Photoluminescence and Resonance Raman Spectroscopy of MOCVD Grown

GaAs/AlGaAs Core-Shell Nanowires

A Thesis

Submitted to the Faculty

of

Drexel University

by

Oren D. Leaer

in partial fulllment of the

requirements for the degree

of

Doctorate of Philosophy

February 2013 © Copyright 2013 Oren D. Leaer.

Figure 5.1 is reproduced from an article copyrighted by the American Physical Society and used with their permission. The original article may be found at: http://link.aps.org/doi/10.1103/PhysRevB.80.245324

Regarding only gure 5.1, the following notice is included as part of the terms of use: Readers may view, browse, and/or download material for temporary copying purposes only, provided these uses are for noncommercial personal purposes. Except as provided by law, this material may not be further reproduced, distributed, transmitted, modied, adapted, performed, displayed, published, or sold in whole or part, without prior written permission from the American Physical Society.

The rest of this work is licensed under the terms of the Creative Commons Attribution-ShareAlike license Version 3.0. The license is available at: http://creativecommons.org/licenses/by-sa/3.0/. i

Dedications

To my family: my parents, my sister, and Bubby and Dan.

Your support made this possible. Thank you. ii

Acknowledgments

It is with great pleasure that I am able to thank and acknowledge the individuals and organizations that helped me in the course of my graduate studies. I should begin by thanking my advisor, Dr. Spanier, and my committee, Drs. Livneh, May, Shih, Taheri, and Zavaliangos for supporting me through this rather long process. Working for Dr. Spanier helped me grow as a scientist and learn things that extend well beyond the lab. And Dr. Spanier, thank you again for the opportunities for collaboration and travel. I would like to additionally thank Dr. May for allowing me to take advantage of his open oce door and drop in and ask questions on a number of occasions. And, Tsachi, thank you for your support and advice; I don’t know how this dissertation would have been completed without your help. I would like to acknowledge my funding sources: I was supported by the GAANN fellowship from the Department of Education, the Dean’s fellowship, and the National Science Foundation (NSF) EAPSI fellowship for summer 2011. The work presented here was part of an ongoing collaboration with our Italian colleagues who I had the pleasure of visiting in Lecce, . Nico Lovergine and Paola Prete supplied the nanowires that are the topic of this work. In addition to providing us with samples, they graciously shared their data and allowed me to reproduce it in gure 2.2. Adriano Cola hosted me at the lab in Lecce and made sure that I saw the lab’s fantastic surroundings. In the lab, I learned to use the FIB from Enrico Melissano. Thanks to Anna Fontcuberta i Morral and the American Physical Society for granting permission to reproduce the gure that appears in gure 5.1. I also gratefully acknowledge the help of Dr. Craig Johnson and Stephanie Johnson in acquiring the TEM micrograph presented in gure 2.4 A number of people at Drexel enriched my graduate experience. I would like to thank Dr. Barsoum for the chance to be part of the MAX-phase collaboration; I feel fortunate to have been part of such a productive collaboration. Volker Presser generously provided access to the Renishaw Raman system at the William Keck Institute; in addition to taking the time to iii

train me on the instrument, he also took the time to help nd the right conguration for the instrument. My colleagues in the lab – Chris Hawley, Jennifer Atchison, Mohammad Islam, Maria Torres, Sean Chen, Stephanie Johnson, Stephen Nonnenmann, and Terrence McGuckin – helped keep standards high and worked to keep things fun as we toiled away and I hope to collaborate with them in the future. Kevin Siegl and Alex Juskovic both taught me quite a bit when I mentored them. Judy Trachtman, Dorilona Rose, Keiko Nakazawa, Yeneeka Long, and Sarit Kunz helped me navigate the university bureaucracy. Andrew Marx provided various acts of technical support over the years. Scott Currie was a great help in building my experiments, as were Mark Shiber and the sta at the Drexel machine shop. And thank you to Dr. Knight for your continued eorts to keep our labs safe. I was fortunate to be able to take classes from Nader Engheta, Andrea Liu, and Vaclav Vitek at the University of Pennsylvania. I am grateful for how welcome I was made to feel in those classes. The highlight of my time as a graduate student was spending a few months at the Australian National Fabrication Facility’s New South Wales Node (ANFF-NSW) in Sydney, Australia as an EAPSI fellow during the summer of 2011. I would like to thank the NSF and the Australian Academy of Sciences (AAS) for the funding and logistical support. Meaghan O’Brien of the AAS did a wonderful job of helping with many of the arrangements in Australia. I am deeply indebted to my host at the ANFF-NSW Facility, Prof. Andrew Dzurak, and to Rosie Hicks for helping to arrange my visit. Dr. Fay Hudson generously shared her nanofabrication expertise with me. And it was a great experience working with the sta at the ANFF-NSW: Frank Wright, Gordon Bates, Warren McKenzie, Joanna Szymanska, and El van Zeijl. Prof. Chennupati Jagadish of the Australian National University graciously hosted me when I visited for a day. He and his team showed me great hospitality and were extremely generous with their technical knowledge. Evan Malone and the sta at NextFab Studio were a great help with building experimental equipment and also provided a pleasant work environment. Matt Blaze, Andrew, Sandy, Perry, and Sam made me feel at home in their lab. Leslie Mitts saw that I needed a place to sit

Acknowledgments iv

and write and invited me to work from the SBDC. And Babak, Dev, and Q kindly let me camp out in their lab space. A number of people helped make Sydney feel like home. I was privileged to live with Oliver and Beun who shared their home with me. Dov Rosenfeld helped me nd the ANFF and introduced me to his family in Sydney: Shula, Peter, and Ronny took good care of me while I was there. I would like to thank my friends Tim Verstynen and Jesse Engel for giving me some much-needed perspective. Dan and Ashlee, Ian and Alexis, Billy and Mimi, Micah, Cookie, Adriana, and Rich Grote hosted me at dierent times while I attended conferences. Nick Vacirca and Andrew Bohl helped keep the atmosphere at Drexel light. Linda Anderson and Howard Baker helped me get through this intact. Terrence, Amy, Zoë, and Calyx were patient with me as a housemate for an intended few months that turned into much longer. Matt Robinson and my father helped with the editing. Any mistakes that remain are mine, but there are denitely fewer thanks to their help. And I am extremely grateful to my family for not only putting up with me through my rather long periods of graduate school related crankiness, but also being consistently supportive and helping me see this through to completion. v

Table of Contents

List of Figures ...... vii Abstract ...... xi 1. Introduction ...... 1 1.1 Semiconductor Heterostructures...... 1 1.1.1 The Double Heterostructure Laser...... 3 1.1.2 Quantum Well Devices...... 5 2. Fundamentals of Core-Shell Nanowires ...... 8 2.1 Our Samples and Hypothesis...... 8 2.2 VLS Growth of Nanowires ...... 12 2.3 Impurities and Defects ...... 15 2.3.1 Carbon...... 15 2.3.2 Wurtzite and Planar Defects...... 17 3. Photoluminescence of GaAs/AlGaAs Nanowires ...... 18 3.1 Introduction ...... 18 3.2 Experimental Procedures...... 19 3.2.1 Sample Preparation...... 19 3.2.2 Experimental Conditions...... 20 3.3 Photoluminescence Results...... 21 3.3.1 PL From GaAs Wafer...... 21 3.3.2 Bare GaAs Nanowires...... 22 3.3.3 Core-Shell Nanowires...... 23 3.4 Interpretation of Results and Relevant Eects...... 28 3.4.1 Bandgap Narrowing ...... 28 3.4.2 Donor-Acceptor Pair Luminescence...... 29 3.4.3 The Role of Hydrogen ...... 30 3.5 Conclusions from PL ...... 31 4. Resonance Raman of GaAs/AlGaAs Nanowires ...... 34 4.1 Raman Spectroscopy...... 34 4.2 Experimental Procedures...... 39 4.3 Results and Interpretation...... 44 4.3.1 Summary of Results...... 44 4.3.2 Incoming vs. Outgoing Resonance...... 47 vi

4.3.3 The Missing TO Mode ...... 49 4.3.4 Variation of the LO Peak Position...... 51 5. Single Nanowire Raman Spectroscopy ...... 55 5.1 Experimental Rationale...... 55 5.2 Procedure...... 57 5.3 Theoretical Considerations...... 59 5.4 Results and Discussion ...... 68 5.5 Conclusions ...... 71 6. Conclusions, Open estions, and Future Work ...... 72 6.1 Conclusions ...... 72 6.2 Open Questions and Future Work ...... 73 Appendix A: Finite Curvature-Mediated Ferroelectricity ...... 75 Bibliography ...... 80 Vita ...... 93 vii

List of Figures

1.1 In a semiconductor homostructure, band bending drives oppositely charged car- riers in opposite directions as shown in (a). In a heterostructure, the combination of quasi-elds and external elds can be tuned to only aect one type of carrier as in (b), or drive both electrons in holes in the same direction as in (c)..... 2

1.2 The band diagram for a double heterostructure laser (a) under moderate forward bias and (b) under high forward bias. Under high forward bias, a well forms, enhancing the trapping of carriers helping create the necessary population inversion for lasing. Based on Kroemer [3,6]...... 3

1.3 The layer structure of the heterostructure used for the discovery of the fractional quantum Hall eect, showing the two dimensional electron gas on the GaAs side of the GaAs/AlGaAs [14–16]. After Lin et al. [17]...... 6

2.1 SEM image of SAGA06 nanowires still on the growth substrate. Courtesy of Enrico Melissano – CNR IMM Lecce...... 8

2.2 PL from dierent batches of core-shell nanowires, labeled by V:III precursor ratios – conditions were otherwise identical during growth. These plots reproduce plots from reference 20. Data provided courtesy of Paola Prete and Nico Lovergine.. 9

2.3 Sample output from a self-consistent Schrödinger-Poisson solver for a radial heterostructure, showing the bandstructure (top) and the calculated electron density (bottom) [Kevin Siegl, Oren Leaer, Jonathan Spanier – unpublished]. 11

2.4 TEM micrograph of SAGA06 nanowires. Courtesy of Stephanie Johnson and Craig Johnson...... 12

3.1 Some of the states and the transitions between them that contribute to the observed PL signal (based on a gure from Gilliland [91])...... 18

3.2 Photoluminescence from a semi-insulating GaAs wafer, acquired with the sample held at 78 K, showing the band edge of GaAs and carbon impurity band. . . . . 21

3.3 PL from bare GaAs nanowires (sample NW40) measured at 4.2 K with peaks labeled with tted energies. The vertical gray line is at 1.519 eV, marking the expected value for the bandgap of GaAs at 4.2 K. The annotated peak separation corresponds to the energy of an LO phonon. Excitation power was 0.023 mW. 23 viii

3.4 PL from core-shell nanowires SAGA10 measured at 4.2 K...... 24

3.5 PL from core-shell nanowires SAGA11 measured at 4.2 K...... 25

3.6 Peak position vs. power for the spectra plotted in gures 3.4 and 3.5 (blue squares correspond to the SAGA10 data and orange circles correspond to the SAGA11 data). Peak position was calculated as a weighted mean as described in the text. 27

4.1 Feynman diagram for the Raman process. Based on Fig. 7.28 in Yu and Cardona [2]. 34

4.2 A schematic of the transitions involved in incoming and outgoing resonances for a Stokes process...... 37

4.3 A schematic of the band structure of GaAs, showing the direct band gap between the conduction band and the light and heavy hole bands (Eg) and the split-o band gap (∆SO)...... 37

4.4 Initial data replotted by transforming wavenumbers to absolute energy; wavenum- ber w measured using excitation wavelength λ, E = E(λ) − E(w)...... 41

4.5 An example of the RRS data being t in OriginPro...... 42

4.6 RRS data from sample SAGA06, measured at 78 K...... 43

4.7 SAGA11 RRS data ...... 44

4.8 NW40 RRS data ...... 45

4.9 GaAs Wafer RRS...... 46

4.10 The squares show the amplitude of the RRS signal from SAGA06 measured at 78 K, plotted as a function of outgoing energy, i.e. Eex − ELO. The solid line shows the PL signal measured under the same conditions...... 47

4.11 Selected spectra showing the lack of a TO peak in the core-shell nanowire data. This eect was independent of the excitation power used, the spectra shown were selected only because they include the full region from 250 cm−1 to 310 cm−1 in a single contiguous scan, rather than being broken up into two separate scans as was often the case...... 49

4.12 LO and TO phonons...... 50

4.13 (a) The laser line measured at a number of dierent excitation energies, measured through closed slits so as to avoid damaging the PMT. (b) The calculated oset for the laser at the dierent wavelengths...... 52

LISTOFFIGURESLISTOFFIGURES ix

4.14 Diagram of instrument...... 54

5.1 The dierence in stacking orders for (a) zinc-blend GaAs and (b) wurtzite GaAs leads to additional peaks in the Raman spectrum of wurtzite GaAs due to (c) the folding of the Brillouin zone. (Reprinted with permission from Zardo et al., Phys. Rev. B 80, 245324 (2009) Copyright 2009 American Physical Society)[ 129]. 56

5.2 Diagram of experimental setup, the line with the arrows represents the path of the laser light, which includes segments in which incoming and scattered light overlap...... 57

5.3 Polarized Raman spectra from a single nanowire, grouped by the position of the polarization analyzer, with the angle of the λ/2-waveplate labeled, less a constant oset of 50°...... 58

5.4 Schematic showing the scattering conguration and illustrating the vectors used in the calculation of the expected Raman scattering intensity...... 60

5.5 Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = eˆi, where φ is the angle between the [211¯ ] and xˆ1...... 65

5.6 Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration where eˆs is set orthogonal to eˆi, where φ is the angle between the [211¯ ] and xˆ1...... 66

5.7 Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = xˆ3, where φ is the angle between the [211¯ ] and xˆ1...... 67

5.8 Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = xˆ2, where φ is the angle between the [211¯ ] and xˆ1...... 67

5.9 (a) Raman from a single SAGA06 nanowire. (b) Image of the nanowire from the microscope’s CCD...... 68

5.10 Zoomed in plot of the single SAGA06 Raman data, showing an apparent luminescent background, and a clearly resolved TO phonon peak, but no sign of the LO phonon peak...... 69

5.11 Solid lines are cos4(θ ) ts to the data measured in the (a) parallel and (b) perpendicular congurations...... 70

LISTOFFIGURESLISTOFFIGURES x

A.1 Polarization vs. distance from inner radius for the size nanowires used in the study, illustrating the eect of strain in increasing the polarization of thin ferroelectric shells...... 75

A.2 Waterfall plot of the polarization across the thinnest shell, with color representing the temperature used in the calculation of the polarization...... 76

LISTOFFIGURESLISTOFFIGURES xi

Abstract Photoluminescence and Resonance Raman Spectroscopy of MOCVD Grown GaAs/AlGaAs Core-Shell Nanowires Oren D. Leaer Jonathan E. Spanier, Ph.D.

This work describes experiments conducted on GaAs/AlGaAs core-shell nanowires. The

GaAs/AlGaAs materials system has been a major platform in the development of heterostruc-

ture devices. Heterostructure nanowires grown using bottom-up growth techniques are a

means to creating high quality nanoscaled heterostructures with a variety of potential

uses, most notably in optoelectronics. The GaAs/AlGaAs nanowires that were the subject

of this study were provided to us by partners in Italy, who specialize in MOCVD growth

of heterostructures. These nanowires were hypothesized to have elds at the interface

between core and shell that were inuencing the bandstructure. We used low temperature

photoluminescence (PL), resonance Raman, and polarized Raman spectroscopies to character-

ize the nanowires and understand the inuence of the shell on the electronic structure of

the nanowires. The PL and resonance Raman experiments were carried out on a custom

instrument constructed as part of carrying out this study.

