2+ in a metastable, tetrahedral geometry.

63 Nanoparticles can undergo facile chemical transformations into derivative materials with morphological retention, and this is leading to a rapidly expanding reaction library for nanoparticles that has analogies to the large and diverse library of reactions available to organic chemists. Concepts that underpin the construction of large molecules – chemoselective and regiospecific reactions, orthogonal reactivity, substituent effects, protection-deprotection strategies – are starting to be demonstrated for multi- component nanostructures, adding new predictive synthesis capabilities and powerful guidelines for implementing multi-step routes to complex nanostructures. Metal nanowire synthons, for example, can be fabricated straightforwardly in rigid, porous templates using electrodeposition. In Chapter 5, an alternative and simple solution chemistry strategy for incorporating multi-metal components of controllable length and composition into template-grown metal nanowires is described, combined with a straightforward application of physical protection.218 Template-confined metal nanowires react with metal salt solutions under reducing conditions, which ensures that site-specific transformation is initiated at the nanowire tips. The nanowire sidewalls are protected from the reaction mixture by the template. When this approach is leveraged, either one or both tips can be converted into multi-metal phases.

As mentioned in section 1.1, colloidal hybrid nanoparticles contain multiple nanoscale domains fused together by solid-state interfaces, and they represent an emerging class of multi-functional lab-on-a-particle architectures that underpin future advances in solar energy conversion, fuel cell catalysis, medical imaging and therapy, and electronics. The complexity of these “artificial molecules” is ultimately limited by the lack of a mechanism-driven design framework. In Chapter 6, we show that known

64 chemical reactions can be applied in a predictable and stepwise manner to build complex hybrid nanoparticle architectures that include M-Pt-Fe3O4 (M = Au, Ag, Ni, Pd) heterotrimers, MxS-Au-Pt-Fe3O4 (M = Pb, Cu) heterotetramers, and higher-order

194 oligomers based on the heterotrimeric Au-Pt-Fe3O4 building block. The reaction toolkit applies solid-state nanoparticle analogues of chemoselective reactions, regiospecificity, coupling reactions, and molecular substituent effects to the construction of exceptionally complex hybrid nanoparticle oligomers.

When the working principles of molecular synthesis and the practical capabilities of nanochemistry are integrated, a total synthesis framework for hybrid inorganic nanostructures emerges. Complex multi-domain nanostructures with precisely defined linkages can be constructed in a stepwise manner by sequential application of the available nanoparticle reactions and transformations.

1.7 References

1) N. Tian, Z. Y. Zhou, S. G. Sun, Y. Ding, Z. L. Wang, Science 2007, 316, 732.

2) T. Yu, D. Y. Kim, H. Zhang, Y. N. Xia, Angew. Chem. Int. Ed. 2011, 50, 2773.

3) X. Q. Huang, Z. P. Zhao, J. M. Fan, Y. M. Tan, N. F. Zheng, J. Am. Chem. Soc.

2011, 133, 4718.

4) M. S. Jin, H. Zhang, Z. X. Xie, Y. N. Xia, Angew. Chem. Int. Ed. 2011, 50, 7850.

5) N. Tian, Z. Y. Zhou, N. F. Yu, L. Y. Wang, S. G. Sun, J. Am. Chem. Soc. 2010,

132, 7580.

65 6) J. A. Zhang, M. R. Langille, M. L. Personick, K. Zhang, S. Y. Li, C. A. Mirkin, J.

Am. Chem. Soc. 2010, 132, 14012.

7) L. Zhang, J. W. Zhang, Q. Kuang, S. F. Xie, Z. Y. Jiang, Z. X. Xie, L. S. Zheng,

J. Am. Chem. Soc. 2011, 133, 17114.

8) Y. J. Deng, N. Tian, Z. Y. Zhou, R. Huang, Z. L. Liu, J. Xiao, S. G. Sun, Chem.

Sci. 2012, 3, 1157.

9) X. G. Han, M. S. Jin, S. F. Xie, Q. Kuang, Z. Y. Jiang, Y. Q. Jiang, Z. X. Xie, L.

S. Zheng, Angew. Chem. Int. Ed. 2009, 48, 9180.

10) H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith, H. M. Cheng, G.

Q. Lu, Nature 2008, 453, 638.

11) H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, J. Zou, H. M.

Cheng, G. Q. Lu, J. Am. Chem. Soc. 2009, 131, 4078.

12) W. C. Huang, L. M. Lyu, Y. C. Yang, M. H. Huang, J. Am. Chem. Soc. 2012,

134, 1261.

13) B. Lang, R. W. Joyner, G. A. Somorjai, Surf. Sci. 1972, 30, 454.

14) M. A. Vanhove, G. A. Somorjai, Surf. Sci. 1980, 92, 489-518.

15) Z. Quan, Y. Wang, J. Fang, Acc. Chem. Res 2013, 46, 191.

16) M. H. Huang, P. H. Lin, Adv. Funct. Mater. 2012, 22, 14.

17) Z. Y. Zhou, N. Tian, J. T. Li, I. Broadwell, S. G. Sun, Chem. Soc. Rev. 2011, 40,

4167.

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

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

19) D. Ho, X. L. Sun, S. H. Sun, Acc. Chem. Res. 2011, 44, 875.

66 20) J. Xie, G. Liu, H. S. Eden, H. Ai, X. Y. Chen, Acc. Chem. Res. 2011, 44, 883.

21) F. X. Redl, K. S. Cho, C. B. Murray, S. O'Brien, Nature 2003, 423, 968.

22) Z. Chen, S. O'Brien, ACS Nano 2008, 2, 1219.

23) M. I. Bodnarchuk, M. V. Kovalenko, W. Heiss, D. V. Talapin, J. Am. Chem. Soc.

2010, 132, 11967.

24) J. Henzie, M. Grunwald, A. Widmer-Cooper, P. L. Geissler, P. D. Yang, Nat.

Mater. 2012, 11, 131.

25) D. V. Talapin, J. S. Lee, M. V. Kovalenko, E. V. Shevchenko, Chem. Rev. 2010,

110, 389.

26) L. Carbone, P. D. Cozzoli, Nano Today 2010, 5, 449.

27) R. Costi, A. E. Saunders, U. Banin, Angew. Chem. Int. Ed. 2010, 49, 4878.

28) P. D. Cozzoli, T. Pellegrino, L. Manna, Chem. Soc. Rev. 2006, 35, 1195.

29) C. Wang, C. J. Xu, H. Zeng, S. H. Sun, Adv. Mater. 2009, 21, 3045.

30) M. Casavola, R. Buonsanti, G. Caputo, P. D. Cozzoli, Eur. J. Inorg. Chem. 2008,

837.

31) A. J. Mieszawska, R. Jalilian, G. U. Sumanasekera, F. P. Zamborini, Small 2007,

3, 722.

32) J. Q. Hu, Y. Bando, D. Golberg, J. Mater. Chem. 2009, 19, 330.

33) X. M. Feng, G. Q. Hu, J. Q. Hu, Nanoscale 2011, 3, 2099.

34) R. G. Chaudhuri, S. Paria, Chem. Rev. 2012, 112, 2373.

35) R. Agarwal, Small 2008, 4, 1872.

36) M. Haruta, S. Tsubota, T. Kobayashi, H. Kageyama, M. J. Genet, B. Delmon, J.

Catal. 1993, 144, 175.

67 37) M. Haruta, Catal. Today 1997, 36, 153.

38) M. G. Walter, E. L. Warren, J. R. McKone, S. W. Boettcher, Q. X. Mi, E. A.

Santori, N. S. Lewis, Chem. Rev. 2010, 110, 6446.

39) I. Mora-Sero, J. Bisquert, J. Phys. Chem. Lett. 2010, 1, 3046.

40) S. Y. Reece, J. A. Hamel, K. Sung, T. D. Jarvi, A. J. Esswein, J. J. H. Pijpers, D.

G. Nocera, Science 2011, 334, 645.

41) K. Maeda, K. Domen, J. Phys. Chem. Lett. 2010, 1, 2655.

42) P. V. Kamat, J. Phys. Chem. Lett. 2012, 3, 663.

43) G. I. Koleilat, X. H. Wang, E. H. Sargent, Nano Lett. 2012, 12, 3043.

44) J. B. Gao, C. L. Perkins, J. M. Luther, M. C. Hanna, H. Y. Chen, O. E. Semonin,

A. J. Nozik, R. J. Ellingson, M. C. Beard, Nano Lett. 2011, 11, 3263.

45) B. H. Meekins, P. V. Kamat, J. Phys. Chem. Lett. 2011, 2, 2304.

46) C. L. Choi, A. P. Alivisatos, Annu. Rev. Phys. Chem. 2010, 61, 369.

47) E. Duguet, A. Desert, A. Perro, S. Ravaine, Chem. Soc. Rev. 2011, 40, 941.

48) R. Costi, A. E. Saunders, E. Elmalem, A. Salant, U. Banin, Nano Lett. 2008, 8,

637.

49) R. Costi, G. Cohen, A. Salant, E. Rabani, U. Banin, Nano Lett. 2009, 9, 2031.

50) S. Kumar, M. Jones, S. S. Lo, G. D. Scholes, Small 2007, 3, 1633.

51) S. Kim, B. Fisher, H. J. Eisler, M. Bawendi, J. Am. Chem. Soc. 2003, 125, 11466.

52) C. Galland, Y. Ghosh, A. Steinbruck, M. Sykora, J. A. Hollingsworth, V. I.

Klimov, H. Htoon, Nature 2011, 479, 203.

53) C. Galland, Y. Ghosh, A. Steinbruck, J. A. Hollingsworth, H. Htoon, V. I.

Klimov, Nat. Commun. 2012, 3:908 doi: 10.1038/ncomms1916.

68 54) L. Amirav, A. P. Alivisatos, J. Phys. Chem. Lett. 2010, 1, 1051.

55) K. P. Acharya, R. S. Khnayzer, T. O'Connor, G. Diederich, M. Kirsanova, A.

Klinkova, D. Roth, E. Kinder, M. Imboden, M. Zamkov, Nano Lett. 2011, 11,

2919.

56) P. Li, Z. Wei, T. Wu, Q. Peng, Y. D. Li, J. Am. Chem. Soc. 2011, 133, 5660.

57) H. F. Yin, C. Wang, H. G. Zhu, S. H. Overbury, S. H. Sun, S. Dai, Chem.

Commun. 2008, 4357.

58) C. Wang, H. Daimon, S. H. Sun, Nano Lett. 2009, 9, 1493.

59) C. Wang, H. F. Yin, S. Dai, S. H. Sun, Chem. Mater. 2010, 22, 3277.

60) A. Figuerola, A. Fiore, R. Di Corato, A. Falqui, C. Giannini, E. Micotti, A.

Lascialfari, M. Corti, R. Cingolani, T. Pellegrino, P. D. Cozzoli, L. Manna, J. Am.

Chem. Soc. 2008, 130, 1477.

61) M. I. Bodnarchuk, M. V. Kovalenko, H. Groiss, R. Resel, M. Reissner, G. Hesser,

R. T. Lechner, W. Steiner, F. Schaffler, W. Heiss, Small 2009, 5, 2247.

62) S. Deka, A. Falqui, G. Bertoni, C. Sangregorio, G. Poneti, G. Morello, M. De

Giorgi, C. Giannini, R. Cingolani, L. Manna, P. D. Cozzoli, J. Am. Chem. Soc.

2009, 131, 12817.

63) C. D. Keating, M. J. Natan, Adv. Mater. 2003, 15, 451.

64) T. R. Kline, M. L. Tian, J. G. Wang, A. Sen, M. W. H. Chan, T. E. Mallouk,

Inorg. Chem. 2006, 45, 7555.

65) W. F. Paxton, K. C. Kistler, C. C. Olmeda, A. Sen, S. K. St Angelo, Y. Y. Cao, T.

E. Mallouk, P. E. Lammert, V. H. Crespi, J. Am. Chem. Soc. 2004, 126, 13424.

69 66) T. R. Kline, W. F. Paxton, T. E. Mallouk, A. Sen, Angew. Chem. Int. Ed. 2005,

44, 744.

67) W. Wang, L. A. Castro, M. Hoyos, T. E. Mallouk, ACS Nano 2012, 6, 6122.

68) A. K. Salem, J. Chao, K. W. Leong, P. C. Searson, Adv. Mater. 2004, 16, 268.

69) J. C. Love, A. R. Urbach, M. G. Prentiss, G. M. Whitesides, J. Am. Chem. Soc.

2003, 125, 12696.

70) B. D. Smith, D. J. Kirby, C. D. Keating, Small 2011, 7, 781.

71) D. J. Pena, J. K. N. Mbindyo, A. J. Carado, T. E. Mallouk, C. D. Keating, B.

Razavi, T. S. Mayer, J. Phys. Chem. B. 2002, 106, 7458.

72) S. R. Nicewarner-Pena, R. G. Freeman, B. D. Reiss, L. He, D. J. Pena, I. D.

Walton, R. Cromer, C. D. Keating, M. J. Natan, Science 2001, 294, 137.

73) S. E. Brunker, K. B. Cederquist, C. D. Keating, Nanomedicine-UK 2007, 2, 695.

74) R. L. Stoermer, K. B. Cederquist, S. K. McFarland, M. Y. Sha, S. G. Penn, C. D.

Keating, J. Am. Chem. Soc. 2006, 128, 16892.

75) F. Liu, J. Y. Lee, W. J. Zhou, J. Phys. Chem. B. 2004, 108, 17959.

76) F. Liu, J. Y. Lee, W. J. Zhou, Adv. Funct. Mater. 2005, 15, 1459.

77) F. Liu, J. Y. Lee, W. J. Zhou, Small 2006, 2, 121.

78) M. Casavola, V. Grillo, E. Carlino, C. Giannini, F. Gozzo, E. F. Pinel, M. A.

Garcia, L. Manna, R. Cingolani, P. D. Cozzoli, Nano Lett. 2007, 7, 1386.

79) R. Buonsanti, V. Grillo, E. Carlino, C. Giannini, F. Gozzo, M. Garcia-Hernandez,

M. A. Garcia, R. Cingolani, P. D. Cozzoli, J. Am. Chem. Soc. 2010, 132, 2437.

80) A. H. Latham, R. S. Freitas, P. Schiffer, M. E. Williams, Anal. Chem. 2005, 77,

5055.

70 81) J. S. Beveridge, J. R. Stephens, A. H. Latham, M. E. Williams, Anal. Chem. 2009,

81, 9618.

82) B. Kowalczyk, I. Lagzi, B. A. Grzybowski, Curr. Opin. Colloid. In. 2011, 16,

135-148.

83) G. Chen, Y. Wang, L. H. Tan, M. X. Yang, L. S. Tan, Y. Chen, H. Y. Chen, J.

Am. Chem. Soc. 2009, 131, 4218.

84) J. S. Beveridge, M. R. Buck, J. F. Bondi, R. Misra, P. Schiffer, R. E. Schaak, M.

E. Williams, Angew. Chem. Int. Ed. 2011, 50, 9875.

85) Buffat, Ph.; Borel, J-P. Phys. Rev. A. 1976, 13, 2287.

86) Alivisatos, A.P. J. Phys. Chem. 1996, 100, 13226.

87) Klimov, V.I. (editor) Nanocrystal Quantum Dots, Second Edition. Taylor and

Francis Group LLC: Boca Raton, 2010.

88) Eustis, S.; El-Sayed, M.A. Chem. Soc. Rev. 2006, 35, 209.

89) LaMer, V.K.; Dinegar, R.H. J. Am. Chem. Soc. 1950, 72, 4847.

90) Sugimoto, T. Monodispersed Particles, Elsevier: Amsterdam, 2001.

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

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

4630.

93) Kwon, S.G.; Piao, Y.; Park, J.; Angappane, S.; Jo, Y.; Hwang, N-M.; Park, J-G.;

Hyeon, T. J. Am. Chem. Soc. 2007, 129, 12571.

94) Green, M. J. Mater. Chem. 2010, 20, 5797.

95) Yin, Y.; Alivisatos, A.P. Nature 2005, 437, 664.

96) Peng, X. Adv. Mater. 2003, 15, 459.

71 97) Yu, W.W.; Wang, Y.A.; Peng, X. Chem. Mater. 2003, 15, 4300.

98) Chen, Y.; Johnson, E.; Peng, X. J. Am. Chem. Soc. 2007, 129, 10937.

99) Xie, R.; Li, Z.; Peng, X.; J. Am. Chem. Soc. 2009, 131, 15457.

100) Sowers, K.L.; Swartz, B.; Krauss, T.D.; Chem. Mater. 2013 dx.doi.org / 10.1021 /

cm400005cl.

101) Wang, Z. L. J. Phys. Chem. B 2000, 104, 1153.

102) Germain, V.; Li, J.; Ingert, D.; Wang, Z.L.; Pileni, M.P. J. Phys. Chem. B, 2003,

107, 8717.

103) Elechiguerra, J.L.; Reyes-Gasgab, J.; Yacaman, M.J. J. Mater. Chem. 2006, 16,

3906.

104) Aherne, D.; Ledwith, D.M.; Gara, M.; Kelly, J.M. Adv. Funct. Mater. 2008, 18,

2005.

105) K. C. Nicolaou, D. Vourloumis, N. Winssinger, P. S. Baran, Angew. Chem. Int.

Ed. 2000, 39, 44.

106) W. A. Smit, A. F. Bochkov, R. Caple, Royal Society of Chemistry (Great

Britain), Organic synthesis: the science behind the art, Royal Society of

Chemistry, Cambridge, UK, 1998.

107) K. C. Nicolaou, C. R. H. Hale, C. Nilewski, H. A. Ioannidou, Chem. Soc. Rev.

2012, 41, 5185.

108) E. J. Corey, X.-M. Cheng, The logic of chemical synthesis, J. Wiley: New York,

1989.

109) E. J. Corey, Chem. Soc. Rev. 1988, 17, 111.

110) S. D. Roughly, A. M. Jordan, Med. Chem. 2011, 54, 3451.

72 111) B. M. Trost, Science 1983, 219, 245.

112) R. A. Shenvi, D. P. O'Malley, P. S. Baran, Acc. Chem. Res. 2009, 42, 530.

113) N. A. Afagh, A. K. Yudin, Angew. Chem. Int. Ed. 2010, 49, 262.

114) P. G. M. Wuts, T. W. Greene, T. W. Greene, Greene's protective groups in

organic synthesis, 4th ed., Wiley-Interscience: Hoboken, N.J., 2007.

115) B. M. Trost, Science 1991, 254, 1471.

116) T. Newhouse, P. S. Baran, R. W. Hoffmann, Chem. Soc. Rev. 2009, 38, 3010.

117) T. Gaich, P. S. Baran, J. Org. Chem. 2010, 75, 4657.

118) D. V. Talapin, Y. D. Yin, J. Mater. Chem. 2011, 21, 11454.

119) Y. Vasquez, A. E. Henkes, J. C. Bauer, R. E. Schaak, J. Solid State Chem. 2008,

181, 1509.

120) G. D. Moon, S. Ko, Y. Min, J. Zeng, Y. N. Xia, U. Jeong, Nano Today 2011, 6,

186.

121) Y. D. Yin, R. M. Rioux, C. K. Erdonmez, S. Hughes, G. A. Somorjai, A. P.

Alivisatos, Science 2004, 304, 711.

122) Y. D. Yin, C. K. Erdonmez, A. Cabot, S. Hughes, A. P. Alivisatos, Adv. Funct.

Mater. 2006, 16, 1389.

123) A. E. Henkes, R. E. Schaak, Chem. Mater. 2007, 19, 4234.

124) H. T. Zhang, D. H. Ha, R. Hovden, L. F. Kourkoutis, R. D. Robinson, Nano Lett.

2011, 11, 188.

125) R. E. Schaak, A. K. Sra, B. M. Leonard, R. E. Cable, J. C. Bauer, Y. F. Han, J.

Means, W. Teizer, Y. Vasquez, E. S. Funck, J. Am. Chem. Soc. 2005, 127, 3506.

73 126) J. C. Bauer, X. Chen, Q. S. Liu, T. H. Phan, R. E. Schaak, J. Mater. Chem. 2008,

18, 275.

127) B. M. Leonard, M. E. Anderson, K. D. Oyler, T. H. Phan, R. E. Schaak, ACS

Nano 2009, 3, 940.

128) F. Pinna, Catal. Today 1998, 41, 129.

129) I. T. Sines, R. E. Schaak, J. Am. Chem. Soc. 2011, 133, 1294.

130) D. H. Son, S. M. Hughes, Y. D. Yin, A. P. Alivisatos, Science 2004, 306, 1009.

131) P. H. C. Camargo, Y. H. Lee, U. Jeong, Z. Q. Zou, Y. N. Xia, Langmuir 2007, 23,

2985.

132) R. D. Robinson, B. Sadtler, D. O. Demchenko, C. K. Erdonmez, L. W. Wang, A.

P. Alivisatos, Science 2007, 317, 355.

133) S. E. Wark, C. H. Hsia, D. H. Son, J. Am. Chem. Soc. 2008, 130, 9550.

134) J. M. Luther, H. M. Zheng, B. Sadtler, A. P. Alivisatos, J. Am. Chem. Soc. 2009,

131, 16851.

135) B. Sadtler, D. O. Demchenko, H. Zheng, S. M. Hughes, M. G. Merkle, U.

Dahmen, L. W. Wang, A. P. Alivisatos, J. Am. Chem. Soc. 2009, 131, 5285.

136) P. K. Jain, L. Amirav, S. Aloni, A. P. Alivisatos, J. Am. Chem. Soc. 2010, 132,

9997.

137) H. B. Li, M. Zanella, A. Genovese, M. Povia, A. Falqui, C. Giannini, L. Manna,

Nano Lett. 2011, 11, 4964.

138) M. Casavola, M. A. van Huis, S. Bals, K. Lambert, Z. Hens, D. Vanmaekelbergh,

Chem. Mater. 2012, 24, 294.

74 139) H. Li, R. Brescia, R. Krahne, G. Bertoni, M. J. P. Alcocer, C. D'Andrea, F.

Scotognella, F. Tassone, M. Zanella, M. De Giorgi, L. Manna, ACS Nano 2012, 6,

1637.

140) F. Dawood, R. E. Schaak, J. Am. Chem. Soc. 2009, 131, 424.

141) J. Park, H. Zheng, Y. W. Jun, A. P. Alivisatos, J. Am. Chem. Soc. 2009, 131,

13943.

142) M. Saruyama, Y. G. So, K. Kimoto, S. Taguchi, Y. Kanemitsu, T. Teranishi, J.

Am. Chem. Soc. 2011, 133, 17598.

143) Y. H. Wei, R. Klajn, A. O. Pinchuk, B. A. Grzybowski, Small 2008, 4, 1635.

144) H. M. Wu, O. Chen, J. Q. Zhuang, J. Lynch, D. LaMontagne, Y. Nagaoka, Y. C.

Cao, J. Am. Chem. Soc. 2011, 133, 14327.

145) G. Krylova, L. J. Giovanetti, F. G. Requejo, N. M. Dimitrijevic, A. Prakapenka,

E. V. Shevchenko, J. Am. Chem. Soc. 2012, 134, 4384.

146) Y. N. Xia, Y. J. Xiong, B. Lim, S. E. Skrabalak, Angew. Chem. Int. Ed. 2009, 48,

60.

147) C. M. Cobley, M. Rycenga, F. Zhou, Z. Y. Li, Y. N. Xia, Angew. Chem. Int. Ed.

2009, 48, 4824.

148) M. J. Mulvihill, X. Y. Ling, J. Henzie, P. D. Yang, J. Am. Chem. Soc. 2010, 132,

268.

149) B. Lim, Y. N. Xia, Angew. Chem. Int. Ed. 2011, 50, 76.

150) Y. G. Sun, Y. N. Xia, J. Am. Chem. Soc. 2004, 126, 3892.

151) X. M. Lu, H. Y. Tuan, J. Y. Chen, Z. Y. Li, B. A. Korgel, Y. Xia, J. Am. Chem.

Soc. 2007, 129, 1733.

75 152) S. E. Skrabalak, L. Au, X. D. Li, Y. N. Xia, Nat. Protoc. 2007, 2, 2182.

153) H. Zhang, M. S. Jin, J. G. Wang, W. Y. Li, P. H. C. Camargo, M. J. Kim, D. R.

Yang, Z. X. Xie, Y. Xia, J. Am. Chem. Soc. 2011, 133, 6078.

154) N. J. Halas, S. Lal, W. S. Chang, S. Link, P. Nordlander, Chem. Rev. 2011, 111,

3913.

155) Neiderberger, M.; Colfen, H. Phys. Chem. Chem. Phys. 2006, 8, 3271.

156) D. Seo, C. Il Yoo, J. Jung, H. Song, J. Am. Chem. Soc. 2008, 130, 2940.

157) E. V. Shevchenko, M. I. Bodnarchuk, M. V. Kovalenko, D. V. Talapin, R. K.

Smith, S. Aloni, W. Heiss, A. P. Alivisatos, Adv. Mater. 2008, 20, 4323.

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

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

159) J. H. Gao, G. L. Liang, B. Zhang, Y. Kuang, X. X. Zhang, B. Xu, J. Am. Chem.

Soc. 2007, 129, 1428.

160) G. Q. Zhang, H. Y. Fang, H. R. Yang, L. A. Jauregui, Y. P. Chen, Y. Wu, Nano

Lett. 2012, 12, 3627.

161) L. Manna, E. C. Scher, A. P. Alivisatos, J. Am. Chem. Soc. 2000, 122, 12700.

162) Z. A. Peng, X. G. Peng, J. Am. Chem. Soc. 2001, 123, 1389.

163) T. Mokari, E. Rothenberg, I. Popov, R. Costi, U. Banin, Science 2004, 304, 1787.

164) S. Kudera, L. Carbone, M. F. Casula, R. Cingolani, A. Falqui, E. Snoeck, W. J.

Parak, L. Manna, Nano Lett. 2005, 5, 445.

165) Carbone, L. et al., Nano Lett. 2007, 7, 2942.

166) G. Dukovic, M. G. Merkle, J. H. Nelson, S. M. Hughes, A. P. Alivisatos, Adv.

Mater. 2008, 20, 4306.

76 167) G. Menagen, D. Mocatta, A. Salant, I. Popov, D. Dorfs, U. Banin, Chem. Mater.

2008, 20, 6900.

168) L. Carbone, A. Jakab, Y. Khalavka, C. Sonnichsen, Nano Lett. 2009, 9, 3710.

169) X. H. Li, J. Lian, M. Lin, Y. T. Chan, J. Am. Chem. Soc. 2011, 133, 672.

170) S. Chakrabortty, J. A. Yang, Y. M. Tan, N. Mishra, Y. T. Chan, Angew. Chem.

Int. Ed. 2010, 49, 2888.

171) X. Chen, Y. Lou, A. C. Samia, C. Burda, Nano Lett. 2003, 3, 799.

172) A. Aharoni, T. Mokari, I. Popov, U. Banin, J. Am. Chem. Soc. 2006, 128, 257.

173) L. Manna, E. C. Scher, L. S. Li, A. P. Alivisatos, J. Am. Chem. Soc. 2002, 124,

7136.

174) D. V. Talapin, I. Mekis, S. Gotzinger, A. Kornowski, O. Benson, H. Weller, J.

Phys. Chem. B. 2004, 108, 18826.

175) S. Deka, A. Quarta, M. G. Lupo, A. Falqui, S. Boninelli, C. Giannini, G. Morello,

M. De Giorgi, G. Lanzani, C. Spinella, R. Cingolani, T. Pellegrino, L. Manna, J.

Am. Chem. Soc. 2009, 131, 2948.

176) A. Eychmuller, A. Mews, H. Weller, Chem. Phys. Lett. 1993, 208, 59.

177) D. Schooss, A. Mews, A. Eychmuller, H. Weller, Phys. Rev. B 1994, 49, 17072.

178) Z. C. Xu, Y. L. Hou, S. H. Sun, J. Am. Chem. Soc. 2007, 129, 8698.

179) W. L. Shi, H. Zeng, Y. Sahoo, T. Y. Ohulchanskyy, Y. Ding, Z. L. Wang, M.

Swihart, P. N. Prasad, Nano Lett. 2006, 6, 875.

180) W. L. Shi, Y. Sahoo, H. Zeng, Y. Ding, M. T. Swihart, P. N. Prasad, Adv. Mater.

2006, 18, 1889.

181) T. Chen, G. Chen, S. X. Xing, T. Wu, H. Y. Chen, Chem. Mater. 2010, 22, 3826.

77 182) T. Ung, L. M. Liz-Marzan, P. Mulvaney, Langmuir 1998, 14, 3740.

183) H. T. Zhang, J. Ding, G. M. Chow, Z. L. Dong, Langmuir 2008, 24, 13197.