PL gave us a more complete picture of the inuence of defect states on the electronic

properties and the signicance of donor-acceptor-pair (DAP) luminescence in the previously

observed spectral character. The contrast in spectral character between bare GaAs nanowires

and the core-shell nanowires led to an alternative hypothesis that the role of the shell in

determining the spectral characteristics was largely through the thermal history imparted xii

on the cores when they were heated to the shell’s growth temperature; at the elevated

temperatures, hydrogen incorporated into the cores as they were grown diused out.

Hydrogen and carbon are impurities that are dicult to avoid in MOCVD growth, because

they are byproducts of the precursor breakdown and may be incorporated into the nanowires

during growth. Prior to exposure to the elevated temperatures, hydrogen compensated the

carbon which is useful as a p-type dopant in GaAs.

Further investigation of the nanowires using resonance Raman spectroscopy showed

strong coupling between the LO-phonon and the bound exciton state, particularly in the

core-shell nanowires, supporting the new hypothesis. Polarized Raman spectroscopy was

used to investigate the possibility of wurtzite GaAs being present in what were otherwise

zinc-blende GaAs structures.

Abstract

1

Chapter 1: Introduction

The principal applications of any suciently new and innovative technology have always been – and will continue to be – applications created by that technology.

Kroemer’s Lemma of New Technology [1]

1.1 The Signicance of Semiconductor Heterostructures

Heterostructure nanowires, the subject of our research, fall under the broader category

of semiconductor heterostructures; so to contextualize our research, some description of

semiconductor heterostructures is in order. To illustrate some of the properties that make

heterostructures useful and the AlGaAs-GaAs material system particularly attractive to work with, we will discuss the double heterostructure (DH) laser and the quantum well

heterostructure. Both of these examples also share the features that the heterostructure

is used to control the spatial distribution of carriers in the device and that they were

breakthroughs enabled by improved control of semiconductor growth, which were achieved

using GaAs/AlGaAs in both cases.

GaAs/AlGaAs heterostructures have been of particular importance in the development of

heterostructure devices, largely because there is such a close match in lattice constant

between GaAs and AlAs, reducing the eect of strain as an issue to contend with during

crystal growth. Because the lattice spacings of the family of ternary Alx Ga1−x As compounds vary so little, while the bandgaps vary signicantly, heterostructures based on GaAs/AlGaAs 2 have often been the rst grown for various types of devices [2, 3]. Most other combinations of III-V compounds have a signicant lattice mismatch that if not accounted for in the design, may lead to the formation of defects in the material.

-qF (a)

+qF

(b)

+qFh -qFe

(c)

+qFh

Figure 1.1: In a semiconductor homostructure, band bending drives oppositely charged carriers in opposite directions as shown in (a). In a heterostructure, the combination of quasi-elds and external elds can be tuned to only aect one type of carrier as in (b), or drive both electrons in holes in the same direction as in (c). (This gure is based on Kroemer [4].)

Nanowire heterostructures are a progression in the continued development of semicon- ductor heterostructures, so before attempting to describe semiconductor heterostructure nanowires, it will be helpful to rst describe semiconductor heterostructure thin lms.

Semiconductor heterostructure devices in use today are built around thin lms. The growth techniques used for nanowires are based on those developed for thin lms [5]. Of particular relevance for nanowire growth are molecular beam epitaxy (MBE) and metal organic chemical

Chapter1: Introduction 1.1SemiconductorHeterostructures 3 vapor deposition (MOCVD).

The feature that denes a semiconductor heterostructure is that the composition of the

semiconductor changes as a function of position – in other words, it includes two or more

semiconductors grown in contact with each other. This is in contrast to homostructures, which can have varied doping, but consist of a single material throughout [3].

1.1.1 The Double Heterostructure Laser

Electrons Electrons

Donors Donors Light Light

Acceptors Acceptors

Holes Holes (a) (b)

Figure 1.2: The band diagram for a double heterostructure laser (a) under moderate forward bias and (b) under high forward bias. Under high forward bias, a well forms, enhancing the trapping of carriers helping create the necessary population inversion for lasing. Based on Kroemer [3,6].

Ideas for devices based on heterostructures signicantly predate their realization. For

example, Shockley was issued a patent in 1951 that mentions the idea of using a wide bandgap

material emitter in a transistor to improve device performance [7], an idea that was more

fully developed by Kroemer [8]. The barrier to their realization was the challenge of growing

lms that combined dierent materials without creating defects [3,9].

In a semiconductor homostructure, i.e. a single semiconductor material, an electric eld

can be externally applied or develop from a doping gradient, but either way the bands remain

parallel. This means that any electrons or holes are acted on with equal and opposite forces,

Chapter1: Introduction 1.1SemiconductorHeterostructures 4

so electrons are subjected to a force Fe = −qE and holes are accelerated by Fh = +qE.A

gradient in the composition has the eect of creating a “quasi-electric eld” arising from the

gradient of the bandgap. The important thing about the quasi-electric eld is that it is

inherently dierent for the valence and conduction bands, thus the forces are now described

by Fe = −qEc and Fh = +qEv [4].

This important feature, illustrated in gure 1.1, is key to the function of a variety of

devices because it allows independent control of electrons and holes. An important application

of this concept was the double heterostructure (DH) laser [3,9, 10].

The DH laser is a particularly compelling historical example of semiconductor heterostruc-

tures, because it was the rst device that achieved something using a heterostructure that

couldn’t be achieved using conventional semiconductor devices. In 1963, Herbert Kroemer

conceived of the DH laser as a means of defeating the limitations of solid state lasers – at the

time, solid state lasers needed to be operated at low temperatures and in short pulses, because

they would otherwise melt due to the high injection current required. He had been thinking

for some time about using heterostructures as a means of controlling the drift of carriers in a

semiconductor device. He was not the rst to think about using heterostructures to improve

transistor performance, but his in-depth analysis enabled him to propose the DH laser as a

means of building a solid state laser that could operate at room temperature continuously

(CW) [1,3].

Kroemer’s general insight about heterostructures was that they are a means of producing what he called “quasi-electric elds.” The DH laser was devised as a means of achieving a

room temperature operating CW solid-state laser by overcoming the need for extremely high

injector current required by previous designs in order to create a population inversion; the

Chapter1: Introduction 1.1SemiconductorHeterostructures 5

high current necessitated their operation at liquid nitrogen temperatures or in short low

duty-cycle pulses [1, 6]. The DH laser design has been extremely important technologically;

it is the basis of the lasers used in CD and DVD players, ber optic communications, as well as other applications [3]. One of the interesting details of the history of the DH laser

is the lack of interest from his management Kroemer received in response to his initial

proposal to attempt to fabricate a DH laser; it is in part in reference to this history that

Kroemer formulated his lemma [1]. The wider bandgap regions around the active region of

the DH laser are used to prevent the carriers from escaping the active region into opposing

contacts [6, 11].

Much of the early heterostructure literature focused on the Ge-GaAs system because their

lattice constants are quite close to each other. It turns out that the Ge-GaAs system has

complications that arise from the interface between a polar and non-polar materials making it

challenging to create device quality heterostructures using the Ge-GaAs system [12].

In addition to optoelectronic devices, heterostructures have been extremely valuable for

use in high-speed electronic devices. The heterostructure bipolar transistor (HBT), for

example, is used in a variety of RF applications. The HBT uses gradients in the bandgap of the

device to inuence the drift of minority carriers in the device to decrease their transit times,

leading to a faster device [13].

1.1.2 Quantum Well Devices

Quantum structures are a further development in the progress of heterostructure devices. In

these structures, carriers are conned to a well that is smaller in at least one dimension than

the Bohr exciton radius [2]. Carriers conned in a quantum well take on a set of discrete

Chapter1: Introduction 1.1SemiconductorHeterostructures 6

AlGaAs (Si-Doped)

AlGaAs (Undoped)

GaAs (Undoped)

GaAs (Substrate) Ef

Figure 1.3: The layer structure of the heterostructure used for the discovery of the fractional quantum Hall eect, showing the two dimensional electron gas on the GaAs side of the GaAs/AlGaAs heterojunction [14–16]. After Lin et al. [17]. states resulting from the particle-in-a-box like nature of the connement [18]. Additionally, their density of states is modied by the dimensionality of their connement, making their physics distinctly dierent from carriers in bulk.

Quantum well structures are also particularly useful for the creation of high mobility devices because they allow the separation of carriers from dopants through the technique of modulation doping [19]. High mobility carriers conned in quantum wells show additionally interesting physics, in particular, the fractional quantum Hall eect was discovered using a

GaAs-AlGaAs heterostructure [14–16]. One quantum well structure in particular is of interest to us here, because its band structure was the basis for the band structure originally proposed for the nanowires we studied.

The layered structure in gure 1.3 is a schematic of the modulation doped structure

Chapter1: Introduction 1.1SemiconductorHeterostructures 7

used by Tsui et al., which was grown with a 1 µm base of undoped GaAs, 500 Å of undoped

AlGaAs, and 600 Å of Si doped AlGaAs. In some of their structures, they also used a thin layer

of doped GaAs on top as a protective layer. The higher electron anity of GaAs leads to

the carriers from the ionized dopants in the AlGaAs accumulating in GaAs side of the

undoped GaAs/AlGaAs heterojunction. The resulting structure achieved very high mobility

by having a high quality interface and separation of carriers from dopants [14]. It was with

this structure in mind that we formulated our hypothesis about our nanowires that had no

intentional doping of their cores and Si in their AlGaAs shells.

Chapter1: Introduction 1.1SemiconductorHeterostructures 8

Chapter 2: Fundamentals of Core-Shell Nanowires

Figure 2.1: SEM image of SAGA06 nanowires still on the growth substrate. Courtesy of Enrico Melissano – CNR IMM Lecce.

2.1 Our Samples and Hypothesis

The samples we studied were GaAs-core—AlGaAs-shell nanowires grown using metal organic

chemical vapor deposition (MOCVD), which is also sometimes referred to as metal organic vapor phase epitaxy (MOVPE). They were provided to us by Nico Lovergine and Paola Prete who run a lab in Lecce, Italy, that specializes in MOCVD growth of III-V semiconductors. The

nanowires we studied were part of a series of nanowires grown with varying V:III precursor

ratios, from which we were able to use core-shell nanowires from three dierent batches

and bare GaAs nanowires from one growth. The parameters used for the growths of the 9 ) s t

i 5 : 1 n U

y r a r t i

b 1 0 : 1 r

A (

y t i s n

e 2 0 : 1 t n I

3 0 : 1

1 . 4 0 1 . 4 5 1 . 5 0 1 . 5 5 1 . 6 0 E n e r g y ( e V )

Figure 2.2: PL from dierent batches of core-shell nanowires, labeled by V:III precursor ratios – conditions were otherwise identical during growth. These plots reproduce plots from reference 20. Data provided courtesy of Paola Prete and Nico Lovergine. nanowires are listed in table 2.1. As will be explained below, at the point that we became involved, it was hypothesized that the interface between core and shell aected the electronic properties of the nanowires.

The signicant dierence between these nanowires and MOCVD grown GaAs/AlGaAs nanowires described in the literature is in the choice of arsenic precursor. Arsine (AsH3) is the precursor used for most MOCVD grown nanowires in which arsenic is the group V element [21, 22]. For the nanowires we studied, tertiarybutylarsine (TBA, C4H11As) was used instead [20, 23, 24]. TBA has a lower decomposition temperature than arsine, so it is utilized more eciently allowing growth at lower temperatures and with a lower V:III ratio. TBA is

Chapter2: Fundamentals 2.1OurSamplesandHypothesis 10

also safer than arsine; the LC50 (median lethal concentration) for arsine is 20 ppm at four

hours exposure time, while for TBA it is over 500 ppm and it has a lower vapor pressure – a

signicant reason to prefer it if the quality of growth products is not adversely aected [25].

TBA has been used successfully as a precursor for MOVPE growth of high quality GaAs

lms with low densities of incorporated impurities [26–29]. In optimized growth conditions

using TBA, it has been shown that the lower bound for the inclusion of carbon acceptor

doping is limited by the breakdown of the trimethylgallium (TMG) [30]. Thus, based on the

growth of GaAs lms, there is no reason to assume that the additional carbon from the TBA would necessarily be incorporated into the growth product.

Table 2.1: Excerpt from sample growth log, growth times are specied in minutes [Lovergine and Prete, private comm.]

Sample Core T(◦C) time V:III Shell T(◦C) time V:III Cap T(◦C) time V:III NW40 GaAs 400 90 5 ------SAGA06 GaAs 400 60 20 AlGaAs 650 4 20 GaAs 650 1 20 SAGA10 GaAs 400 60 20 AlGaAs 650 10 20 - - - - SAGA11 GaAs 400 60 5 AlGaAs 650 4 5 GaAs 650 1 5

The hypothesis for this work was predicated on the knowledge that carbon from the

breakdown of the precursors can result in p-type GaAs and that there was Si contamination

of Al source that could lead to n-type shells. It was hypothesized that these impurities created

the conditions for band bending and possibly a quantum well. The plan for our experiments was developed based largely on accounts such as Jiang et al. [31], Lauhon et al. [32], Lu et al.

[33], and Li et al. [34], which focus on the heterostructure properties of the nanowires, rather

than emphasize the challenges of nanowire growth.

Self-consistent Schrödinger-Poisson calculations suggested that with suciently high

doping of core and shell, it was possible for a quantum well to form at the interface. These

Chapter2: Fundamentals 2.1OurSamplesandHypothesis 11

Figure 2.3: Sample output from a self-consistent Schrödinger-Poisson solver for a radial heterostructure, showing the bandstructure (top) and the calculated electron density (bottom) [Kevin Siegl, Oren Leaer, Jonathan Spanier – unpublished]. calculations were performed using a solver for a radially symmetric heterostructure written by Kevin Siegl, an REU summer student in our lab, based on a published description of such a solver [35, 36]. Figure 2.3 shows the output of one such calculation, with an assumed doping of 1 × 1017 cm−3 in the core and 2 × 1018 cm−3 in the core.

The remainder of this chapter will provide more detail about nanowires and their growth, followed by a more in-depth discussion of some of the related issues to provide context for the following chapters where we discuss our experimental results.

Chapter2: Fundamentals 2.2VLSGrowthofNanowires 12

Figure 2.4: TEM micrograph of SAGA06 nanowires. Courtesy of Stephanie Johnson and Craig Johnson.

2.2 VLS Growth of Nanowires

Nanowires are structures that are on a scale of tens of nanometers or less in cross section with lengths of hundreds of nanometers up to many microns. Semiconductor nanowire

heterostructures are nanowires composed of more than one semiconductor material.

Nanowires can be produced by either top-down or bottom-up methods. Top-down methods

start with bulk material of the desired composition and remove material leaving behind

the nanowires; bottom-up methods follow nature’s method of building structures up

by self-assembly. The semiconductor industry uses top-down methods and the rapid

improvement of semiconductor devices has been closely tied to advancement fabrication

methods. Bottom-up methods oer potential advantages over what can be achieved through

Chapter2: Fundamentals 2.2VLSGrowthofNanowires 13

top-down fabrication; at the nanoscale, bulk materials are no longer especially uniform, and

bottom-up methods can be used to fabricate structures with geometries not achievable using

top-down methods [5, 37].

Nanowires have been demonstrated in a variety of applications including: p-n junctions [38,

39], LEDs [40, 41], lasers [42–44], photodetectors [45], single molecule detectors [46], and

photovoltaics [47–52]. Nanowire heterostructures have been fabricated to contain quantum wells [32, 33, 53], tunnel diodes [54], and work as high-mobility transistors [31, 34].

Most nanowires, including the nanowires that were the subject of this study, are grown

from the vapor phase starting with seed particles, which are most often Au [5]. One of the

drawbacks to the use of a catalyst is the possibility of contamination by the catalyst materials;

this has been documented with the contamination of Si nanowires by their Au catalyst

particles. Au in Si acts as a deep level trap, degrading the performance of the material for any

device applications [5, 55]. There are methods for growing nanowires from vapor without the

use of a catalyst; these methods avoid the issues associated with catalysts, but generally

require some element of top-down control, such as lithographically patterning a substrate

before nanowire growth [56–60]. Other methods of growing nanowires include solution

phase epitaxy (SPE) and liquid phase epitaxy (LPE), but they are applicable to a limited range

of materials and not widely used for nanowire growth [5].

The origins of nanowire growth techniques lie with observations of spontaneous BeO

microwhisker formation in the presence of metal droplets reported by Edwards and Happel

in 1962[ 61]. Application of this method to Si was reported by Wagner and Ellis in 1964 with the rst proposed mechanism, which they named the Vapor–Liquid–Solid (VLS)

mechanism [62].

Chapter2: Fundamentals 2.2VLSGrowthofNanowires 14

For a single component material, like Si, with a Au seed, the VLS mechanism is relatively

simple. It assumes that precursor molecules in the vapor phase (SiCl4 in the case of Wagner

and Ellis) preferentially stick to the liquid droplet formed from the melted seed particle and

breakdown forming a solution of Si-Au. The Si-Au solution becomes super saturated with Si

as more precursors breakdown and the temperature of the growth system is lowered towards

the Si-Au eutectic point. The super saturated solution then spontaneously precipitates out

Si leading to the one-dimensional growth, with the seed particle taking on the eutectic

composition [62].

There is some complication to the model, because the role played by the seed particle is

not actually that of a catalyst, thought it is often discussed as if it were. A catalyst raises

a reaction rate by lowering the activation energy through its presence, while not being

consumed itself. There have been several studies that have shown that the activation energy

for the breakdown of the precursors in the presence of the seed is approximately the same as without it [5, 23, 62].

Additionally, the role of the phase of the seed particle is unclear, but it may be unimportant

because the Vapor–Solid–Solid (VSS) mechanism has been observed in which the seed

particle remains solid. The VSS mechanism was observed using in-situ TEM by Kodambaka

et al. where they observed nanowires growing via VLS and VSS simultaneously in the same

growth environment [63].

The VLS model becomes more complex in dealing with binary compounds; in part

because the solubility of the components can be vastly dierent from each other, making the

explanation of supersaturation driven growth less plausible [5, 64]. The solubility of the

group V elements tends to be negligible in a Au seed particle, while the low melting points of

Chapter2: Fundamentals 2.2VLSGrowthofNanowires 15

Ga and In are generally below the growth temperature leading to the issue that there is no

solubility limit for the two liquid materials. The insoluble elements can be assumed to reach

the growth surface through surface diusion. However the explanation of why the soluble

element precipitates out is unclear [5].

After axial growth, during which the VLS (or similar) mechanism dominates, the growth

conditions may be changed to promote epitaxial growth of material on the surface of the

nanowire [32, 37, 65]. This is the technique used to grow the nanowires that we studied.

There have been a number of papers in the last few years detailing methods for growing

high quality nanowires using MOCVD; although, because arsine is the generally preferred

precursor, nearly all the papers describe the use of arsine rather than TBA. For our purposes,

a high quality nanowire is one that is straight (unkinked), untapered, free of planar defects, a

controlled crystal structure, and with a controlled dopant concentration [22].

2.3 Impurities and Defects in MOCVD Grown Nanowires

2.3.1 Carbon

Carbon is a byproduct of the MOCVD precursor chemistry, making it a challenging impurity to

eliminate from growth product; but it can be a desirable impurity in controlled concentrations,

because it is a good p-type dopant in GaAs: it has low diusivity compared to other p-type

dopants, it can achieve high doping concentrations, and it can be included just by changing

growth conditions [66–71]. For the nanowires in our study, tertiarybutylarsine (TBA) was

chosen over the more commonly used AsH3, introducing more carbon to the reaction. But

this shouldn’t necessarily lead to large amounts of carbon in the nanowires: in lm growth,

TBA has been used to produce high quality GaAs with low carbon content, and there is

Chapter2: Fundamentals 2.3ImpuritiesandDefects 16

evidence that use of TBA does not signicantly increase the carbon content of lms nearly as

much as the choice of TMG as a Ga precursor [26, 27, 72, 73]. In lm MOCVD growth, carbon

doping can be controlled by using a combination of gallium precursors, such as TMG with

triethylgallium (TEG) [74–76].

The growth rate of the nanowires in our study provides good reason to expect that carbon

is incorporated into the nanowires; it is low enough that the nanowires are tapered as seen in

gure 2.4. Tapering is evidence of sidewall growth, which occurs during slow growth of GaAs

nanowires; and sidewall deposition is the route by which carbon is incorporated into GaAs

nanowires; it doesn’t enter through the catalyst particle because the solubility of carbon in

gold is very low [77–79]. There are also a number of possible planar defects visible in the

gure 2.4, but without further imaging, it is impossible to distinguish them from other

phenomena, such as strain, that can appear as imaging artifacts.

In growth of GaAs using AsH3 and TMG as precursors, carbon in the growth product is a

byproduct of TMG – the only C containing precursor – which is obvious from the chemical

equation describing the growth,

AsH3 + Ga(CH3)3 −−→ GaAs + 3 CH4. (2.1)

Subtler is how carbon is incorporated in the GaAs, because it is via an intermediary product

of the breakdown of TMG, which explains how the carbon comes to be incorporated into the

As-sites. In what is believed to be the primary route for carbon incorporation, as the TMG

undergoes pyrolysis, Ga-bound carbene (GaCH2) forms as a reaction product, which can be

adsorbed from both ends – the carbon can occupy an As-site or the gallium can occupy its

Chapter2: Fundamentals 2.3ImpuritiesandDefects 17 own site [80–82].

2.3.2 Wurtzite and Planar Defects

Planar defects are in part related to the issue that wurtzite GaAs is a possible growth product,

rather than only zinc blende GaAs as is the case for bulk GaAs. Wurtzite is a crystal structure

that diers from zinc blende only in the stacking order of the layers, so a single plane of wurtzite in zinc blende would essentially be a stacking fault.

One of the complications of growing GaAs as a nanowire is that unlike in the bulk case where GaAs only forms in the zinc blende structure, GaAs can form a wurtzite (WZ) or zinc

blende (ZB) structure when it is nanoscaled [83–85]. In fact, GaAs nanowires can have WZ

and ZB segments mixed along the length of a single nanowire [86, 87]. Control over which

phase nanowires grow in has been a recent theme in several recent publications [88–90].

Chapter2: Fundamentals 2.3ImpuritiesandDefects 18

Chapter 3: Photoluminescence of GaAs/AlGaAs Nanowires

3.1 Introduction

Conduction Band Free Exciton

Donor Levels

CB-A Band to Band Hot Electron D-A (F,X)

Acceptor Level

Valence Band

Figure 3.1: Some of the states and the transitions between them that contribute to the observed PL signal (based on a gure from Gilliland [91]).

Photoluminescence spectroscopy (PL) – like cathodoluminescence, thermoluminescence,

or electroluminescence – is an emission spectroscopy where carriers are excited from the valence band to the conduction band and the radiative recombination of carriers is observed.

In the case of PL, the carriers are excited into the conduction band by light, but otherwise, the

physics are the same as in other luminescence processes. Luminescence processes can

be viewed as having three steps: excitation, thermalization, and radiative recombination. 19

Thermalization means that the excited carriers reach thermal equilibrium and relax to the

lowest energy states available to them, thus “forgetting” their histories, meaning that the

resulting spectrum should be essentially independent of the energy of the excitation used –

the exception to this is hot luminescence, where carriers that are not fully thermalized

recombine [2, 92].

Because of the thermalization, PL spectra can contain a signicant amount of information

about defect states lying within the bandgap of a material. In a perfect sample of direct

bandgap material, where the transition is allowed, we would expect carriers to accumulate at

the band edges and then recombine radiatively. At low temperatures, when dopants are not

ionized, the excited carriers can interact with the dopants, creating peaks in the spectrum that

can dominate the band edge luminescence [2, 91, 92].

Though our initial hypothesis was that the interface plays a role in the electronic

properties of the nanowires, we will interpret the results in this chapter as being due to the

impurity content of the GaAs. As we will explain below, high concentrations of dopants can

also modify the band structure of a material, causing the bandgap to shrink or renormalize. We will show that most of the features in our PL spectra are accounted for by donor-acceptor pair

(DAP) luminescence, a defect band due to carbon, and the renormalization of the bandgap.

3.2 Experimental Procedures

3.2.1 Sample Preparation

Nanowires were removed from growth substrates by brief ultrasonic agitation while immersed

in isopropyl alcohol (IPA). Nanowires were then deposited on substrates by placing drops of

IPA-nanowire suspension on SiOx on degenerately doped Si substrates that were placed

Chapter3:PL 3.2ExperimentalProcedures 20

on a hot plate set at ≈100 ◦C to evaporate o the IPA. Substrates were then placed on the

cryostat’s cold nger with a thin layer of Apezion-N thermal grease holding them in place.

All data was collected in a back scattering conguration. Laser power was measured

using a calibrated sensor (Newport 883-SL) that was placed at one leg of a beam-splitter.

Samples were aligned using a camera that was placed in the optical train for alignment

purposes and then the sample position and focus were optimized by maximizing the collected

PL intensity. See gure 4.14 for a diagram of the instrument.

3.2.2 Experimental Conditions Temperature Selection

At room temperature, the PL from GaAs nanowires is strongly quenched by thermally

activated defects. Early on in our experiments, we observed this eect anecdotally, nding it

much easier to measure PL from nanowires cooled to 78 K using liquid nitrogen; therefore we

conducted all our experiments at below room temperature. Titova et al. investigated this

eect in detail, showing that the strong quenching occurred above 120 K [93].

For PL, it is preferable to cool samples to 4.2 K using liquid helium. This improves PL

in several ways: thermal broadening is reduced, non-radiative relaxation mechanisms

(e.g. phonons) are suppressed, and carriers are frozen on impurities [2]; thus, we conducted

some of our PL experiments at 4.2 K. The reason not all experiments were carried out at 4.2 K was that we had limited access to liquid helium.

Laser Line Selection

The main selection requirement for the excitation energy used for the PL experiments was that it was above the bandgap energy of the core and below that of the shell. At low

Chapter3:PL 3.2ExperimentalProcedures 21

temperature, the bandgap of the AlGaAs shell is approximately 2 eV, thus any laser line over

1.52 eV from the Ti:Sa laser meets this requirement [24].

3.3 Photoluminescence Results

3.3.1 PL From GaAs Wafer

7 8 K ) s t i n U

y r a r t

i b r A (

y t i s n

e 1 . 4 9 0 5 7 t n I

1 . 4 6 1 . 4 8 1 . 5 0 1 . 5 2 1 . 5 4 1 . 5 6 E n e r g y ( e V )

Figure 3.2: Photoluminescence from a semi-insulating GaAs wafer, acquired with the sample held at 78 K, showing the band edge of GaAs and carbon impurity band.

In order to conrm the operation of the instrument, we used a piece of GaAs wafer as a

reference sample. The PL spectrum from the GaAs wafer (gure 3.2) shows two main peaks:

the band edge at 1.508 eV and the peak at 1.491 eV which we attribute to the carbon impurity

band [94].

Chapter3:PL 3.3PhotoluminescenceResults 22

A common approximation for the bandgap of GaAs is the Varshni equation,

aT2 E = E(T=0 K) − , (3.1) g T + b with E(T=0 K) = 1.519 eV, a = 5.405 × 10−4 eV/K, and b = 204 K (for the Γ transition) [94,

95]. With the given values, at 78 K, equation (3.1) predicts a bandgap of 1.507 eV, which

agrees well with the peak in gure 3.2.

Measuring a bulk wafer sample was a means to conrm the operation of our instrument with a sample that was somewhat similar to the samples of interest (the nanowires), but also

much easier to measure by virtue of giving a very strong signal and being large enough that it was easy to align to. After we conrmed the operation of the instrument, we continued onto

measuring the nanowire samples.

3.3.2 Bare GaAs Nanowires

PL from the bare nanowires (NW40) measured at 4.2 K resolved three peaks in the spectrum,

as shown in gure 3.3. The highest energy peak appears at 1.512 eV, while the strongest

peak is at 1.494 eV. The expected position for the band edge is 1.519 eV, which we interpret

as meaning that the band edge has been depressed to 1.512 eV. The peak at 1.494 eV, we

interpret as being carbon impurity band, which also helps account for the energy of the band

edge. A large carbon content helps account for the depressed band edge energy, because the

band edge can be reduced by suciently high p-doping by carbon [96]. The peak at 1.457 eV

is a LO phonon replica of the band-edge peak [97]. Phonon replicas arise from a coupling

between excitons and phonons, in this case between a bound-exciton and the LO phonon,

Chapter3:PL 3.3PhotoluminescenceResults 23

1 . 4 9 4 4 e V

) 4 . 2 K s t

i 7 9 6 n m e x c i t a t i o n n U

y r a r t

i 3 6 . 5 m e V b r

A (

y t i s

n 1 . 4 5 7 9 2 e V e t X 1 0 1 . 5 1 1 7 1 e V n I

1 . 4 2 1 . 4 4 1 . 4 6 1 . 4 8 1 . 5 0 1 . 5 2 1 . 5 4 1 . 5 6 E n e r g y ( e V )

Figure 3.3: PL from bare GaAs nanowires (sample NW40) measured at 4.2 K with peaks labeled with tted energies. The vertical gray line is at 1.519 eV, marking the expected value for the bandgap of GaAs at 4.2 K. The annotated peak separation corresponds to the energy of an LO phonon. Excitation power was 0.023 mW. where the exciton scatters a phonon before recombining [98]. This consistent with strong

coupling between LO phonons and carbon acceptors, which has been documented before in

GaAs as noted in the discussion below [97, 99].

3.3.3 Core-Shell Nanowires

Figures 3.4 and 3.5 show low temperature PL from nanowires SAGA10 and SAGA11. Data was normalized by dividing by product of the excitation power and signal integration time,

and the minimum value was subtracted o to reference the curves to a common baseline.

Chapter3:PL 3.3PhotoluminescenceResults 24

3 80x10 Power (mW) 0.0058 0.0059 0.0062 60 0.0072 0.017 0.024 0.036 0.065 0.070 0.10 40 0.118 0.15 0.19 counts/mW/s 0.20 0.290

20

0 1.40 1.42 1.44 1.46 1.48 1.50 1.52 1.54 Energy (eV)

Figure 3.4: PL from core-shell nanowires SAGA10 measured at 4.2 K.

Because not all of the spectra covered the same energy range, the normalization is imperfect, but the alternative of not removing the background amplies the eect of the constant background (dark signal) in the spectra measured with lower power excitation, making the comparison between spectra worse.

The peak of the PL from the core-shell nanowires varies as a function of the excitation power and it is consistently lower energy than the band-edge of bulk GaAs at 4.2 K. The power dependence of the peak position shows the trend of increasing peak energy with increasing power. Figure 3.6 plots the peak position as a function of the excitation intensity for the data plotted in gures 3.4 and 3.5. The peak position was calculated as a weighted

Chapter3:PL 3.3PhotoluminescenceResults 25

Power (mW) 120x103 0.0055 0.011 0.035 0.152 100 0.22 0.24 0.25 80 0.27 0.28 0.36 0.38 60 0.47 counts/mW/s 0.56 0.76 1.19 40

20

0 1.40 1.42 1.44 1.46 1.48 1.50 1.52 1.54 Energy (eV)

Figure 3.5: PL from core-shell nanowires SAGA11 measured at 4.2 K.