184) L. D. Qin, S. Park, L. Huang, C. A. Mirkin, Science 2005, 309, 113.

185) X. Chen, Y. M. Jeon, J. W. Jang, L. Qin, F. Huo, W. Wei, C. A. Mirkin, J. Am.

Chem. Soc. 2008, 130, 8166.

186) L. D. Qin, S. L. Zou, C. Xue, A. Atkinson, G. C. Schatz, C. A. Mirkin, Proc. Natl.

Acad. Sci. USA 2006, 103, 13300.

187) L. Qin, M. J. Banholzer, J. E. Millstone, C. A. Mirkin, Nano Lett. 2007, 7, 3849-

3853.

188) M. McEachran, D. Keogh, B. Pietrobon, N. Cathcart, I. Gourevich, N. Coombs,

V. Kitaev, J. Am. Chem. Soc. 2011, 133, 8066.

189) Lu, X.; Au, L.; McLellan, J.; Li, Z-Y.; Marquez, M.; Xia, Y. Nano Lett. 2007, 7,

1764.

190) X. Guo, Q. Zhang, Y. H. Sun, Q. Zhao, J. Yang, ACS Nano 2012, 6, 1165.

191) Y. M. Lee, M. A. Garcia, N. A. F. Huls, S. H. Sun, Angew. Chem. Int. Ed. 2010,

49, 1271.

192) C. George, D. Dorfs, G. Bertoni, A. Falqui, A. Genovese, T. Pellegrino, A. Roig,

A. Quarta, R. Comparelli, M. L. Curri, R. Cingolani, L. Manna, J. Am. Chem.

Soc. 2011, 133, 2205.

193) F. Gao, S. Mukherjee, Q. Z. Cui, Z. Y. Gu, J. Phys. Chem. C. 2009, 113, 9546.

194) Buck, M.R.; Bondi, J.F.; Schaak, R.E. Nature Chem. 2012, 4, 37.

195) R. Lavieville, Y. Zhang, A. Casu, A. Genovese, L. Manna, E. Di Fabrizio, R.

Krahne, ACS Nano 2012, 6, 2940.

78 196) A. Figuerola, I. R. Franchini, A. Fiore, R. Mastria, A. Falqui, G. Bertoni, S. Bals,

G. Van Tendeloo, S. Kudera, R. Cingolani, L. Manna, Adv. Mater. 2009, 21, 550.

197) A. Figuerola, M. van Huis, M. Zanella, A. Genovese, S. Marras, A. Falqui, H. W.

Zandbergen, R. Cingolani, L. Manna, Nano Lett. 2010, 10, 3028.

198) T. Teranishi, M. Saruyama, M. Nakaya, M. Kanehara, Angew. Chem. Int. Ed.

2007, 46, 1713.

199) M. Saruyama, M. Kanehara, T. Teranishi, Chem. Commun. 2009, 2724.

200) N.R. Jana, L. Gearheart, S. O. Obare, C. J. Murphy, Langmuir 2002, 18, 922.

201) Z. H. Bao, Z. H. Sun, M. D. Xiao, H. J. Chen, L. W. Tian, J. F. Wang, J. Mater.

Chem. 2011, 21, 11537.

202) Y. Ridelman, G. Singh, R. Popovitz-Biro, S. G. Wolf, S. Das, R. Klajn, Small

2012, 8, 654.

203) J. K. Sahoo, M. N. Tahir, F. Hoshyargar, B. Nakhjavan, R. Branscheid, U. Kolb,

W. Tremel, Angew. Chem. Int. Ed. 2011, 50, 12271.

204) N. S. Lewis, D. G. Nocera, Proc. Natl. Acad. Sci. USA 2007, 104, 20142.

205) M. A. Hines, P. Guyot-Sionnest, J. Phys. Chem.-US 1996, 100, 468.

206) X. G. Peng, M. C. Schlamp, A. V. Kadavanich, A. P. Alivisatos, J. Am. Chem.

Soc. 1997, 119, 7019.

207) Y. Chen, J. Vela, H. Htoon, J. L. Casson, D. J. Werder, D. A. Bussian, V. I.

Klimov, J. A. Hollingsworth, J. Am. Chem. Soc. 2008, 130, 5026.

208) J. T. Zhang, Y. Tang, K. Lee, O. Y. Min, Science 2010, 327, 1634.

209) R. B. Merrifield, J. Am. Chem. Soc. 1963, 85, 2149.

79 210) A. A. Tseng, Nanofabrication: Fundamentals and Applications, World Scientific:

Singapore 2008.

211) D. Steinigeweg, M. Schutz, M. Salehi, S. Schlucker, Small 2011, 7, 2443.

212) L. Bai, X. J. Ma, J. F. Liu, X. M. Sun, D. Y. Zhao, D. G. Evans, J. Am. Chem.

Soc. 2010, 132, 2333.

213) P. H. Qiu, C. B. Mao, Adv. Mater. 2011, 23, 4880.

214) X. J. Ma, Y. Kuang, L. Bai, Z. Chang, F. Wang, X. M. Sun, D. G. Evans, ACS

Nano 2011, 5, 3242.

215) X. Xu, S. Stottinger, G. Battagliarin, G. Hinze, E. Mugnaioli, C. Li, K. Mullen, T.

Basche, J. Am. Chem. Soc. 2011, 133, 18062.

216) Buck, M.R.; Sines, I.T.; Schaak, R.E. Chem. Mater. 2010, 22, 3236.

217) Buck, M.R.; Biacchi, A.J.; Popczun, E.J.; Schaak, R.E. Chem. Mater. 2013 DOI:

10.1021/cm4009656.

218) Anderson, M.E.; Buck, M.R.; Sines, I.T.; Oyler, K.D.; Schaak, R.E. J. Am. Chem.

Soc. 2008, 130, 14042.

Chapter 2

Liquid-Phase Synthesis of Uniform Cube-Shaped GeTe Microcrystals

2.1 Introduction

Chalcogenide alloys exhibiting a thermally-induced, reversible, amorphous-to- crystalline phase change are important for a wide range of technologies.1,2 The large difference in reflectivity between the amorphous and crystalline states enables the operation of common optical data storage media, such as CDs and DVDs.3-5 The amorphous-to-crystalline transition is also accompanied by a significant decrease in resistivity, which is the basis of emerging and next-generation technologies such as phase change random access memory (PCRAM),6,7 phase change-based logic,8 and cognitive computing.9 Widespread use of phase change technologies that operate via differential resistivity, however, still relies critically on optimizing material parameters that govern the speed, cyclability, and power consumption of devices.7,10,11

Many technologically relevant phase change materials are located on the ternary

Ge-Sb-Te phase diagram, including alloys along the pseudobinary GeTe-Sb2Te3 tie line,

11 Ge15Te85, and doped Sb2Te. Recent synthetic routes to these materials have largely focused on sputter deposition and lithographical patterning of phase-change films.7,12-16

Nanoscale phase change materials have been important targets, particularly as platforms for investigating the influence of reduced dimensionality on crystallization properties,15-18 melting temperature,19,20 thermal bi-stability,21,22 and material degradation.23 Among the

Adapted with permission from Chem. Mater. 22, 3236, “Liquid-Phase Synthesis of Uniform Cube-Shaped GeTe Microcrystals” Copyright 2010 American Chemical Society. 81 nanoscale materials that have been made, single-crystalline phase change nanowires fabricated by chemical vapor deposition (CVD) via vapor-liquid-solid (VLS) growth have received the most attention.24-30 Electrical contacts for interrogating the resistive phase change behavior can be attached directly to the ends of the nanowires, and the melting temperature required to achieve the crystalline-to-amorphous transition has been shown to scale directly with nanowire diameter.

Solution-based synthetic routes represent an attractive alternative for depositing films and generating dimensionally-controlled nanoscale materials, including for quantum-confined binary antimonide and telluride . However, there have been few reports describing liquid-phase routes to chalcogenide phase change materials.

Milliron and co-workers deposited GeSbSe phase change films from solutions containing

Ge-Se and Sb-Se precursors, which had been prepared by dissolving the bulk metal chalcogenides in hydrazine.31 Korgel and co-workers reacted diphenylgermane and a trioctylphosphine-tellurium complex (TOP-Te) in supercritical hexane, which generated micron-scale GeTe octahedra decorated with vertically-aligned arrays of Te nanowires.32

GeTe nanowires were formed in high-boiling organic solvents via seeded growth from Bi nanoparticles, but the product contained large amounts of excess Te.33 More recently, a

GeI2-TOP adduct and TOP-Te were reacted at 250°C in the presence of an activating alkylthiol to produce amorphous GeTe nanoparticles.34 Interestingly, in-situ X-ray diffraction measurements while heating revealed a dramatic increase in crystallization temperature above the bulk value, which was strongly dependent on particle size.

Here, an important addition to the small number of reports on solution routes to

Ge-based phase change materials is described: a bench-top, liquid-phase synthesis of

82 crystalline GeTe via reduction of GeI2 with tert-butylamine borane (TBAB) in the presence of TOP-Te at 180 ºC. The GeTe product consists of highly faceted microcrystals having a cube-shaped morphology with a narrow size distribution. These microcrystals can be deposited on planar substrates to form highly textured films. The effects of exposure to an electron beam, including electron beam induced fragmentation and partial amorphization, are also reported. Finally, we show preliminary evidence that this reaction chemistry can be leveraged to convert pre-formed, chemical vapor- deposited, amorphous Ge thin films into GeTe.

2.2 Experimental Details

2.2.1 Materials

All chemicals were used as received. Germanium(II)iodide (GeI2, 99.99+%), trioctylphosphine (TOP, tech. 90%), tert-butylamine borane complex (TBAB, powder

97%) and octyl ether (99%) were purchased from Aldrich. Tellurium powder (99.99%,

-325 mesh) was obtained from Alfa Aesar. All syntheses were carried out under Ar using standard Schlenk techniques, and work-ups were performed in air.

2.2.2 Synthesis of GeTe microcrystallites

A clear, yellow stock solution 0.4 M in TOP-Te complex was first prepared by dissolving 512 mg Te powder in 10 mL TOP at 260 °C. GeI2 (66 mg) was dissolved in

2.5 mL TOP via ultrasonication and mixed with 2.5 mL TOP-Te solution in a 25 mL

83 round-bottom flask equipped with a reflux condenser, thermometer adapter, rubber septum, and vacuum adapter. The yellow solution was heated to 120 °C under vacuum for 30 min to remove air and moisture, and then heated to 180 °C under an Ar blanket.

TBAB (52 mg), a mild reducing agent, was dissolved in 1 mL warmed (~100 °C) octyl ether, and swiftly injected into the reaction flask. The introduction of TBAB produced an immediate color change from yellow to brown-black. After aging at 180 °C for 90 min, the product mixture was cooled to room temperature. The precipitate was separated by centrifugation from a transparent, pale yellow supernatant, washed several times with 3:1 toluene/ethanol and dried under vacuum, yielding a grey powder.

2.2.3 Conversion of vapor-deposited a-Ge thin films to GeTe

2 To remove any native GeOx surface species, 0.5-1.0 cm pieces of the vapor- deposited a-Ge film were briefly sonicated in aqueous base (0.01 M NaOH), rinsed with water, and rinsed with ethanol. The washed film was immediately transferred to a 20 mL, septum-capped vial, which was then sealed and evacuated. Under a blanket of Ar, 2-

3 mL of TOP-Te stock solution was added to the vial. The vial was heated until the a-Ge film rapidly changed appearance from shiny to black (230-260 °C), which indicated that a reaction with TOP-Te had occurred. The vessel was cooled, and the converted film was rinsed thoroughly with hexanes and ethanol.

84 2.2.4 Characterization

Powder X-Ray diffraction (XRD) data were collected using a Bruker D8 Advance

X-ray diffractometer equipped with CuKα radiation, and lattice parameters were determined using Chekcell. Scanning electron microscopy (SEM) images and energy- dispersive X-ray spectroscopy (EDS) data were acquired from a FEI Quanta 200

Environmental SEM operating in low vacuum mode. Differential scanning

(DSC) was performed on a TA Instruments Q600 SDT under flowing Ar at a heating rate of 10 °C/min. Transmission electron microscopy (TEM) images and selected area electron diffraction (SAED) patterns were obtained from a JEOL 1200 EX II operating at

80 kV. Samples were prepared by suspending the washed crystallites in toluene and drop-casting onto Formvar-coated copper TEM grids.

2.3 Results and Discussion

The product formed upon reducing GeI2 with TBAB in the presence of TOP-Te at

180 ºC is crystalline rhombohedral GeTe, as shown by XRD in Figure 2-1. This rhombohedral phase is slightly distorted from the rocksalt structure, and is the GeTe polymorph that is typically observed at low temperatures (< 400 °C).35,36 Refinement of the lattice constants yielded a = 5.984(4) Å and α = 88.63(5)º in space group R3m (No.

85

Figure 2-1: Experimental and simulated XRD patterns for dried powders and drop-cast films of rhombohedral GeTe. Drop-casting the microcrystallites produces preferred orientation along <002>, indicating a faceted, non-spherical morphology.

160), which agreed well with literature reports (a = 5.996 Å, α = 88.18º).36 The intensities of powdered samples agree well with those predicted from the simulated XRD pattern, also shown in Figure 2-1. However, when XRD samples were prepared by suspending the product crystallites in toluene and drop-casting them onto a Si substrate, significant preferred orientation was observed (Figure 2-1). The enhanced relative intensity of (002) and (004) reflections in the drop-cast samples suggests that the crystallites are non-spherical with exposed {002} facets.

Figure 2-2(a) shows an SEM image of a typical GeTe product. The crystallites appear uniform in size and highly faceted with a cube-shaped morphology. The edge

86 lengths of the cubes are 1.0 ± 0.2 µm, as summarized in the size distribution histogram in

Figure 2-2(b). EDS element mapping clearly shows the presence of both Ge and Te

Figure 2-2: (a) SEM micrograph of the GeTe product, which consists of highly-faceted, micron- sized crystallites with a cube-shaped morphology. The particle shape is consistent with the preferred orientation observed by XRD. (b) Size distribution histogram for the GeTe crystallites, which have an average edge length of 1.0 ± 0.2 µm.

uniformly distributed throughout each particle [Figure 2-3(c,d)]. The EDS spectrum in

Figure 2-3(a) indicates a 58:42 ratio of Ge:Te, which is in agreement with the nominal

1:1 ratio of Ge:Te expected in the product (within experimental error due to the semi- quantitative nature of EDS analysis), although we cannot rule out the possible presence of amorphous Ge or GeO2 impurities on the particle surface or within its interior.

Prior reports of GeTe microcrystallites, synthesized without added surface stabilizers in supercritical fluid32 or using CVD,37 were in the form of octahedra with exposed {111} facets. The octahedral morphology has been rationalized on the basis of surface energy minimization, because the fewest number of dangling bonds results from

87 the most densely-packed {111} planes in the distorted rocksalt structure.38 Chemical additives such as octanol, hexadecanethiol, and isoprene have been reported to influence

Figure 2-3: (a) EDS spectrum and (b) SEM image of the GeTe microcrystals, along with the associated EDS element mapping data showing uniform distribution of (c) Ge and (d) Te throughout each particle.

the GeTe particle growth and resultant mophology.32 In the present study, the exposed facets are the {002} family of planes rather than {111}. Figure 2-4 shows two views of the GeTe crystal structure: one highlighting the (111) surface, and the other highlighting the (002) surface. The {111} faces are composed entirely of a hexagonal close-packed arrangement of either Te or Ge atoms, and in contrast, the {002} faces have a rocksalt- like arrangement of alternating Ge and Te atoms. As described by Wang for metallic fcc crystals, the cube-shaped morphology likely results from faster growth in the <111> directions relative to <002>.39 We hypothesize that one or more components of the

88 reaction mixture (TOP, I-, octyl ether) has a significant affinity for the {002} planes.

While this process yields uniform crystallites, the size is difficult to control, which could result from the growth-dominated crystallization kinetics of the Ge-Te systems.40

Figure 2-4: Two views of the rhombohedral GeTe crystal structure, with the (002) [top] and (111) [bottom] surfaces highlighted. In contrast to the {002} faces, which consist of an alternating arrangement of Ge and Te atoms, the {111} faces are composed of hexagonally close- packed Te (or Ge) atoms.

A representative TEM image of some of the smaller (~500 nm) GeTe crystallites is shown in Figure 2-5(a). Despite the thickness of the particles and the corresponding high contrast, it is clear that the crystallites are largely faceted cubes with truncated corners, which is consistent with particle morphologies observed by SEM in Figure 2-2.

Interestingly, upon exposure to the electron beam, significant fracturing occurs. Figure

89 2-5(b) shows a representative TEM image of the fractured products, which are well dispersed and correspond to highly crystalline GeTe based on the SAED pattern (Figure

2-5(c). Electron beam induced fragmentation (EBIF) is a well-known process that results from the transfer of thermal energy and electrical charge to certain samples, particularly those containing constituent elements of disparate melting points (MPGe = 938 ºC, MPTe

= 450 ºC).40-44 The GeTe fragmentation products are consistent in size, dispersion, and crystallinity with other systems that have been reported to exhibit EBIF behavior.40,43,44

Figure 2-5: TEM images of (a) a collection of some of the smaller GeTe crystallites, showing generally truncated and faceted cube-shaped morphologies and (b) a region of smaller crystallites produced during imaging via electron beam induced fragmentation. The SAED pattern in (c) corresponds to the smaller fragmented nanocrystals from (b) and indicates highly crystalline rhombohedral GeTe. The SAED pattern in (d) is from a region of crystalline GeTe particles similar in morphology and size to those shown in (a) that began to become amorphous after continued exposure to the electron beam. (e) DSC trace for the GeTe microcrystallites, showing an endotherm at 717 °C corresponding to the melting of GeTe.

90 Accessing the amorphous state is critical for applications that exploit the optical and electrical properties associated with reversible phase change. One approach for inducing amorphization involves melting that results from Joule heating.45 Figure 2-5(e) shows DSC data for the GeTe microcrystallites. An endotherm is observed at 717 ºC, which agrees well with the melting point of bulk GeTe (720 ºC).46 Also, exposure of

GeTe to the electron beam during imaging shows evidence of melting and amorphization.

Figure 2-5(d) shows an SAED pattern corresponding to a crystalline sample that shows signs of becoming amorphous during imaging: a few diffraction spots indicative of crystalline regions are evident, but the majority of the sample generates diffuse rings that are consistent with amorphous GeTe.

Importantly, GeI2 and TOP-Te do not react to form GeTe in the absence of the reducing agent, TBAB. The introduction of TBAB initiated a rapid color change from yellow to dark brown, which is presumably correlated to the reduction of Ge2+. This observation suggests that the reduction of Ge2+ to Ge0 may be an essential step in the formation of GeTe, which led us to hypothesize that pre-fabricated Ge0 nanostructures may also convert to GeTe when treated with TOP-Te. Indeed, initial experiments show that a-Ge thin films (50-100 nm thickness), when deposited on either glass or (111)- oriented Si substrates, transform into GeTe films when immersed in hot TOP-Te solution.

Figure 2-6 shows a representative SEM image of a 50 nm thick a-Ge film on (111)-Si, after treatment with TOP-Te solution at 250 ºC. EDS-generated elemental mapping confirms the presence of Ge and Te, which are uniformly distributed throughout the converted film, and are clearly supported by the Si substrate. Notably, conversion of

91

Figure 2-6: (a) SEM image of a 50 nm thick a-Ge film on (111)-Si, after treatment with TOP-Te solution at 250 ºC. EDS-generated elemental mapping confirms the presence of (c) Ge and (d) Te, which are uniformly distributed throughout the converted film, and clearly supported by (b) the Si substrate.

a-Ge films on glass resulted in delamination of the film from the substrate during heating.

These initial studies provide preliminary evidence of an alternative route to GeTe nanostructures: one that merges solution-phase reactions with conventional “top-down” physical deposition and lithography techniques.

2.4 Conclusions

In summary, we have reported a low temperature liquid-phase synthesis of rhombohedral GeTe microcrystals by reacting GeI2 and TOP-Te at 180 °C in the presence of TBAB. The microcrystals are uniform (1.0 ± 0.2 µm) and cube-shaped, and

92 can be deposited onto planar substrates to produce highly textured 002-oriented films.

During TEM imaging, the particles undergo electron beam induced fragmentation and, in some cases, partial amorphization. This work provides insight into solution-mediated morphology control, which is rare for Ge-based phase change materials. It also targets a size regime that is often difficult to access as high-quality uniform particles using solution chemistry methods, but one that is important for accessing larger-scale features that directly map onto applications that require near-infrared optical responses. Initial attempts to convert a-Ge thin films to GeTe using TOP-Te solution were also successful, which implies a promising, alternative approach for synthesizing phase change films and patterned nanostructures.

2.5 References

(1) Raoux, S. Annu. Rev. Mater. Res. 2009, 39, 25.

(2) Raoux, S.; Welnic, W.; Ielmini, D. Chem. Rev. 2010, 110, 240.

(3) Yamada, N.; Ohno, E.; Nishiuchi, K.; Akahira, N.; Takao, M. J. Appl. Phys. 1991,

69, 2849.

(4) Yamada, N. MRS Bull. 1996, 21, 48.

(5) Wuttig, M.; Yamada, N. Nat. Mater. 2007, 6, 824.

(6) Hudgens, S.; Johnson, B. MRS Bull. 2004, 29, 829.

(7) Lankhorst, M.H.R.; Ketelaars, B.W.S.M.M.; Wolters, R.A.M. Nat. Mater. 2005,

4, 347.

93 (8) Lyke, J. In Phase Change Materials: Science and Applications; Raoux, S.,

Wuttig, M., Eds.; Springer: Berlin, Heidelberg, New York 2008, Chapter 18, pp

409-430.

(9) Ovshinsky, S.R. Jpn. J. Appl. Phys. 2004, 43, 4695.

(10) Wuttig, M. Nat. Mater. 2005, 4, 265.

(11) Lencer, D.; Salinga, M.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Wuttig, M.

Nat. Mater. 2008, 7, 972.

(12) Chen, Y.C. et al. Electron Devices Meeting 2006, IEDM '06. International, 1-4.

(13) Kim, S.M.; Shin, M.J.; Choi, D.J.; Lee, K.N.; Hong, S.K.; Park, Y.J. Thin Solid

Films 2004, 469, 322.

(14) Zhang, Y.; Wong, H-S.P.; Raoux, S.; Cha, J.N. ; Rettner, C.T. ; Krupp, L.E. ;

Topuria, T. ; Milliron, D.J. ; Rice, P.M. ; Jordan-Sweet, J.L. Appl. Phys. Lett.

2007, 91, 013104.

(15) Raoux, S.; Rettner, C.T.; Jordan-Sweet, J.L.; Kellock, A.J.; Topuria, T.; Rice,

P.M.; Miller, D.C. J. Appl. Phys. 2007, 102, 094305.

(16) Zhang, Y. ; Raoux, S.; Krebs, D. ; Krupp, L.E. ; Topuria, T.; Caldwell, M.A. ;

Milliron, D.J. ; Kellock, A. ; Rice, P.M. ; Jordan-Sweet, J.L.; Wong, H-S.P. J.

Appl. Phys. 2008, 104, 074312.

(17) Martens, H.C.F.; Vlutters, R.; Prangsma, J.C. J. Appl. Phys. 2004, 95, 3977-3983.

(18) Wang, W.J.; Shi, L.P.; Zhao, R.; Lim, K.G.; Lee, H.K.; Chong, T.C.; Wu, Y.H.

Appl. Phys. Lett. 2008, 93, 043121.

(19) Raoux, S.; Shelby, R.M.; Jordan-Sweet, J.L.; Munoz, B.; Salinga, M.; Chen, Y-

C.; Shih, Y-H.; Lai, E-K. ; Lee, M-H. Microelectron. Eng. 2008, 85, 2330.

94 (20) Sun, X. ; Yu, B. ; Ng, G. ; Meyyappan, M. J. Phys. Chem. C 2007, 111, 2421.

(21) Gotoh, T, ; Sugawara, K. ; Tanaka, K. Jpn. J. Appl. Phys. 2004, 43, 818.

(22) Lee, S-H.; Jung, Y.; Agarwal, R. Nano Lett. 2008, 8, 3303.

(23) Lai, S. Electron Devices Meeting, IEDM Tech. Dig. 2003, IEEE International,

10.1.1-10.1.4.

(24) Meister, S.; Peng, H.; McIlwrath, K.; Jarausch, K.; Zhang, X.F.; Cui, Y. Nano

Lett. 2006, 6, 1514.

(25) Yu, D.; Wu, J.; Gu, Q.; Park, H. J. Am. Chem. Soc. 2006, 128, 8148.

(26) Jung, Y.; Lee, S-H.; Ko, D-K.; Agarwal, R. J. Am. Chem. Soc. 2006, 128, 14026.

(27) Lee, S-H.; Ko, D-K.; Jung, Y.; Agarwal, R. Appl. Phys. Lett. 2006, 89, 223116.

(28) Lee, S-H.; Jung, Y.; Agarwal, R. Nat. Nanotech. 2007, 2, 626.

(29) Lee, J.S.; Brittman, S.; Yu, D.; Park, H. J. Am. Chem. Soc. 2008, 130, 6252.

(30) Jennings, A.T.; Jung, Y.; Engel, J.; Agarwal, R. J. Phys. Chem. C 2009, 113,

6898.

(31) Milliron, D.J.; Raoux, S.; Shelby, R.; Jordan-Sweet, J. Nat. Mater. 2007, 6, 352.

(32) Tuan, H-Y.; Korgel, B.A. Cryst. Growth Design 2008, 8, 2555.

(33) Lee, M-K.; Kim, T.G.; Ju, B-K.; Sung, Y-M. Cryst. Growth Design 2009, 9, 938.

(34) Caldwell, M.A.; Raoux, S.; Wang, R.Y.; Wong, H-S.P.; Milliron, D.J. J. Mater.

Chem. 2010, 20, 1285.

(35) Edwards, A.H.; Pineda, A.C.; Schultz, P.A.; Martin, M.G.; Thompson, A.P.;

Hjalmarson, H.P.; Umrigar, C.J. Phys. Rev. B 2006, 73, 045210.

(36) Goldak, J.; Barrett, C.S.; Innes, D.; Youdelis, W. J. Chem. Phys. 1966, 44, 3323.

95 (37) Chung, H-S.; Jung, Y.; Kim, S.C.; Kim, D.H.; Oh, K.H.; Agarwal, R. Nano Lett.

2009, 9, 2395.

(38) Wang, S.G.; Tian, E.K.; Lung, C.W. J. Phys. Chem. Solids 2000, 61, 1295.

(39) Wang, Z.L. J. Phys. Chem. B 2000, 104, 1153.

(40) Raoux, S.; Cheng, H-Y.; Caldwell, M.A.; Wong, H-S.P. App. Phys. Lett. 2009,

95, 071910.

(41) Ru, Q. Appl. Phys. Lett. 1997, 71, 1792.

(42) Lu, B. ; Wang, C. ; Zhang, Y. Appl. Phys. Lett. 1997, 70, 717.

(43) Li, Y.; Wang, J.; Deng, Z.; Wu, Y.; Sun, X.; Yu, D.; Yang, P. J. Am. Chem. Soc.

2001, 123, 9904.

(44) Pan, Z.M.; Dai, Z.R.; Wang, Z.L. Appl. Phys. Lett. 2002, 80, 309.

(45) Pyrz, W.D. ; Park, S. ; Vogt, T. ; Buttrey, D.J. J. Phys. Chem. C 2007, 111,

10824.

(46) Yu, B. ; Sun, X. ; Ju, S. ; Janes, D.B. ; Meyyappan, M. IEEE Trans. Nanotech.

2008, 7, 496.

(47) Schlieper, A. ; Feutelais, Y. ; Fries, S.G.; Legendre, B.; Blachnik, R. Calphad

1999, 23, 1.

Chapter 3

Polymer-Assisted Synthesis of Colloidal Germanium Telluride Nano- Octahedra, Nanospheres, and Nanosheets

3.1 Introduction

Nanostructures of the binary IV-VI semiconductor, germanium telluride (GeTe), exhibit a variety of fundamentally interesting and potentially useful size-dependent properties. GeTe belongs to a class of so-called phase change materials1 that undergo a reversible, amorphous-to-crystalline structural transformation that is accompanied by a significant change in electrical resistance, which underpins the operation of emerging non-volatile phase change memory technologies2-4 and future neuromimetic computing applications.5,6 Several important properties of GeTe, such as the crystallization and melting temperature, scale favorably for such applications at reduced dimensions.