P ξ ξ · I (ξ ) mean of the data, i.e. P where ξ is the energy and I (ξ ) is the intensity of the PL at ξ I (ξ ) that energy. It seems worth highlighting that the data presented in gure 3.6 is consistent for the most part, showing a fairly consistent trend over more than two orders of magnitude of excitation intensity – especially considering that the data was taken on two separate days, on dierent samples, with some inherent uncertainties about the actual power and number of nanowires in the spot due to the drift associated with the instrument.

The PL from SAGA11 (gure 3.5) shows a peak at approximately 1.44 eV. We only found this peak in PL from the SAGA11 nanowires. It is likely that this is signicant; SAGA11 were the core-shell nanowires grown with the lowest V:III precursor ratio among the samples of

Chapter3:PL 3.3PhotoluminescenceResults 26 core-shell nanowires we studied here. The lower ratio increases the likelihood of wurtzite growth, leading us to hypothesize that the peak at 1.44 eV is a result of the that form between zinc blende and wurtzite sections [89].

Chapter3:PL 3.3PhotoluminescenceResults 27

1.500

1.495

1.490

1.485

1.480 Weighted Mean Energy (eV)

1.475

1.470 5 6 7 8 9 2 3 4 5 6 7 8 9 2 3 4 5 6 7 8 9 0.01 0.1 1 Power (mW)

Figure 3.6: Peak position vs. power for the spectra plotted in gures 3.4 and 3.5 (blue squares correspond to the SAGA10 data and orange circles correspond to the SAGA11 data). Peak position was calculated as a weighted mean as described in the text.

Chapter3:PL 3.3PhotoluminescenceResults 28

3.4 Interpretation of Results and Relevant Eects

Table 3.1: Summary of peak assignments for dominant PL peaks.

Peak Position (eV) Assignment ≈1.49 Carbon defect band ≈1.485-1.51 DAP — peak increases in energy with increasing power ≈1.51 Band edge

The decreased band edge energy indicates a high level of doping in the nanowires [100, 101].

The bare GaAs nanowires have a PL spectrum matching that of p-doped GaAs, grown at

400 ◦C [102]. A sign of how much carbon is incorporated is that our spectra are completely

missing the free-exciton peak, which is clear when our spectra are compared with spectra

from high quality GaAs – this applies to both the bare GaAs nanowires and the core-shell

nanowires [79].

3.4.1 Bandgap Narrowing in Highly Doped Semiconductors

The nature of semiconductors is that a small impurity content leads to a signicant change in

their electrical properties through the addition of free carriers. At higher doping concentrations,

the band structure changes as is observed through a number of eects. The Burstein-Moss shift

is an eect observed in the absorption spectrum of highly doped semiconductors [103, 104].

The Burstein-Moss shift appears as a blue shift in the absorption spectrum of highly doped

semiconductors, while the bandgap of the semiconductor is in fact reduced by the high

concentration of dopants. While the Burstein-Moss shift is explained by the lling of states,

the reduction of the bandgap is instead related the introduction of states by the dopants.

Qualitatively, this can be argued from the standpoint that the bandgap of a semiconductor

arises from a description of the material as a perfect crystal. The introduction of impurities

Chapter3: PL 3.4InterpretationofResultsandRelevantEffects 29

creates disorder in the material, which will in eect be a perturbation on the model of the

material as a perfect crystal. The change in the shape of the band structure of doped GaAs is

related to band tails, also known as Urbach tails for Franz Urbach who rst documented their

presence in semiconductors [92, 105–107].

Another description is that at low concentrations, acceptors (or donors) create isolated

states within the bandgap. At higher concentrations, the wavefunctions of the acceptor states

begin to overlap, creating a band of energies, that eventually will overlap with the band edges;

in addition the decreasing the band edge, these eects modify the peak shape and broaden the

peaks [2, 107, 108].

3.4.2 Donor-Acceptor Pair Luminescence

Donor-acceptor pair transitions result from carriers created by optical excitation being

trapped at ionized donor (D+) and acceptor (A–) sites, producing D0 and A0 sites. When an

electron transitions from the neutral donor to the neutral acceptor, a photon is released; this

process can be written:

0 0 + − D + A −−→ }ω + D + A . (3.2)

The energy of }ω should clearly depend on the bandgap and the binding energies of the

carriers to the dopants, ED and EA. But more subtly, it is also dependent on the distance

between the donor and acceptor, because the Coulomb interaction between the ionized

dopants lowers the energy of the nal state of the system. Assuming a separation of R

2 between the donor and acceptor, the potential energy of the pair is −e /εr R, where e is the

charge of an electron and εr is the relative dielectric constant. The photon in equation (3.2)

Chapter3: PL 3.4InterpretationofResultsandRelevantEffects 30 has energy e2 }ω = Eg − ED − EA + . (3.3) εr R

What is not apparent from the above, is that R can only take on discrete values because the

donors and acceptors sit on particular lattice sites. At low enough temperatures (<2 K), the

DAP spectrum, in many semiconductors, can be resolved to lines corresponding to dierent

separations [2, 109].

Under low intensity excitation, there will be some average separation between donors

and acceptors participating in DAP luminescence. As the excitation intensity increases,

more donors and acceptors will participate, thus lowering the average separation between

participating pairs because the volume of the sample is unchanged; with the decreased

average separation, the DAP peak will increase in energy, until the intensity is high enough

that the sample reaches saturation [2, 110].

3.4.3 The Role of Hydrogen

We have been able to explain the decreased band edge energy in bare GaAs and core-shell

nanowires as resulting for carbon acting as a p-type dopant in the nanowires. But it isn’t

sucient to explain the dierences between the types of nanowire: the band edge of the

GaAs in the core-shell nanowires appears to be lower than the band edge of the bare GaAs

nanowires and the overall shapes of the spectra are very dierent. Both eects can be

explained in relation to the behavior of hydrogen in these nanowires.

Hydrogen is the second of the two elemental impurities that must be contended with in

MOCVD growth of GaAs as a consequence of the precursor chemistry. Pyrolysis of the

precursors creates elemental hydrogen during the growth, which can then incorporate into the

Chapter3: PL 3.4InterpretationofResultsandRelevantEffects 31

GaAs, either on an interstitial site, or as a complex with an adsorbed carbon atom [111, 112].

In either case, in carbon doped GaAs, the hydrogen has the same net eect of compensating

the p-doping, by either contributing an electron when it’s on an interstitial site or by forming

a neutral C−H complex that can be detected by infrared spectroscopy [113–116].

Heating GaAs above 530 ◦C quickly recovers much of the hole population compensated

by hydrogen [117, 118]. When annealed under an inert atmosphere at 600 ◦C, the hole

concentration can approach the carbon concentration within a few minutes in GaAs lms

over one micron thick [119, 120]. The high temperatures reactivate the carbon, by breaking

the C−H bonds and causing the hydrogen to diuse out [117, 118]. But above 600 ◦C, carbon

related defects begin to form in GaAs, lowering the carrier mobility [117, 118, 121].

Our core-shell nanowires, were heated to 650 ◦C for the growth of the shells. The growth

of the shells also sees the formation of atomic hydrogen through the breakdown of precursors, which can potentially diuse into the nanowires, but the rate of diusion into the material is

much smaller than the rate at which it diuses out [119, 120, 122].

3.5 Conclusions from PL

By comparing the peak positions in the bare GaAs wires to results from the literature, we can

tentatively attribute the strongest peak to carbon-bound exciton emission [79, 91, 123]. And

also just by comparison, the luminescence from the core-shell nanowires appears dominated

by donor-acceptor-pair (DAP) emission [2, 79]. The peak at 1.44 eV lines up with emission

from a heterojunction that forms at the interface between zinc blende and wurtzite GaAs [89].

As we will show below, these tentative assignments based strictly on comparison with

published results provide a good starting point for us to interpret our results.

Chapter3:PL 3.5ConclusionsfromPL 32

In order to explain our results, there are a number of factors we could focus on, most

related to defects. Control of defects in nanowires has been the topic of many recent papers

on nanowires, which have examined many factors in the growth of nanowires, including

growth temperature [124], precursor ow rate [79, 88, 125], seed particle [57, 126], and

substrates [127, 128]. In addition, we can look further, to the extensive literature on the

growth of GaAs lms for more insight on the formation of defects in GaAs. Point defects and

planar defects are common in III-V nanowires; in MOCVD grown nanowires, they can be

largely controlled via the growth parameters alone. Point defects are in the form of vacancies, which are never desirable, and atomic impurities, which can be a positive feature when the

goal is to dope the nanowires. Carbon and hydrogen are present during the nanowire growth:

the precursor chemistry has hydrogen and carbon in addition to the desired growth elements,

and H2 is used as a carrier gas during the growth. However, whether they are incorporated

into the nanowires depends on the growth conditions.

From the literature, we expect that wurtzite is most likely in nanowires grown at the

lowest V:III precursor ratios [56, 88, 89, 129]. Even though most of the literature discussing

MOCVD grown nanowires is based on experiments using AsH3 as a precursor instead of TBA

and ideal conditions depend on the growth system making comparisons between systems

inexact, we can use the estimate that TBA is ve times more ecient in delivering arsenic to

the reaction to have a rough comparison of the growth parameters with what has been

published [25, 88]. Based on the vefold estimate, the V:III precursor ratio of 5:1 with TBA would be approximately equivalent to a 25:1 ratio with AsH3, which is still much lower than

the ratios used by others to achieve defect free GaAs nanowires; for example, Joyce et al. use

ratios of 46:1 to 93:1 to produce nanowires free of planar defects [78].

Chapter3:PL 3.5ConclusionsfromPL 33

Other than precursor ratio, two parameters that aect the quality of the nanowires are

temperature and growth rate. The temperature of the core growth was 400 ◦C, which is

higher than the temperature found by others to minimize planar defects in zinc blende

nanowires, 375 ◦C [124]. Our core-shell nanowires are approximately 8 µm long: the cores were grown for 60 minutes, therefore, we estimate the growth rate as ≈ 130 nm/min.A

comparison of our spectra to those for GaAs nanowires grown at similar growth rates reveals

a great deal of similarity [79].

Based on the above, the story seems to be that the bare-GaAs nanowires have a signicant

amount of carbon in them, but are compensated by hydrogen. The core-shell nanowires

have much less hydrogen as a result of the high temperatures used for shell growth.

So, we interpret the spectra from the bare GaAs nanowires as being dominated by the

free-electron—neutral-acceptor peak (e, A0)[130]; and the spectra from the core-shell

nanowires as being dominated by donor-acceptor-pair (DAP) luminescence.

Chapter3:PL 3.5ConclusionsfromPL 34

Chapter 4: Resonance Raman of GaAs/AlGaAs Nanowires

4.1 Raman and Resonance Raman Spectroscopy

Figure 4.1: Feynman diagram for the Raman process. Based on Fig. 7.28 in Yu and Cardona [2].

The Raman eect is named after C. V. Raman, who won the Nobel Prize in Physics in 1930

for the discovery in 1928. Raman described the eect as an optical analog for Compton

scattering. In the rst order Raman process, three events occur: a photon is absorbed creating

an excited state in the material, the excited state interacts with an elementary excitation, and

a photon is emitted as the excited state relaxes. Though it is not a very strong eect, the

Raman eect is an extremely powerful analytical tool, providing a way to measure many

types of non-optical excitations in a material [131, 132].

The power of Raman spectroscopy is that it is a means to measure the energies of

elementary excitations in a material, which do not themselves participate directly in radiative

transitions. Most commonly, Raman spectroscopy is used to measure the energy of phonons

(quantized vibrations) in a material. Not all phonons will Raman scatter and which do scatter

depends on the experimental geometry. The dependence on experimental geometry is a result

of the symmetry of the process that leads to selection rules, however these rules are often

relaxed under resonant conditions. 35

Figure 4.1 shows a Feynman diagram for the rst order Raman eect. A Feynman diagram

is a graph, the edges of which are called propagators and represent (quasi–)particles that

interact with each other at the vertices, with each vertex contributing a Hamiltonian to a

product of perturbed Hamiltonians [2, 133, 134]. Figure 4.1 shows a photon (dashed line on

the left), being absorbed to form an electron-hole pair (double line), which emits a phonon

(wavy line), and nally emits a photon (dashed line on the right) [2]. This corresponds to the

equation:

2 X 0 0 hi | HeR | n ihn | He-ion | nihn | HeR | ii P ≈ × δ [}ωi − }ω0 − }ωs] , (4.1) [ ω − (E − E )][ ω − ω − (E 0 − E )] n,n0 } i n i } i } 0 n i

0 where n and n are intermediate states with energies En and En0 respectively, HeR and He-ion

are the Hamiltonians for the interaction of an electron with electromagnetic radiation and

of an electron with a phonon respectively, }ωi and }ωs correspond to the energy of the

incoming and scattered photons respectively, }ω0 is the energy of the elementary excitation

(e.g. phonon), and the delta function ensures that the sum is over combinations of states that

conserve energy [2].

Resonance Raman spectroscopy (RRS) is an experimental technique that utilizes Raman

spectroscopy by modifying the Raman cross-section by tuning the excitation energy to

resonance with energy transitions in the material. Resonance Raman spectroscopy can be

used to measure the Raman spectrum of a sample as a function of the excitation energy; or

RRS can be used to enhance the cross-section so that weak Raman peaks can be observed

experimentally, such as electronic transitions between states in a quantum well or plasmon

excitations [135–137]. Raman scattering is an inherently low eciency process compared

Chapter4: ResonanceRaman 4.1RamanSpectroscopy 36 with luminescence, which is something that Raman himself pointed out as evidence that it was a scattering process, stating “That the eect is a true scattering and not a uorescence is

indicated in the rst place by its feebleness in comparison with the ordinary scattering, and

secondly by its polarisation, which is in many cases quite strong and comparable with the

polarisation of the ordinary scattering” [138]. The expression for the Raman process involves

the product of three Hamiltonians, each of which are small; but it is possible to minimize the

denominator in equation (4.1) by tuning the excitation wavelength (}ωi) to be close to

a transition in the material, that is, exciting the sample resonantly. This is the basis of

resonance Raman spectroscopy (RRS).

Near a discrete energy level, Ea, the Raman scattering probability is approximately

2 h0 | HeR(ωs) |aiha | He-ion |aiha | HeR(ωs) | 0i P ≈ + C , (4.2) (Ea − }ωi − iΓa)(Ea − }ωs − iΓa)

where the term iΓa is a damping constant included to account for the nite lifetime of the

intermediate state. The lifetime (τ ) is related to Γa by Γa = }/τ . The non-resonant contribution

to the scattering is accounted for in the constant (C), and the summation and delta functions

have been removed by summing over the intermediate state [2].

In the denominator of equation (4.2), there are two terms in the denominator, leading to

two resonant conditions. When the incoming light is in resonance with Ea it is an incoming

resonance; when the scattered light is in resonance with Ea, it is an outgoing resonance.

These two conditions are illustrated in gure 4.2. The equation for the cross-section suggests

that the incoming and outgoing resonances should be of similar magnitude, as is the case in

some high quality samples [139–142]. Deviations from symmetric resonances are the result of

Chapter4: ResonanceRaman 4.1RamanSpectroscopy 37 a variety of factors, mostly related to material defects and Raman processes mediated by exciton states [2, 143]. For example, Kusch et al. observed a very symmetric resonance in

GaAs nanowires grown using MBE [142].

En }ω0 Ea Ea }ω0 En

}ωi }ωs }ωi }ωs

Incoming Outgoing resonance resonance

Figure 4.2: A schematic of the transitions involved in incoming and outgoing resonances for a Stokes process.

E CB

Eg

0 HH ∆SO LH

SO k0

Figure 4.3: A schematic of the band structure of GaAs, showing the direct band gap between the conduction band and the light and heavy hole bands (Eg) and the split-o band gap (∆SO).

Our motivation for using resonance Raman spectroscopy was based on our hypothesis that growth conditions aected the bandstructure of the nanowires. Based on the assumption that

Chapter4: ResonanceRaman 4.1RamanSpectroscopy 38 we were looking for changes in the band structure, this was an appropriate choice of method,

particularly there was a hope that we were working with a quantum well structure [137].

The band structure of GaAs has two direct transitions from the lowest point in the

conduction band, which is at Γ = 0. One is to the valence band and the other is to the split-o

band; the split-o band is the result of the coupling between electron spin and orbital angular

momentum via the spin-orbit interaction [2]. At 4 K the direct bandgap of GaAs has energy,

Eg = 1.517 eV; while the split-o band is of energy, Eg + ∆SO = 1.851 eV [94].

One of the strengths of RRS is that it can be applied to individual nanostructures, or even

used to map structure within a single nanowire [86]. There are some recent examples of RRS

being used to characterize GaAs nanowires published by Ketterer et al., where they used

RRS to measure the bandgap and split-o band energies for zinc-blend and wurtzite GaAs

nanowires [144, 145]. Our experimental conditions made it dicult to estimate the number of

nanowires in the laser spot with any accuracy, because we were most successful measuring

RRS from mats of nanowires that placed a large number of randomly oriented nanowires in

the laser spot. The optics were also non-ideal – because the microscope objective we used is a

plan view objective, the cryostat window cause a signicant amount of spherical aberration –

making accurate calculation of the spot size dicult. But based on our experience with the

system, the spot size of the laser was approximately 5 µm, so our lowest estimates are that

hundreds of nanowires were in the laser spot.

Generally Eg +∆SO is the preferred transition to use for RRS, though exciting at Eg

directly is a viable technique [137, 146]. The primary disadvantage of exciting near the

bandgap is that the luminescence from the direct bandgap will then overlap with the Raman

spectrum. This makes the detection of the Raman spectrum more dicult because of how

Chapter4: ResonanceRaman 4.1RamanSpectroscopy 39 weak a process Raman scattering is, particularly when it is non-resonant. The background

luminescence, imparts the requirements that the detector used have a large dynamic range

and that subsequent data analysis include the removal of the background luminescence.

4.2 Experimental Procedures

Sample preparation was identical to the preparation described in section 3.2.1, with the

nanowires dispersed on SiOx on Si substrates which were then mounted on the cold nger

of the cryostat. For each Raman spectrum making up the resonance Raman spectra, the

following procedure was used:

1. The CCD was placed in optical train, blocking the monochromator input.

2. Using the grating mounted to the rotation stage, the laser was realigned to the sample

by observing the laser with the CCD.

3. The laser’s power was measured using the Newport 883-SL sensor and recorded in

LabSpec.

4. The laser wavelength was recorded in LabSpec.

5. The CCD was removed from the optical train and then the measurement was started.

Prior to acquiring any spectra, there were several hours of warm-up and stabilization time

for the instrumentation, allowing the lasers time to warm-up, the cryostat time to reach a

stable temperature, and the mechanical parts of the system to settle.

The analysis procedure began with each individual Raman spectrum t using a Lorentzian with a linear baseline correction (f (x) = fLOR(x) + m · x); the linear baseline was added to

Chapter4: ResonanceRaman 4.2ExperimentalProcedures 40 account for the background uorescence that overlapped the Raman spectrum as a result of our choice to excite the samples at Eg rather than the split-o resonance (Eg +∆SO). The

Lorentzian was chose as the tting function because it is the solution to a damped oscillator equation. Fits were calculated initially in IgorPro, but then recalculated in Origin.

The form of the Lorentzian t was:

2A w fLOR(x) = y0 + 2 2 , (4.3) π 4(x − x0) + w

from which the parameters A, w, and x0 were taken as the peak area, line width, and peak position respectively. These were then plotted as a function of the excitation energy, with the error bars based on the standard error from the t parameters.

Figure 4.6 shows the output of the analysis procedure applied to the data taken using the

SAGA06 nanowires. The set of three plots shows the magnitude, width, and position of the

LO Raman peak plotted as a function of the excitation energy.

Chapter4: ResonanceRaman 4.2ExperimentalProcedures 41

Excitation Energy (eV) 1.625 1.6165 1.6081 1.5998 1.5916 1.5835 1.5754 1.5635 1.5556 1.5498 1.5479 1.545 1.5421 1.5402 1.5383 1.5345 1.5326 1.5307 1.5269 1.525 1.5231 1.5194 1.5176 1.5157 1.512 1.5083 1.5065 1.5047 1.501 1.4992 1.4974 1.4938 1.492 1.4902 1.4866 Intensity (Arbitrary Units)

1.46 1.48 1.50 1.52 1.54 1.56 1.58 Energy (eV)

Figure 4.4: Initial data replotted by transforming wavenumbers to absolute energy; wavenumber w measured using excitation wavelength λ, E = E(λ) − E(w).

Chapter4: ResonanceRaman 4.2ExperimentalProcedures 42 ) s t i n U

y r a r t i b r A (

y t i s n e t n I

2 8 0 2 8 5 2 9 0 2 9 5 3 0 0 3 0 5 3 1 0 R a m a n S h i f t ( c m - 1 )

Figure 4.5: An example of the RRS data being t in OriginPro.

Chapter4: ResonanceRaman 4.2ExperimentalProcedures 43

Laser Excitation Wavelength (nm) 820 800 780 760

1000 800 600 400 200

LO peak mag. (Arb. Units) 0 )

-1 2.0

1.5

1.0

0.5 LO peak width (cm 0.0 293.6 ) -1 293.4 293.2 293.0 292.8 292.6 292.4 LO peak position (cm

1.48 1.52 1.56 1.60 1.64 Laser Excitation Energy (eV)

Figure 4.6: RRS data from sample SAGA06, measured at 78 K.

Chapter4: ResonanceRaman 4.2ExperimentalProcedures 44

4.3 Results and Interpretation

4.3.1 Summary of Results

Following the previously described analysis, results in gures 4.6 to 4.9, showing the output

of the RRS for SAGA06, SAGA11, NW40, and the GaAs wafer.

3 14x10 12 10 8 6 4 2

LO Peak Mag. (Arb. Units) 0

4 ) -1 3

2

1 LO Peak Width (cm 0

296.0 ) -1 295.8 295.6 295.4 295.2 295.0 294.8 LO Peak Position (cm 294.6

1.48 1.50 1.52 1.54 1.56 Excitation Energy (eV)

Figure 4.7: SAGA11 RRS data

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 45

300 200 100 0 (Arb. Units) LO Peak Mag. -100

3.0 )

-1 2.0

(cm 1.0

LO Peak Width 0.0 293.4

) 293.2 -1 293.0 (cm 292.8 LO Peak Pos.

120 80 40

(Arb. Units) 0 TO Peak Mag.

0.6

) 0.4 -1 0.2 (cm 0.0 TO Peak Width

270.2 ) 270.1

-1 270.0 269.9 (cm 269.8

TO Peak Pos. 269.7

1.48 1.50 1.52 1.54 1.56 Excitation Energy (eV)

Figure 4.8: NW40 RRS data

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 46

120

80

40 (Arb. Units) LO Peak Mag.

0 293.7 293.6 293.5 ) -1 293.4 (cm 293.3 LO Peak Pos. 293.2

160

120

80 (Arb. Units) TO Peak Mag. 40

270.5 270.4 )

-1 270.3

(cm 270.2 TO Peak Pos. 270.1

1.48 1.50 1.52 1.54 1.56 Excitation Energy (eV)

Figure 4.9: GaAs Wafer RRS

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 47

4.3.2 Incoming vs. Outgoing Resonance ) s t i n U

. b r ) s A t ( i

y n t i U s

n y e r t a n I r t i b r 0 0 2 5 5 0 7 5 0 0 2 5 A . 5 . 5 . 5 . 5 . 6 . 6 (

1 1 1 1 1 1 y t

i E x c i t a t i o n E n e r g y ( e V ) s n e t n I

0 0 0 0 0 4 6 4 8 5 0 5 2 5 4 1 . 1 . 1 . 1 . 1 . O u t g o i n g E n e r g y ( e V )

Figure 4.10: The squares show the amplitude of the RRS signal from SAGA06 measured at 78 K, plotted as a function of outgoing energy, i.e. Eex − ELO. The solid line shows the PL signal measured under the same conditions.

The RRS spectra from the core-shell nanowires show a peak from the outgoing resonance, with little if any contribution from the incoming resonance. The dominance of the outgoing

resonance is notable, but not altogether surprising. There are a number of published resonance

Raman studies of GaAs – including a study of GaAs nanowires – that observe both incoming

and outgoing resonances of the same order of magnitude. There has been debate in the

literature centered on the asymmetry between the two resonances; however, we conclude

that the purely outgoing resonance that we observe is consistent with the conclusion that the

nanowires contain a large number of point defects and strongly supports the conclusion that

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 48 the resonance is with an excitonic state [139–142, 147, 148].

An eect closely related to the coupling of an outgoing resonance to an exciton state, is

the cascade eect, which manifests as the appearance of multiple phonon peaks that can be

of comparable intensity to the luminescent background, when the excitation source is

tuned a multiple of ELO above certain transitions. The cascade eect was rst observed

in CdS from which large numbers (n ≈ 10) of phonon peaks can be observed at room

temperature [149, 150]. This eect also has been observed in GaAs nanowires; in both cases it

is due to the strong coupling between the bound exciton state and the LO phonon [2, 151].

In the cascade model, carriers are excited to a virtual state and then relax through a series

of allowed states, rather than through virtual states, making it competitive with luminescence,

because it avoids intermediary virtual states [2, 152, 153]. The cascade model has been used

to explain a number of exciton related spectral phenomena [143, 154]. It has been invoked in

arguing that at an outgoing resonance with an excitonic state, resonance Raman is equivalent

to hot-luminescence where photons are absorbed to a resonant state and then re-radiated;

though this stance is somewhat controversial, at outgoing resonances, resonant Raman

scattering is very similar to a luminescent process [155, 156]. The cascade model can also be

used, substituting elastic scattering by impurities rather than an LO phonon, to help explain

the strength of outgoing over incoming resonance in doped materials [2, 157].

This correlates with our PL results (gure 3.3), where we observed a phonon replica peak

resulting from coupling between the bound exciton state that dominated the spectrum and

LO phonons.

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 49

GaAs Wafer

Bare GaAs

Intensity (Arbitrary Units) Nanowires

Core-Shell Nanowires 310300290280270260250 Raman Shift (cm−1)

Figure 4.11: Selected spectra showing the lack of a TO peak in the core-shell nanowire data. This eect was independent of the excitation power used, the spectra shown were selected only because they include the full region from 250 cm−1 to 310 cm−1 in a single contiguous scan, rather than being broken up into two separate scans as was often the case.

4.3.3 The Missing TO Mode

None of the core-shell nanowire resonance Raman data has a detectable TO phonon peak.

This is a somewhat surprising result; in non-resonant Raman of GaAs, the TO phonon is the

stronger mode. It does appear in the RRS spectra from the GaAs wafer and bare GaAs

nanowire samples, showing that it was not necessarily an instrumental artifact. And in

non-resonant Raman spectra (gure 5.9), only the TO mode is observed in the core-shell

nanowires.

The lack of a TO peak in the resonant spectra seems to be an extreme example of

the phenomena of the LO peak being enhanced in carbon doped GaAs. The LO mode is

preferentially enhanced over the TO mode because of the strong coupling between carbon

acceptors and excitons, through the Fröhlich mechanism [97, 99, 158].

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 50

(a)

(b)

Figure 4.12: LO and TO phonons

In polar materials, such as GaAs, longitudinal phonons couple with electric elds

more strongly than transverse phonons. This is because a longitudinal phonon involves

planes of atoms changing distance from each other (gure 4.12a), which induces a charge

transfer between the atoms giving rise to a macroscopic electric elds, while transverse

phonons involve planes of atoms moving within their planes, but because their distance

does not substantially change relative to each other (gure 4.12b), no net polarization is

induced [2, 158, 159].

More formally, the electric eld resulting from an LO phonon gives rise to an electric eld,

ELO = −FuLO, where: q   − 2 −1 − −1 F = 4πN µωLO ε∞ ε0 . (4.4)

In equation (4.4), uLO is the atomic displacement due to the LO phonon, ε∞ and ε0 are the

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 51

high and low frequency limits for the dielectric constant, N is the number of atoms per unit volume, and µ is the reduced mass of the ions [2, 158]. The Fröhlich interaction is responsible

for a number of eects near resonance that lead to otherwise forbidden scattering [153, 160].

4.3.4 Variation of the LO Peak Position

Another eect we saw, but could not satisfactorily explain, was an apparent change in the

peak position of the LO phonon in the core-shell nanowires, but not the bare GaAs nanowires.

A possible explanation is that there is wurtzite present in the nanowires because wurtzite has

a slightly dierent resonant energy and due to strain may have a slightly dierent peak

position [144]. However, this would not explain why it was observed only in the core-shell

nanowires. Another possibility, is that it originates with the emission of acoustic phonons by

excitons near the outgoing resonance energy, which has been shown to give rise to a peak

splitting at a dierent resonance [99].

Because the resonance peak for the core-shell nanowires is at at a dierent energy than

anything else we measured, it is tempting to explain away the change in the LO peak position

as stemming from an experimental artifact. In order to check this possibility, the laser was

measured at a number of dierent energies, then the oset for each energy was calculated as P P a center of mass (i.e. ∆ = (y · x)/ y), with the results shown in gure 4.13. Thus we have

some certainty that it is a real eect – although it is somewhat disconcerting that the eect

seems to be independent of temperature, but it may also be that we are not sampling enough

energies to identify what is changing.

There are some weaknesses with the hypothesis that the change in LO peak position is a

result of wurtzite mixed in the nanowires. The LO phonon energies are not very dierent

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 52

Excitation Energy (eV) 1.5655 1.5615 1.5556 1.544 1.5402 1.5275 1.5269 1.525 0.25 1.5192 1.5157 1.5138 1.5078 1.5019 0.20 ) -1

0.15 Offset (cm 0.10 Intensity (Arbitrary Units) 0.05

0.00 -10 -5 0 5 10 1.50 1.52 1.54 1.56 -1 Raman Shift (cm ) Excitation Energy (eV) (a) (b)

Figure 4.13: (a) The laser line measured at a number of dierent excitation energies, measured through closed slits so as to avoid damaging the PMT. (b) The calculated oset for the laser at the dierent wavelengths.

between wurtzite GaAs and zinc-blende GaAs. However, there is some uncertainty as to what

ensemble eects come into play because these measurements sampled a large number of

nanowires. The dierence in bandgap between the phases is less than 0.5 meV [145]. If there were wurtzite, there would be interface strain, which could in part account for the shift in

energy. Also, this eect does not appear in the bare GaAs nanowires, which should contain at

least as much wurtzite as the core-shell nanowires. This eect appears in all the core-shell

nanowires, independent of growth conditions. Therefore, this suggests that the cause may be

connected to the way that the core-shell nanowires dier from the bare nanowires. As was

discussed in the previous chapter, the core-shell nanowires contain approximately the same

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 53 amount of carbon, but because of the elevated temperatures used for the shell growth, they contain more active carbon. So, it makes sense to look at eects associated with carbon doping in GaAs to identify whether any may account for the eect we observe. However, the only similar eect that we could nd, was scattering by longitudinal acoustic (LA) phonons enhanced by carbon doping documented by Huang and Ulbrich. This is an attractive explanation at rst sight, but it predicts a much larger shift than we observed: following their calculations predicts a 15 cm−1 shift [99].