Significantly elevated crystallization temperatures (relative to bulk), for example, have been observed in low-dimensional GeTe thin films7 and amorphous GeTe nanoparticles,8,9 which prevent spontaneous crystallization events that can erase amorphous bits of stored memory. Also, the energy input required for amorphization of the crystalline phase is minimized at low melting temperatures, and depressed melting has been observed in GeTe thin films7 and nanowires.10,11 GeTe is also among the simplest known ferroelectrics,12 and persistent, monodomain polar ordering was recently observed in sub-10 nm GeTe nanocrystals at room temperature, with multiple polarization domains emerging in larger size nanoparticles.13

Adapted with permission from Chem. Mater. DOI:10.1021/cm4009656, “Polymer-Assisted Synthesis of Colloidal Germanium Telluride Nano-Octahedra, Nanospheres and Nanosheets” Copyright 2013 American Chemical Society. 97 Interest in its dimensionally-dependent properties has led to the preparation of nanoscale GeTe by a variety of high-energy input methods, such as sputter deposition of thin films14-16 that are amenable to lithographic patterning,17,18 and vapor-liquid-solid

(VLS) growth of high-aspect ratio nanowires.10,11,19,20 Although colloidal synthetic routes represent a more versatile, lower-cost, and potentially scalable alternative for synthesizing solution-processable, morphology-controlled nanomaterials,21 and have been used to make other IV-VI tellurides,22-24 there have been few reports of solution-grown

GeTe nanostructures. The most successful of these involve reaction between

Ge(II)bis(hexamethyldisilylamide) (Ge[N(SiMe3)2]2) and tellurium-trioctylphosphine complex (TOP-Te) in the presence of surface-stabilizing agents, and rely critically on the immediate, rapid cooling and quenching of the reaction mixture. Arachchige et al.

injected Ge[N(SiMe3)2]2 into a mixture of TOP-Te containing dodecanethiol and oleylamine surfactants at 220 °C, and after one min of reaction, well-dispersed 4 nm

GeTe nanocrystals were formed.9 At longer reaction times (2-3 min), however, the nanocrystals showed a strong tendency to coalesce into irregular-shaped crystallites.

Polking et al. reported the same reaction in several solvent-surfactant mixtures, and obtained high-quality 8 and 17 nm GeTe nanocrystals, when the reaction was cooled

25 II rapidly after 1.5 to 2 min. The reaction of Ge[N(SiMe3)2]2 (and other Ge sources) with

TOP-Te, in many solvent/surfactant combinations, forms insoluble GeTe precipitates at longer reaction times (>5 min).26,27 Collectively, these observations suggest the presence of a strong, attractive force between crystalline GeTe particles that is only weakly mediated by molecular surface-passivating ligands. Unlike other IV-VI compounds,28,29 there is a priori no oriented attachment mechanism that has been observed for solution-

98 grown GeTe nanostructures, at the short timescales required for quenching, that has led to higher-order control over size and shape.

Polymers have been used extensively as surface-stabilizing and shape-directing agents in the colloidal synthesis of metals30-33 and metal chalcogenides,34-37 either as alternatives to, or combined with, molecular surfactants. In addition to having Lewis basic groups that passivate dangling surface bonds, such polymers sterically hinder and prevent nanoparticle aggregation, and impart colloidal stability. Here, we report the

reaction of Ge[N(SiMe3)2]2 with TOP-Te in the presence of either poly(vinylpyrrolidone)

(PVP), or a hydrophobic derivative poly(1-vinylpyrrolidone)-graft-(1-triacontene) (PVP- t), which effectively mediate the growth of GeTe nanoparticles without the need for rapid quenching of the reaction mixture. The stability that is conferred to the nanoparticles by the polymer affords an additional degree of freedom with respect to achieving morphology control, primarily because avoiding irreversible agglomeration and precipitation are no longer the focus. We leverage this by studying the effects of charged carboxylate species on the polymer-stabilized reaction, which has allowed us to synthesize several novel GeTe nanostructures, including GeTe nano-octahedra,

amorphous GexTe1-x nanospheres, and single-crystal GeTe nanosheets. (Figure 3-1 concisely summarizes the chemical details and the products obtained, as a complement to the order of presentation in the Results and Discussion.) These previously unreported nanostructures and the chemical insights into their formation represent a clear starting point from which composition- and morphology-controlled GeTe and related materials may be tailored for phase change memory applications, and for studying nanoscale ferroelectricity.

99 3.2 Experimental Details

3.2.1 Materials

All chemicals were used as received. Ge(II)bis(hexamethyldisilylamide)

(Ge[N(SiMe3)2]2, >95%) was purchased from Gelest, Inc. Trioctylphosphine (TOP, tech.

90%), 1-octadecene (ODE, tech. 90%), poly(1-vinylpyrrolidone)-graft-(1-triacontene)

(PVP-t), and sodium octanoate (NaOct, ≥99%) were purchased from Sigma-Aldrich.

Tellurium powder (99.99%, -325 mesh), oleic acid (OLAC, tech. 90%), and nonanoic acid (NONAC, 97%) were obtained from Alfa Aesar. Poly(vinylpyrrolidone) (PVP,

Avg. MW=40,000) was purchased from TCI America. A clear, yellow stock solution of

0.5 M TOP-Te was prepared by dissolving 1.27 g Te powder in 20 mL TOP by heating at

260-280 °C, stored in an Ar-filled glovebox, and used for all syntheses of GeTe nanostructures. All syntheses were carried out under Ar using standard Schlenk techniques, and work-ups were performed in air.

3.2.2 Synthesis of faceted GeTe nanoparticles

A colorless solution, approx. 2.5 wt% PVP-t, was prepared by dissolving 125 mg

PVP-t in 6 mL OLAC in a 50 mL round-bottom flask equipped with a reflux condenser, thermometer adapter, rubber septum, and vacuum adapter. The solution was first heated to 120 °C under vacuum for 30 min to remove air and moisture, and then 120 mg (0.3 mmol) Ge[N(SiMe3)2]2 dissolved in 0.5 mL hexanes was added to the contents of the flask, which remained colorless despite the orange-yellow color of Ge[N(SiMe3)2]2.

100 After the hexanes was distilled off under flowing Ar, the solution was heated to 230 °C under an Ar blanket. The reaction was initiated by rapid injection of TOP-Te (0.3 mL,

0.15 mmol), forming an opaque, dark brown colloidal dispersion in 1-2 min, and was allowed to proceed at 230 °C for 30 min. The product mixture was cooled to room temperature by removing the heating source, diluted with an equal volume of toluene, and GeTe nanoparticles were separated by centrifugation at 12,000 rpm. Several re- dispersion/centrifugation cycles in toluene were used to purify the product, which was dispersed in toluene for sample preparation.

3.2.3 Amorphous GexTe1-x nanospheres and single-crystal GeTe nanosheets

The experimental setup and process are identical to that used to prepare faceted

GeTe nanoparticles. PVP (67 mg, 0.6 mmol) and NaOct (50 mg, 0.3 mmol) were dissolved in 5-6 mL NONAC and degassed at 110 °C under vacuum. (Heating at reduced pressures also promotes dissolution of PVP in NONAC, presumably due to removal of water.) Under an Ar blanket at 110 °C, Ge[N(SiMe3)2]2 (120 mg) was added via syringe, and the colorless solution gradually turned pale yellow as it was then heated to 250 °C. TOP-Te (0.3 mL) was injected at 250 °C, and ~5 min were required for a dark-brown colloidal dispersion to form (in the absence of NaOct, the color change occurs in ~1 min). After 20 min the product mixture was cooled, diluted with an equal volume of hexanes, and centrifuged (12,000 rpm). Several re-dispersion/centrifugation cycles in ethanol were used to purify the products, which were dispersed in ethanol for

101 sample preparation. For preparation of GeTe nanosheets, the amount of NaOct was increased to 250 mg (1.5 mmol), and ~15 min were required for a color change to occur.

3.2.4 Characterization

Powder X-ray diffraction (XRD) data were collected using a Bruker D8 Advance

X-ray diffractometer equipped with Cu Kα radiation. Transmission electron microscopy

(TEM) images were obtained from a JEOL 1200 EX II operating at 80 kV. High- resolution TEM (HRTEM) images and selected area electron diffraction (SAED) patterns were collected using a JEOL 2010 LaB6 microscope operating at an accelerating voltage of 200 kV. Scanning transmission electron microscopy, including high angle annular dark field (HAADF) imaging, coupled with energy-dispersive X-ray spectroscopy

(STEM/EDS), was performed using a JEOL 2010F field-emission microscope outfitted with an EDAX solid-state X-ray detector. ES Vision software (Emispec) was used for

STEM/EDS data processing and the Ge and Te L-level transitions, which do not appreciably overlap, were chosen for quantitative STEM/EDS analysis. Samples were prepared by suspending the washed crystallites in toluene or ethanol and drop-casting onto Formvar-coated copper TEM grids.

Field emission scanning electron microscopy (FESEM) images were collected with an FEI NanoSEM 630 operating under high vacuum (1.7 × 10-6 Torr). Beam deceleration (BD), a feature that reduces electron beam energy by applying a bias to the sample stage, was used to reduce charging and enhance surface sensitivity in the images.

Accelerating voltages used ranged from 2 to 3 kV and landing energy (energy of the

102 decelerated electrons as they impact the sample) ranged from 0.7 to 1.5 kV. The samples were also studied using energy dispersive X-ray spectroscopy (SEM/EDS). Characteristic x-rays were collected with an X-Max silicon drift detector (Oxford Instruments) inside the FESEM. Elements were selected and quantified using Aztec Energy Analyzer

Software (Oxford Instruments). Statistical analyses reported in the text were performed by counting the number of particles that have a particular morphology, or fall within a certain size regime, relative to the total number of particles counted in either TEM or

SEM images (n, typically 200–300), expressed as a percentage.

AFM measurements were made on a Bruker Dimension Icon (Bruker, Billerica,

MA, USA). Peak force tapping mode with Scanalyst air probes (Bruker, nom. spring constant of 0.4 N ⁄ m, nom. deflection sensitivity of 60 nm/V) was used to obtain the sample topography. The linear scanning rate was set between 0.25 and 1.0 Hz, scan resolution set to 512 samples per line, and peak force set point between 0.5 and 1 nN.

3.3 Results and Discussion

3.3.1 Polymer-Assisted Synthesis of Faceted GeTe Nanoparticles

Injection of TOP-Te into a 1-octadecenyl solution of Ge[N(SiMe3)2]2 at 230 °C induces an instantaneous color change from orange-yellow to dark brown and formation of an insoluble precipitate. As shown by the XRD pattern in Figure 3-2(b), the precipitate is crystalline rhombohedral GeTe (space group R3m), with peak positions and intensities that correspond well to the simulated pattern in Figure 3-2(a). However, under

103

Figure 3-1: Summary of primary chemical variables and corresponding products.

otherwise identical reaction conditions, when the solvent is replaced by a 2.5 wt. % solution of PVP-t in ODE, a stable colloid is obtained. The isolated brown/black solid is

104 amorphous by XRD [Figure 3-2(c)] after 1 h of reaction at 230 °C, and even after 12 h

[Figure 3-2(d)], the largely amorphous powder contains only a small amount of crystalline Te. These results strongly suggest that PVP-t provides a barrier to the crystallization of GeTe that occurred in the reaction corresponding to Figure 3-2(b). The

Figure 3-2: (a) Simulated and (b) experimental powder XRD patterns corresponding to rhombohedral GeTe, prepared by reacting (Ge[N(SiMe3)2]2) and TOP-Te in ODE at 230 °C. Under otherwise identical conditions, when a 2.5 wt. % solution of PVP-t in ODE is used as the solvent, amorphous solids are obtained even after (c) 1 h and (d) 12 h of heating, suggesting that the polymer prevents crystallization of GeTe. (e) When 2.5 wt. % PVP-t in OLAC is used, the XRD pattern shows only crystalline GeTe, indicating that the reaction is affected significantly by the solvent.

amorphous solids obtained in Figures 3-2(c) and 3-2(d) consist of tiny, structurally disordered particulates that contain Ge and Te (Figure 3-3). In the absence of PVP-t, they would otherwise be transformed into crystalline GeTe nuclei by a pathway that

105 involves coalescence or fusion. By the same rationale, the polymer likely acts as a steric blockade, which does not hinder Ge[N(SiMe3)2]2 and TOP-Te from reacting, but does prevent the colloidal, amorphous Ge- and Te-containing particulates from interacting with each other.

Figure 3-3: TEM image of the amorphous solids obtained by reacting Ge[N(SiMe3)2]2 and TOP- Te in ODE, in the presence of PVP-t. The product consists primarily of small particulates with average sizes that mostly fall within the range of 3-6 nm, although a few are larger. Scale bar = 20 nm.

Replacing the non-coordinating ODE solvent with OLAC, a carboxylic acid, has a significant effect on the rate of reaction and the product obtained. While the injection of

TOP-Te induces an instantaneous color change in ODE solvent, the reaction appears to proceed much slower in OLAC, requiring 1-2 min for the pale yellow solution to transform into a stable, dark brown colloidal dispersion. The XRD pattern shown in

Figure 3-2(e) reveals that the product is not amorphous, and consists only of rhombohedral GeTe, without any crystalline impurities.

106 It is well-established that group 14 bis(trimethylsilylamide) complexes, including

Ge[N(SiMe3)2]2, react with a wide variety of protogenic molecules (HA) to form derivative complexes by

M[N(SiMe!)!]! + ! HA → M[N(SiMe!)!]!!!A! + ! HN(SiMe!)! (Reaction 3-1)

- 38-42 due to the basicity of the N(SiMe3)2 ligand. When Ge[N(SiMe3)2]2 is introduced into

ODE at 120 °C, the orange-yellow color of Ge[N(SiMe3)2]2 persists in the resulting solution. Mixing Ge[N(SiMe3)2]2 with OLAC, however, results immediately in a colorless solution, indicating a possible ligand exchange that is similar to Reaction 3-1. It is therefore likely that Ge[N(SiMe3)2]2 reacts with OLAC via protonolysis to form an oleate-type complex, which has significantly different reactivity with TOP-Te than does

Ge[N(SiMe3)2]2 in ODE. In many nanocrystal systems, including some binary chalcogenides, the rate at which the “monomer” species that lead to nanocrystal nuclei are consumed is strongly influenced by the stability of the complexes that form in the surrounding ligand environment.43,44 In general, when monomer is consumed quickly, a large number of smaller-sized nuclei are formed, and when monomer is consumed more slowly, a smaller number of larger nuclei are formed. We therefore propose that the putative oleate complexes formed upon mixing Ge[N(SiMe3)2]2 with OLAC have a significantly lower activity than Ge[N(SiMe3)2]2 in ODE, leading to a slower reaction with TOP-Te, and accordingly, to a smaller number of larger, crystalline GeTe nuclei.

In Figure 3-4 are representative TEM images of the product formed by reacting

Ge[N(SiMe3)2]2 with TOP-Te in OLAC in the presence of PVP-t, which show a narrowly

107

Figure 3-4: (a) Representative TEM image of faceted GeTe nanoparticles with (b) a higher- magnification image highlighting the octahedral morphology and (c) the corresponding electron diffraction pattern.

dispersed mixture of dense, highly faceted nanoparticles. The dominant morphology appears to be octahedra (85%, n = 275), and the edge lengths range from approximately

70 to 100 nm. The SAED pattern shown in Figure 3-4(c), which corresponds well to rhombohedral GeTe, exhibits an intense diffraction ring that can be assigned to overlap of the {022} and {022} reflections. Preferred orientation caused by slow drying of the

{111}-terminated, octahedral-shaped particles on the TEM grid, however, may also be responsible for additional intensity. Indeed, HRTEM images of the nanoparticles show a hexagonal array of atoms (Figure 3-5), which is characteristic of the {111} planes of

108 GeTe. The lattice planes perpendicular to the image plane have a d-spacing of 2.1 ± 0.1

Å, which is consistent with the crystal structure, and correspond well to the {022} planes in rhombohedral GeTe (2.09 Å). The highly faceted, polyhedral-shaped GeTe nanoparticles can be more clearly seen in the SEM images shown in Figure 3-6(a). EDS

[Figure 3-6(b)] confirms that Ge and Te are present in a 52:48 atomic percent ratio, which agrees, within experimental error, with the expected 1:1 stoichiometry of Ge:Te.

Figure 3-5: (a) HRTEM image of a representative GeTe nanoparticle, with an enlarged region in (b), corresponding to region enclosed by the red dashed box in (a), showing lattice fringes matching those of GeTe {022}. (c) The crystal structure of rhombohedral GeTe (purple = Ge, yellow = Te) viewed along its [111] axis, highlighting the atomic arrangement and interatomic distances that correspond well to the image shown in (b).

109

Regardless of whether Ge[N(SiMe3)2]2 and TOP-Te are reacted in the coordinating or non-coordinating solvent, the inclusion of the polymer PVP-t mediates the growth of GeTe, and gives rise to colloidally-stable nanocrystallites that can be processed and stored in non-polar liquids. We also found that PVP-t could be replaced with the hydrophilic parent polymer, PVP, with the restriction that the OLAC solvent also be replaced with a carboxylic acid that is sufficiently amphiphilic, in order to maintain compatibility with the reactants and the polymer. Specifically, we chose nonanoic acid (NONAC) as a solvent, which is a liquid at room temperature, and has a boiling temperature (250 °C) that is convenient for the synthesis of GeTe.

Figure 3-6: (a) Representative SEM image of faceted GeTe nanoparticles and (b) the corresponding EDS spectrum.

When TOP-Te is injected into a NONAC solution containing Ge[N(SiMe3)2]2 and

PVP at 250 °C, highly faceted GeTe nanoparticles are formed that can be processed and stored in polar liquids such as alcohol or water. The size and shape distribution of the resulting crystallites, when low concentrations of PVP are present (2:1 PVP:Ge2+, defined as the molar ratio of N-vinylpyrrolidone (111 g/mol) to Ge[(NSiMe3)2]2), is similar to

110

Figure 3-7: TEM image of faceted GeTe nanoparticles, synthesized in the presence of PVP in NONAC at 250°C and at a PVP:Ge2+ ratio of 2.

Figure 3-8: (a) Powder XRD data for GeTe nanoparticles synthesized in the presence of PVP in NONAC at 250°C and at a PVP:Ge2+ ratio of 10. The peak positions correspond well to the simulated pattern, but the diffraction intensities from the {111}, {111 }, {222}, and {222} planes are enhanced, suggesting preferred orientation. (b) TEM image of the corresponding GeTe nanoparticles, showing that many have grown into plate-like, 2D shapes, which is likely the cause of the preferred orientation in (a).

111 that obtained using PVP-t, albeit less uniform (Figure 3-7). Unlike when PVP-t is used, where a range of concentrations of PVP-t in OLAC was observed to produce no perceivable differences in the final product, PVP in higher concentrations (10:1-20:1

PVP:Ge2+) in NONAC gives rise to preferred orientation in the XRD pattern. The relative intensities of the {111}, {111}, {222} and {222} reflections shown in Figure 3-

8(a) are markedly enhanced. This observation is consistent with the corresponding TEM images shown in Figure 3-8(b), where a significant number of platelets with hexagonal and triangular shapes can be seen. Although GeTe adopts a face-centered cubic (fcc)

NaCl-type structure above ~400 °C, the symmetry-breaking distortion that characterizes the low-temperature rhombohedral phase gives rise to alternating long (weak) and short

(strong) Ge-Te bonds along its <111> crystallographic axis,45,46 which is known to cause stacking faults and several types of twinning.47 Assisted by the presence of PVP, and perhaps the nonanoate ion, these planar defects may promote two-dimensional (2D) growth and give rise to plate-like crystals. Similar-looking 2D triangular and hexagonal platelets have been observed in fcc metals that exhibit a strong tendency to form stacking faults and twin boundaries, such as Ag and Au. Formation of those metallic plates has been ascribed to the structural asymmetry conferred by such planar defects, combined with surface-ligand interactions.48-51

3.3.2 Synthesis of amorphous GexTe1-x alloy nanospheres

Increasing the ratio of PVP to Ge[N(SiMe3)2]2 beyond 10 does not further alter the distribution of shapes shown in Figure 3-8(b), which suggests that free nonanoate ion

112 may play a role in formation of 2D GeTe nanostructures. Irrespective of the PVP concentration, addition of sodium octanoate (NaOct), a source of free carboxylate ion which is structurally similar to nonanoate, had dramatic and unexpected effects on the reaction between Ge[N(SiMe3)2]2 and TOP-Te. While a color change from pale yellow to dark brown ordinarily took 1-2 min in NONAC solvent, anywhere from 5 to 15 min was required for the same color change when NaOct was included. This observation suggests that NaOct decreases the reactivity of one or both reactants, significantly slowing the formation of GeTe. Although a small number of faceted GeTe nanoparticles are still observed, including some triangular and hexagonal plates, at NaOct:Ge2+ ratios ranging from 0.25 to 1, the dominant product emerged as the spheres shown in Figure 3-

9(a). TEM images show that the spherical particles range in size from about 50 to 100 nm in diameter (davg = 90 ± 10 nm, n = 270). SAED patterns acquired from groups of the spheres [inset, Figure 3-9(a)] consistently show broad, featureless rings, indicating that the particles are amorphous, and HRTEM images [Figure 3-9(b)] of the nanospheres do not show features that are consistent with atomic-scale ordering. HAADF imaging coupled with STEM/EDS was used to determine the elemental composition of individual nanospheres. The L-level emission intensity from Ge and Te, plotted in Figure 3-9(c) as a function of position across a representative nanosphere, shows that Ge and Te are homogeneously distributed throughout each particle. These measurements indicate that the atomic percents of Ge and Te present in each particle are 30% and 70%, respectively, which agrees with the high Z-contrast in the annular dark-field images (the nanospheres appear very bright) due to Te-rich composition.

113

Figure 3-9: (a) Representative TEM image of GexTe1-x alloy nanospheres with corresponding SAED pattern (inset), showing that the particles are amorphous, and (b) HRTEM image (with an enlargement in the inset, corresponding to the region enclosed by the dashed red box) highlighting the lack of atomic ordering for a representative nanosphere. (c) STEM/EDS line scan spectrum across a GexTe1-x nanosphere (right) as indicated on the HAADF image (left), showing a homogeneous distribution of Ge and Te where x ≈ 0.3.

The formation of amorphous GexTe1-x (where x ≈ 0.3) alloy nanospheres is an unusual and unexpected result. Although amorphous GeTe nanoparticles have been synthesized previously, the particle sizes were in the sub-5 nm regime, where crystallization temperatures are elevated, and the Ge:Te ratios were near 1:1.8,9 The most

114 closely-related nanomaterials reported are non-stoichiometric GexTe1-x alloy thin films with compositions ranging from x = 0.3 - 0.7, which had been prepared by co-sputtering from elemental Ge and Te sources in the appropriate ratio.52-54 Notably, the crystallization time and temperature of these films were found to be strongly dependent on the composition, which has important implications for GeTe-based phase change memory technology. The annealing temperature of 250 °C used to prepare GexTe1-x alloy nanospheres is above that at which crystallization occurs for those Te-rich thin films, and at reaction times of 20-30 min, it is surprising that we did not observe precipitation of crystalline Te.54 Also, our approach yields Te-rich nanospheres, in spite of the 2:1

Ge[N(SiMe3)2]2:TOP-Te ratio employed in the reaction, which leaves much of the Ge precursor unreacted. A 4:1 Ge[N(SiMe3)2]2:TOP-Te ratio yielded a nearly identical product. Implied previously, the emergence of the nanospheres is most likely caused by the presence of NaOct, and it is possible that NaOct suppresses the reactivity of the Ge precursor, but does not limit the activity of TOP-Te. By that rationale, intermediate Ge- and Te-containing species that form initially may react with TOP-Te at a faster rate than they react with Ge precursor, giving rise to Te-rich species. For example,

Ge[N(SiMe3)2]2 reacts with elemental Te to form [Ge{N(SiMe3)2}2(µ-Te)]2 complexes

55 that are highly reactive. The reaction pathway that leads to GexTe1-x alloy nanospheres, however, is still under investigation.

115 3.3.3 Synthesis of single-crystal 2D GeTe nanosheets

When Ge[N(SiMe3)2]2 and TOP-Te were reacted in the presence of higher concentrations of NaOct (NaOct:Ge2+ > 5), the results changed drastically. In addition to

GexTe1-x alloy nanospheres, a large number of 2D nanosheets emerge, which can be seen in the TEM images in Figure 3-10. The 2D nanostructures represent approximately 45% of a typical sample (n = 250), with the remainder corresponding to the amorphous nanospheres. The nanosheets have lateral dimensions that range from approximately

175-900 nm, and exhibit diverse 2D morphologies, including triangles, truncated triangles, quadrilaterals, and hexagons. In some cases they are folded at the edges. Figure

3-11(a) shows a powder XRD pattern from a representative sample containing a mixture of 2D nanosheets and GexTe1-x nanospheres, which is consistent with the pattern expected for rhombohedral GeTe, but with preferred orientation. Only broad, weak reflections for

{111}, {111}, {222} and {222} can be seen, which suggests that the 2D nanostructures are in fact crystalline GeTe, considering that the GexTe1-x nanospheres are amorphous.

SAED was used to examine the crystal structure of individual nanosheets, including the triangular nanosheet shown in the inset of Figure 3-11(a), and all nanosheets consistently showed bright diffraction spots corresponding to the {022} planes in rhombohedral GeTe

[Figure 3-11(b)]. Collectively, the diffraction data indicate that the nanosheets are 2D single crystals of GeTe that are lying flat, oriented with {111} planes parallel to the sample stage. The HAADF/STEM/EDS linescan spectra presented in Figure 3-12 confirm that Ge and Te are uniformly distributed throughout each nanosheet in a 1:1 ratio, within the limits of experimental error due to the semi-quantitative nature of EDS.

116

Figure 3-10: (a) TEM image of the GeTe product formed in the presence of excess NaOct. In addition to GexTe1-x nanospheres, the product mixture contains a large number of 2D nanosheets. (b-h) Representative TEM images of the nanosheets, showcasing the diverse 2D morphologies that emerge. The product includes (b) triangles, (c) quadrilaterals, (d,e) hexagons, (f) other polyhedral shapes, and (g-h) folded, multi-layered structures. Unlabeled scale bars are 100 nm.

Figure 3-11: (a) XRD pattern corresponding to the product shown in Figure 6a, confirming the presence of rhombohedral GeTe, with preferred orientation in the [111] direction. (b) SAED pattern of the single triangular nanosheet shown in the inset of (a). The {022} diffraction spots indicate a [111] axis perpendicular to the TEM image, which is consistent with the preferred orientation observed in (a).

117

Figure 3-12: STEM/EDS line scan and corresponding HAADF images for a hexagonal (top) and triangular (bottom) GeTe nanosheet. Ge and Te are uniformly distributed throughout each nanosheet in an approximate 1:1 ratio.

AFM was used to measure the thickness of the nanosheets, which ranged from approximately 5.5 to 13 nm (Figure 3-13).

Many IV-VI semiconductors (GeS, GeSe, SnS, SnSe) adopt a layered crystal structure, with weak van der Waals forces holding the 2D, covalently bonded lamellae together. Such materials exhibit a structurally derived, intrinsic tendency to form anisotropic 2D structures, and many such nanostructures have been solution-grown.56-58

The polar distortion along a <111> axis, which accompanies the cubic-to-rhombohedral phase transition in GeTe, weakens the Ge-Te bonding in these planes.45 This is structurally related to the weaker interlayer bonding that exists in layered solids, and therefore represents the most likely driving force for the formation of GeTe nanosheets, which indeed appear to grow with the polar <111> axis perpendicular to the sheet plane.

Considering that GeTe nanosheets form only in the presence of NaOct, it may be that octanoate ion binds preferentially to the {111} surfaces, which would slow growth along

118 the polar axis. The presence of planar defects, as discussed earlier, may also play a role in the growth of these 2D nanostructures.

Figure 3-13: Representative AFM data characterizing the thicknesses of the colloidal GeTe nanosheets. The line profiles labeled [1], [2], and [3] in (b) correspond to the lines indicated in the AFM image in (a). The vertical position markers in (b) are represented by color-coded dots in (a), which highlight the height differential between the nanosheets and the substrate. The thicknesses of the nanosheets [1], [2], and [3] are 6.7 nm, 13.1 nm, and 7.9 nm, respectively. The image in panel (a) includes groups of GexTe1-x alloy nanospheres, which appear white because their height is taller than that of the maximum z-axis scale. (c) High-magnification AFM image and corresponding height profile of a multi-layered GeTe nanosheet with a maximum height of 5.6 nm.

Particle growth was investigated by studying the products of the Ge[N(SiMe3)2]2 /

TOP-Te reaction at short reaction times (< 1 min after a color change was observed), and we found some evidence of a coalescence pathway that may have led to GeTe

119 nanosheets. In Figures 3-14(a) and 3-14(b) can be seen many small nanoparticles (5 to

30 nm) that appear to have the same transparency and planarity as the GeTe nanosheets.