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 54 Head Power Meter Mirrors CCD Camera Sample Coldfinger Objective Lens Cryostat Diagram of instrument. Lens Focusing Ti:SaLaser Grating on Figure 4.14: Rotation Stage Monochromator Laser + Has Been PMT Simplified Ar Optical Path

Chapter4: ResonanceRaman 4.3ResultsandInterpretation 55

Chapter 5: Single Nanowire Raman Spectroscopy

In order to better understand the resonance Raman spectra, we took additional Raman

measurements on a single nanowire, using a commercial Raman system – a Renishaw inVia

spectrometer, which we were kindly granted access to by Volker Presser and the W. M. Keck

Institute for Attouidic Nanotube Based Probes. This provided us a way to better understand

the missing TO peak in the resonance Raman spectra and look for evidence of wurtzite

in the nanowires. The Renishaw inVia system we used is integrated with a microscope

making it easy see the sample being measured and is stable enough to enable single nanowire

measurements that take several several hours.

5.1 Experimental Rationale

Because wurtzite and zinc blende structures have dierent crystal symmetries, their Raman

tensors dier from each other. In bulk, GaAs only forms the zinc blende structure, but in

nanowires it can take on the wurtzite structure, which is the structure of a number of III-V

materials including GaN. Wurtzite GaAs (w-GaAs) consists of two types of bilayers stacked

(ABABAB...), while in the zinc blende phase it consists of three bilayers (ABCABC...) as is

illustrated in gure 5.1.

Because the wurtzite unit cell length is the [0001] direction is twice that of the zinc

blende unit cell in the [111] direction, the Brillouin zone has half the length in k-space, so the

phonon dispersion curves are in eect “folded,” as is seen with the dispersion curves of 56

Figure 5.1: The dierence in stacking orders for (a) zinc-blend GaAs and (b) wurtzite GaAs leads to additional peaks in the Raman spectrum of wurtzite GaAs due to (c) the folding of the Brillouin zone. (Reprinted with permission from Zardo et al., Phys. Rev. B 80, 245324 (2009) Copyright 2009 American Physical Society)[129]. superlattice structures. This gives rise to new Raman peaks corresponding to the phonon modes at zone center, illustrated in gure 5.1c [129].

Chapter5: SingleNanowire... 5.1ExperimentalRationale 57

5.2 Procedure

Sample

Objective

λ/2 waveplate Analyzer

Dispersion, detection, etc Microscope

Figure 5.2: Diagram of experimental setup, the line with the arrows represents the path of the laser light, which includes segments in which incoming and scattered light overlap.

Nanowires were dispersed in isopropyl alcohol, drops of which were then placed on a

pyrex slide that was heated to evaporate o the solvent. The nanowires were then located

optically using the microscope integrated into the spectrometer; an example image is shown

in gure 5.9b.

Polarization was controlled using a λ/2-waveplate in the optical train positioned so that

both the light going to the microscope and the scattered light would pass through it, so it would return to the spectrometer in the same polarization regardless of the λ/2-waveplate

position to ensure that the experiment was not aected by detector polarization sensitivity.

Chapter5:SingleNanowire... 5.2Procedure 58 ) ) s s t t

i 9 0 i n n U U

. . b b r 6 0 r A A ( (

y y t t i i

s 3 0 s n n e e t t n n I I

0

2 0 0 2 2 0 2 4 0 2 6 0 2 8 0 3 0 0 2 0 0 2 2 0 2 4 0 2 6 0 2 8 0 3 0 0 R a m a n S h i f t ( c m - 1 ) R a m a n S h i f t ( c m - 1 )

Figure 5.3: Polarized Raman spectra from a single nanowire, grouped by the position of the polarization analyzer, with the angle of the λ/2-waveplate labeled, less a constant oset of 50°.

Scattered light passed through an analyzer so that either parallel or perpendicularly polarized scattered light was selectively collected.

The λ/2-waveplate was previously aligned to the mount so that the angle of the waveplate’s axis was known so that by rotating it θ/2, the polarization of the light rotated by θ. The scattered light was passed through a linear polarizer, which served as an analyzer by ltering one polarization at a time.

Chapter5:SingleNanowire... 5.2Procedure 59

5.3 Theoretical Considerations

The expected polarization dependence of the Raman scattering can be calculated from the

Raman tensor, R, according to the relation

2 Is ∝ eˆi · Re · eˆs , (5.1)

where Is is the intensity of the Raman scattered radiation and eˆi and eˆs are unit vectors in the

direction of the incident and scattered radiation, respectively – in the notation of Wu et al.

[161]. For the zinc-blende crystal structure, we can write Re for optical phonons with atomic

displacements in the h100i, h010i, and h001i crystallographic directions, corresponding to the vectors eˆ1, eˆ2, and eˆ3, as [2, 132, 161]:

      0 0 0 0 0 d 0 d 0             Re(eˆ1) = 0 0 d , Re(eˆ2) = 0 0 0 , Re(eˆ3) = d 0 0 . (5.2)             0 d 0 d 0 0 0 0 0

In the backscattering conguration, conservation of momentum requires that the wavevector of the light and the scattering phonon be parallel. Assuming that the direction of

the scattering is in the eˆ1 direction, we require LO phonons to have q in the eˆ1 direction,

necessitating atomic displacements parallel to q, so the intensity of the Raman scattering

follows

2 LO Is ∝ eˆs · Re(eˆ1) · eˆi . (5.3)

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 60

xˆ1

0 φ xˆ1 211¯ 

[111] xˆ3

 ¯  011 laser xˆ2 0 θ xˆ2

Figure 5.4: Schematic showing the scattering conguration and illustrating the vectors used in the calculation of the expected Raman scattering intensity.

While the atomic displacements for a TO phonon are orthogonal to q, so we have:

2 2 TO Re(eˆ2) Re(eˆ3) Is ∝ eˆs · √ · eˆi + eˆs · √ · eˆi . (5.4) 2 2

In our experimental setup, the nanowire was selected because its orientation was closely

aligned to the polarization of the laser without any rotation from the waveplate. In the

laboratory frame of reference, we consider the scattering to be along the xˆ1 direction. For

convenience, we can pick xˆ3 to be the direction corresponding to the polarization of the

unrotated laser light. We then choose xˆ2 so that it is orthogonal to both xˆ1 and xˆ3. The

scattering conguration is illustrated in gure 5.4.

To translate between the laboratory coordinates (xˆm) and the crystal coordinates (eˆm), we

must make some assumptions about the crystallographic orientation of our nanowire, because we do not have TEM data from the same nanowire, which would allow us to unambiguously

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 61

determine its orientation. However, based on previous work by our collaborators, as well as

more general aspects of GaAs nanowire growth, we expect that the nanowire’s growth

direction is h111i and that the facets are of the 211¯ family [22, 23]. Thus, we will begin by  calculating the Raman tensor for scattering from a 211¯ surface, using the orientation      xˆ1 = 211¯ , xˆ2 = 011¯ , and xˆ3 = [111]. Later, we will account for the possibility that the

nanowire is not sitting ush on a face, so that it is rotated about the growth axis. From xˆm  we can build the matrices

     −2 1   −2 1 1  √ 0 √  √ √ √   6 3  6 6 6  1 −1 1  T  −1 1  Be = √ √ √  and Be =  0 √ √  . (5.5)  6 2 3  2 2  1 1 1   1 1 1  √ √ √  √ √ √  6 2 3 3 3 3

Which allow us to calculate the Raman tensor in the xˆm basis by:

   0 T  1  Re (xˆ1) = Be ·  √ −2Re(eˆ1) + Re(eˆ2) + Re(eˆ3)  · Be (5.6a)  6     0 T  1  Re (xˆ2) = Be ·  √ − Re(eˆ2) + Re(eˆ3)  · Be (5.6b)  2      0 T  1  Re (xˆ3) = Be ·  √ Re(eˆ1) + Re(eˆ2) + Re(eˆ3)  · Be. (5.6c)  3 

We describe the incident polarization by its angle θ with respect to xˆ1, thus:

   0      eˆi =  θ  . (5.7) sin    cosθ

The experimental design was intended to produce the same polarization for the incident

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 62 and scattered light, by having the light pass through the waveplate twice – once en route to the sample to rotate the polarization and then again as it was directed back to the spectrometer to counter-rotate the polarization, as depicted in gure 5.2 – so in the parallel conguration the scattered polarization is aligned with the incident polarization, i.e. eˆs = eˆi.

In the perpendicular conguration, the collected light is polarized perpendicularly to the incident polarization, thus:

     0   0      k   ⊥   eˆ = eˆi =  θ  and eˆ =  θ  . (5.8) s sin  s  cos      cosθ − sinθ

0 Substituting Re (xˆm) for Re(eˆm) in equations (5.3) and (5.4), will enable us to calculate the expected scattering as a function of angle. We can calculate these by substituting equations (5.2) and (5.5) into equation (5.6):

 √ √   √  − 6 0 − 3  0 6 0      0 1  √  0 1 √ √  Re (xˆ1) = ·  0 6 0  , Re (xˆ2) = ·  6 0 − 3 , 3   3    √   √  − 3 0 0   0 − 3 0   √  (5.9) − 3 0 0    0 1  √  Re (xˆ3) = ·  0 − 3 0 . 3    √   0 0 2 3

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 63

Using equations (5.7) to (5.9) we nd for the parallel conguration:

√ k 6 2 eˆs · Re(xˆ1) · eˆi = sin θ , (5.10a) 3 √ k −2 3 eˆs · Re(xˆ2) · eˆi = √ cosθ sinθ , (5.10b) 3 √ √ − 3 2 3 eˆk · Re(xˆ ) · eˆ = sin2 θ + cos2 θ. (5.10c) s 3 i 3 3

And for the perpendicular conguration:

√ ⊥ 6 eˆs · Re(xˆ1) · eˆi = sinθ cosθ , (5.11a) 3√ ⊥ − 3 eˆs · Re(xˆ2) · eˆi = , (5.11b) √3 3 eˆ⊥ · Re(xˆ ) · eˆ = cosθ sinθ. (5.11c) s 3 i 3

If the scattering is not in fact from a 211¯ face, because the nanowire is rotated about  the growth axis, we can account for that by rotating our basis appropriately to account for the

misalignment between the laboratory and crystallographic coordinate systems. First, it is

useful to generalize our notation a bit, so continuing in the notation used by Wu et al. [161],   we describe the general transformation between the lab coordinate system xˆ1, xˆ2, xˆ3 and

 0 0 0  the crystallographic coordinate system xˆ1, xˆ2, xˆ3 .

0 X X 0 xˆk = αkmeˆm, xˆk = βkmxˆm, (5.12) m m

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 64 where k,m = 1, 2, 3. From this we can dene matrices:

  α α α   11 12 13   Be = α21 α22 α23 (5.13a)     α31 α32 α33

    β β β  cos φ − sin φ 0  11 12 13       Te = β21 β22 β23 = sin φ cos φ 0 . (5.13b)         β31 β32 β33  0 0 1

The matrix Te represents a rotation about the xˆ3 axis by an angle φ, which we take to be the   angle between xˆ1 and 211¯ .

0 We can calculate the Raman tensor in an arbitrary direction xˆk by:

0 X Re(xˆk ) = αkmRe(eˆm). (5.14) m

0 And if we transform into the coordinate system xˆk , we have: n o

0 0 X −1 Re (xˆk ) = αkmBe · Re(eˆm) · Be . (5.15) m

Using this we can generate a set of plots of the intensity of the Raman peak given dierent values of φ. Due to symmetry, we need only examine the range 0◦ ≤ φ ≤ 30◦ for rotations

about the xˆ3 axis.

However, for reasons that are not clear, the instrument instead behaved as if the light only

passed through the waveplate once – as determined by measuring the response of a Si wafer with a known orientation [Livneh, private communication]. Thus, as we analyze the data,

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 65

TO:

LO:

φ = 0◦ φ = 10◦ φ = 20◦ φ = 30◦

Figure 5.5: Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = eˆi, where φ is the angle between the [211¯ ] and xˆ1.

k rather than assume that eˆs = eˆi for the parallel conguration, we will instead assume that the polarization of the scattered light is unaected by the rotation of the waveplate, i.e.

    0 0     k   ⊥   eˆ = xˆ3 = 0 and eˆ = xˆ2 = 1 . (5.16) s   s       1 0

Repeating the calculations as before, we nd for the parallel conguration:

xˆ3 · Re(xˆ1) · eˆi = 0, (5.17a) √ 2 3 xˆ3 · Re(xˆ2) · eˆi = cosθ , (5.17b) 3√ − 3 xˆ · Re(xˆ ) · eˆ = sinθ. (5.17c) 3 3 i 3

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 66

TO:

LO:

φ = 0◦ φ = 10◦ φ = 20◦ φ = 30◦

Figure 5.6: Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration where eˆs is set orthogonal to eˆi, where φ is the angle between the [211¯ ] and xˆ1.

And for the perpendicular conguration:

√ 6 xˆ2 · Re(xˆ1) · eˆi = sinθ , (5.18a) 3√ − 3 xˆ2 · Re(xˆ2) · eˆi = cosθ , (5.18b) √3 − 3 xˆ · Re(xˆ ) · eˆ = sinθ. (5.18c) 2 3 i 3

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 67

TO:

LO:

φ = 0◦ φ = 10◦ φ = 20◦ φ = 30◦

Figure 5.7: Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = xˆ3, where φ is the angle between the [211¯ ] and xˆ1.

TO:

LO:

φ = 0◦ φ = 10◦ φ = 20◦ φ = 30◦

Figure 5.8: Plots of the predicted dependence of the Raman signal on the polarization angle of the incoming light for the conguration eˆs = xˆ2, where φ is the angle between the [211¯ ] and xˆ1.

Chapter5: SingleNanowire... 5.3TheoreticalConsiderations 68

5.4 Results and Discussion ) s t i n U

. b r A (

y t i s n e t n I

1 0 0 2 0 0 3 0 0 4 0 0 5 0 0 R a m a n S h i f t ( c m - 1 )

(a) (b)

Figure 5.9: (a) Raman from a single SAGA06 nanowire. (b) Image of the nanowire from the microscope’s CCD.

Figure 5.9a shows a typical Raman spectrum of a single SAGA06 nanowire (shown in

gure 5.9b) measured using a 633 nm laser as the excitation source. The features at low wavenumber end of the spectrum (< 150 cm−1) are artifacts from the lter used to block the

back-reected (i.e. elastically scattered) laser light from reaching the CCD. The feature from

≈ 190 cm−1 to ≈ 290 cm−1 (1.935 eV to 1.923 eV) appears to be a luminescent band. This

feature has nearly the same energy as the X -valley of the bandstructure of GaAs; though this

may be coincidental, it leads us to speculate that this luminescence is from the X-valley

because the excitation energy is nearly resonant with it (Eex = 1.959 eV and EX = 1.90 eV).

Additional experiments would be necessary to clarify the origin of this feature, such as

observing the eect of temperature on the luminescent band or a PLE experiment using a

laser that could be tuned in the 1.9 eV range.

A point of sharp contrast between the Raman taken out of resonance and the resonance

Chapter5: SingleNanowire... 5.4ResultsandDiscussion 69

2 6 8 c m - 1 ) s t i n U

y r a r t i b r

A (

y t i s n e t n I

1 5 0 2 0 0 2 5 0 3 0 0 R a m a n S h i f t ( c m - 1 )

Figure 5.10: Zoomed in plot of the single SAGA06 Raman data, showing an apparent luminescent background, and a clearly resolved TO phonon peak, but no sign of the LO phonon peak.

Raman spectra is the presence of a TO phonon peak and the absence of an LO phonon peak in the non-resonance data. Figure 5.10 shows the region between 150 cm−1 and 300 cm−1 of the data shown in gure 5.9a, which more clearly shows the position of the one Raman peak seen.

H Because we did not observe the LO mode or the E2 mode observed by Zardo et al. at approximately 256 cm−1 in nanowires with 30% wurtzite content, it seems unlikely that there is much, if any, wurtzite in the nanowires we measured [129].

The polarization dependence of the Raman intensity plotted in gure 5.11 appears to depend on the alignment of the polarization with the nanowire, because it clearly does not correspond to any of the patterns predicted in gures 5.5 to 5.8. This type of polarization sensitivity was rst observed in carbon nanotubes [162]. It is known that the response of nanowires to light is very dependent on polarization and their emission of light is strongly

Chapter5: SingleNanowire... 5.4ResultsandDiscussion 70

(a) (b)

Figure 5.11: Solid lines are cos4(θ ) ts to the data measured in the (a) parallel and (b) perpendicular congurations.

polarized when the diameter of the nanowire is much less than the wavelength of light in

question (λ  d)[163, 164].

Given λ  d, the electric eld is reduced inside the nanowire for light polarized

2ε0 perpendicular to the axis of the nanowire, such that E⊥ = δE0, where δ = [164, 165]. εGaAs + ε0 The ratio of parallel to perpendicularly polarized Raman intensities is calculated by, 2 Ik − I⊥ 1 − δ ρ = ; while the predicted value for ρ is given by, ρ = 2 [163, 164]. For GaAs Ik + I⊥ 1 + δ

at room temperature the dielectric constant for λ = 633 nm is εGaAs ≈ 14.8, leading to an

expected value for ρ of 0.97 [166].

4 The data in gure 5.11 are t to r (θ ) = r0 + A cos (θ − θ0), chosen based on the work of

Wu et al. [161]. Because each polarization measurement takes several hours, during which

time there is some drift in the sample position and the intensity of the Raman signal is very sensitive to the position of the sample, the overall shape of the polarization response

Chapter5: SingleNanowire... 5.4ResultsandDiscussion 71

measured is distorted. The lobes in gure 5.11 should be symmetric; the asymmetry is purely

an experimental artifact that limits the utility of the ts as a source for parameters to use in

the calculation of δ.

5.5 Conclusions

Through this experiment we did not nd any evidence of the presence of wurtzite in the

nanowires, which we would have potentially observed as additional peaks in the Raman

spectra, or as a dierent symmetry in the polarization response. This may be in part because

the only nanowires we were able to study in this manner were SAGA06, which were grown with a V:III precursor ratio of 20:1, the highest precursor ratio of the nanowires we studied

(table 2.1), and hence the least likely to contain wurtzite [88].

We were able to account for the missing TO peak discussed in section 4.3.3; because we

see it out of resonance, its absence was most likely an issue of instrument sensitivity. This

is in itself is not surprising; the resonance instrumentation was not sensitive enough to

measure the Raman spectra out of resonance from the nanowires. Nevertheless, it remains

striking how much the LO phonon was apparently favored in the resonance spectra from the

core-shell nanowires.

The polarization response of the nanowires showed that the Raman response of the cores

of the core-shell nanowires behaved as if they were still small in diameter compared to the

excitation wavelength, despite the presence of the shell giving them a total diameter on the

order of 100 nm.

Chapter5:SingleNanowire... 5.5Conclusions 72

Chapter 6: Conclusions, Open Questions, and Future Work

6.1 Conclusions

Using PL, resonance Raman, and single nanowire Raman spectroscopies, we have enriched

our understanding of the properties of the nanowires we studied. While we reject our initial

hypothesis that the interface between core and shell plays a signicant role in determining the

electronic properties of the nanowires, we have achieved a profoundly deeper understanding

of our samples. Instead of the interface being signicant, the concentration of impurities

resulting from the growth conditions plays the most signicant role, with the thermal history

playing a large role in dierentiating the properties of the bare GaAs nanowires from the

core-shell nanowires.

Our experimental accomplishments include:

• Low temperature PL with varied incident power, which allowed us to identify donor-

acceptor-pair luminescence as a major contribution to the PL spectra of the core-shell

nanowires. Our spectra were well enough resolved that we could more clearly identify

the origin of spectral features that had been part of spectra observed before from

similar nanowires.

• Resonance Raman spectra of the dierent types of nanowires that revealed dierences

in their resonant behavior. These experiments, along with the PL, were conducted on a

custom instrument built at Drexel. The combination of features – low temperatures,

tunable excitation, high resolution – our instrument was capable of is generally not 73

available commercially. Our resonance Raman experiments showed us that the LO

phonon was coupling especially strongly to the bound exciton state in the core-shell

nanowires.

• An experiment using polarized Raman spectroscopy on a single nanowire allowed us to

look for wurtzite based on scattering symmetry and polarization. This experiment also

helped resolve some questions regarding the TO phonon mode that were raised by the

resonance Raman experiment, showing that the lack of a TO phonon peak in the

resonance spectra of the core-shell nanowires was most likely an artifact of the low

sensitivity of the instrument.

6.2 Open Questions and Future Work

The overarching aim for this study was to support the rational design of nanowire structures

for device applications, by supplying growers with information about their product. In this

context, the relative value of possible future experiments is easier to assess.

The only additional experiments on the samples studied that seem advisable, would be

higher quality TEM microscopy in order to more accurately assess the quality of the nanowires

from the standpoint of crystalline defects. Correlating these results with spectroscopy would

greatly assist in the interpretation of the spectroscopic features.

The limits in quality for nanowires grown using tertiarybutylarsine remain unknown.

Most of the work to discover those limits depends on the growers, but we can assist by being

aware of the challenges they face.

Until the nanowire growth results in materials of controlled doping concentration, as we

characterize future samples, it is premature to worry about the more exotic phenomena that

Chapter6: Conclusions 6.2OpenestionsandFutureWork 74

may occur at the interface between core and shell. In this sense too, while it is interesting that

the LO phonon energy appears to move, it would be more valuable to hold o from further

investigating it, until similar experiments could be repeated with better controlled nanowires.

That said, in order to better resolve the origin of the phenomena, it would be interesting to

conduct resonance Raman experiments on single nanowires in order to understand which

eects were artifacts of the ensemble measurement.

Chapter6: Conclusions 6.2OpenestionsandFutureWork 75

Appendix A: Finite Curvature-Mediated Ferroelectricity

Figure A.1: Polarization vs. distance from inner radius for the size nanowires used in the study, illustrating the eect of strain in increasing the polarization of thin ferroelectric shells.

This appendix will highlight my contributions to two published papers: Nonnenmann et al. [OL8] and Johnson et al. [OL10] (see publications list, page 93). From the standpoint of the simulation the critical dierence between the two papers, was that Johnson et al.

[OL10] required agreement between more than one type of material and inclusion of the magneto-elastic parameters in the strain. Both papers are attached as a part of this appendix.

Both of these papers discussed core-shell nanowires with a ferroelectric shell and a metallic core. Experimentally, it was shown that the ferroelectric properties of these 76

Figure A.2: Waterfall plot of the polarization across the thinnest shell, with color representing the temperature used in the calculation of the polarization.

nanowires were enhanced over what would be expected from thin lms of the same material.

My contribution to the work was to provide numerical models that helped explain the

experimental results as a consequence of the strain imposed on the shells by the high degree

of curvature imparted by their nite size. In general, surface eects lead to a decrease in the

polarization of ferroelectric lms, an eect that becomes more signicant the thinner they are.

The experimental results that these models supported, showed that very thin ferroelectric

lms could have their polarization enhanced when they were highly curved.

The models were implemented in Matlab. The main numerical task, solving a dierential

equation was achieved using code based on a non-linear nite dierence algorithm described

by Burden and Faires [167]. The high degree of curvature, imposes a stress eld on the

Appendix A: Finite Curvature-Mediated Ferroelectricity 77 system ! ! a2 b2 b2 a2 σ (r ) = 1 ∓ p − 1 ∓ p . (A.1) rr ,φφ b2 − a2 r 2 aρ b2 − a2 r 2 bρ

The ferroelectric polarization of a shell is calculated through the minimization of the

Gibbs free energy, written in terms of the Landau stiness, renormalized to account for

the eect of strain on the eective material constants, Aˆ(r ) = A − 2Q11σrr (r ) − 2Q12σφφ (r ),

A = A0(T − TC ), the material constants A0, B, and C; and g to account for the energetic cost

of variation of Pr near surface and a depolarizing eld Ed . The Gibbs free energy is given by

b " ˆ # A 2 B 4 C 6 1 2 2 G = Pr (r ) + Pr (r ) + Pr (r ) + g(∇Pr ) − Ed (r )Pr (r ) rdr + Pr dS. (A.2) a 2 4 6 2 S

The Gibbs free energy is minimized through the use of the Euler-Lagrange equation

2 ˆ 3 5 g∇ Pr (r ) = APr (r ) + BPr (r ) + CPr (r ) − Ed (r ).

The algorithms for calculating the coercive eld and the eect of the stress on the Curie

temperature (TC) for the ferroelectric shells were designed to be ecient by feeding the

previous answer in as the initial guess as a parameter (i.e. temperature or applied eld) was

changed. At some point, the only stable solution for the polarization is either the trivial

solution Pr = 0 (for T > TC) or only one sign for Pr will be a solution. Because we eventually

needed to deal with more than just one type of material, the original Matlab scripts were

turned into a more robust set of tools. And because the algorithms required thousands

of polarization proles be calculated, the code was carefully optimized in critical areas.

Figures A.1 and A.2 illustrate some of the results of these calculations.

Figure A.1 plots the results of the calculation of the polarization for thin ferroelectric

Appendix A: Finite Curvature-Mediated Ferroelectricity 78

shells in strained and unstrained states for shell thicknesses corresponding to the sizes that were investigated experimentally. The peak values of polarization curves for the strained

shells show an increase with decreasing shell thickness, because the thinner shells correspond

to smaller radius of curvature samples and strain increases as the inner radius shrinks. The

peak values for the unstrained shells have decreasing maxima as the shells get thinner,

because of the depolarizing eld at the surface.

Figure A.2 shows the polarization across a 7 nm shell as a function of temperature. At the

highest temperature plotted, the shell has become paraelectric and the polarization is zero.

In the printed version this appendix contains reprints of Nonnenmann et al. [OL8] and

Johnson et al. [OL10] (see publications list, page 93). Because the copyright to the articles

remains with the publishers, links are provided here rather than copies of the articles.

Nonnenmann et al. can be found at http://dx.doi.org/10.1021/nl903384p.

Johnson et al. can be found at http://dx.doi.org/10.1063/1.3657152.

Appendix A: Finite Curvature-Mediated Ferroelectricity

87

Bibliography

[1] Herbert Kroemer. How the double-heterostructure laser idea got started. IEEE LEOS Newsletter, August 2007. [2] Peter Y. Yu and . Fundamentals of Semiconductors: Physics and Materials Properties. Springer, May 2005. [3] Herbert Kroemer. Nobel lecture: Quasielectric elds and band osets: teaching electrons new tricks. Reviews of Modern Physics, 73(3):783–793, October 2001. [4] Herbert Kroemer. Quasi-electric and quasi-magnetic elds in non-uniform semiconductors. RCA Review, 18:332–342, 1957. [5] Kimberly A. Dick. A review of nanowire growth promoted by alloys and non-alloying elements with emphasis on Au-assisted III-V nanowires. Progress in Crystal Growth and Characterization of Materials, 54(3-4):138–173, September 2008. [6] H. Kroemer. A proposed class of hetero-junction injection lasers. Proceedings of the IEEE, 51 (12):1782–1783, 1963. [7] William Shockley. Circuit element utilizing semiconductive material, September 1951. U.S. Patent: 2,569,347. [8] H. Kroemer. Theory of a wide-gap emitter for transistors. Proceedings of the IRE, 45(11): 1535–1537, 1957. [9] Z. I. Alferov. The history and future of semiconductor heterostructures. Semiconductors, 32(1): 1–14, 1998. [10] Zhores I. Alferov. Nobel lecture: The double heterostructure concept and its applications in physics, electronics, and technology. Reviews of Modern Physics, 73(3):767–782, October 2001. [11] Herbert Kroemer. Solid state radiation emitters, March 1967. U.S. Patent: 3,309,553. [12] W. A. Harrison, E. A. Kraut, J. R. Waldrop, and R. W. Grant. Polar heterojunction interfaces. Physical Review B, 18(8):4402–4410, October 1978. [13] H. Kroemer. Heterostructure bipolar transistors and integrated circuits. Proceedings of the IEEE, 70(1):13–25, 1982. [14] D. C. Tsui, H. L. Stormer, and A. C. Gossard. Two-dimensional magnetotransport in the extreme quantum limit. Physical Review Letters, 48(22):1559, May 1982. [15] H. L. Stormer, A. Chang, D. C. Tsui, J. C. M. Hwang, A. C. Gossard, and W. Wiegmann. Fractional quantization of the hall eect. Physical Review Letters, 50(24):1953, June 1983. 88

[16] H. L. Stormer, Z. Schlesinger, A. Chang, D. C. Tsui, A. C. Gossard, and W. Wiegmann. Energy structure and quantized hall eect of two-dimensional holes. Physical Review Letters, 51(2):126, July 1983.

[17] B. J. F. Lin, D. C. Tsui, M. A. Paalanen, and A. C. Gossard. Mobility of the two-dimensional electron gas in GaAs-Alx Ga1−x As heterostructures. Applied Physics Letters, 45(6):695–697, 1984. [18] R. Dingle, W. Wiegmann, and C. H. Henry. Quantum states of conned carriers in very thin Alx Ga1−x As-GaAs-Alx Ga1−x As heterostructures. Physical Review Letters, 33(14):827–830, September 1974.

[19] R. Dingle, H. L. Stormer, A. C. Gossard, and W. Wiegmann. Electron mobilities in modulation-doped semiconductor heterojunction superlattices. Applied Physics Letters, 33(7): 665–667, October 1978.

[20] Paola Prete, Nico Lovergine, Ilio Miccoli, Fabio Marzo, Joan S. Burger, Giancarlo Salviati, and Laura Lazzarini. On the luminescence of VLS-grown GaAs-AlGaAs core-shell nanowires and its dependence on MOVPE growth conditions. MRS Online Proceedings Library, 1206, 2009.

[21] C. P. T. Svensson, W. Seifert, M. W. Larsson, L. R. Wallenberg, J. Stangl, G. Bauer, and L. Samuelson. Epitaxially grown GaP/GaAs1−x Px /GaP double heterostructure nanowires for optical applications. Nanotechnology, 16(6):936–9, June 2005.

[22] Hannah J. Joyce, Qiang Gao, H. Hoe Tan, C. Jagadish, Yong Kim, Jin Zou, Leigh M. Smith, Howard E. Jackson, Jan M. Yarrison-Rice, Patrick Parkinson, and Michael B. Johnston. III-V semiconductor nanowires for optoelectronic device applications. Progress in Quantum Electronics, 35(2-3):23–75, March 2011.

[23] P. Paiano, P. Prete, N. Lovergine, and A. M. Mancini. Size and shape control of GaAs nanowires grown by metalorganic vapor phase epitaxy using tertiarybutylarsine. Journal of Applied Physics, 100(9):094305–094305–4, November 2006.

[24] P. Prete, F. Marzo, P. Paiano, N. Lovergine, G. Salviati, L. Lazzarini, and T. Sekiguchi. Luminescence of GaAs/AlGaAs core-shell nanowires grown by MOVPE using tertiarybutylarsine. Journal of Crystal Growth, 310(23):5114–5118, November 2008.

[25] Thomas M. Mirandi. Arsine Source Replacement for the Growth of Gallium Arsenide via MOCVD. The Toxics Use Reduction Institute University of Massachusetts Lowell, August 1993.

[26] R. M. Lum, J. K. Klingert, and M. G. Lamont. Use of tertiarybutylarsine in the metalorganic chemical vapor deposition growth of GaAs. Applied Physics Letters, 50(5):284–286, February 1987.

[27] S. P. Watkins, Darlene M. Brake, and G. Haacke. Transport measurements and donor spectra of very high purity GaAs grown using tertiarybutylarsine and triethylgallium. Journal of Applied Physics, 75(6):2952–2956, March 1994.