The edges of many nanosheets at this stage of the reaction appear rough, suggesting that the smaller, planar fragments may fuse together along their edges to form the larger 2D

Figure 3-14: (a,b) High-magnification TEM images of GeTe nanosheets taken at early stages of reaction, which suggest that growth may occur by laterally-oriented fusion of smaller particles. (c) HRTEM image of a 2D GeTe quadrilateral (inset), showing well-defined edges and vertices, but an atomically disordered face. The nanosheets consistently exhibited such disorder, despite the single-crystallinity revealed by electron diffraction. (d) TEM image of several 2D GeTe triangles that have higher-contrast, triangular domains in their centers. This observation suggests that, in rare cases, 2D GeTe nanostructures may grow laterally from a dense GeTe seed.

120 structures. Nearly all of the GeTe nanosheets, even those with well-defined, smooth edges and vertices, appear to have porous faces, which also suggests their formation by attachment of smaller fragments. Interestingly, HRTEM images in Figure 3-14(c) indicate that the surfaces of GeTe nanosheets lack any evidence of atomic ordering, and that this “roughness” masks the single-crystalline ordering beneath. We cannot rule out the possibility, however, that surface oxidation was responsible for the observed disorder.

A previous report has shown that clean GeTe (111) surfaces, that were freshly prepared under ultra-high vacuum conditions, spontaneously react with oxygen upon exposure to air.59

Remarkably, a small fraction (1-2 %) of GeTe nanosheets appear to form via a seeded-growth mechanism. A dense, triangular domain can be clearly seen at the center of several triangular GeTe nanosheets in Figure 3-14(d), suggesting that the nanosheets grew outwardly from the dense seed. In addition to the bright {022} diffraction spots expected for the nanosheets [see Figure 3-11(b)], additional reflections for rhombohedral

GeTe {111}, {002}, and {222} are clearly present in SAED of a representative “seeded” nanosheet (Figure 3-15). The additional reflections appear to be due to the dense, central domain, which is presumably thicker than the rest of the nanosheet. The presence of these two-domain nanostructures implies that the seeded-growth strategies used in other anisotropic metal60-62 and semiconducting63,64 nanocrystal systems, could perhaps be leveraged to synthesize other anisotropic GeTe nanostructures. To our knowledge, these nanosheets represent the first examples of free-standing, colloidal, 2D GeTe, and as such, they add to the rapidly growing library of 2D metal chalcogenide nanosheets65-67 that are of interest for applications in energy storage68-70 and optoelectronic devices.71-76

121

Figure 3-15: (a) TEM image and (b) SAED pattern for a representative GeTe nanosheet with a dense, triangular domain at its center, which implies seed-mediated growth. In addition to the {02-2} reflections expected for the GeTe nanosheets, additional diffraction spots are present in (b), which are likely due to the central domain. All the reflections in (b) correspond well to the crystal structure of rhombohedral GeTe. The scale bar in (a) corresponds to 100 nm.

3.4 Conclusions

Despite their fundamental and technological relevance, the solution-phase synthesis of morphology-controlled GeTe nanostructures remains challenging. In our experience, the primary synthetic obstacle is to mediate those strong, attractive, interparticle forces that provoke agglomeration, uncontrolled growth, and precipitation, which preclude higher-order shape control. We have approached this problem by using soluble polymers (PVP and PVP-t) to mediate the reaction between Ge[N(SiMe3)2]2 and

TOP-Te, and have discovered that the polymers indeed provide colloidal stabilization during growth of GeTe nanostructures at longer reaction times than previously reported.

We have leveraged this additional degree of freedom, which has allowed us to study the dramatic effects that free carboxylate ligands have on the reaction kinetics and GeTe morphology.

122 By varying the free carboxylate concentration, we have observed the formation of several GeTe nanostructures that have not been previously reported, including GexTe1-x alloy nanospheres and single-crystal 2D GeTe nanosheets. Although some key insights have been revealed, additional studies, which are ongoing, are needed in order to fully understand the reaction pathway that leads to these new and unexpected morphologies.

The relevance of novel GeTe nanostructures to phase-change memory technologies is intriguing, considering the advantages of solution-phase processing and the ease with which anisotropic shapes with such dimensions can be “wired-up” to study melting and crystallization behavior. Such morphology-controlled GeTe nanostructures could also continue to provide an ideal platform for correlating theory and experiment in quantum- confined ferroelectric materials.

3.5 References

(1) Lencer, D.; Salinga, M.; Grabowski, B.; Hickel, T.; Neugebauer, J.; Wuttig, M.

Nature Mater. 2008, 7, 972.

(2) Raoux, S. Annu. Rev. Mater. Res. 2009, 39, 25.

(3) Raoux, S.; Welnic, W.; Ielmini, D. Chem. Rev. 2010, 110, 240.

(4) Siegrist, T.; Merkelbach, P.; Wuttig, M. Annu. Rev. Condens. Matter Phys. 2012,

3, 215.

(5) Kuzum, D.; Jeyasingh, R.G.D.; Lee, B.; Wong, H-S.P. Nano Lett. 2012, 12, 2179.

(6) Wright, C.D.; Hosseini, P.; Vasquez Diosdado, J.A. Adv. Funct. Mater. 2013,

DOI: 10.1002/adfm.201202383.

123 (7) Raoux, S.; Shelby, R.M.; Jordan-Sweet, J.L.; Munoz, B.; Salinga, M.; Chen, Y-

C.; Shih, Y-H.; Lai, E-K. ; Lee, M-H. Microelectron. Eng. 2008, 85, 2330.

(8) Caldwell, M.A.; Raoux, S.; Wang, R.Y.; Wong, H-S.P.; Milliron, D.J. J. Mater.

Chem. 2010, 20, 1285.

(9) Arachchige, I.U.; Soriano, R.; Malliakas, C.D.; Ivanov, S.A.; Kanatzidis, M.G.

Adv. Funct. Mater. 2011, 21, 2737.

(10) Sun, X.; Yu, B.; Ng, G.; Meyyappan, M. J. Phys. Chem. C 2007, 111, 2421.

(11) Lee, S-H.; Ko, D-K.; Jung, Y.; Agarwal, R. Appl. Phys. Lett. 2006, 89, 223116.

(12) Steigmeier, E.F.; Harbeke, G. Solid State Commun. 1970, 8, 1275.

(13) Polking, M.J.; Urban, J.J.; Milliron, D.J.; Zheng, H.; Chan, E.; Caldwell, M.A.;

Raoux, S.; Kisielowski, C.F.; Ager, J.W., III; Ramesh, R.; Alivisatos, A.P. Nano

Lett. 2011, 11, 1147.

(14) Bruns, G.; Merkelbach, P.; Schlockerman, C.; Salinga, M.; Wuttig, M.; Happ,

T.D.; Philipp, J.B.; Kund, M. Appl. Phys. Lett. 2009, 95, 043108.

(15) Raoux, S.; Munoz, B.; Cheng, H-Y.; Jordan-Sweet, J.L. Appl. Phys. Lett. 2009,

95, 143118.

(16) Santala, M.K.; Reed, B.W.; Raoux, S.; Topuria, T.; LaGrange, T.; Campbell, G.H.

Phys. Status Solidi B 2012, 249, 1907.

(17) Milliron, D.J.; Raoux, S.; Shelby, R.; Jordan-Sweet, J. Nature Mater. 2007, 6,

352.

(18) Zhang, Y.; Wong, H-S.P.; Raoux, S.; Cha, J.N. ; Rettner, C.T. ; Krupp, L.E.;

Topuria, T.; Milliron, D.J.; Rice, P.M. ; Jordan-Sweet, J.L. Appl. Phys. Lett. 2007,

91, 013104.

124 (19) Yu, D.; Wu, J.; Gu, Q.; Park, H. J. Am. Chem. Soc. 2006, 128, 8148

(20) Jennings, A.T.; Jung, Y.; Engel, J.; Agarwal, R. J. Phys. Chem. C 2009, 113,

6898.

(21) Talapin, D.V.; Lee, J-S.; Kovalenko, M.V.; Shevchenko, E.V. Chem. Rev. 2010,

110, 389.

(22) Murphy, J.E.; Beard, M.C.; Norman, A.G.; Ahrenkiel, S.P.; Johnson, J.C.; Yu, P.;

Micic, O.I.; Ellingson, R.J.; Nozik, A.J. J. Am. Chem. Soc. 2006, 128, 3241.

(23) Urban, J.J.; Talapin, D.V.; Shevchenko, E.V.; Murray, C.B. J. Am. Chem. Soc.

2006, 128, 3248.

(24) Kovalenko, M.V.; Heiss, W.; Shevchenko, E.V.; Lee, J-S.; Schwinghammer, H.;

Alivisatos, A.P.; Talapin D.V. J. Am. Chem. Soc. 2007, 129, 11354.

(25) Polking, M.J.; Zheng, H.; Ramesh, R.; Alivisatos, A.P. J. Am. Chem. Soc. 2011,

133, 2044.

(26) Tuan, H-Y.; Korgel, B.A. Cryst. Growth Design 2008, 8, 2555.

(27) Buck, M.R.; Sines, I.T.; Schaak, R.E. Chem. Mater. 2010, 22, 3236.

(28) Schliehe, C.; Juarez, B.H., Pelletier, M.; Jander, S.; Greshnykh, D.; Nagel, M.;

Meyer, A.; Foerster, S.; Kornowski, A.; Klinke, C.; Weller, H. Science 2010, 329,

550.

(29) Koh, W-K.; Bartnik, A.C.; Wise, F.W.; Murray, C.B. J. Am. Chem. Soc. 2010,

132, 3909.

(30) Tao, A.R.; Habas, S.; Yang, P. Small 2008, 4, 310.

(31) Li, C.; Shuford, K.L.; Chen, M.; Lee, E. J.; and Cho S.O. ACS Nano 2008, 2,

1760.

125 (32) Wiley, B.; Sun, Y.; Mayers, B.; Xia, Y. Eur. J. Inorg. Chem. 2005, 11, 454.

(33) Wang, F.; Tang, R.; Yu, H.; Gibbons, P.C., Buhro, W.E. Chem. Mater. 2008, 20,

3656.

(34) Nenadovic, M.T.; Comor, M.I.; Vasic, V.; and Micic, O.I. J. Phys. Chem. 1990,

94, 6390.

(35) Zhou, H.S.; Honma, I.; Komiyama, H.; Haus, J.W. J. Phys. Chem. 1993, 97, 895.

(36) Zhang, G.; Kirk, B.; Jauregui, L.A.; Yang, H.; Xu, X.; Chen, Y.P.; Wu, Y. Nano

Lett. 2012, 12, 56.

(37) Zhang, G.; Fang, H.; Yang, H.; Jauregui, L.A.; Chen, Y.P.; Wu, Y. Nano Lett.

2012, 12, 3627.

(38) Lappert, M.F.; Protchenko, A.V.; Power, P.P.; Seeber, A.L., Metal Amide

Chemistry; John Wiley & Sons: Hoboken 2008, pp 279-295.

(39) Caulton, K.G.; Hubert-Pfalzgraf, L. G. Chem. Rev. 1990, 90, 969.

(40) Gerung, H.; Boyle, T.J.; Tribby, L.J.; Bunge, S.D.; Brinker, C.J.; Han, S.M. J.

Am. Chem. Soc. 2006, 128, 5244.

(41) Green, R.A.; Moore, C.; Rheingold, A.L.; Weinert, C.S. Inorg. Chem. 2009, 48,

7510.

(42) Boyle, T.J.; Tribby, L.J.; Ottley, L.A.M.; Han, S.M. Eur. J. Inorg. Chem. 2009,

36, 5550.

(43) Yu, W.W.; Peng, X. Angew. Chem. Int. Ed. 2002, 41, 2368.

(44) Yu, W.W.; Wang, Y.A.; Peng, X. Chem. Mater. 2003, 15, 4300.

(45) Goldak, J.; Barrett, C.S.; Innes, D.; Youdelis, W. J. Chem. Phys. 1966, 44, 3323.

(46) Rabe, K. M.; Joannopoulos, J. D. Phys. Rev. B 1987, 36, 6631.

126 (47) Synkers, M.; Delavignette, P.; Amelinckx, S. Mat. Res. Bull. 1972, 7, 831.

(48) Germain, V.; Li, J.; Ingert, D.; Wang, Z.L.; Pileni, M.P. J. Phys. Chem. B,

2003,107, 8717.

(49) Elechiguerra, J.L.; Reyes-Gasgab, J.; Yacaman, M.J. J. Mater. Chem. 2006, 16,

3906.

(50) Aherne, D.; Ledwith, D.M.; Gara, M.; Kelly, J.M. Adv. Funct. Mater. 2008, 18,

2005.

(51) Zhang, Q.; Li, N.; Goebl, J.; Lu, Z.; Yin, Y. J. Am. Chem. Soc. 2011, 133, 18931.

(52) Raoux, S.; Cheng, H-Y.; Caldwell, M.A.; Wong, H-S.P. Appl. Phys. Lett. 2009,

95, 07190.

(53) Yan, N., Liu, X.Q.; Zhang, L.; Zhang, Z.; Han, X.D. Chem. Phys. Lett. 2013, 556,

108.

(54) Carria, E.; Mio, A.M.; Gibilisco, S.; Miritello, M.; Bongiorno, C.; Grimaldi,

M.G.; Rimini, E. J. Electrochem. Soc. 2012, 159, H130.

(55) Hitchcock, P.B.; Jasim, H.A.; Lappert, M.F.; Leung, W-P.; Rai, A.K.; Taylor,

R.E. Polyhedron 1991, 10, 1203.

(56) Vaughn, D.D., II; Patel, R.J.; Hickner, M.A.; Schaak, R.E. J. Am. Chem. Soc.

2010, 132, 15170.

(57) Vaughn, D.D., II; In, S-I.; Schaak, R.E. ACS Nano. 2011, 5, 8852.

(58) Deng, Z.; Cao, D.; He, J.; Lin, S.; Lindsay, S.M.; Liu, Y. ACS Nano 2012, 6,

6197.

(59) Yashina, L.V.; Puttner, R.; Neudachina, V.S.; Zyubina, T.S.; Shtanov, V.I.;

Poygin, M.V. J. Appl. Phys. 2008, 103, 094909.

127 (60) Jana, N. R.; Gearheart, L.; Murphy, C. J. Adv. Mater. 2001, 13, 1389.

(61) Ye, X.; Jin, L.; Caglayan, H.; Chen, J.; Xing, G.; Zheng, C.; Doan-Nguyen, V.;

Kang, Y.; Engheta, N.; Kagan, C.R.; Murray, C.B. ACS Nano, 2012, 6, 2804.

(62) Skrablak, S.E.; Xia, Y. ACS Nano, 2009, 3, 10.

(63) Manna, L.; Milliron, D.J.; Meisel, A,; Scher, E.C.; Alivisatos, A.P. Nat. Mater.

2003, 2, 382.

(64) Deka, S.; Miszta, K.; Dorfs, D.; Genovese, A.; Bertoni, G.; Manna, L. Nano Lett.

2010, 10, 3770.

(65) Son, J.S.; Yu, J.H., Kwon, S.G; Lee, J.; Joo, J.; Hyeon, T. Adv. Mater. 2011, 23,

3214.

(66) Bouet, C.; Tessier, M.D.; Ithurria, S.; Nadal, B.; Mahler, B.; Dubertret, B. Chem.

Mater. 2013, DOI: 10.1021/cm303786a

(67) Butler, S.Z., et al, ACS Nano 2013, DOI: 10.1021/nn400280c.

(68) Seo, J-W.; Jang, J-T.; Park, S-W.; Kim, C.; Park, B.; Cheon, J. Adv. Mater. 2008,

20, 4269.

(69) Zhang, Y.; Lu, J.; Shen, S.; Xu, H.; Wang, Q. Chem. Commun. 2011, 47, 5226.

(70) Vaughn, D.D., II; Hentz, O.D.; Chen, S.; Wang, D.; Schaak, R.E Chem. Commun.

2012, 48, 5608.

(71) Choi, J.; Kang, N.; Yang, H.Y.; Kim, H.J.; Son, S.U. Chem. Mater. 2010, 22,

3586.

(72) Kirkeminde, A.; Ruzicka, B.A.; Wang, R.; Puna, S.; Zhao, H.; Ren, S. ACS Appl.

Mater. Interfaces 2012, 4, 1174.

128 (73) Yin, Z.; Li, H.; Li, H.; Jiang, L.; Shi, Y.; Sun, Y.; Lu, G.; Zhang, Q.; Chen, X.;

Zhang, H. ACS Nano, 2012, 6, 74.

(74) Xue, D-J.; Tan, J.; Hu, J-S.; Hu, W.; Guo, Y-G.; Wan, L-J. Adv. Mater. 2012, 24,

4528.

(75) Hu, P.; Wen, Z.; Wang, L.; Tan, P.; Xiao, K. ACS Nano, 2012, 6, 5988.

(76) Li, L.; Chen, Z.; Hu, Y.; Wang, X.; Zhang, T.; Chen, W.; Wang, Q. J. Am. Chem.

Soc. 2013, 135, 1213.

Chapter 4

Investigating the Thermal Decomposition of Co(II) Oleate for Synthesis of Colloidal Wurtzite CoO Nanocrystals

4.1 Introduction

The thermally-induced decomposition of solid-state metal carboxylate salts was once an important reaction for generating symmetrically-substituted ketones via ketonic decarboxylation, which liberates the ketone as the major product, along with CO2 and the corresponding metal oxide.1 High-yield ketonization reactions involving base-catalyzed decarboxylation of carboxylic acids have since replaced metal carboxylate thermolysis as the preferred methods for preparing ketones.2 This is primarily because such thermolysis reactions proceed by complicated pathways, which effects poorly on ketone yields. In addition to the target ketone, CO2, and the metal oxide, thermal decomposition of carboxylate salts often evolves a complex by-product mixture of lower, asymmetric ketones, , , CO, H2O, H2, carboxylic acids, and other hydrocarbon fragments.3 Detailed analytical studies have strongly implied the formation of radical intermediates during the decomposition of carboxylate salts,3-6 that may be generated by thermolytic cleavage of either metal-oxygen (M-O) or carbon-oxygen (C-O) bonds by

RCOOM → RCOO·+ M· or RCOOM → RC·O+ O·M (Reaction 4-1)

130

The pathways by which these highly reactive species propagate the decomposition reaction are generally unknown, and are likely responsible for the product diversity.

Despite its shortcomings as a route to ketones, the thermal decomposition of metal carboxylates has garnered renewed interest as a pathway to high-quality metal oxide nanocrystals. Fatty acid salts of 3d transition metals, in particular, thermolyze in high-boiling (typically, 300-330 °C) organic solvents to generate colloidal nanocrystals in

7 high yield. Gram-scale quantities of uniformly-sized Fe3O4 nanocrystals, for example, can be produced in a facile, one-step, reaction by thermolysis of Fe(III) cis-9- octadecanoate (Fe-oleate) in non-coordinating solvents. Metal-carboxylate decomposition has also been applied successfully to the synthesis of Mn(II), Fe(II), Fe(III), Co(II),

Co(II,III), and Zn(II) oxide nanocrystals.8-13 Several fatty acid salts can be decomposed in the same solution to yield mixed-metal oxide nanocrystals such as CoFe2O4, MnFe2O4,

14,15 and NiFe2O4. In several cases, the reaction conditions, including the incorporation of surface stabilizers and other chemical additives, have enabled researchers to rigorously control the size, shape, and uniformity of the nanocrystals.16,17 Nanostructured transition metal oxides, including colloidal nanocrystals, have size- and shape-dependent magnetic, optical, and electrical properties that are promising for applications in biology,18-20 catalysis,21-24 and energy.25-27

The thermal decomposition of metal-carboxylate solutions is often sensitive to subtle differences in experimental conditions, which can have deleterious effects on reproducibility and utility. Although scalable quantities of monodisperse, inverse-spinel

7 Fe3O4 nanocrystals were prepared using Fe(III)oleate, for example, subsequent reports

131 indicated that body-centered cubic Fe and wustite-phase FeO were the dominant products of this reaction.28-31 The thermolysis of Fe(III) oleate solutions was later studied using mass spectrometry, and many volatile by-products, including H2, CO, CO2, and various hydrocarbons, were detected during decomposition.32 The reducing power of dissolved

H2 or CO was likely responsible for the generation of Fe and lower-valent iron oxides.

This series of reports demonstrates the unpredictable influence that side-products can have on the formation of nanocrystals. Although the conditions that may have regulated the generation of such side-products include the solvent and heating rate, the cause for discrepancy remains unclear.

The temperature at which thermal decomposition occurs is also known to be sensitive to subtle, structural variability in the metal-carboxylate complexes. For example, the presence of coordinated solvent or water molecules influences the coordination geometry of carboxylate ligand and the stability of the complex. Bronstein et al. have shown that Fe(III) oleate exhibits decomposition temperatures ranging from

300-380 °C, that depend on whether the complex is purified by washing with water, methanol, or acetone.31 They reported a range of decomposition temperatures that also depended on how the complex was dried. When decomposition occurs at higher temperatures, side-reactions may also play a dominant role in the processes that effect nanocrystal growth, such as reduction, oxidation, and etching.

In the seminal paper by Peng, et al. describing nanocrystal synthesis of several 3d transition metal oxides, a significant difference in reactivity between carboxylate salts of different metals was identified.8 Stearate and oleate salts of Mn(II), Fe(II), Fe(III), and

Zn(II) decomposed readily in non-coordinating solvents at 300-320 °C. Fatty

132 carboxylate salts of Cr(III), Co(II) and Ni(II), however, were sluggish to react at similar temperatures in the same solvents. Metal oxide nanocrystals were produced from the latter precursors only when the decomposition pathway was changed by the addition of

“activating” agents. Nucleophilic amines and alcohols, for example, destabilize the metal-carboxylate complexes by attacking the ligand carbonyl group, and eliminate either amides or esters.8,10,12 In contrast to this report, it was later found that solutions of Co(II) oleate [Co(OL)2] straightforwardly decomposed into colloidal, pencil-shaped, CoO nanocrystals that had a metastable, wurtzite-type crystal structure [wz-CoO].11 This called into question the “inert” behavior of those [Co(OL)2] solutions, and the need for activating additives.

Nanostructured cobalt oxides are promising materials for applications in energy storage,25 sensing,34-37 catalysis,38-40 and magnetism.41 Binary cobalt(II) oxide typically crystallizes in the cubic, NaCl-type structure, but hexagonal wz-CoO has been synthesized by several groups.11,42-45 Interest in wz-CoO began with the discovery of room-temperature ferromagnetism in dilute, solid-solutions of wz-CoO in a wz-ZnO host lattice.46 Studies that have highlighted the magnetic and optical properties of nanostructured wz-CoO have followed.11,43,47,48 Recently, wz-CoO nanorods were demonstrated as a suitable host for Cu2+ dopants, which remarkably increased the stability of the (metastable) wz-CoO phase.49 Preliminary studies showed enhanced electrochemical capacitance and cycling performance for Cu2+-doped wz-CoO, which is promising for lithium-ion battery applications.

In this work, we have investigated the colloidal synthesis of wz-CoO nanostructures via thermal decomposition of [Co(OL)2] complex. Importantly, we have

133 discovered an isolable impurity that is inherent to the conventional synthesis of

[Co(OL)2]. Furthermore, we have discovered that [Co(OL)2] does not thermally decompose within the expected temperature range, when it is purified of the impurity. In contrast, we have synthesized pencil-shaped wz-CoO nanorods when no measures were taken to remove the impurity. Comparative analysis of purified [Co(OL)2] and the separated impurity, using Fourier Transform (FTIR), suggests that the impurity contains a free hydroxide functionality, which may be acting as an activating agent. We have successfully synthesized size- and shape-controlled wz-CoO nanocrystals, using purified [Co(OL)2] solutions, by incorporating additives that mimic the action of the hydroxide impurity.

By employing a two-stage heating process under controlled atmosphere, we have also determined that the distillation of volatile side-products is a critical step in the thermal decomposition of purified [Co(OL)2]. This has enabled us to prepare colloidal wz-CoO nanorods from pure [Co(OL)2] solutions, without the need for any chemical additives. This research disseminates detailed empirical knowledge about the thermal decomposition of an important metal carboxylate salt, which may be representative of other viable metal oxide nanocrystal precursors. It also represents our ongoing effort to acquire general insights into those molecular processes that promote nucleation of kinetically-stable, non-equilibrium solids. Eventually, such insights may be used to rationally synthesize inorganic nanostructures that feature metal ions in metastable geometries, like wz-CoO.50-54

134 4.2 Experimental Details

4.2.1 Materials

All chemicals were used as received. Cobalt(II) chloride hexahydrate

(CoCl2ž6H2O, 98%), 1-octadecene (ODE, tech. 90%), oleic acid (OLAC, tech. 90%) and tetramethylammonium hydroxide pentahydrate (TMAOH, ≥97%) were purchased from

Sigma-Aldrich. Sodium oleate [Na(OL)] was purchased from TCI America. All syntheses were carried out under Ar using standard Schlenk techniques, and work-ups were performed in air.

4.2.2 Preparation of crude Co(OL)2

4.76 g CoCl2ž6H2O (20 mmol) and 12.18 g Na(OL) (40 mmol) were loaded into a

1-neck, 250 mL round-bottom flask, equipped with a reflux condenser and magnetic stir bar. At room temperature, 15 mL ethanol, 20 mL distilled, deionized water, and 35 mL hexanes were added to the mixture of powders. The reaction vessel was sealed under a blanket of Ar, placed in a water bath pre-heated to 70 °C and was allowed to react for 4 h while stirring vigorously. The cooled product mixture was transferred to a separatory funnel, the purple-colored organic layer was washed three times with distilled, deionized water, and the aqueous phase was discarded. The washing procedure resulted in a cloudy, purple, organic phase. The solvent was evaporated at room temperature under vacuum, yielding Co(OL)2 as a waxy, dark-purple solid, which was immediately transferred to a refrigerated (8-12 °C) Ar-filled glovebox for storage.

135

4.2.3 Purification of crude Co(OL)2

In certain iterations of our experiments, the cloudy, purple, organic phase was further purified by centrifugation. An off-white colored solid was separated from a clear, purple solution upon centrifugation. The solvent was evaporated at room temperature under vacuum, yielding Co(OL)2 as a waxy, dark-purple solid, which was immediately transferred to an Ar-filled glovebox for storage (also refrigerated, 8-12 °C). The off- white colored solid was also dried under vacuum and stored in an inert atmosphere for analysis.

4.2.4 Synthesis of pencil-shaped wz-CoO nanocrystals.

250 mg of crude Co(OL)2 (0.4 mmol) was added to 6 mL degassed ODE in a 50 mL round-bottom flask, equipped with a reflux condenser and thermometer adapter.

Under an Ar blanket, the mixture was heated from room temperature to reflux (~310 °C) at approximately 2 °C/min. Between 100 °C and 250 °C, the increasing temperature caused a gradual color change from purple to blue to dark blue. Beginning at roughly

200 °C, significant amounts of a white-gray vapor were evolved from the blue solution, which subsided at about 250 °C. At the reflux temperature, the color gradually changed from dark blue to dark green over the course of about 15 min, indicating the nucleation of wz-CoO nanocrystals. The reaction was cooled after 30 min of heating, and 2-propanol was added to precipitate the nanocrystals. Centrifugation (7,000 rpm) was applied to collect the green-colored precipitate, and several re-dispersion/centrifugation cycles in hexanes/ethanol were used to purify the product.

136 4.2.5 Synthesis of cone- and golf tee-shaped wz-CoO nanocrystals

For cone-shaped nanocrystals, 250 mg of purified Co(OL)2 (0.4 mmol) and 9-18 mg (0.05-0.10 mmol) TMAOH were added to 6 mL of degassed ODE in a 50 mL round- bottom flask, equipped with a reflux condenser and thermometer adapter. The heating process and isolation procedure are identical to those used to synthesize pencil-shaped wz-CoO. For golf tee-shaped nanocrystals, we replaced the TMAOH with 30-60 mg

(0.10-0.20 mmol) Na(OL), and kept other variables the same. In cases where Na(OL) was added, we noticed that the color change from dark blue to green occurred more rapidly - in 3 to 5 min. We also found it useful to perform at least two re- dispersion/centrifugation cycles in chloroform/acetone to purify the nanocrystals, when

Na(OL) was used.

4.2.6 Synthesis of bullet-shaped wz-CoO nanocrystals

250 mg of purified Co(OL)2 (0.4 mmol) was added to 6 mL of a degassed ODE solution in a 50 mL round-bottom flask, equipped with a reflux condenser, thermometer adapter, and flow-control adapter. The mixture was heated to 240 °C under flowing Ar, and the temperature was maintained there for 2 h. During this time, clear, colorless distillate could be collected in an external vial. After 2 h, an Ar blanket was applied, and the dark blue solution was heated at a rate of 2 °C/min to the reflux temperature. The reaction was allowed to proceed for 20 min more, during which time the color changed from dark blue to dark green. The isolation procedure was identical to that used for pencil-shaped wz-CoO nanocrystals.

137 4.2.7 Characterization

Powder X-ray diffraction (XRD) data were collected using a Bruker D8 Advance

X-ray diffractometer equipped with Cu Kα radiation. Transmission electron microscopy

(TEM) images and selected area electron diffraction (SAED) patterns were obtained from a JEOL 1200 EX II operating at 80 kV. Samples were prepared by suspending the washed nanocrystals in hexanes or toluene, and drop-casting onto Formvar-coated copper

TEM grids. Fourier Transform Infrared (FTIR) spectra were collected using a Bruker

Vertex V70 Spectrometer (Bruker Optics Billerica MA) equipped with an MVP-Pro diamond single reflection ATR accessory (Harrick Scientific Pleasantville NY). 400 scans at 6 cm-1 resolution were averaged for each sample using a DTGS detector and scan frequency of 5 kHz. In all cases, the spectrum of the clean diamond crystal was used as the reference spectrum. All spectral manipulations were performed using OPUS 5.5

(Bruker Optics, Billerica MA).

4.3 Results and Discussion

4.3.1 Preparation, purification, and FTIR analysis of Co(OL)2 precursors

The conventional preparation of transition metal oleates involves the reaction of

Na(OL) with the appropriate metal chloride in a biphasic (aqueous/organic) solvent mixture. During the reaction, which is usually carried out at reflux (60-70 °C), metal- oleate complex is rapidly extracted into the organic phase, and NaCl remains dissolved in the aqueous phase. Following reference 11, we reacted CoCl2ž6H2O and Na(OL) in a

138 mixture of hexanes, ethanol, and water at 70 °C. After the product was cooled, a purple- colored organic layer separated from a clear, colorless aqueous layer. The colorless liquid was discarded, and the purple organic layer was washed several times with water in a separatory funnel. Although the organic and aqueous phases clearly separated after each washing, we noticed that this process caused the organic layer to become noticeably cloudy. At first, we thought that the cloudiness was due to slow or incomplete phase separation, likely caused by a Co(OL)2-stabilized, water-in-oil emulsion. After several hours, however, a light-colored precipitate settled at the bottom of the organic phase, leaving behind a clear, purple solution. In subsequent iterations of Co(OL)2 synthesis, the precipitated “impurity” was removed from the purple Co(OL)2 solution using centrifugation. After the solution was decanted and hexanes evaporated, the product was vacuum-dried overnight, which yielded Co(OL)2 as the expected waxy, purple, solid.

Although the precipitated impurity appeared an off-white color after centrifugation, a tan- colored, waxy solid remained after drying.

The FTIR spectra of Co(OL)2 and the impurity are presented in Figures 4-1(c) and

4-1(d), respectively, and the spectra for free OLAC and Na(OL) [Figures 4-1(a) and 4-

1(b)] are included as a basis for comparison. All four spectra exhibit similarly- shaped

-1 absorbance in the 2800-3000 cm range, characteristic of CH3 and CH2 stretching bands, including the =C-H stretch at 3025 cm-1, which are due to the cis-9-octadecenoic (oleic) chains present in each molecule. In the 1400-1750 cm-1 range, the FTIR spectrum of

OLAC is dominated by a strong peak at 1710 cm-1, which is characteristic of C=O stretching. Instead of a strong C=O stretch, Na(OL) exhibits peaks at 1420 and 1560

-1 - cm , which respectively correspond to the symmetric [νsym(COO )] and asymmetric

139

- 55 [νasym(COO )] stretching modes of carboxylate ion. Importantly, these features of the

Na(OL) spectrum are representative of metal-carboxylate bonding that is entirely ionic.

-1 Peaks in the Na(OL) spectrum at 1447 and 1465 cm are consistent with asymmetric CH3 bending and symmetric CH2 bending.

Figure 4-1: FTIR absorbance spectra of (a) OLAC, (b) Na(OL), (c) purified Co(OL)2, and (d) an impurity that was isolated by centrifugation during preparation of Co(OL)2.

Carboxylate ligands can coordinate to ionic metal centers in a variety of ways, which include unidentate, chelating, and bridging modes. Each mode of coordination causes variance in the equivalence of the two oxygen atoms, which is observable in FTIR spectra. It is possible to infer the coordination mode of a carboxylate complex by

- - comparing the frequency difference (Δ) between νsym(COO ) and νasym(COO ) to the case

55,56 of ionic bonding; in this case ΔNa(OL). The FTIR spectrum of Co(OL)2 [Figure 4-1(c)]

140

- is very similar to that of Na(OL), however, exhibiting strong signals for νsym(COO ) at

-1 - -1 1415 cm and νasym(COO ) at 1562 cm . These similarities suggest that metal-ligand bonding in Co(OL)2 is largely ionic in character, which is consistent with conclusions drawn by other researchers for the chemically-related, cobalt(II) stearate complex.57

-1 Unlike Na(OL), Co(OL)2 absorbs weakly at 1715 cm , which suggests the presence of either free or weakly bound OLAC. It is not uncommon for the synthesis of metal oleates to result in acidic salts, unless measures are taken to remove the free OLAC, such as

Sohxlet extraction or washing with organic solvents.31,58

The FTIR spectrum of the Co(OL)2 impurity [Figure 4-1(d)] is nearly identical to that of Co(OL)2, with two important differences. First, a sharp doublet centered at ~3560

-1 cm is observed, that is not present in the spectrum of Co(OL)2, which is characteristic of

“free” (not hydrogen-bonded) O-H stretching found in metal-hydroxide complexes.55,59,60

Secondly, the relative intensities of the peaks in the 2800-3000 cm-1 range, when compared to the intensities of signals in the 1400-1750 cm-1 range, are diminished significantly. When Figures 4-1(c) and 4-1(d) are compared, this second observation suggests that there is simply less of the cis-9-octadecenoic moiety in the impurity than in

Co(OL)2. The emergence of hydroxide groups, and the reduced presence of oleyl chains implies strongly that the impurity is structurally related to Co(II) mono-oleate hydroxide complex [Co(OL)2-x(OH)x ] (referred to as “Co(OL)2-x(OH)x impurity” hereafter). Mono- oleate hydroxide complexes have been observed in the preparation of other transition metal oleates (MnII, FeIII) using the described approach, and could be derived from the in situ hydrolysis of metal-oleate bonds during synthesis.13,31

141 4.3.2 Pencil-shaped wz-CoO nanocrystals

The thermal decomposition of Co(OL)2, which reportedly occurs in refluxing 1- octadecene (ODE) solvent (~320 °C),11 depended critically upon whether or not the putative Co(OL)2-x(OH)x impurity was removed during preparation of Co(OL)2. In cases where Co(OL)2-x(OH)x impurity was removed via centrifugation, thermolysis of Co(OL)2 did not take place in refluxing ODE. Rather, the solution remained a stable blue color, even after 90 min of reflux. Between 90 and 120 min at this temperature, partial decomposition of Co(OL)2 was evidenced by a slow color change from blue to blue/gray- brown, which resulted in an insoluble, tan precipitate upon cooling that could not be characterized. In contrast, when no measures were taken to remove the Co(OL)2-x(OH)x impurity, or to purify Co(OL)2 beyond washing with water, we sometimes observed thermal decomposition of Co(OL)2 into a green-colored colloidal dispersion.

Thermolysis of “crude” Co(OL)2 in refluxing ODE resulted in pencil-shaped nanorods of wz-CoO, as evidenced by XRD and TEM (Figure 4-2), which is consistent with a previous report.11 This result, however, was achievable only very sporadically using the same Co(OL)2 precursor sample from experiment to experiment, under identical reaction conditions (250 mg Co(OL)2 in 6 mL ODE, ramp rate 2 °C/min). In the absence of the

Co(OL)2-x(OH)x impurity, Co(OL)2 was sluggish to react at elevated temperatures, suggesting that the impurity was necessary for Co(OL)2 thermolysis. Decomposition to wz-CoO was not consistently reproducible, implying that the Co(OL)2-x(OH)x impurity may be distributed inhomogeneously within the Co(OL)2 precursor.

142

Figure 4-2: (a) Experimental XRD patterns for colloidal wz-CoO nanocrystals, along with a simulated pattern for reference. (b) Large-area TEM image of the nanocrystal product, which consists of uniform nanorods with pencil-point tips. A higher magnification image is shown in the inset of (a). These results were obtained by thermally decomposing Co(OL)2 samples that contained the [Co(OL)2-x(OH)x] impurity. The scale bar in the inset of (a) is 20 nm, and the scale bar in (b) is 50 nm.

4.3.3 Cone- and golf tee-shaped wz-CoO nanocrystals

Co(OL)2 precursor that had not been purified by centrifugation, wherein Co(OL)2- x(OH)x impurity was observed in FTIR spectra, sometimes, albeit infrequently, decomposed into wz-CoO nanocrytals. This led us to hypothesize that the function of

Co(OL)2-x(OH)x could be mimicked by including controlled amounts of other hydroxide- containing additives. Indeed, when judicious quantities of TMAOH were added to the reaction mixture, we consistently obtained wz-CoO nanocrytals via thermolysis of

Co(OL)2 that did not contain Co(OL)2-x(OH)x. Figures 4-3(a-c) show TEM images and a

SAED pattern of wz-CoO nanocrytals that have a stout, conical shape, where the average distance from the apex of the particle to the base is 23 ± 2 nm, and the basal facet measures 29 ± 2 nm. Notably, the particles tended to assemble into pairs with the bases

143 in contact with one another, which has been observed previously for pyramidal- and conically-shaped wz-ZnO nanocrytals.12 This may be due to polar attractions that arise from the non-equivalence of (001) and (001) planes in wurtzite crystals, which are respectively Co2+ and O2- terminated [Figure 4-3(d)]. These results were reproducible when either a 2:1 or 4:1 Co(OL)2:TMAOH molar ratio were used. When larger quantities of TMAOH were added, insoluble precipitates often resulted, that were primarily NaCl-type CoO by XRD.

Figure 4-3: (a) Low-magnification and (b) higher-magnification TEM images of stout, cone- shaped nanocrystals that were formed by thermolysis of Co(OL)2 in the presence of TMAOH. It appears that the nanocones assembled into pairs, with the bases in contact with one another. (c) SAED of the nanocrystals in (a) and (b), which shows diffraction rings that can be assigned to wz- CoO. (d) Representation of the wz-CoO crystal structure depicting the non-equivalent (001) and (001) planes, and the polar attractions that may have caused the nanocones to assemble pairwise. Oxygen atoms are colored red and cobalt atoms are colored blue. The scale bars in (a) and (b) are 50 nm and 10 nm, respectively.

Interestingly, wz-CoO nanocrystals are unstable in the presence of OLAC. When we attempted to direct the nanocrystal shape by injecting a sub-stoichiometric volume of

OLAC after 20 min at reflux, the dark-green, colloidal solution changed (back) to a vivid

144 blue color in ~1 min, indicating the dissolution of wz-CoO nanocrystals and the formation of Co(OL)2. Peng et al. reported previously that wz-ZnO nanocrystals also readily dissolve in OLAC.12 This observation may be critical to understanding the solution- mediated thermolysis of Co(OL)2, and adds several levels of complexity to designing a reproducible procedure for generating nanocrystalline wz-CoO. For example, the FTIR

-1 spectrum of Co(OL)2 in Figure 4-1(c) shows a significant C=O signal at 1710 cm , implying the presence of free or weakly bound OLAC, which may hinder formation of wz-CoO. It is possible that the hydroxide-containing additives, either Co(OL)2-x(OH)x or

TMAOH, play a role by deprotonating and neutralizing excess OLAC. This pathway may be complicated by the water of hydration that is associated with purple-colored

Co(OL)2, because amphoteric water can also react with either OLAC, TMAOH, or

Co(OL)2-x(OH)x. The preceding observations - 1) that wz-CoO dissolves in the presence of (the weak acid) OLAC into Co(OL)2, and 2) that strongly basic, hydroxide-containing additives are necessary to induce Co(OL)2 decomposition into wz-CoO - suggest that complex, acid-base equilibria are involved in the thermolysis of Co(OL)2. Moreover, it is reasonable to assume that the formation of wz-CoO is favored in a basic environment.

We have therefore continued studying the influence of basic additives to the reaction mixture.

Adding controlled amounts of sodium oleate [Na(OL)], which contains the conjugate base of OLAC, promoted decomposition of Co(OL)2 in ODE. XRD of the green-colored colloidal product [Figure 4-4(a)] shows broad reflections, with peak positions that are consistent with wz-CoO. The relative intensity of the {010} reflection is enhanced, suggesting preferred orientation of the crystallites. TEM revealed the

145

Figure 4-4: (a) Experimental XRD and SAED (inset) patterns for the wz-CoO nanocrystals shown in (b), with a simulated XRD pattern for reference. (b) TEM image of the nanocrystals, which have rod-shaped domains protruding from a conical base. The nanocrystals in (b) were formed by thermolysis of Co(OL)2 in the presence of Na(OL), at Co(OL)2:Na(OL) molar ratio of 4. When Co(OL)2:Na(OL) of 2 was used, the rod-shaped domains were longer, which can be seen in (c). All scale bars are 50 nm.

formation of nanocrystals with a unique shape, that consist of high aspect ratio domains protruding from a conically-shaped base, similar to a golf tee [Figure 4-4(b)]. The similarity of the bases to the stout, wz-CoO nanocones shown in Figure 4-3, suggests that the conical portion forms at early stages of nanocrystal growth, and that the rod-shaped domains grow from the apex of each cone. The slight preferred orientation observed in

Figure 4-4(a) is consistent with the elongated shapes. The anisotropic morphology is likely caused by preferential growth of the {002} planes in wz-CoO, which increases the likelihood that the nanocrystals become oriented with {010} planes parallel to the XRD

11,12 sample stage. When a 4:1 molar ratio of Co(OL)2:Na(OL) was used, the average

146 distance from the basal facet to the tip of the “nanotees” was 35 ± 4 nm, the bases measured 18 ± 2 nm in diameter, and the nanorod domains had diameters 7.9 ± 0.7 nm.

Using a Co(OL)2:Na(OL) of 2, the nanorod portion of the particles was significantly longer, resulting in nanotees with an average length of 62 ± 8 nm [Figure 4-4(c)]. The differences in base diameter and nanorod diameter between the two samples was negligible.

Interestingly, the color change from blue to green occurred gradually, over the course of ~ 15 min, when a Co(OL)2:Na(OL) ratio of 4 was used. When Na(OL) was present in higher concentration [Co(OL)2:Na(OL) = 2], the color change happened much faster, in 3-5 minutes. It seems likely that the addition of Na(OL) plays a key role in the formation of reactive monomers that ultimately nucleate wz-CoO, and that the resulting monomer production is proportional to the amount of Na(OL) present. Thus, relatively high concentrations of Na(OL) allow the rod portions of the CoO nanostructures to grow longer. In those cases where higher concentrations of Na(OL) were used, however, a significant fraction of the CoO nanocrystals appeared only as short cones (or triangles), without a rod-shaped domain [Figure 4-4(c)]. Others in the same sample had multiple rods protruding from a single cone. The mechanistic role of Na(OL) in the generation of

Co-O monomers, which presumably occurs by a reaction with Co(OL)2, is currently under investigation. A fatty acid - conjugate base pair can react with one another via

2 ketonic decarbonylation, to form a symmetrically substituted ketone and CO2. It is possible that Na(OL) and excess OLAC react and are consumed in this way, promoting the thermolysis of Co(OL)2 by preventing the dissolution of wz-CoO by OLAC.

147 4.3.4 Bullet-shaped wz-CoO nanocrystals

Thermal decomposition of Co(OL)2 into wz-CoO was also achievable without any additives, when a two-stage heating process was used. Heating Co(OL)2 dissolved in

ODE at a constant rate (2 °C/min) to reflux in Ar atmosphere, as described earlier, did not reproducibly result in thermal decomposition. Under these conditions, however, we consistently observed significant quantities of white vapor being liberated from the solution (filling the vessel headspace) between 210-250 °C. This observation gave us the idea that an equilibrium concentration of those volatiles may be dissolved in the reaction mixture, and that a component of those dissolved gases may be preventing thermolysis of

Co(OL)2.

We designed a two-stage heating profile, wherein flowing Ar was applied in the temperature range 60-240 °C. The temperature was held at 240 °C under Ar flow for two hours, with the intention removing those volatiles from the reaction mixture by distillation. Indeed, we collected 0.5-1.0 mL of clear, colorless liquid distillate. The temperature of the reaction was then increased to reflux (310 °C) at a controlled rate, and the blue solution gradually changed to dark green, indicating thermal decomposition of

Co(OL)2. In Figure 4-5, TEM images of the product are presented, which consists of nanorods with average dimensions 49 ± 7 nm in length and 6.9 ± 0.7 in width. Many have an interesting bullet-like shape, consisting of two distinct domains that have slightly different diameters (the average diameter was calculated from measurements that were made at the length-bisector of each nanorod), and a pointed tip. The SAED pattern shown in Figure 4-5(c) is consistent with the crystal structure of wz-CoO.

148

Figure 4-5: (a) Large-area and (b) higher-magnification TEM images of wz-CoO nanocrystals with a bullet-like shape, along with a corresponding SAED pattern in (c). The scale bar in (a) is 100 nm and the scale bar in (b) is 20 nm.

For transition-metal carboxylates that contain coordinated water, such as the tetrahydrates of Co(II) acetate [Co(CH3COO)2ž4H2O] and Ni(II) acetate

[Ni(CH3COO)2ž4H2O], a key step in solid-state thermal decomposition is

61-63 dehydration. [Co(CH3COO)2ž4H2O] and [Ni(CH3COO)2ž4H2O] are isostructural, having two axial, unidentate, acetate ligands, and four, coordinated equatorial water

64,65 molecules. [Co(CH3COO)2ž4H2O] is pink, and turns blue upon loss of its water, like many hydrated Co(II) complexes. Co(OL)2 is purple, notably, and most likely represents a case of partial hydration.58 Interestingly, the FTIR spectra (Figure 4-6) of

149

Figure 4-6: FTIR absorbance spectra of [Co(CH3COO)2ž4H2O] and Co(Ol)2, which highlight the - - νsym(COO ) and νasym(COO ) stretches in each compound.

- - [Co(CH3COO)2ž4H2O] and Co(OL)2 are similar with respect to νsym(COO ), νasym(COO ),

-1 -1 -1 -1 -1 and Δ, [ΔCo(OL)2 = (1567 cm - 1415 cm ) = 152 cm , ΔCoOAc = (1554 cm - 1392 cm ) =

-1 162 cm ] suggesting that unidentate oleate ligand may be found in Co(OL)2. Thermal displacement of water from either [Co(CH3COO)2ž4H2O] or [Ni(CH3COO)2ž4H2O] is known to cause concomitant evolution of acetic acid by hydrolysis of surface acetate groups.62,63 At elevated temperatures, this process gives rise to dehydrated, basic metal acetate, M(CH3COO)2-x(OH)x, (M = Co, Ni) which has been identified as the key intermediate that decomposes into oxides. During decomposition of M(CH3COO)2- x(OH)x, acetone was also a major product, presumably formed via ketonic decarboxylation.

150

Figure 4-7: FTIR absorbance spectrum of a distillate (top), collected from Co(OL)2 solution in ODE over a heating period of 2 hours at 240 °C. The FTIR spectrum of ODE solvent (bottom) is included as a reference. Peaks marked with an asterisk (*), in the distillate spectrum, do not appear in the spectrum for ODE.

Assuming that the coordination sphere of Co(OL)2 resembles that of

[Co(CH3COO)2ž4H2O], we thought it was possible that Co(OL)2 must also be dehydrated for thermolysis to occur. At 210 °C, the vapors that evolve from Co(OL)2 in ODE may be related to thermal displacement of water molecules, and/or side-products of that process. We collected a substantial volume of distillate by flowing Ar over our reaction at 240 °C, but we could not unambiguously identify any new species in the liquid using

FTIR. The spectrum in Figure 4-7 is identical to that of the ODE solvent, with the exception of peaks at 1740, 1130 and 1075 cm-1. The signal at 1740 cm-1 corresponds well to the C=O stretch in esters, and the three bands in the 1000-1300 cm-1 range could be characteristic of ester C-O stretches.55 Further investigation is necessary to propose a pathway that results in ester formation.

151

Figure 4-8: FTIR absorbance spectra of Co(OL)2, before (bottom) and after (top) heating in the solid-state for 2 hours at 240 °C. The spectra highlight the disappearance of a peak 1705 cm-1, and the appearance of a new peak at 1725 cm-1. The spectral changes suggest that free carboxylic acid is consumed, and a ketone is generated, as a result of heating.

We also heated Co(OL)2 in the solid state to 240 °C, maintained that temperature for two hours, and analyzed the before-and-after FTIR spectra (Figure 4-8).

Disappearance of the C=O peak at 1710 cm-1, and emergence of a new peak at 1725 cm-1, indicates the consumption of free OLAC, and the formation of ketones. This result suggests that Co(OL)2, like many other metal carboxylates in the solid state, may decompose by pathways that involve ketonic decarboxylation. The thermal decomposition of Co(OL)2 in solution likely follows a different pathway; one that is influenced by the presence of solvent, volatile and soluble intermediates, the practical handling of atmosphere, and complex acid-base equilibria.

152 4.4 Conclusions

The solution-phase, thermal decomposition of transition metal carboxylates is a promising chemical transformation for producing scalable quantities of high-quality metal oxide nanocrystals. Unfortunately, subtle experimental details can profoundly impact the outcome of such reactions, including the choice of solvent, the presence of impurities, precursor preparation, and atmosphere handling. Moreover, metal carboxylate thermolysis is likely propagated by the formation of reactive radical species, which makes the pathways by which decomposition occurs difficult to analyze. Reports on this subject have been contradictory, and synthetic protocols are often irreproducible.

We were motivated in particular by the case of Co(OL)2 thermolysis into wz-CoO nanocrystals, which feature Co2+ in a metastable, tetrahedral geometry. First of all, the conventional, biphasic synthesis of Co(OL)2 produces a side-product, which we have proposed to be Co(OL)2-x(OH)x. We have also found that colloidal wz-CoO nanocrystals dissolve quickly in the presence of the weak acid, OLAC. The tendency of Co(OL)2 to decompose into wz-CoO in the presence of the basic Co(OL)2-x(OH)x impurity, but not from purified Co(OL)2 precursor - along with the tendency of protogenic OLAC to impede the thermolysis - led us to hypothesize that acid-base interactions may play a key role in the reaction. Indeed, basic additives such as TMAOH and Na(OL) allowed us to reproducibly synthesize wz-CoO nanocrystals with interesting shapes, including nano-

“cones” and “tees”. Simplified Co(OL)2 thermolysis reactions, without any additives, could also generate uniform, elongated wz-CoO nanocrystals. We achieved this result by

153 designing a two-stage heating process, whereby volatile by-products were removed via distillation, prior to reaching the decomposition temperature.

The molecular processes that lead to nucleation of colloidal inorganic nanostructures are often complex. Reproducible synthetic routes to industrial-scale quantities of morphology-controlled nanocrystals are highly desirable, however, and efforts to understand the mechanisms are ongoing. The thermal decomposition of metal carboxylate solutions is an especially simple case of transformation, requiring only one reactant, solvent, and heat, and is a worthwhile process to study in depth. Our studies of

Co(OL)2 decomposition have allowed us to make several insights which may be general, or representative of other carboxylate precursors that yield important nanocrystal targets.

4.5 References

1) Davis, R.; Schultz, H.P. J. Org. Chem. 1962, 27, 854.

2) Renz, M. Eur. J. Org. Chem. 2005, 6, 979.

3) Sola Akanni, M.; Okoh, E.K., Burrows, H.D.; Ellis, H.A. Thermochimica Acta.

1992, 208, 1.

4) Reed, R.I. J. Chem. Phys. 1953, 21, 377.

5) Rosenberg, A.S.; Stepanov, V.R. Russ. Chem. Bull. 1996, 45, 1336.

6) Hites, R.A.; Biemann, K. J. Am. Chem. Soc. 1972, 94, 5772.

7) Park, J.; An, K.; Hwang, Y.; Park, J-G.; Noh, H-J.; Kim, J-Y.; Park, J-H.; Hwang,

N-M.; Hyeon, T. Nature Mater. 2004, 3, 891.

8) Jana, N.R.; Chen, Y.; Peng, X. Chem. Mater. 2004, 16, 3931.

154 9) Choi, S-H.; Kim, E-G.; Park, J.; An, K.; Lee, N.; Kim, S.C.; Hyeon, T. J. Phys.

Chem. B 2005, 109, 14792.

10) Chen, Y.; Johnson, E.; Peng, X. J. Am. Chem. Soc. 2007, 129, 10937.

11) An, K.; Lee, N.; Park, J.; Kim, S.C.; Hwang, Y.; Park, J-G.; Kim, J-Y.; Park, J-H.;

Han, M.J.; Yu, J.; Hyeon, T. J. Am. Chem. Soc. 2006, 128, 9753.

12) Chen, Y.; Kim, M.; Lian, G.; Johnson, M.B.; Peng, X. J. Am. Chem. Soc. 2005,

127, 13331.

13) Puglisi, A.; Mondini, S.; Cenedese, S.; Ferretti, A.M.; Santo, N.; Ponti, A. Chem.

Mater. 2010, 22, 2805.

14) Bao, N.; Shen, L.; Wang, Y.; Padhan, P.; Gupta. A. J. Am. Chem. Soc. 2007, 129,

12374.

15) Bao, N.; Shen, L.; An, W.; Padhan, P.; Turner, H.; Gupta, A . Chem. Mater. 2009,

21, 3458.

16) Park, J.; Lee, E.; Hwnag, N-M.; Kang, M.; Kim, S.C.; Hwang, Y.; Park, J-G.;

Noh, H-J.; Kim, J-Y.; Park, J-H.; Hyeon, T. Angew. Chem. Int. Ed. 2005, 44,

2872.

17) Kovalenko, M.V.; Bodnarchuk, M.I.; Lechner, R.T.; Hesser, G.; Schaffler, F.;

Heiss, W. J. Am. Chem. Soc. 2007, 129, 6352.

18) Lee, J.H.; Huh, Y.M.; Jun, Y.; Seo, J.; Jang, J.; Song, H.T.; Kim, S.; Cho, E.J.;

Yoon, H.G.; Suh, J.S.; Cheon, J. Nat. Med. 2007, 13, 95.

19) Ho, D.; Sun, X.L.; Sun, S. Acc. Chem. Res. 2011, 44, 875.

20) Xie, J.; Liu, G.; Eden, H.S.; Ai, H.; Chen, X.Y. Acc. Chem. Res. 2011, 44, 883.

155 21) Han, X.G.; Jin, M.S.; Xie, S.F.; Kuang, Q.; Jiang, Z.Y.; Jiang, Y.Q.; Xie, Z.X.;

Zheng, L.S. Angew. Chem. Int. Ed. 2009, 48, 9180.

22) Yang, H.G.; Sun, C.H.; Qiao, S.Z.; Zou, J.; Liu, G.; Smith, S.C.; Cheng, H.M.;

Lu, G.Q.; Nature 2008, 453, 638.

23) Yang, H.G.; Liu, G.; Qiao, S.Z.; Sun, C.H.; Jin, Y.G.; Smith, S.C.; Zou, J.;Cheng,

H.M.; Lu, G.Q. J. Am. Chem. Soc. 2009, 131, 4078.

24) Huang, W.C.; Lyu, L.M.; Yang, Y.C.; Huang, M.H. J. Am. Chem. Soc. 2012, 134,

1261.

25) Poizot, P.; Laruelle, S.; Grugeon, S.; Dupont, L.; Tarascon J-M. Nature 2000,

407, 496.

26) O’Regan, B.; Grätzel, M. Nature 1991, 353, 737.

27) Chen, X.; Mao, S.S. Chem. Rev. 2007, 107, 2891.

28) Bodnarchuk, M.I., Kovalenko, M.V.; Groiss, H.; Resel, R.; Reissner; M.; Hesser,

G.; Lechner, R.T.; Steiner, W.; Schaffler, F.; Heiss, W. Small 2009, 5, 2247.

29) Shavel, A.; Rodriguez-Gonzales, B.; Spasova, M.; Farle, M.; Liz-Marzan, L.M.

Adv. Funct. Mater. 2007, 17, 3870.

30) Pinchon, B.P. et al. Chem. Mater. 2011, 23, 2886.

31) Bronstein, L.M.; Huang, X.; Retrum, J.; Schmucker, A.; Pink, M.; Stein, B.D.;

Dragnea, B. Chem. Mater. 2007, 19, 3624.

32) Kwon, S.G.; Piao, Y.; Park, J.; Angappane, S.; Jo, Y.; Hwang, N-M.; Park, J-G.;

Hyeon, T. J. Am. Chem. Soc. 2007, 129, 12571.

33) Joo, J.; Kwon, S.G.; Yu, J.H.; Hyeon, T. Adv. Mater. 2005, 17, 1873.

156 34) Cao, A-M.; Hu, J-S.; Liang, H-P.; Song, W-G.; Wan, L-J.; He, X-L.; Gao, X-G.;

Xia, S-H. J. Phys. Chem. B 2006, 110, 15858.

35) Choia, K. Kima, H.R.; Kim, K.M.; Liu, D.; Cao, G.Z.; Lee, J.H. Sens. Actuators,

B 2010, 146, 183.

36) Sun, C.; Su, X.T.; Xiao, F.; Niu, C.; Wang, J. Sens. Actuators, B 2011, 157, 681.

37) Jiao, Q.; Fu, M.; You, C.; Zhao, Y.; Li, H. Inorg. Chem. 2012, 51, 11513.

38) Hu, L.; Peng, Q.; Li, Y. J. Am. Chem. Soc. 2008, 130, 16136.

39) Koza, J.A.; He, Z.; Miller, A.S.; Switzer, J.A. Chem. Mater. 2012, 24, 3567.

40) Xie, X.; Li, Y.; Liu, Z-Q.; Haruta, M.; Shen, W. Nature 2009, 458, 746.

41) Skumryev, V.; Stoyanov, S.; Zhang, Y.; Hadjipanayis, G.; Givord, D.; Nogues, J.

Nature 2003, 423, 850.

42) Redman, M. J.; Steward, E. G. Nature 1962, 193, 867.

43) Risbud, A.S.; Snedeker, L.P.; Elcombe, M.M.; Cheetham, A.K.; Seshadri, R.

Chem. Mater. 2005, 17, 834.

44) Grimes, R. W.; Lagerlof, K. P. D. J. Am. Ceram. Soc. 1991, 74, 270.

45) Nam, K.M.; Shim, J.H.; Han, D-W.; Kwon, H.S.; Kang, Y-M.; Li, Y.; Song, H.;

Seo, W.S.; Park, J.T. Chem. Mater. 2010, 22, 4446.

46) Ueda, K.; Tabata, H.; Kawai, T. Appl. Phys. Lett. 2001, 79, 988.

47) Seo, W. S.; Shim, J. H.; Oh, S. J.; Lee, E. K.; Hur, N. H.; Park, J. T. J. Am. Chem.

Soc. 2005, 127, 6188.

48) White, M.A.; Ochsenbein, S.T.; Gamelin, D.R. Chem. Mater. 2008, 20, 7107.

49) Nam, K.M.; Choi, Y.C.; Jung, S.C.; Kim, Y-I.; Jo, M.R.; Park, S.H.; Kang, Y-M.;

Han, Y-K.; Park, J.T. Nanoscale 2012, 4, 473.

157 50) Sines, I.T.; Misra, R.; Schiffer, P, Schaak, R.E. Angew. Chem. Int. Ed. 2010, 49,

4638.

51) Norako, M.E.; Greaney, M.J.; Brutchey, R.L. J. Am. Chem. Soc. 2012, 134, 23.

52) Singh, A.; Geaney, H.; Laffir, F.; Ryan, K.M. J. Am. Chem. Soc. 2012, 134, 2910.

53) Soriano, R.B.; Malliakas, C.D.; Wu, J.; Kanatzidis, M.G. J. Am. Chem. Soc.

2012, 134, 3228.

54) Soriano, R.B.; Arachchige, I.U., Malliakas, C.D.; Wu, J.; Kanatzidis, M.G. J.

Am. Chem. Soc. 2013, 135, 738.

55) Nakamoto, K. Infrared and Raman Spectra of Inorganic and Coordination

Compounds, Part B: Applications in Coordination, Organometallic, and

Bioinorganic Chemistry. John Wiley and Sons, Inc: Hoboken, 2009.

56) Deacon, G.B.; Phillips, R.J. Coord. Chem. Rev. 1980, 33, 227.

57) Kambe, H.; Mita, I. Bull. Chem. Soc. Japan 1961, 34, 1797.

58) Kambe, H.; Mita, I. Bull. Chem. Soc. Japan 1961, 34, 1790.

59) Shieh, S.R.; Duffy, T.S. Phys. Rev. B. 2002, 66, 134301.

60) Wang, X.; Andrews, L. J. Phys. Chem. A 2006, 110, 10035.

61) Grimes, W.R.; Fitch, A.N. J. Mater. Chem. 1991, 1, 461.

62) Wanjun, T.; Donghua, C. Chem. Pap. 2007, 61, 329.

63) De Jesus, J.C.; Gonzalez, I.; Quevedo, A.; Puerta, T. J. Mol. Catal A: Chem.

2005, 228, 283.

64) van Niekirk, J.N.; Schoening, F.R.L. Acta Cryst. 1953, 6, 609.

65) Nicolai, B.; Kearley, G.J., Johnson, M.R.; Fillaux, F.; Suard, E. J. Chem. Phys.

1998, 109, 9062.

Chapter 5

Engineering Solid-State Complexity into Segmented Metal Nanowires Using Site-Selective Solution Chemistry

5.1 Introduction

One-dimensional (1D) heterostructured nanowires are used extensively as platforms for studying physical phenomena at nano- and micrometer length scales, including locomotion,1-3 self- and directed assembly,4-7 photoconductivity,8 magneto- optics,9 and interfacial electron transport.10 Nanowires with chemically distinct segments that are adjoined in a linear array can be conveniently synthesized by sequentially electrodepositing different materials into a rigid, porous template, such as an anodic aluminum oxide (AAO) membrane. Template-confined electrodeposition is amenable to a wide range of metals, semiconductors, and conducting polymers, and typically yields nanowires with uniform widths and controllable lengths.11-13 In general, 1D nanostructures are promising active components and interconnects for future electronic devices, because of their anisotropic shape. Accordingly, diverse template-grown nanowire-based sensors,14 molecular electronics,15,16 biological detection17-19 and separation systems20, and photovoltaics21 have been fabricated.

The compositional complexity of the individual domains in segmented nanowires is limited by the efficacy of electrochemical co-deposition methods.22-24 Simultaneous

Adapted with permission from J. Am. Chem. Soc. 130, 14042 “On-Wire Conversion Chemistry: Engineering Solid-State Complexity into Striped Metal Nanowires Using Solution Chemistry Reactions” Copyright 2008 American Chemical Society.

159 deposition of multiple elements (that have different reduction potentials) often requires system-specific optimization to obtain stoichiometry-controlled segments, which is largely based on empirical trial and error. Chapter 1 (Section 1.4) described a synthetic strategy that our group and others have been developing, which treats pre-formed nanoparticles as “synthons”, and enables their chemical transformation into derivative compounds. Metal nanoparticles, in particular, can be converted into alloys and intermetallic phases via metallurgical, diffusion-mediated reactions.25-32 Here we demonstrate that the same diffusion reactions can be applied to AAO template-grown metal nanowires. Single-metal nanowire synthons are thereby transformed into segmented nanowires with length-controlled, multi-metal domains, which represents a robust alternative to electrochemical co-deposition.

We also show that the backside coating of Ag on the AAO membrane, which is inherent to the electrodeposition process, can be leveraged as a protecting group for one end of the nanowires. One terminus of each template-confined nanowire is physically protected, which has allowed us to demonstrate site-selective transformation of the solution-exposed end. Moreover, the sidewalls of membrane-bound nanorods are also protected from the reaction mixture, which ensures that the transformation is initiated at the nanowire tips. As such, this work is an important contribution to the small number of reports that apply the strategies commonly used to synthesize complex molecular targets, to the construction of multi-domain inorganic nanostructures - in this case, the use of a protecting group.33-36

160 5.2 Experimental Details

5.2.1 Nanowire electrodeposition and on-wire conversion

Pt nanowires were electrodeposited following standard methods published elsewhere.11-13 Anodic aluminum oxide membranes (Anodisc 25, Whatman Scientific,

0.2 µm) were used as the template. A Pt counter-electrode was used and the working electrode was a thin Ag film (Ag backing) on the membrane, which is deposited thermally and electrochemically (Ag Cyless R RTU plating solution, Technic, 1.65 mA).

The platinum nanowires are electrodeposited from Pt TP RTU plating solution (Technic) at constant current of 5 mA with a growth rate of ~2 µm/hr. The silver backing can be removed by exposure to 8 M nitric acid either before or after the conversion reaction, depending on the desired site-specific modification.

To generate PtPb intermetallic segments, lead acetate trihydrate

(Pb(CH3COO)2·3H2O, Alfa Aesar 99.999%) was dissolved in tetraethylene glycol (TEG,

Alfa Aesar 99%) and the nanowire-embedded membrane was immersed in the solution.

The reaction was heated to reduce Pb(CH3COO)2·3H2O, and to induce diffusion of Pb into the Pt nanowire. Precursor concentrations ranged from 10-100 mM, and reaction temperatures were generally between 200ºC and 300ºC. Typically, solution volumes were 10 mL, and one-sixth of a membrane was used for each reaction. The reaction conditions and PtPb segment dimensions are as follows: (a) Short PtPb tips (~0.4 µm,

Figure 3a) on Pt nanowires were fabricated using 10 mM Pb(CH3COO)2·3H2O at 220 ºC for 1 h. (b) Longer PtPb segments (> 1 µm, Figure 2) on Pt nanowires were prepared using 30 mM Pb(CH3COO)2·3H2O at 300 ºC for 1 hr. (c) Pt nanowires were entirely

161 converted to PtPb (Figure 3b) while confined within the AAO template, using 60 mM

Pb(CH3COO)2·3H2O at 300 ºC for 1 h. (d) Released Pt nanowires were fully converted to

PtPb nanowires (Figure 5e) using 60 mM Pb(CH3COO)2·3H2O at 190 ºC for 1 hr. (e) A single PtPb segment (~1.1 µm, Figure 3c) was fabricated on the exposed terminus of a Pt nanowire (while the Ag backing was maintained on the membrane) using 60 mM

Pb(CH3COO)2·3H2O at 190 ºC for 1 hr. After the chemical reactions, the AAO membrane was dissolved in 3M NaOH. The resulting nanowires were collected via centrifugation, washed thoroughly with water, and dispersed in ethanol.

5.2.2 Characterization

Powder X-ray diffraction (XRD) data was collected on a Bruker D8 Advance X- ray diffractometer using Cu Kα radiation. Scanning electron microscopy (SEM), energy dispersive X-ray spectroscopy (EDS) and elemental mapping were performed on a FEI

Quanta 200 environmental SEM operating in high vacuum mode, and on a JEOL JSM

5400 SEM.

5.3 Results and Discussion

A cartoon schematic is shown in Figure 5-1, which shows the process for transforming single-component metal nanowires into multi-segmented nanowires with multi-metal components. Metal nanowires are first grown by galvanostatic electrodeposition into Ag-backed AAO membranes. Nanowire-containing membrane

162

Figure 5-1: Metal nanowires in AAO membrane reacted either before (a-c) or after (d-f) Ag backing removed. Nanowires undergo site-specific modifications (b,e) and are then released from the membrane (c,f).

fragments are then immersed into a tetraethylene glycol (TEG) metal salt solution for conversion. With the Ag backing intact, one end of the nanowire is physically protected from the metal salt solution, which enables selective conversion of the unprotected end of the nanowire. With the Ag backing removed, both tips of the nanowires react. Upon heating, the metal salt reduces and diffuses into the exposed ends of the nanowires. The

AAO membrane can then be dissolved to release the nanowires. For proof-of-concept purposes, we focused initially on introducing intermetallic PtPb segments into Pt nanowires, since reaction conditions that promote solid-solid diffusion are established for this system. Pt-based intermetallics, including PtPb, also exhibit a range of properties that may be useful to incorporate into striped metal nanowires, including catalytic activity28,37 superconductivity38 and ferromagnetism.39

Figure 5-2 shows a SEM image, along with EDS element mapping data, for a segmented PtPb/Pt/PtPb nanowire formed by heating an AAO membrane (Ag backing removed) containing Pt nanowires with dimensions 300 nm x 4 µm, in a 30 mM solution of Pb(CH3COO)2·3H2O at 300 °C for 1 h. EDS-generated element mapping clearly

163

Figure 5-2: (a) SEM micrograph with EDS element maps for (b) Pt, (c) Pb, and (d) SEM image with Pt and Pb overlay. (e) XRD data for as-grown Pt nanowires in the AAO membrane, and released PtPb/Pt/PtPb nanowires after conversion. Inset: EDS spectrum for PtPb/Pt/PtPb nanowires. shows the presence of Pt throughout the entire nanowire, with a higher concentration in the center [Figure 5-2(b)]. In contrast, Pb is detected primarily at the ends of the nanowire [Figure 5-2(c)]. XRD data shown in Figure 5-2(e) confirm that the nanowire synthon consists only of Pt, and that the chemically converted products contain a mixture of Pt and NiAs-type PtPb, along with a small impurity of a Pt3Pb alloy phase. It is worth noting that the Pt XRD peaks in the PtPb/Pt/PtPb nanowire are sharper than in the Pt nanowire template, implying that the crystallinity of the nanowires was improved during the solution-phase annealing. EDS spectra taken on an ensemble of nanowires show that both Pt and Pb are present exclusively [inset, Figure 5-2(e)], with more Pt than Pb as expected. Ensemble XRD and EDS data, coupled with single-wire imaging, provide good evidence that the nanowires consist of a Pt center with intermetallic PtPb tips

(rather than a compositionally variable Pt1-xPbx alloy).

164

Figure 5-3: EDS-generated element maps for (a) PtPb/Pt/PtPb nanowires with short tips, and (b) fully-converted PtPb nanowires, which were both prepared by conversion of template-confined Pt nanowires. c) Anisotropic PtPb/Pt nanowires were made in-membrane, with the Ag backing intact. All scale bars are 2 µm.

The reaction temperature and precursor concentration affect the diffusion- mediated reactions, and provide control over segment lengths. Compared to the nanowires in Figure 5-2, it is possible to fabricate Pt nanowires with shorter PtPb tips, for example, by lowering either the reaction temperature, the Pb(C2H3O2)2·3H2O concentration, or both. Such mild conditions were used to fabricate the 1D PtPb/Pt/PtPb heterostructures in Figure 5-3(a), where template-confined Pt nanowires were reacted in a

10 mM Pb(C2H3O2)2·3H2O solution at 220 °C for 1 hr. In contrast, the entire Pt nanowire was converted to PtPb at higher temperatures and Pb(C2H3O2)2·3H2O concentrations [Figure 5-3(b); 60 mM, 300 °C, 1 hr]. When the Ag backing is left on during the conversion reaction, one terminus is physically protected from the reaction mixture, and Pb diffuses site-selectively into the exposed end of each Pt nanowire. As a result, anisotropic PtPb/Pt two-segment nanowires are formed [Figure 5-3(c)]. The Pt nanowires can also be transformed either before or after release from the membrane. The former produces segmented PtPb/Pt/PtPb nanowires that are uniform in diameter [Figure

5-4(a-c)], with surface roughness close to that of AAO membrane [Figure 5-4(d)]. The

165 nanowire sidewalls are protected by the membrane, which ensures that diffusion proceeds inward from the exposed tips. In contrast, conversion of released Pt nanowires generates structures with significantly rougher-looking surfaces [Figure 5-4(e)].

Figure 5-4: A series of SEM images at increasing magnifications for PtPb / Pt / PtPb nanowires that were prepared by reacting template-confined Pt nanowires in (a) 30 mM Pb2+ at 300 ºC, and (b,c) 10 mM Pb2+ at 220 ºC. These large field images demonstrate the sample-wide uniformity of nanowires that were transformed in-membrane. SEM images of fully-converted PtPb nanowires reacted either (d) in the membrane, or (e) after release. The smooth morphology of the wire in (d) demonstrates the effectiveness of the membrane for protecting the sidewalls of the nanowires, which ensures that the shape and morphology of the nanowire are conserved during the chemical conversion process.

5.4 Conclusions

Metallurgical, diffusion-mediated, solution reactions have been used to transform template-grown metal nanowire synthons into 1D heterostructures with multi-metal segments. Site-specific reactions are observed when the electrodeposited nanowires are converted in-membrane, which leverages the AAO template and the Ag backing as protecting groups, and the reaction parameters permit control over segment length and

166 composition. PtPb intermetallic compounds were chosen as initial targets, but similar reactions convert metal nanoparticles to oxides,32 chalcogenides,31,32 and phosphides.29,30

Related reactions such as galvanic replacement40 and ion exchange41,42 also work under similar conditions. This work represents a convenient toolkit that can be developed further for engineering complex solid-state components into 1D nanowire heterostructures, which is an important contribution to the burgeoning total synthesis framework for synthesizing multi-domain inorganic nanostructures.

5.5 References

1) Paxton, W.F.; Kistler, K.C.; Olmeda, C.C.; Sen, A.; St. Angelo, S.K.; Cao, Y.Y.;

Mallouk, T.E.; Lammert, P.E.; Crespi, V.H. J. Am. Chem. Soc. 2004, 126, 13424.

2) Kline, T.R.; Paxton, W.F.; Mallouk, T.E.; Sen, A. Angew. Chem. Int. Ed. 2005,

44, 744.

3) Wang, W.; Castro, L.A.; Hoyos, M.; Mallouk, T.E. ACS Nano 2012, 6, 6122.

4) Salem, A.K.; Chao, J.; Leong, K.W.; Searson, P.C. Adv. Mater. 2004, 16, 268.

5) Love, J.C.; Urbach, A.R.; Prentiss, M.G.; Whitesides, G.M. J. Am. Chem. Soc.

2003, 125, 12696.

6) Smith, B.D.; Kirby, D.J.; Keating, C.D. Small 2011, 7, 781.

7) Smith, B.D.; Mayer, T.S.; Keating, C.D. Annu. Rev. Phys. Chem. 2012, 63, 241.

8) Pena, D.J.; Mbindyo, J.K.N.; Carado, A.J.; Mallouk, T.E.; Keating, C.D.; Razavi,

B.; Mayer, T.S. J. Phys. Chem. B. 2002, 106, 7458.

9) Zhang, Y.; DaSilva, M.; Ashall, B.; Doyle, G.; Zerulla, D.; Sands, T.D.; Lee,

167 G.U. Langmuir 2011, 27, 15292.

10) Agarwal, R. Small 2008, 4, 1872.

11) A. J. Mieszawska, R. Jalilian, G. U. Sumanasekera, F. P. Zamborini, Small 2007,

3, 722.

12) T. R. Kline, M. L. Tian, J. G. Wang, A. Sen, M. W. H. Chan, T. E. Mallouk,

Inorg. Chem. 2006, 45, 7555.

13) Hust, S.J.; Payne, E.K.; Qin, L.; Mirkin, C.A. Angew. Chem. Int. Ed. 2006, 45,

2672.

14) Li, M.; Bhiladvala, R.B.; Morrow, T.J.; Sioss, J.A. Lew, K-K.; Redwing, J.M.;

Keating, C.D.; Mayer, T.S. Nature Nanotech. 2008, 3, 88.

15) Chen, X.; Yeganeh, S.; Qin, L.; Li, S.; Xue, C.; Braunschweig, A.B.; Schatz,

G.C.; Ratner, M.A.; Mirkin, C.A. Nano Lett. 2009, 9, 3974.

16) Chen, X.; Jeon, Y-M.; Jang, J-W.; Qin, L.; Huo, F.; Wei, W.; Mirkin, C.M. J. Am.

Chem. Soc. 2008, 130, 8166.

17) Nicewarner-Pena, S.R.; Freeman, R.G.; Reiss, B.D.; He, L.; Pena, D.J.; Walton,

I.D.; Cromer, R.; Keating, C.D.; Natan, M.J. Science 2001, 294, 137.

18) Brunker, S.E.; Cederquist, K.B.; Keating, C.D. Nanomedicine-UK 2007, 2, 695.

19) Stoermer, R.L.; Cederquist, K.B.; McFarland, S.K.; Sha, M.Y.; Penn, S.G.;

Keating, C.D. J. Am. Chem. Soc. 2006, 128, 16892.

20) Oh, B-K.; Park, S.; Millstone, J.E.; Lee, S.W.; Lee, K-B.; Mirkin, C.A. J. Am.

Chem. Soc. 2006, 128, 11825.

21) Hernandez-Pagan, E.A.; Wang, W.; Mallouk, T.E. ACS Nano 2011, 5, 3237.

22) Huang, H.Y.; Okumura, H.; Hadjipanayis, G.C.; Weller, D. J. Appl. Phys. 2002,

168 91, 6869.

23) Singh, K.V.; Martinez-Morales, A.A.; Avan, G.T.S.; Bozhilov, K.N.; Ozkan, M.

Chem. Mater. 2007, 19, 2446.

24) Mosby, J.M.A.; Prieto, A.L. J. Am. Chem. Soc. 2008, 130, 10656.

25) Vasquez, Y. Henkes, A.E.; Bauer, J.C.; Schaak, R.E. J. Solid State Chem. 2008,

181, 1509.

26) Chou, N.H.; Schaak, R.E.; J. Am. Chem. Soc.2007, 129, 7339.

27) Cable, R.E.; Schaak, R.E.; Chem. Mater. 2007, 19, 4098.

28) Bauer, J.C.; Chen, X.; Liu, Q.; Phan, T-H.; Schaak, R.E. J. Mater. Chem. 2008,

18, 275.

29) Henkes, A.E.; Vasquez, Y.; Schaak, R.E. J. Am. Chem. Soc. 2007, 129, 1896.

30) Henkes, A.E.; Schaak, R.E. Inorg. Chem. 2008, 47, 671.

31) Jeong, U.; Camargo, P.H.C.; Lee, Y.H.; Xia, Y. J. Mater. Chem. 2006, 16, 3893.

32) Yin, Y.; Rioux, R.M.; Erdonmez, C.K.; Hughes, S.; Somorjai, G.A.; Alivisatos,

A.P. Science 2004, 304, 711.

33) Leonard, B.M.; Anderson, M.E.; Oyler, K.D.; Phan, T.H.; Schaak, R.E. ACS

Nano 2009, 3, 940.

34) Bao, Z.H. Sun, Z.H., Xiao, M.D.; Chen, H.J.; Tian, L.W.; Wang, J.F. J. Mater.

Chem. 2011, 21, 11537.

35) Ridelman, Y.; Singh, G.; Popovitz-Biro, R.; Wolf, S.G.; Das, S.; Klajn, R. Small

2012, 8, 654.

36) Sahoo, J.K.; Tahir, M.N.; Hoshyargar, F.; Nakhjavan, B.; Branscheid, R.; Kolb,

U.; Tremel, W. Angew. Chem. Int. Ed. 2011, 50, 12271.

169 37) Alden, L.R.; Roychowdhury, C.; Matsumoto, F.; Han, D.K.; Zeldovich, V.B.;

Abruna, H.D.; Disalvo, F.J. Langmuir 2006, 22, 10465.

38) Matthias, B.T. Phys. Rev. B. 1953, 92, 874.

39) Weller, D.; Doerner, M.F. Annu. Rev. Mater. Sci. 2000, 30, 611.

40) Sun, Y.; Xia, Y. J. Am. Chem. Soc. 2004, 126, 3892.

41) Son, D.H.; Hughes, S.M.; Yin, Y.; Alivisatos, A.P. Science 2004, 306, 1009.

42) Camargo, P.H.C.; Lee, Y.H.; Jeong, U.; Zou, Z.; Xia, Y. Langmuir 2007, 23,

2985.

Chapter 6

A Total Synthesis Framework for the Construction of High-Order Colloidal Hybrid Nanoparticles

6.1 Introduction

Colloidal hybrid nanoparticles1-3 are emerging as an important family of multi- functional nanoscale materials that facilitate diverse applications in solar energy conversion,4,5 cancer therapy,6 fuel cell catalysis,7,8 biological imaging,9,10 electronics,11 and magnetism.12 These unique hybrid lab-on-a-particle systems are characterized by solid-state interfaces,13 which support direct electronic and magnetic communication between components that can lead to synergistic effects not observed for physical mixtures or nanostructures assembled using molecular or biological linkers. Constructing these hybrid nanoparticles requires pre-made particles to fuse together or, more commonly, one particle to nucleate directly on the surface of another.14,15 More sophisticated architectures are needed as the applications of these materials continue to expand, and the growing disconnect between design and synthesis becomes a critical limitation. Here, we demonstrate a rational and stepwise approach for constructing hybrid nanoparticle oligomers that conceptually parallels the total synthesis framework used by chemists to construct large organic molecules.

Chemists approach the synthesis of complex organic molecules by identifying simple molecules that can be rationally modified or linked together by sequentially Adapted with permission from Nature Chem. 4, 37 “A total-synthesis framework for the construction of high-order colloidal hybrid nanoparticles.” Copyright 2011 Macmillan Publishers Limited. 171 applying a set of known chemical reactions, guided by mechanistic considerations.16

Reactions are often carried out chemoselectively and orthogonally, occurring at one pre- determined location on a molecule that may have multiple reaction sites. When applied in a stepwise manner, large and complex molecules can be constructed. Colloidal hybrid nanoparticles have been described as solid-state analogues of molecules: the constituent nanocrystals behave as ‘artificial atoms’ that are linked together to form “artificial molecules”.17 A total synthesis framework for these artificial molecules would merge strategies from molecular organic synthesis with needs that underpin the advancement of complex colloidal nanostructures, providing the tools to deterministically construct target hybrid nanoparticles in a rational and stepwise manner.

6.2 Experimental Details

6.2.1 Materials

Pentacarbonyl(0)iron [Fe(CO)5], silver(I) acetate [Ag(OAc), anhydrous], copper(II) chloride [CuCl2, anhydrous], lead(II) oxide [PbO], hydrogen tetrachloroaurate(III) trihydrate [HAuCl4·3H2O], platinum(II) 2,4-pentanedionate

[Pt(acac)2], nickel(II)2,4- pentanedionate [Ni(acac)2], sulfur powder [-325 mesh] and oleic acid [OLAC, tech, 90%] were purchased from Alfa Aesar. Copper(I) acetate

[Cu(OAc)], palladium(II) 2,4-pentanedionate [Pd(acac)2], 1-octadecene [tech, 90%], trioctylamine [98%], and diphenyl ether [≥99%] were purchased from Sigma-Aldrich.

Oleylamine [OLAM, C18 content 80-90%] was purchased from Acros Organics. Sodium

172 oleate was purchased from TCI Chemicals. All chemicals were used as-received without further purification.

6.2.2 Synthesis of Pt nanoparticles19

100 mg Pt(acac)2, 1 mL OLAC, and 1 mL OLAM were mixed with 10 mL 1- octadecene in a 50 mL 3-neck round-bottom flask equipped with a thermometer adapter, reflux condenser, flow-control adapter, and rubber septum. The mixture was heated under a flow of Ar to 90-100 °C to obtain a clear, yellow solution. The solution was heated under a blanket of Ar to 190 °C, and 0.1 mL Fe(CO)5 dissolved in 1 mL hexanes was swiftly injected to the flask, immediately turning the solution black. The colloidal solution was further heated to 200 °C, left at that temperature for 1 h, and then allowed to cool to room temperature. Pt nanoparticles were precipitated from the product mixture by adding 2-propanol, followed by centrifugation (5,000-7,000 rpm, 5 min). The nanoparticles were redispersed in hexanes solution containing OLAC and OLAM (a typical stock solution was prepared by mixing 0.2 mL OLAC and 0.2 mL OLAM in 20 mL hexanes) to ensure colloidal stability/processablility. Ethanol was added thereafter to precipitate the nanocrystals, followed by centrifugation and redispersion in hexanes. The washing process was repeated once, and the particles were stored in 4 mL hexanes for use as seeds.

173

20 6.2.3 Synthesis of Pt-Fe3O4 heterodimers

1 mL OLAC was mixed with 20 mL 1-octadecene in a 100 mL 3-neck round- bottom flask equipped with a thermometer adapter, reflux condenser, flow-control adapter, and rubber septum. The mixture was heated under a flow of Ar for 30 min at 120

°C to remove air and moisture. Under an Ar blanket, 0.14 mL Fe(CO)5 was injected at

120 °C. After 5-10 minutes, 1 mL OLAM was added, followed by ~20 mg Pt nanoparticles (2 mL hexane dispersion). The mixture was heated slowly (5-7 °C/min) to

300 °C and left at that temperature for 30 min. To further oxidize any oxygen-deficient

Fe species to Fe3O4, the product mixture was heated to ~80 °C while bubbling air through the black/brown colloidal solution overnight. The mixture was subsequently heated to

~200 °C for an additional 30 min and allowed to cool to room temperature. Pt-Fe3O4 heterodimers were precipitated from the product mixture by adding 2-propanol, followed by centrifugation. The nanoparticles were redispersed in hexanes solution containing

OLAC and OLAM, and centrifuged to remove any insoluble material. Ethanol was added thereafter to precipitate the nanocrystals, followed by centrifugation and redispersion in hexanes. The washing process was repeated once, and the particles were stored in 4-5 mL hexanes.

6.2.4 Synthesis of Au-Pt-Fe3O4 heterotrimers

25 mg HAuCl4·3H2O was dissolved in 5 mL 1-octadecene containing 0.5 mL

OLAM in a 25 mL 3-neck round-bottom flask equipped with a thermometer adapter, reflux condenser, and rubber septum. The red-orange solution was heated under a flow

174 of Ar to 60 °C, and 10-12 mg Pt-Fe3O4 heterodimers (1 mL hexanes dispersion) were added via syringe. The temperature was held at 60 °C for 4 h, yielding a dark purple/black colloidal solution. Similar results were obtained by heating 10 mg

HAuCl4·3H2O and 0.2 mL OLAM in 5 mL 1-octadecene at 90 °C for 2 hours. Au-Pt-

Fe3O4 heterotrimers were precipitated from the product mixture by adding 2-propanol, followed by centrifugation. The nanoparticles were redispersed in hexanes solution containing OLAC and OLAM (a typical stock solution was prepared by mixing 0.2 mL

OLAC and 0.2 mL OLAM in 40 mL hexanes) to ensure colloidal stability/processablility.

Ethanol was added thereafter to precipitate the nanocrystals, followed by centrifugation and redispersion in hexanes. The washing process was repeated once, and the particles were stored in 2 mL hexanes for use as seeds.

The high yield and uniformity of our Au-Pt-Fe3O4 product can be ascribed to optimization of the seed-to-precursor ratio and reduction kinetics. While holding the absolute concentrations of HAuCl4 and oleylamine constant, we have found that lower- than-optimal Pt-Fe3O4 seed concentrations give rise to an undesirable number of freestanding Au particles in the final product. Higher seed concentrations result in large amounts of unreacted Pt-Fe3O4 heterodimers and significant amounts of insoluble precipitate. Heterogeneous nucleation of Au on Pt-Fe3O4 also seems to be favored by faster reduction of HAuCl4 by oleylamine. Upon decreasing the absolute concentrations of HAuCl4 and oleylamine or lowering the reaction temperature, which should slow down reduction, we observed many free Au particles and unreacted seeds. This suggests that slower reduction of HAuCl4 results in larger, stable Au nuclei that may not benefit energetically by growing from the Pt surface.

175

6.2.5 Synthesis of Ag-Pt-Fe3O4 heterotrimers

12.5 mg Ag(OAc) was dissolved in 10 mL toluene containing 0.740 mL OLAM in a 50 mL 3-neck round-bottom flask, and 10-12 mg Pt-Fe3O4 heterodimers were added at room temperature. The mixture was heated to 60 °C under an Ar blanket and held at this temperature overnight (14-16 hours), yielding a yellow-brown colloidal solution.

The particles were precipitated from the cooled product mixture by adding ethanol, centrifuged and redispersed in hexanes solution containing OLAC and OLAM. This precipitation/centrifugation process was repeated twice, and the particles were stored in hexanes. An identical process was used to prepare Ag-Fe3O4 and Ag-Pt heterodimers, using Fe3O4 and Pt seeds, respectively. Fe3O4 nanoparticles were prepared according to a published procedure.54

6.2.6 Synthesis of Ni-Pt-Fe3O4 heterotrimers

20 mg Ni(acac)2 was dissolved in a mixture of 2 mL OLAM and 8 mL trioctylamine in a 50 mL 4-neck round-bottom flask equipped with a thermometer adapter, reflux condenser, flow control adapter and rubber septum. The solution was heated under vacuum at 120 °C for 30 minutes to remove air and moisture. 10-12 mg Pt-

Fe3O4 heterodimers were injected under flowing Ar, and the hexanes was allowed to distill out of the solution. The mixture was heated slowly (~3°C/min) under a blanket of

Ar to 210 °C, and held at this temperature for 1 hour. The particles were precipitated from the cooled product mixture by adding ethanol, centrifuged and redispersed in

176 hexanes solution containing OLAC and OLAM. This precipitation/centrifugation process was repeated twice, and the particles were stored in hexanes.

6.2.7 Synthesis of Pd-Pt-Fe3O4 heterotrimers

30 mg Pd(acac)2 was dissolved in 5 mL 1-octadecene containing 1.0 mL OLAM and 0.25 mL OLAC in a 25 mL 4-neck round-bottom flask, and the solution was heated under vacuum at 60 °C for 30 minutes. 10-12 mg Pt-Fe3O4 heterodimers were injected under flowing Ar, and the hexanes was allowed to distill out of the solution. The mixture was heated under a blanket of Ar to 120 °C, and held at this temperature overnight. The particles were isolated from the cooled product mixture and purified using the same procedure as that for Au-Pt-Fe3O4 heterodimers, and stored in hexanes.

6.2.8 Synthesis of Cu9S5-Au-Pt-Fe3O4 heterotetramers

11.1 mg Cu(OAc) and 0.070 mL OLAC (or 62.5 mg Cu(II) oleate, made using a published procedure55) were added to 5 mL trioctylamine in a 25 mL 4-neck round- bottom flask. The mixture was heated under vacuum to 80 °C for 30 minutes, producing a clear, dark green solution. Au-Pt-Fe3O4 heterotrimers (1 mL hexanes dispersion) were added under flowing Ar, and the hexanes were allowed to distill out of the solution.

Under an Ar blanket, 0.5 mL of a 0.1 M sulfur solution (prepared by dissolving 8 mg sulfur powder in 2 mL diphenyl ether and 0.5 mL OLAM) was injected at 80 °C, the temperature was increased to 120 °C, and the reaction was allowed to proceed for 1 h.

177 The particles were precipitated from the cooled product mixture by adding ethanol, centrifuged and redispersed in hexanes solution containing OLAC and OLAM. This precipitation/centrifugation process was repeated twice, and the particles were stored in hexanes. Cu9S5-Au heterodimers were prepared using an identical process, using 10-12 mg Au nanocrystals as seeds.

6.2.9 Synthesis of PbS-Au-Pt-Fe3O4 heterotetramers

22 mg PbO and 0.1 mL OLAC were added to 5 mL trioctylamine in a 25 mL 4- neck round-bottom flask equipped with a thermometer adapter, reflux condenser, flow- control adapter, and rubber septum. The mixture was heated under Ar flow to 100 °C, producing a clear, colorless solution. Au-Pt-Fe3O4 heterotrimers (1 mL hexanes dispersion) were added under flowing Ar, and the hexanes was allowed to distill out of the solution. Under an Ar blanket, 0.33 mL of a 0.1 M sulfur solution was injected at

100 °C, and the reaction was allowed to proceed for 1 h. After the product mixture was allowed to cool, the particles were isolated and purified using the same procedure as that for CuxS-Au-Pt-Fe3O4 heterotetramers, and stored in hexanes.

6.2.10 Synthesis of higher-order Au(Pt-Fe3O4)n oligomers

5 mL trioctylamine was heated under Ar flow to 120 °C in a 25 mL 3-neck round- bottom flask equipped with a thermometer adapter, reflux condenser, flow-control adapter, and rubber septum. Au-Pt-Fe3O4 heterotrimers (1 mL hexanes dispersion) were

178 added via syringe, and the hexanes was allowed to distill out of the solution. Under an

Ar blanket, 6 mg sulfur powder dissolved in 1.6 mL diphenyl ether and 0.4 mL OLAM was injected via syringe, the temperature was increased to 150 °C, and the reaction was allowed to proceed for 1 h. After the product mixture was allowed to cool, the particles were isolated and purified using the same procedure as that for CuxS-Au-Pt-Fe3O4 heterotetramers, and stored in hexanes.

6.2.11 Characterization

TEM images were obtained from a JEOL 1200 EX II operating at 80 kV.

HRTEM images, SAED patterns, and EDS spectra were collected using a JEOL JEM-

2010 LaB6 microscope operating at 200 kV. Elemental mapping was performed on a

JEOL EM-2010F STEM operating at 200 kV with a typical probe size of 0.7 nm and an

EDAX (division of AMETEK) solid state X-ray detector. Data analysis was carried out using ESVision software (copyright Emispec, 2004). Samples for TEM and STEM were prepared by drop-casting hexane-dispersed nanoparticles onto Formvar-coated copper or nickel TEM grids. Lattice fringes were measured from the fast-fourier transform (FFT) of HRTEM images, using Gatan Digital MicrographTM software. Powder XRD data were collected using a Bruker D8 Advance X-ray diffractometer equipped with Cu Kα radiation. XPS analyses were performed on a monochromatic Al Kα source instrument

(Kratos, Axis Ultra, England) operating at 14 kV and 20 mA for an X-ray power of 280 watts. Spectra were collected with a photoelectron take off angle of 90° from the sample surface plane, energy steps of 0.10 eV, and pass energy of 20 eV. All XPS spectra were

179 referenced to the C 1s peak with a binding energy of 284.5 eV. UV-visible absorption spectra were collected using an Ocean Optics DH-2000-BAL with quartz cuvettes. ICP-

AES was performed on a Perkin-Elmer Optima 5300, using synthetic standards from

High Purity Standards to calibrate results. “Morphological yield” of hybrid particles is defined as the number of particles having a particular morphology relative to the total number of particles counted in TEM images (typically 200-300), expressed as a percentage.

6.3 Results and Discussion

6.3.1 M-Pt-Fe3O4 heterotrimers (M=Au, Ag, Ni, Pd)

We chose colloidal Pt-Fe3O4 heterodimers as a model system, since they are part of a class of Fe3O4 / noble metal hybrid nanoparticles that are stable, readily synthesized, and among the most studied because of their dual plasmonic-magnetic18 and catalytic- magnetic8 functionalities. Pt seed particles [Figure 6-1(a)] were synthesized by reducing

19 Pt(acac)2 with Fe(CO)5 at 190 °C, and the Pt-Fe3O4 heterodimers were synthesized by

20 thermal decomposition of Fe(CO)5 in 1-octadecene containing the Pt seed particles. The transmission electron microscopy (TEM) image in Figure 6-1(b) shows Pt-Fe3O4 heterodimers with Pt and Fe3O4 domains having average diameters of 5.1 ± 0.6 nm and

14 ± 1 nm, respectively. The heterodimer morphology was obtained in approximately

90% yield (n = total number of particles counted = 320), with the remainder consisting of unreacted Pt seeds and Pt / Fe3O4 nanostructures with multiple Fe3O4 domains.

180

Figure 6-1: Stepwise construction of M-Pt-Fe3O4 heterotrimers (M = Ag, Au, Ni, Pd). A schematic showing the multi-step synthesis of M-Pt-Fe3O4 heterotrimers, along with the most significant possible products and their observed frequencies (expressed as the percentage of observed heterotrimers, not total yield), is shown at the top. Representative TEM images show (a) Pt nanoparticle seeds, (b) Pt-Fe3O4 heterodimers, and (c) Au-Pt-Fe3O4, (d) Ag-Pt-Fe3O4, (e) Ni-Pt-Fe3O4, and (f) Pd-Pt-Fe3O4 heterotrimers. Panel (g) shows photographs of (left) a vial containing Au-Pt-Fe3O4 heterotrimers in hexane, which responds to an external Nd-Fe-B magnet, (middle) the same vial with Au-Pt-Fe3O4 heterotrimers in a larger volume of hexanes, and (right) the same vial after precipitation of the heterotrimers with ethanol. The precipitated heterotrimers collect next to the external magnet. All scale bars are 25 nm.

To begin building higher-order hybrid nanoparticles, we studied the reduction of a

1-octadecene solution of HAuCl4 with oleylamine in the presence of Pt-Fe3O4 heterodimers at mild (60-90 °C) temperatures. Notably, colloidal Au-Fe3O4 heterodimers

181 are stable and can be synthesized under similar conditions using Fe3O4 nanocrystals as

21,22 seeds. The heterogeneous nucleation of Au on Pt-Fe3O4 seeds could therefore yield four primary products: Au on Pt, Au on Fe3O4, Au on both the Fe3O4 and Pt ends, and

Au at the interface of both Fe3O4 and Pt (Figure 6-1). While a mixture of products that includes both Au-Fe3O4-Pt and Au-Pt-Fe3O4 linkages might be expected, we instead observe Au-Pt-Fe3O4 as the exclusive heterotrimer product [Figure 6-1(c)], which forms in approximately 85% yield (n = 285). The remainder of the product consists of either unreacted Pt-Fe3O4 or higher-order structures that form from a small number of Fe3O4-Pt-

Fe3O4 seeds that were present in the original Pt-Fe3O4 sample. The Au domains are largely spherical, with average diameters of 8.4 ± 0.6 nm. High-resolution TEM

(HRTEM) images indicate that all three components are crystalline and that they have coherent interfaces (Figure 6-2). Element mapping by scanning TEM (STEM) confirms the spatial distribution of Au, Pt, and Fe in the heterotrimer product (Figure 6-3).

A similar Au-Pt-Fe3O4 heterotrimer was observed previously, along with other morphologically diverse products, using similar reaction conditions23. However, Ag-Pt-

Fe3O4 [Figure 6-1(d)], Ni-Pt-Fe3O4 [Figure 6-1(e)], and Pd-Pt-Fe3O4 [Figure 6-1(f)] heterotrimers have not been previously reported, and they can also be synthesized by reducing Ag(OAc), Ni(acac)2, or Pd(acac)2, respectively, in the presence of the Pt-Fe3O4 seeds. Like the Au-Pt-Fe3O4 system, each reaction results in heterogeneous nucleation of

Ag, Ni, or Pd exclusively onto the Pt domain of Pt-Fe3O4 seeds. While the growth of Pd on Pt is difficult to control using oleylamine as a reducing agent24 and branched nanostructures often form, it is clear that heterogeneous nucleation occurs only on the Pt surface [Figure 6-1(f)]. The morphologically well-defined Ag-Pt-Fe3O4 and Ni-Pt-Fe3O4

182 trimers were obtained in yields of approx. 88% (n = 310) and 70% (n = 287), respectively, with the remainder of the product consisting primarily of unreacted Pt-

Fe3O4 seeds. While there appears to be some preference for a bent vs. linear morphology

– 56% bent vs. 44% linear for Au-Pt-Fe3O4, 84% bent vs. 16 % linear for Ag-Pt-Fe3O4, and 63% bent vs. 37 % linear for Ni-Pt-Fe3O4, possibly due to preferred nucleation at the higher-energy corner sites of the Pt domains – we do not observe nucleation of Au, Ag,

Ni, or Pd on the Fe3O4 domain or across the Pt-Fe3O4 interface. This result demonstrates chemoselective deposition of a variety of metals exclusively onto the Pt ends of Pt-Fe3O4 heterodimers, with no observable side products.

Figure 6-2: HRTEM image of a single Au-Pt-Fe3O4 heterotrimer, highlighting the coherent (111) interfaces. Scale bar is 5 nm.

183

Figure 6-3: A STEM image of a single Au-Pt-Fe3O4 heterotrimer is shown in (a), along with EDS-generated elemental maps corresponding to (b) Fe, (c), Pt, (d), Au, and (e) superimposed Fe (red), Pt (green), and Au (blue). Scale bar is 5 nm.

Representative data from powder X-ray diffraction (XRD), energy dispersive X- ray spectroscopy (EDS), and selected area electron diffraction (SAED) for all four heterotrimer systems are shown in Figure 6-4. The powder XRD data in Figure 6-4(a) show broad peaks corresponding to nanocrystalline Pt and Fe3O4 in all of the M-Pt-Fe3O4 systems, along with evidence of Au, Ag, Ni, and Pd in the appropriate samples. The EDS spectra in Figure 6-4(b) show the presence of Pt and Fe in ratios that are qualitatively consistent with those expected for the M-Pt-Fe3O4 heterotrimers, as well as signals for

Au, Ag, Ni, and Pd. The SAED patterns in Figure 6-4(c) show distinct diffraction spots that correspond to each of the crystalline phases in the representative heterotrimer systems. UV-Visible absorption spectra for 8.3 ± 0.9 nm Au nanoparticles (image shown in Figure 6-5) and the Au-Pt-Fe3O4 heterotrimers show absorbance maxima of 524 and

530 nm, respectively [Figure 6-4(d), bottom]. The spectrum for the Au-Pt-Fe3O4 heterotrimers is only slightly broadened and red-shifted relative to the free Au nanoparticles, indicating that the Au-Pt interface does not significantly alter the surface plasmon resonance of the Au domain. Similarly, the UV-Visible absorption spectrum for the Ag-Pt-Fe3O4 heterotrimers shows features that are consistent with the spectra of the

184

Figure 6-4: Characterization data for M-Pt-Fe3O4 heterotrimers (M = Au, Ag, Ni, Pd). (a) Powder XRD data for all of the M-Pt-Fe3O4 heterotrimers and the Pt-Fe3O4 heterodimer for comparison. (b) EDS spectra for all of the M-Pt-Fe3O4 heterotrimers. For the Au-Pt-Fe3O4 and Ag-Pt-Fe3O4 samples, a Ni TEM grid was used, and the Ni and Cu signals originate from this grid. For the Ni-Pt-Fe3O4 and Pd-Pt-Fe3O4 samples, a Cu TEM grid was used, and the Cu signal originates from this grid. (c) SAED patterns for all of the M-Pt-Fe3O4 heterotrimers show distinct diffraction spots or rings for each of the components. (d) UV-visible absorption spectra for the Au-Pt-Fe3O4 and Ag-Pt-Fe3O4 heterotrimers, also showing reference spectra for Au and Ag nanoparticles and the Pt-Fe3O4 heterodimers.

constituent Ag and Pt-Fe3O4 nanoparticles [Figure 6-4(d), top]. Finally, as shown in

Figure 6-1(g), the entire Au-Pt-Fe3O4 sample responds to an external magnet, both while

185 suspended in hexanes and after precipitation as a powder. This suggests that no significant amount of Pt or Au is present that is not attached to Fe3O4.

Figure 6-5: Au nanoparticles (8.3 ± 0.9 nm) prepared by heating 50 mg HAuCl4 dissolved in 20 mL OLAM for 15 minutes at 90 °C. The UV-Visible absorbance spectrum corresponding to this sample is shown in Figure 6-4(d).

6.3.2 Insights into M-Pt-Fe3O4 heterotrimer formation

The observation of heterogeneous nucleation of Au, Ag, Ni, and Pd that is exclusively chemoselective with respect to the Pt surface is interesting, especially considering that colloidal hybrid nanostructures with stable M-Fe3O4 (M = Au, Ag, Ni,

Pd) heterojunctions are known.18,20,21,25-27 Lattice matching and surface energies are often used to rationalize site-selective heterogeneous nucleation events in related metal-metal oxide systems, typically for growth of metal domains over single-component metal oxide nanorods.28,29 However, these simple arguments may not be sufficient to fully explain the observed chemoselectivity in the heterotrimer systems. For example, the d-spacing between adjacent (111) planes in Fe3O4 (d111 = 4.85 Å) is approximately two times that of

186 the (111) spacing of Au (d111 = 2.35 Å) or Ag (d111 = 2.36 Å), and this epitaxial relationship is recognized to be a driving force for the formation and long-term stability of those interfaces.18,20 Although well-defined Au-Pt heterostructures with coherent (111)

30 planes have been reported, the lattice mismatch between Au (or Ag) and Pt (d111 = 2.27

Å) is slightly higher (3.5 %) than between Au (or Ag) and Fe3O4 (3.1 %). This, coupled with the even greater lattice mismatch between Ni(111) (d111 = 2.03 Å) and Pt(111), makes it difficult to rationalize the preferential M-Pt linkages in the M-Pt-Fe3O4 heterotrimers on the basis of epitaxy.

Figure 6-6(a-d) shows a series of control experiments designed to probe the chemoselectivity of heterogeneous nucleation for a prototype system, Ag-Pt-Fe3O4.

When Fe3O4 nanoparticles are reacted with a toluene solution of Ag(OAc) in oleylamine

25 at 60 °C, Ag-Fe3O4 heterodimers form [Figure 6-6(a)]. When Pt nanoparticle seeds are reacted with Ag(OAc) under identical conditions, Ag-Pt heterostructures form (Figure

6-6(b)]. While their morphology is different than that of the Ag-Fe3O4 particles, a bimetallic Ag-Pt heterostructure is clearly evident, and this demonstrates the propensity for Ag to nucleate on Pt nanoparticle surfaces. When a physical mixture of the Fe3O4 and

Pt nanoparticles is reacted with Ag(OAc) at 60 °C, both Ag-Fe3O4 heterodimers and Ag-

Pt heterostructures form [Figure 6-6(c)]. Ag nucleates indiscriminately on both the Fe3O4 and Pt seed particles, with no evidence of chemoselectivity or a preference for either

Fe3O4 or Pt. However, similar to the data presented in Figure 6-1, Pt-Fe3O4 heterodimers react with Ag(OAc) under identical conditions to form exclusively Ag-Pt-Fe3O4 heterotrimers, with no evidence of Ag nucleating on the Fe3O4 domains [Figure 6-6(d)].

187

Figure 6-6: Chemoselective nucleation in the Ag-Pt-Fe3O4 heterotrimer system. (a-d) TEM images and corresponding schematics showing a series of control experiments, all carried out under the same reaction conditions and designed to probe the chemoselectivity of heterogeneous nucleation in the prototype Ag-Pt-Fe3O4 system: (a) nucleation of Ag on Fe3O4 nanoparticles, (b) nucleation of Ag on Pt nanoparticles, (c) nucleation of Ag on both Fe3O4 and Pt nanoparticles that are present as a physical mixture, and (d) nucleation of Ag exclusively on the Pt domain of Pt- Fe3O4 nanoparticles that are present as interconnected dimers with a solid-state interface. The insets in (b) and (c) show enlarged views of the regions indicated by boxes, highlighting the lower-contrast Ag regions that nucleate on the higher-contrast Pt particles. The scale bars in the main panels correspond to 20 nm and those in the insets correspond to 5 nm.

Figure 6-7: Pt4f XPS spectra for Pt-Fe3O4 heterodimers and the corresponding Pt seeds, indicating a decreased binding energy of 0.4 eV in the heterodimers.

188

This clearly indicates that the Pt-Fe3O4 heterodimers, with fused Pt and Fe3O4 domains, have significantly different reactivity toward the nucleation of Ag than when the Pt and

Fe3O4 domains remain separate. This result can be considered as a nanocrystal analogue of regiospecificity in molecular systems, where several products with different spatial arrangements of functional domains are possible, but only one product is observed.

The Pt domain in Pt-Fe3O4 heterodimers has been shown to have a higher electron density than single-component Pt particles of the same size, and this has been used to rationalize the enhanced catalytic activity of Pt-Fe3O4 heterodimers relative to isolated Pt particles for the oxygen reduction reaction.8 This enhancement has been attributed to electron transfer from Fe to Pt during heterogeneous nucleation, which involves deposition of Fe0 and subsequent oxidation, and is driven by equilibration of the Fermi energies in Fe (11.1 eV)31 and Pt (8.8 eV).32 There are also many well-known examples of strong metal-support interactions (SMSI) between late transition metal nanoparticles

(including Pt) and reducible metal oxide surfaces.33,34 These interactions are characterized by electron transfer from the support to the metal, heteroatomic interfacial bonds, significant changes in surface chemistry, enhanced catalytic activity, and (in some cases) encapsulation of the metal particles.35-38

Figure 6-7 shows X-ray photoelectron spectroscopy (XPS) data for the Pt-Fe3O4 heterodimers, which indicates that the binding energy of the Pt4f electrons decreases by approx. 0.4 eV (referenced to the C1s peak at 284.5 eV) relative to the isolated Pt nanoparticle seeds. Several factors can contribute to such small shifts in electron binding energies in supported metal nanoparticles.39 Given that the size of the Pt nanoparticles was the same in both samples, the two factors that are most likely to have contributed to

189 the observed peak shifts are differential charging and charge transfer.39 Assuming no change in the position of the C1s reference peak, the small shift of electron binding energies in the Pt-Fe3O4 system relative to the isolated Pt particles is likely attributable to charge transfer to the Pt domain. Similar XPS data have previously been attributed to

8 this phenomenon for Pt-Fe3O4 heterodimers , and it is consistent with the difference in

31 40 the work functions of Pt(111) and Fe3O4(111), which are 5.93 eV and 5.52 eV, respectively.

The Pt domain appears to serve as an electron sink during the formation of Pt-

Fe3O4 heterodimers, and this helps to rationalize the observed chemoselectivity. The electron-rich Pt domain may help to anchor and reduce the cationic , silver, nickel and palladium reactants, as well as enhance the polarizability of the Pt domain to help facilitate greater surface interaction with small nuclei of those metals. This can be considered as a nanocrystal analogue of a molecular substituent effect, where enhanced reactivity at a specific site is driven by electron- transfer phenomena that are similar to what is well known in the heterogeneous catalysis literature (SMSI).33-38 The presence of

Fe3O4 and the electronic interactions across the Fe3O4/Pt interface modify the reactivity of the Pt domain, making it behave differently than isolated Pt nanoparticles in a manner that facilitates chemoselective nucleation and regiospecific M-Pt-Fe3O4 connectivity.

6.3.3 MxSy-Au-Pt-Fe3O4 heterotetramers (M = Cu, Pb)

Using the Au-Pt-Fe3O4 heterotrimers as seeds, we studied the nucleation of a fourth component, copper sulfide. A variety of non-stoichiometric and mixed-valence

190

Figure 6-8: Stepwise construction of MxSy-Au-Pt-Fe3O4 heterotetramers (M = Cu, Pb). A schematic showing the stepwise construction of MxSy-Au-Pt-Fe3O4 heterotetramers, along with the most significant possible products and their observed frequencies (expressed as the percentage of observed heterotetramers, not total yield), is shown at the top. Representative TEM images of Cu9S5-Au-Pt-Fe3O4 heterotetramers are shown in (a) and (b), with enlarged regions in (b) showing lattice fringes that correspond to the Fe3O4 (311), Cu9S5 (200), Cu9S5 (111), Au (111), and Pt (111) planes. Powder XRD data for the Cu9S5-Au-Pt-Fe3O4 and PbS-Au-Pt-Fe3O4 heterotetramers are shown in (c) and (d), respectively, along with a pattern for Cu9S5-Au heterodimers in (c) for comparison. Panel (e) shows a representative TEM image of the PbS-Au- Pt-Fe3O4 heterotetramers. The composite EDS element map for a single PbS-Au-Pt-Fe3O4 heterotetramer particle in panel (f) includes an overlay of Fe (red), Pt (green), Au (blue), and Pb

191

(orange). A STEM image of a single Cu9S5-Au-Pt-Fe3O4 heterotetramer is shown in (g), along with the corresponding EDS element maps for (h) Fe, (i) Pt, (j) Au, (k) Cu, and (l) superimposed Fe, Pt, Au, and Cu. The scale bars correspond to (a) 25 nm, (b,g) 5 nm, and (e) 10 nm.

phases occur in the Cu-S system (e.g. CuxS, 1 ≤ x ≤ 2), and these phases have optical and electrical properties that are useful for photovoltaic, electronic, and sensing

41,42 applications. Au-Pt-Fe3O4 seeds were dispersed in a trioctylamine solution of Cu(II) oleate and reacted with sulfur powder at 80 °C. Again, four primary products are possible (Figure 6-8), but we observe exclusively the linear heterotetramer CuxS-Au-Pt-

Fe3O4, where CuxS nucleates only on the terminal Au particle [Figure 6-8(a)]. The heterotetramer morphology is obtained in ~80% yield (n = 250), with the remainder corresponding to free-standing CuxS particles and unreacted heterodimer and heterotrimer seeds.

When Cu(II) oleate is used as a copper source, XRD data show only Fe3O4, Pt, and Au [Figure 6-9(a)]. This suggests that the CuxS component is amorphous, but its presence is confirmed by EDS [Figure 6-9(b)]. The crystallinity of the CuxS domain was significantly improved when Cu(I) acetate was used as the copper source. The HRTEM image in Figure 6-8(b) shows lattice spacings of 3.26 Å and 2.83 Å, which correspond

43 well to the (111) and (200) planes of cubic digenite, Cu1.8S (or Cu9S5). The powder

XRD data in Figure 6-8(c) also show peaks that are consistent with digenite-type Cu9S5, along with Au, Pt, and Fe3O4. STEM element mapping data further confirms the presence of species that contain Fe, Pt, Au, and Cu, as well as their spatial distribution in the heterotetramer [Figure 6-8(g-l)]. The UV-Visible absorption spectrum is shown in

Figure 6-10, and this indicates that the position of the surface plasmon resonance of the

192

Figure 6-9: (a) Powder XRD pattern corresponding to CuxS-Au-Pt-Fe3O4 tetramers prepared using Cu(II) oleate, indicating the presence of (1) Fe3O4 (2) Au, and (3) Pt. The diffraction pattern is nominally the same as for Au-Pt-Fe3O4 trimers [Fig 6-4(a)], which suggests that the CuxS component is amorphous. (b) EDS spectra for a blank formvar-coated Ni TEM grid (bottom), showing the presence of Cu and Ni, and CuxS-Au-Pt-Fe3O4 tetramers (top) dropcast onto the same grid. The presence of a Cu-containing phase is inferred by the significantly higher Cu signal in the heterotetramer sample.

Figure 6-10: The UV-visible absorption spectrum corresponding to the CuxS-Au-Pt-Fe3O4 heterotetramers (blue), along with that of Au-Pt-Fe3O4 heterotrimers (red) for comparison.

193 Au domain is significantly affected by the presence of the copper sulfide, causing broadening and a red shift from 530 to 580 nm. Similar results are obtained for the nucleation of crystalline PbS on the Au-Pt-Fe3O4 heterotrimer, forming exclusively PbS-

Au-Pt-Fe3O4 [Figure 6-8(e)]. Powder XRD data for the PbS-containing heterotetramer show evidence of all four crystalline components: PbS, Au, Pt, and Fe3O4 [Figure 6-

8(d)].

Control experiments were designed to investigate the site-specific heterogeneous nucleation of Cu9S5 and PbS on Au rather than on the other solvent-exposed surfaces. Au nanocrystals and Pt-Fe3O4 heterodimers were treated in separate flasks with dissolved sulfur (and without any Cu source) under reaction conditions identical to those that generated Cu9S5-Au-Pt-Fe3O4 heterotetramers. The sulfur-soaked Au and Pt-Fe3O4 nanoparticles were then isolated and thoroughly purified to remove all traces of the sulfur-containing reaction medium. Sulfur was detected in both the Au nanocrystal and

Pt-Fe3O4 heterodimer samples by inductively coupled plasma emission spectrometry

(ICP-AES), which suggests that sulfur atoms adsorb to the solvent-exposed surfaces of the Au-Pt-Fe3O4 heterotrimers without preference for the Au or Pt-Fe3O4 domains. When isolated Au nanoparticles were reacted with Cu(I) acetate and sulfur under conditions identical to those that yielded the Cu9S5-Au-Pt-Fe3O4 heterotetramers, TEM images show that Cu9S5-Au heterodimers form [Figure 6-11(a)]. In contrast, when the Pt-Fe3O4 heterodimers were reacted with Cu(I) acetate and sulfur under identical conditions, TEM images indicate that irregularly-shaped CuxSy particles form and that they are clearly isolated and separate from the unreacted Pt-Fe3O4 heterodimers [Figure 6-11(b)].

Because sulfur adsorbs to both the Au and Pt-Fe3O4 subunits, adsorption energies do not

194 likely contribute significantly to the observed chemoselectivity. Rather, the adsorption of sulfur onto Au and the reaction of sulfur with the copper reagent are kinetically competitive processes: under the reaction conditions, sulfur atoms likely adsorb to Au at

+ a faster rate than they react with Cu , thereby localizing Cu9S5 nucleation to the Au surface. Nucleation of Cu9S5 on the Au surface is also favored by known differences in the reactivity and mobility of atomic sulfur species on Au vs. Pt surfaces.44 This mechanistic hypothesis can be extrapolated to the PbS-Au-Pt-Fe3O4 system as well, and is consistent with previous kinetic arguments that have been proposed for the nucleation of sulfide semiconductors on noble metal surfaces.14

Figure 6-11: TEM images corresponding to control experiments design to probe the chemoselectivity of Cu9S5 nucleation on the Au surface of Au-Pt-Fe3O4 seed particles. Cu9S5-Au heterostructures (which include some heterodimers), shown in (a), clearly indicate that heterogeneous nucleation of Cu9S5 on the Au seeds has occurred. When an identical reaction is carried out using Pt-Fe3O4 heterodimers as seeds, shown in (b), a mixture of irregularly shaped CuxSy nanoparticles and unreacted Pt-Fe3O4 seeds is obtained, clearly indicating that the CuxSy and Pt-Fe3O4 components remain separate from one another. The scale bars correspond to 20 nm in (a) and 50 nm in (b).

195 The library of chemical reactions that generate colloidal hybrid nanostructures is continually being expanded upon. However, the ability to carry out the seed-mediated reactions sequentially and in a way that can be predicted in advance based on mechanistic considerations, ultimately leading to the high-yield synthesis of the first reported colloidal heterotetramers with tetrafunctional semiconducting-plasmonic-catalytic- magnetic functionality, is a significant synthetic achievement both in terms of the heterotetramer itself and the total synthesis concept that it represents. The semiconductor-metal, metal-metal, and metal-metal oxide heterojunctions present in these four-component nanostructures are expected to strongly modulate the way each component interacts with light, accommodates charge, and behaves toward catalytic chemical reactions.

6.3.4 Higher-order (Au-Pt-Fe3O4)n hetero-oligomers via coupling

In addition to chemoselective chemical reactions, the total synthesis framework for organic molecules also relies on coupling reactions that link together smaller fragments to form larger molecules and oligomers. It is known that colloidal Au-Fe3O4 heterodimers can be fused together by heating them in the presence of sulfur to form

14 Fe3O4-Au-Fe3O4 dumbbells, and this suggests that a similar coupling reaction could link together higher-order hybrid nanoparticles to form larger oligomers. Indeed, when the

Au-Pt-Fe3O4 heterotrimers are heated to 120-150 °C with a trace amount of sulfur powder in trioctylamine, larger hetero-oligomers are observed. Figure 6-12 shows

196 representative TEM images of typical products, which span a range of oligomeric hybrid nanoparticles that each consist of a minimum six distinct domains.

The Au domains in our Au-Pt-Fe3O4 heterotrimers are significantly larger than those present in Au-Fe3O4 heterodimers (see reference 14, approximately 2 nm in Au-

Fe3O4 compared to 8 nm in Au-Pt-Fe3O4). In the case of Au-Fe3O4 dimer coupling, reasonably uniform Fe3O4-Au-Fe3O4 dumbbells with a well-defined morphology are obtained. In Figure 6-12, it is likely that incomplete, partial diffusion of the coupled Au-

Pt-Fe3O4 heterotrimers led to the random distribution of morphologies. It is reasonable to conclude that these coupling processes are optimized when the solid-solid diffusion distances are minimized, and the surface areas are small enough to self-regulate the morphological outcome (e.g. when the domains that promote coupling are small).

Figure 6-12: Higher order hetero-oligomers based on the Au-Pt-Fe3O4 heterotrimer building block. Top: A schematic showing the sulfur-mediated coupling of Au-Pt-Fe3O4 heterotrimers to form higher-order oligomers. Bottom: Representative TEM images of coupled products, including linear and bent Fe3O4-Pt-Au-Au-Pt-Fe3O4 hexameric oligomers and higher-order branched nanostructures. Cartoon representations of each hetero-oligomer are also shown. Scale bars correspond to 10 nm.

197 6.4 Conclusions

In analogy to complex organic molecules, colloidal nanoparticle hetero-oligomers can be constructed rationally by the sequential and mechanism-driven application of known reactions. These reactions can be rationalized, and therefore in many cases predicted in advance, based on kinetic arguments and quantifiable parameters that include interfacial electronic interactions. Given the diversity of materials that are amenable to such reactions45 and the growing library of chemical transformations that can further modify composition, structure, and morphology,46-48 a total synthesis framework has the potential to conceptually underpin the rational construction of large multi-domain hybrid nanoparticles with unprecedented architectural complexity for applications in solar energy conversion, catalysis, medicine, and . Emerging advances in nanoparticle separation and purification can further improve yield and homogeneity49-52 and the continued development of new classes of reactions, including orthogonal reaction schemes,53-54 can further expand these synthesis-by-design capabilities.

6.5 References

1. Costi, R.; Saunders, A. E.; Banin, U. Angew. Chem. Int. Ed. 2010, 49, 4878.

2. Wang, C.; Xu, C.; Zeng, H.; Sun, S. Adv. Mater. 2009, 21, 3045.

3. Cozzoli, P. D.; Pellegrino, T.; Manna, L. Chem. Soc. Rev. 2006, 35, 1195.

4. Costi, R.; Saunders, A. E.; Elmalem, E.; Salant, A; Banin, U. Nano Lett. 2008, 8,

637.

198 5. Elmalem, E.; Saunders, A. E.; Costi, R.; Salant, A.; Banin, U. Adv. Mater. 2008,

20, 4312.

6. Xu, C.; Wang, B.; Sun, S. J. Am. Chem. Soc. 2009, 131, 4216.

7. Yin, H.; Wang, C.; Zhu, H.; Overbury, S. H.; Sun, S. Chem. Commun. 2008, 36,

4357.

8. Wang, C.; Daimon, H.; Sun, S. Nano Lett. 2009, 9, 1493.

9. Xu, C.; Xie, J.; Ho, D.; Wang, C.; Kohler, N.; Walsh, E.G.; Morgan, J.R.; Chin,

Y.E.; Sun, S. Angew. Chem. Int. Ed. 2008, 47, 173.

10. Choi, J-S.; Jun, Y-W.; Yeon, S-I.; Kim, H. C.; Shin, J-S; Cheon, J. J. Am. Chem.

Soc. 2006, 126, 15982.

11. Mokari, T.; Rothenberg, E.; Popov, I.; Costi, R.; Banin, U. Science 2004, 304,

1787.

12. Figuerola et al.; J. Am. Chem. Soc. 2008, 130, 1477.

13. Kwon, K-W.; Shim, M. J. Am. Chem. Soc. 2005, 127, 10269.

14. Shi, W.; Zeng, H.; Sahoo, Y.; Ohulchanskyy, T. Y.; Ding, Y.; Wang, Z. L.;

Swihart, M.; Prasad, P. N. Nano Lett. 2006, 6, 875.

15. Casavola, M.; Buonsanti, R.; Caputo, G.; Cozzoli, P. D. Eur. J. Inorg. Chem.

2008, 6, 837.

16. Nicolau, K. C.; Vourloumis, D.; Winssinger, N.; Baran, P. S. Angew. Chem. Int.

Ed. 2000, 39, 44.

17. Choi, C.; Alivisatos, A. P. Annu. Rev. Phys. Chem. 2010, 61, 369.

18. Yu, H.; Chen, M.; Rice, P. M.; Wang, S. X.; White, R. L; Sun, S. Nano Lett.

2005, 5, 379.

199 19. Wang, C.; Daimon, H.; Onodera, T.; Koda, T.; Sun, S. Angew. Chem. Int. Ed.

2008, 47, 3588.

20. Wang, C.; Yin, H.; Dai, S.; Sun, S. Chem. Mater. 2010, 22, 3277.

21. Gu, H.; Yang, Z.; Gao, J.; Chang, C. K.; Xu, B. J. Am. Chem. Soc. 2005, 127, 34.

22. Xu, Z.; Hou, Y.; Sun, S. J. Am. Chem. Soc. 2007, 129, 8698.

23. Zhang, H-T.; Ding, J.; Chow, G-M.; Dong, Z-L. Langmuir 2008, 24, 13197.

24. Mazumder, V.; Sun, S. J. Am. Chem. Soc. 2009, 131, 4588.

25. Zhang, L.; Dou, Y-H.; Gu, H-C. J. Colloid Interface Sci. 2006, 297, 660.

26. Choi, S-H.; Na, H. B.; Park, Y. I.; An, K.; Kwon, S. G.; Jang, Y.; Park, M.;

Moon, J.; Son, J. S.; Song, I. C.; Moon, W. K.; Hyeon, T. J. Am. Chem. Soc.

2008, 130, 15573.

27. Jang, Y.; Chung, J.; Kim, S.; Jun, S. W.; Kim, B. H.; Lee, D. W.; Kim, B. M.;

Hyeon, T. Phys. Chem. Chem. Phys. 2011, 13, 2512.

28. Casavola, M.; Grillo, V.; Carlino, E.; Giannini, C.; Gozzo, F.; Pinel, E. F.; Garcia,

M. A.; Manna, L.; Cingolani, R.; Cozzoli, P. D. Nano Lett. 2007, 7, 1386.

29. Casavola, M.; Falqui, A.; Garcia, M. A.; Garcia-Hernandez, M.; Giannini, C.;

Cingolani, R.; Cozzoli, P. D. Nano Lett. 2009, 9, 366.

30. Wang, C.; Tian, W.; Ding, Y.; Ma, Y-Q.; Wang, Z. L.; Markovic, N. M.;

Stamenkovic, V. R.; Daimon, H.; Sun, S. J. Am. Chem. Soc. 2010, 132, 6524.

31. Lide, D. R. CRC Handbook of Chemistry and Physics 87rd Edition, CRC Press:

Boca Raton, FL (2006).

32. Smith, N. Phys. Rev. B 1974, 9, 1365.

200 33. Tauster, S. J.; Fung, S. C.; Baker, R. T. K.; Horsley, J. A. Science 1981, 211,

1121.

34. Tauster, S. J. Acc. Chem. Res. 1987, 20, 389.

35. Goodman, D. W. Catal. Lett. 2005, 99, 1.

36. Horsley, J. A. J. Am. Chem. Soc. 1979, 101, 2870.

37. Ohyama, J.; Yamamoto, A.; Teramura, K.; Shishido, T.; Tanaka, T. ACS Catal.

2011, 1, 187.

38. Qin, Z-H.; Lewandowski, M.; Sun, Y-N.; Shaikhutdinov, S.; Freund, H-J. J.

Phys. Chem. C 2008, 112, 10209.

39. Barr, T. L. Modern ESCA: The Principles and Practice of X-Ray Photoelectron

Spectroscopy, CRC Press: Boca Raton, FL (1994).

40. Weiss, W.; Ranke, W. Prog. Surf. Sci. 2002, 70, 1.

41. Lim, W. P.; Wong, C. T.; Ang, S. L.; Low, H. Y.; Chin, W. S. Chem. Mater.

2006, 18, 6170.

42. Regulacio, M. D.; Ye, C.; Lim, S. H.; Bosman, M.; Polavarapu, L.; Koh, W. L.;

Zhang, J.; Xu, Q-H.; Han, M-Y. J. Am. Chem. Soc. 2011, 133, 2052.

43. JCPDS Card No. 00-004-0861

44. Rodriguez, J. A.; Dvorak, J.; Jirsak, T.; Liu, G.; Hrbek, J.; Aray, Y.; Gonzalez,

C. J. Am. Chem. Soc. 2003, 125, 276.

45. Carbone, L.; Cozzoli, P. D. Nano Today 2010, 5, 449.

46. Son, D. H.; Hughes, S. M.; Yin, Y.; Alivisatos, A. P Science 2004, 306, 1009.

47. Yin, Y.; Rioux, R. M.; Erdonmez, C. K.; Hughes, S.; Somorjai, G. A.; Alivisatos,

A. P. Science 2004, 304, 711.

201 48. Vasquez, Y.; Henkes, A. E.; Bauer, J. C.; Schaak, R. E. J. Solid State Chem.

2008, 181, 1509.

49. Beveridge, J. S.; Buck, M. R.; Bondi, J. F.; Misra, R.; Schiffer, P.; Schaak, R. E.;

Williams, M. E. Angew. Chem. Int. Ed. 2011, 50, 9875.

50. Latham, A. H.; Frietas, R. S.; Schiffer, P.; Williams, M. E. Anal. Chem. 2005, 77,

5055.

51. Beveridge, J. S.; Stephens, J. R.; Latham, A. H.; Williams, M. E. Anal. Chem.

2009, 81, 9618.

52. Chen, G.; Wang, Y.; Tan, L. H.; Yang, M.; Tan, L. S.; Chen, Y.; Chen, H. J. Am.

Chem. Soc. 2009, 131, 4218.

53. Wang, Y.; Chen, G.; Yang, M.; Silber, G.; Xing, S.; Tan, L. H.; Wang, F.; Feng,

Y.; Liu, X.; Li, S.; Chen, H. Nat. Commun. 2010 1:87 doi: 10.1038/ncomms1089

54. Leonard, B. M.; Anderson, M. E.; Oyler, K. D.; Phan, T.-H.; Schaak, R. E.

ACS Nano 2009, 3, 940.

55. Kovalenko, M.; Bodnarchuk, M. I.; Lechner, R. T.; Hesser, G.; Schaffler, F.;

Heiss, W. J. Am. Chem. Soc. 2007, 129, 6352.

56. Choi, S-H.; An, K.; Kim, E-G.; Yu, J. H.; Kim, H.; Hyeon, T. Adv. Funct. Mater.

2009, 19, 1645.

Chapter 7

Summary and Outlook

Natural products are molecules produced by living organisms that have pharmacological or biological value. If biologists or medical experts perceive a natural product to be valuable enough, a synthetic organic chemist may be called upon to prepare the molecule artificially. In these situations, an organic chemist knows precisely what to do. Organic synthesis of complex molecules involves the sequential application of predictive chemical reactions, guided by well-established mechanisms. Chemists approach the synthesis of complex organic molecules by drawing upon a comprehensive library of well-known transformations, which allows them to identify simple molecules and fragments that can be modified straightforwardly, linked together or split apart.

Many useful organic transformations are selective, occurring at one pre-determined location on a molecule that may have multiple reaction sites. There are also robust strategies for manipulating the reactivity of certain parts of a molecule, so that other parts can be reacted selectively, such as chemical protection, orthogonal reactivity, and substituent effects. Collectively, these powerful guidelines and strategies form a conceptual framework known as total synthesis, which allows organic chemists to rationally construct almost any target molecule with precise control over the structure, bonding, connectivity and function.

Current and future technologies based on materials with nanometer-sized components have become ubiquitous in the physical sciences, engineering, biology and

203 medicine. Inorganic materials chemists who are practitioners of colloidal nanochemistry are facing synthetic challenges that are different, but in many ways analogous, to the challenges faced by organic chemists. Colloidal nanochemistry has the potential to enable impressive control over the composition, size, shape, uniformity, and surface properties of crystalline inorganic nanoparticles - which can be considered as the building blocks of nanomaterials - but many synthetic challenges remain before nanoparticle building blocks can be rationally integrated into functional materials and devices. This is precisely analogous to the timeline between natural product discovery and medicinal application. Progress in nanochemistry, however, is somewhat limited by two closely related factors - 1) the lack of an expansive set of colloidal reactions and nanoparticle transformations that can be relied upon as the foundation for 2) a mechanism-driven framework like total synthesis.

The intrinsic properties of nanoparticle building blocks can be finely tuned by controlling size and shape, but spatial organization and assembly are equally important for integrated nanosystems. Newer research is focused on how nanoparticles “talk” to each other, which has revealed that nanomaterial performance requires synergetic communication between the building blocks that allows for physical phenomena such as charge transfer and energy propagation to occur – often across boundaries between different materials. Robust and well-defined junctions between building blocks are therefore desirable, assembled with a programmed connectivity in mind. This translates directly into a synthetic challenge to construct nanoparticles with more sophisticated architectures, as the growing disconnection between what is designed and what can be made becomes a critical limitation.

204 Like the organic chemist who is called upon to prepare a natural product, the nanoparticle practitioner must respond to the designs and synthetic demands of nanoscience. The size-tunable physics of discrete nanoparticles has earned them the nickname “artificial atoms”, but no conceptual synthetic framework yet exists for rationally modifying or linking nanoparticles together into “artificial molecules”. The concepts and research described in this dissertation have attempted to bridge the worlds of organic molecule synthesis and nanoscience, through our efforts to synthesize hybrid, multi-component nanostructures that have spatially engineered inorganic domains connected by solid-state interfaces. The intimate contact between the domains in these objects is analogous to the bonding between atoms in an organic molecule, which can support inter-domain communication for energy conversion, electronics, catalysis, and other applications.

The research in Chapters 2, 3, and 4 improves upon this emerging synthetic framework for inorganic nanostructures, because it describes our efforts to formulate reliable strategies for synthesizing several important single-component nanoparticle targets. As such, those studies enrich the current (and growing) library of reactions that transform molecular precursors into colloidal nanoparticles, which can be considered as an enhancement of the basic capabilities in nanochemistry. The chemical tools for preparing and manipulating single-component nanoparticles define our means to construct more complex, multi-component nanostructures. The objective of Chapters 2 and 3 was to develop such reactions for nanostructured GeTe, which is an important

Group IV-VI semiconductor with fascinating optical, electronic, phase change, and ferroelectric properties. The technological relevance of novel GeTe nanostructures is

205 intriguing, but the solution-phase synthesis of morphology-controlled GeTe remains an underexplored subject. In our experience, the primary synthetic obstacle is to mediate strong, attractive, interparticle forces that provoke agglomeration, uncontrolled growth, and precipitation, which preclude higher-order shape control. We approached these challenges by deeply investigating the reaction chemistry between several Ge and Te precursors, and by qualifying the advantages of using soluble polymers for mediating the growth of GeTe. We observed the formation of several GeTe nanostructures that had not been previously reported, including uniform, micron-sized nanocubes, octahedral nanoparticles, amorphous alloy nanospheres and single-crystal 2D nanosheets.

In Chapter 4, we studied the colloidal synthesis of wz-CoO nanostructures via thermal decomposition of [Co(OL)2] complex. Thermolysis of such 3d transition metal carboxylates is a promising reaction for producing scalable quantities of uniform colloidal metal oxide nanocrystals, but optimized conditions for each salt differ significantly. Previous studies implied that a significantly higher activation barrier exists for thermolysis of [Co(OL)2], when compared to [Fe(OL)3] or [Mn(OL)2], for example.

Perhaps more importantly, reports pertaining to the thermolysis of [Co(OL)2] are not consistent with each other. Nanocrystalline wz-CoO is an interesting synthetic target that features Co2+ ions in a metastable, tetrahedral bonding geometry, and its preparation via decomposition of [Co(OL)2] was found to be very sensitive to impurities - some of which were present (or absent) in the precursor, and some of which were generated in situ. By analyzing and managing these impurities, we had synthesized a variety of size- and shape-controlled wz-CoO nanocrystals.

The current reaction toolkit for synthesizing multi-component inorganic

206 nanostructures is summarized in Chapter 1.5. Chapters 5 and 6 describe some important contributions to that collection of transformations. In Chapter 5, segmented nanowires with multi-metal domains were prepared by reacting template-grown metal nanowire synthons with metal salts under reducing conditions, which represents a novel alternative to preparing such complex targets using electrochemical co-deposition. Importantly, the backside coating of Ag on the AAO membrane, which is inherent to the electrodeposition process, was leveraged as a protecting group for one end of the nanowires. One terminus of each template-confined nanowire was physically protected, which allowed us to demonstrate site-selective transformation of the solution-exposed end. Furthermore, the sidewalls of membrane-bound nanorods were also protected from the reaction mixture, which ensured that site-selective transformation occurred at the nanowire tips.

In Chapter 6 we discussed a nanoparticle analogue of chemoselectivity that was a strong, materials-general driving force for establishing connectivity in a variety of colloidal hybrid nanoparticles. To begin building hybrid structures, for example, we studied the heterogeneous addition (or seeded-growth) of transition metals (M) Au, Ag,

Ni, and Pd from two-component Pt-Fe3O4 heterodimer seeds. Pt-Fe3O4 heterodimers have two chemically distinct surfaces from which M can grow, and we predicted at the outset of these studies that hybrid nanoparticle products with M-Pt-Fe3O4, Pt-Fe3O4-M, and M-Pt-Fe3O4-M connectivities were all conceivable. Instead each reaction produced three-component heterotrimers with only M-Pt-Fe3O4 linkage, where Au, Ag, Ni, and Pd all grew exclusively from the Pt surface, leaving the Fe3O4 surface unreacted. This surprisingly general result was also akin to regiospecificity in molecular systems, because the addition of M gave rise to a structure with only M-Pt-Fe3O4 connectivity.

207 A series of control experiments was designed to probe the chemo- and regioselectivity of seeded-growth for a prototype system, Ag-Pt-Fe3O4, in an effort to elucidate the driving force. Remarkably, we found that Ag grew from a physical mixture of Fe3O4 and Pt nanoparticle seeds to generate both Ag-Fe3O4 and Ag-Pt heterodimers in the same reaction, with no evidence of chemoselectivity or a preference for either Fe3O4 or Pt. Under identical conditions, Ag grew from Pt-Fe3O4 heterodimers to form exclusively Ag-Pt-Fe3O4 heterotrimers, with no evidence of Ag growing from the Fe3O4 domains. This clearly indicated that the hybrid particles with adjoined Pt and Fe3O4 domains had significantly different reactivity toward growth of Ag than when the Pt and

Fe3O4 domains were separate. X-Ray photoelectron spectroscopy (XPS) measurements showed that the electron density of Pt connected to Fe3O4 was significantly higher than isolated Pt, which was likely due to electron transfer from the Fe3O4. The electronic structure and reactivity of the Pt domain were enhanced by linkage to Fe3O4, which was a nanoparticle analogue of a molecular substituent effect.

Finally, our mechanism-driven framework was also used to construct the first known examples of heterotetramers, which are hybrid nanoparticles with four distinct components and functionalities. Heterotrimers Au-Pt-Fe3O4 with three reaction-exposed surfaces, were used as the seeds for addition. The seeded-growth of metal sulfides

(MxSy; M = Cu, Pb) resulted exclusively in linear heterotetramers with Cu9S5-Au-Pt-

Fe3O4 connectivity, where MxSy grew chemoselectively from the Au surface. Sulfur- initiated coupling of Au-Pt-Fe3O4 heterotrimers was also described, which gave rise to higher-order (Au-Pt-Fe3O4)n hetero-oligomers.

Emerging strategies for the “total synthesis” of inorganic nanostructures have the

208 potential to permit chemists to respond to design criteria for next-generation nanomaterials, because they merge concepts from molecular organic synthesis with the needs that underpin advances in . Capabilities in colloidal nanochemistry for preparing both single- and multi-domain nanostructures are continually expanding.

Crystalline inorganic nanoparticles undergo facile chemical transformations into derivative materials with morphological retention, and this is leading to a rapidly expanding toolkit that has analogies to the large and diverse library of reactions available to organic chemists. Strategies that facilitate the construction of large molecules, which include chemoselectivity, regioselectivity, orthogonal reactivity, substituent effects, protection-deprotection and coupling, are starting to be demonstrated for multi- component nanostructures. It is our hope that the research described herein contributes to these predictive synthesis capabilities and powerful guidelines for implementing multi- step routes to new and more complex nanostructures.

VITA

Matthew Robert Buck

Education

Ph.D. Chemistry (2013) The Pennsylvania State Universtity - University Park, PA B.A. Chemistry (2007) The State University of New York at New Paltz - New Paltz, NY B.S. Finance (2001) The State University of New York at New Paltz - New Paltz, NY

Honors and Awards

Penn State Chemistry Department Travel Award (2011, 2010) Harold F. Martin Outstanding Graduate Assistant Teaching Award (2008) Dan Waugh Memorial Teaching Award (2008) Pennsylvania State University Graduate Fellowship (2007-2008) Dr. James Campion Memorial Scholarship in Chemistry (2007)

Selected Publications

[1] Buck, M.R., Biacchi, A.J., Popczun, E.J., Schaak, R.E. Polymer-Assisted Synthesis of Colloidal Germanium Telluride Nano-Octahedra, Nanospheres, and Nanosheets. Chem. Mater. 2013, DOI: 10.1021/cm4009656

[2] Buck, M.R., Schaak, R.E. Emerging Strategies for the Total Synthesis of Inorganic Nanostructures. Angew. Chem. Int. Ed. 2013, DOI: 10.1002/anie.201207240

[3] Buck, M.R., Bondi, J.F., Schaak, R.E. A Total Synthesis Framework for the Construction of High-Order Colloidal Hybrid Nanoparticles. Nature Chem. 2012, 4, 37.

[4] Beveridge, J.S., Buck, M.R., Bondi, J.F., Misra, R., Schiffer, P., Schaak, R.E.,

Williams, M.E. Purification and Magnetic Interrogation of Hybrid Au-Fe3O4 and FePt-

Fe3O4 Nanoparticles. Angew. Chem. Int. Ed. 2011, 50, 9875.

[5] Buck, M.R., Sines, I.T., Schaak, R.E. Liquid-Phase Synthesis of Uniform Cube-Shaped GeTe Microcrystals. Chem. Mater. 2010, 22, 3236.

[6] Anderson, M.E., Buck, M.R, Sines, I.T, Oyler, K.D., Schaak, R.E. On-Wire Conversion Chemistry: Engineering Solid-State Complexity into Striped Metal Nanowires Using Solution Chemistry Reactions. J. Am. Chem. Soc. 2008, 130, 14042.