[28] C. A. Larsen, N. I. Buchan, S. H. Li, and G. B. Stringfellow. GaAs growth using tertiarybutylarsine and trimethylgallium. Journal of Crystal Growth, 93(1-4):15–19, 1988.

Bibliography 89

[29] C. A. Larsen, S. H. Li, N. I. Buchan, and G. B. Stringfellow. Mechanisms of GaAs growth using tertiarybutylarsine and trimethylgallium. Journal of Crystal Growth, 94(3):673–682, March 1989. [30] S. P. Watkins and G. Haacke. Carbon acceptor incorporation in GaAs grown by metalorganic chemical vapor deposition: Arsine versus tertiarybutylarsine. Applied Physics Letters, 59(18): 2263–2265, October 1991. [31] Xiaocheng Jiang, Qihua Xiong, Sungwoo Nam, Fang Qian, Yat Li, and Charles M. Lieber. InAs/InP radial nanowire heterostructures as high electron mobility devices. Nano Letters, 7 (10):3214–3218, October 2007. [32] Lincoln J. Lauhon, Mark S. Gudiksen, Deli Wang, and Charles M. Lieber. Epitaxial core-shell and core-multishell nanowire heterostructures. Nature, 420(6911):57–61, November 2002. [33] Wei Lu, Jie Xiang, Brian P. Timko, Yue Wu, and Charles M. Lieber. One-dimensional hole gas in germanium/silicon nanowire heterostructures. Proceedings of the National Academy of Sciences, 102(29):10046–10051, July 2005. [34] Yat Li, Jie Xiang, Fang Qian, Silvija Gradečak, Yue Wu, Hao Yan, Douglas A. Blom, and Charles M. Lieber. Dopant-free GaN/AlN/AlGaN radial nanowire heterostructures as high electron mobility transistors. Nano Letters, 6(7):1468–1473, July 2006. [35] I-H. Tan, G. L. Snider, L. D. Chang, and E. L. Hu. A self-consistent solution of Schrödinger–Poisson equations using a nonuniform mesh. Journal of Applied Physics, 68(8): 4071–4076, October 1990. [36] Lingquan Wang, Deli Wang, and Peter M. Asbeck. A numerical Schrödinger–Poisson solver for radially symmetric nanowire core-shell structures. Solid-State Electronics, 50(11-12): 1732–1739, November 2006. [37] Mark S. Gudiksen, Lincoln J. Lauhon, Jianfang Wang, David C. Smith, and Charles M. Lieber. Growth of nanowire superlattice structures for nanoscale photonics and electronics. Nature, 415(6872):617–620, February 2002. [38] Yat Li, Fang Qian, Jie Xiang, and Charles M. Lieber. Nanowire electronic and optoelectronic devices. Materials Today, 9(10):18–27, October 2006. [39] E. Tutuc, J. O. Chu, J. A. Ott, and S. Guha. Doping of germanium nanowires grown in presence of PH3. Applied Physics Letters, 89(26):263101–3, December 2006. [40] Jiming Bao, Mariano A. Zimmler, Federico Capasso, Xiaowei Wang, and Z. F. Ren. Broadband ZnO single-nanowire light-emitting diode. Nano Letters, 6(8):1719–1722, August 2006. [41] Katsuhiro Tomioka, Junichi Motohisa, Shinjiroh Hara, Kenji Hiruma, and Takashi Fukui. GaAs/AlGaAs core multishell nanowire-based light-emitting diodes on Si. Nano Letters, 10(5): 1639–1644, May 2010. [42] Michael H. Huang, Samuel Mao, Henning Feick, Haoquan Yan, Yiying Wu, Hannes Kind, Eicke Weber, Richard Russo, and Peidong Yang. Room-temperature ultraviolet nanowire nanolasers. Science, 292(5523):1897–1899, June 2001.

Bibliography 90

[43] Bin Hua, Junichi Motohisa, Yasunori Kobayashi, Shinjiroh Hara, and Takashi Fukui. Single GaAs/GaAsP coaxial core-shell nanowire lasers. Nano Letters, 9(1):112–116, January 2009. [44] Justin C. Johnson, Heon-Jin Choi, Kelly P. Knutsen, Richard D. Schaller, Peidong Yang, and Richard J. Saykally. Single gallium nitride nanowire lasers. Nature Materials, 1(2):106–110, October 2002. [45] Linyou Cao, Joon-Shik Park, Pengyu Fan, Bruce Clemens, and Mark L. Brongersma. Resonant germanium nanoantenna photodetectors. Nano Letters, 10(4):1229–1233, March 2010. [46] Yi Cui, Qingqiao Wei, Hongkun Park, and Charles M. Lieber. Nanowire nanosensors for highly sensitive and selective detection of biological and chemical species. Science, 293(5533): 1289–1292, August 2001. [47] Josef A. Czaban, David A. Thompson, and Ray R. LaPierre. GaAs core-shell nanowires for photovoltaic applications. Nano Letters, 9(1):148–154, January 2009. [48] Bozhi Tian, Thomas J. Kempa, and Charles M. Lieber. Single nanowire photovoltaics. Chemical Society Reviews, 38(1):16, 2009. [49] Ning Han, Fengyun Wang, SenPo Yip, Jared J. Hou, Fei Xiu, Xiaoling Shi, Alvin T. Hui, TakFu Hung, and Johnny C. Ho. GaAs nanowire schottky barrier photovoltaics utilizing Au-Ga alloy catalytic tips. Applied Physics Letters, 101(1):013105–013105–4, July 2012. [50] M. Tchernycheva, L. Rigutti, G. Jacopin, A. de Luna Bugallo, P. Lavenus, F. H. Julien, M. Timofeeva, A. D. Bouravleuv, G. E. Cirlin, V. Dhaka, H. Lipsanen, and L. Largeau. Photovoltaic properties of GaAsP core-shell nanowires on Si(001) substrate. Nanotechnology, 23 (26):265402, July 2012. [51] Sun-Kyung Kim, Robert W. Day, James F. Cahoon, Thomas J. Kempa, Kyung-Deok Song, Hong-Gyu Park, and Charles M. Lieber. Tuning light absorption in core/shell silicon nanowire photovoltaic devices through morphological design. Nano Letters, 12(9):4971–4976, September 2012. [52] Matt Law, Lori E. Greene, Justin C. Johnson, Richard Saykally, and Peidong Yang. Nanowire dye-sensitized solar cells. Nature Materials, 4(6):455–459, 2005. [53] Fang Qian, Megan Brewster, Sung K. Lim, Yichuan Ling, Christopher Greene, Oleg Laboutin, Jerry W. Johnson, Silvija Gradečak, Yu Cao, and Yat Li. Controlled synthesis of AlN/GaN multiple quantum well nanowire structures and their optical properties. Nano Letters, 12(6): 3344–3350, June 2012. [54] Jesper Wallentin, Johan M. Persson, Jakob B. Wagner, Lars Samuelson, Knut Deppert, and Magnus T. Borgström. High-performance single nanowire tunnel diodes. Nano Letters, 10(3): 974–979, March 2010. [55] Jonathan E. Allen, Eric R. Hemesath, Daniel E. Perea, Jessica L. Lensch-Falk, Z. Y. Li, Feng Yin, Mhairi H. Gass, Peng Wang, Andrew L. Bleloch, Richard E. Palmer, and Lincoln J. Lauhon. High-resolution detection of Au catalyst atoms in Si nanowires. Nature Nanotechnology, 3(3): 168–173, 2008.

Bibliography 91

[56] Michael Moewe, Linus C. Chuang, Shanna Crankshaw, Chris Chase, and Connie Chang-Hasnain. Atomically sharp catalyst-free wurtzite GaAs/AlGaAs nanoneedles grown on silicon. Applied Physics Letters, 93(2):023116–023116–3, July 2008.

[57] C. Colombo, D. Spirkoska, M. Frimmer, G. Abstreiter, and A. Fontcuberta i Morral. Ga-assisted catalyst-free growth mechanism of GaAs nanowires by molecular beam epitaxy. Physical Review B, 77(15):155326–5, April 2008.

[58] Keitaro Ikejiri, Jinichiro Noborisaka, Shinjiroh Hara, Junichi Motohisa, and Takashi Fukui. Mechanism of catalyst-free growth of GaAs nanowires by selective area MOVPE. Journal of Crystal Growth, 298:616–619, January 2007.

[59] Jinichiro Noborisaka, Junichi Motohisa, and Takashi Fukui. Catalyst-free growth of GaAs nanowires by selective-area metalorganic vapor-phase epitaxy. Applied Physics Letters, 86(21): 213102–3, May 2005.

[60] Martin Heiss, Eva Riedlberger, Dance Spirkoska, Max Bichler, Gerhard Abstreiter, and Anna Fontcuberta i Morral. Growth mechanisms and optical properties of GaAs-based semiconductor microstructures by selective area epitaxy. Journal of Crystal Growth, 310(6): 1049–1056, March 2008.

[61] P. L. Edwards and R. J. Happel. Beryllium oxide whiskers and platelets. Journal of Applied Physics, 33(3):943–948, March 1962.

[62] R. S. Wagner and W. C. Ellis. Vapor-liquid-solid mechanism of single crystal growth. Applied Physics Letters, 4(5):89–90, March 1964.

[63] S. Kodambaka, J. Terso, M. C. Reuter, and F. M. Ross. Germanium nanowire growth below the eutectic temperature. Science, 316(5825):729–732, May 2007.

[64] E. I. Givargizov. Fundamental aspects of VLS growth. Journal of Crystal Growth, 31(0):20–30, December 1975.

[65] L. J. Lauhon, Mark S. Gudiksen, and Charles M. Lieber. Semiconductor nanowire heterostructures. Philosophical Transactions: Mathematical, Physical and Engineering Sciences, 362(1819):1247–1260, June 2004.

[66] T. F. Kuech and E. Veuho. Mechanism of carbon incorporation in MOCVD GaAs. Journal of Crystal Growth, 68(1):148–156, September 1984.

[67] W. M. Theis, K. K. Bajaj, C. W. Litton, and W. G. Spitzer. Direct evidence for the site of substitutional carbon impurity in GaAs. Applied Physics Letters, 41(1):70–72, July 1982.

[68] Koki Saito, Eisuke Tokumitsu, Takeshi Akatsuka, Motoya Miyauchi, Takumi Yamada, Makoto Konagai, and Kiyoshi Takahashi. Characterization of p-type GaAs heavily doped with carbon grown by metalorganic molecular-beam epitaxy. Journal of Applied Physics, 64(8):3975–3979, October 1988.

Bibliography 92

[69] B. T. Cunningham, L. J. Guido, J. E. Baker, J. S. Major, N. Holonyak, and G. E. Stillman. Carbon diusion in undoped, n-type, and p-type GaAs. Applied Physics Letters, 55(7):687–689, August 1989.

[70] M. J. Ashwin, B. R. Davidson, K. Woodhouse, R. C. Newman, T. J. Bullough, T. B. Joyce, R. Nicklin, and R. R. Bradley. Carbon acceptors passivated with hydrogen and the search for carbon donors in highly doped GaAs:C. Semiconductor Science and Technology, 8(5):625–629, May 1993.

[71] G. E. Höer, H. J. Höer, N. Holonyak, and K. C. Hsieh. Eect of annealing temperature on the hole concentration and lattice relaxation of carbon-doped GaAs and Alx Ga1−x As. Journal of Applied Physics, 72(11):5318–5324, December 1992.

[72] G. Haacke, S. P. Watkins, and H. Burkhard. Metalorganic chemical vapor deposition of high-purity GaAs using tertiarybutylarsine. Applied Physics Letters, 54(20):2029–2031, May 1989.

[73] G. Haacke, S. P. Watkins, and H. Burkhard. Epitaxial growth of high-mobility GaAs using tertiarybutylarsine and triethylgallium. Applied Physics Letters, 56(5):478–480, January 1990.

[74] C. R. Abernathy, S. J. Pearton, R. Caruso, F. Ren, and J. Kovalchik. Ultrahigh doping of GaAs by carbon during metalorganic molecular beam epitaxy. Applied Physics Letters, 55(17): 1750–1752, October 1989.

[75] Naoki Kobayashi, Toshiki Makimoto, and Yoshiji Horikoshi. Abrupt p-type doping prole of carbon atomic layer doped GaAs grown by ow-rate modulation epitaxy. Applied Physics Letters, 50(20):1435–1437, May 1987.

[76] T. F. Kuech, M. A. Tischler, P.-J. Wang, G. Scilla, R. Potemski, and F. Cardone. Controlled carbon doping of GaAs by metalorganic vapor phase epitaxy. Applied Physics Letters, 53(14): 1317–1319, October 1988.

[77] H. Okamoto and T. B. Massalski. The Au-C (gold-carbon) system. Bulletin of Alloy Phase Diagrams, 5(4):378–379, August 1984.

[78] Hannah J. Joyce, Qiang Gao, H. Hoe Tan, Chennupati Jagadish, Yong Kim, Melodie A. Fickenscher, Saranga Perera, Thang Ba Hoang, Leigh M. Smith, Howard E. Jackson, Jan M. Yarrison-Rice, Xin Zhang, and Jin Zou. High purity GaAs nanowires free of planar defects: Growth and characterization. Advanced Functional Materials, 18(23):3794–3800, December 2008.

[79] Hannah J. Joyce, Qiang Gao, H. Hoe Tan, Chennupati Jagadish, Yong Kim, Melodie A. Fickenscher, Saranga Perera, Thang Ba Hoang, Leigh M. Smith, Howard E. Jackson, Jan M. Yarrison-Rice, Xin Zhang, and Jin Zou. Unexpected benets of rapid growth rate for III-V nanowires. Nano Letters, 9(2):695–701, February 2009.

[80] Maurizio Masi, Harsono Simka, Klavs F. Jensen, Thomas F. Kuech, and Robert Potemski. Simulation of carbon doping of GaAs during MOVPE. Journal of Crystal Growth, 124(1-4): 483–492, November 1992.

Bibliography 93

[81] N. I. Buchan, T. F. Kuech, D. Beach, G. Scilla, and F. Cardone. The use of azo-compounds as probes of carbon incorporation of nominally undoped metalorganic vapor phase epitaxy grown GaAs. Journal of Applied Physics, 69(4):2156–2160, February 1991.

[82] Triantallos J. Mountziaris and Klavs F. Jensen. Gas-Phase and surface reaction mechanisms in MOCVD of GaAs with trimethyl-gallium and arsine. Journal of The Electrochemical Society, 138(8):2426–2439, August 1991.

[83] Masanari Koguchi, Hiroshi Kakibayashi, Masamitsu Yazawa, Kenji Hiruma, and Toshio Katsuyama. Crystal structure change of GaAs and InAs whiskers from zinc-blende to wurtzite type. Japanese Journal of Applied Physics, 31(Part 1, No. 7):2061–2065, July 1992.

[84] M. I. McMahon and R. J. Nelmes. Observation of a wurtzite form of gallium arsenide. Physical Review Letters, 95(21):215505, November 2005.

[85] Frank Glas, Jean-Christophe Harmand, and Gilles Patriarche. Why does wurtzite form in nanowires of III-V zinc blende semiconductors? Physical Review Letters, 99(14):146101, October 2007.

[86] Martin Heiss, Sonia Conesa-Boj, Jun Ren, Hsiang-Han Tseng, Adam Gali, Andreas Rudolph, Emanuele Uccelli, Francesca Peiró, Joan Ramon Morante, Dieter Schuh, Elisabeth Reiger, Efthimios Kaxiras, Jordi Arbiol, and Anna Fontcuberta i Morral. Direct correlation of crystal structure and optical properties in wurtzite/zinc-blende GaAs nanowire heterostructures. Physical Review B, 83(4):045303, January 2011.

[87] B. V. Novikov, S. Y. Serov, N. G. Filosofov, I. V. Shtrom, V. G. Talalaev, O. F. Vyvenko, E. V. Ubyivovk, Y. B. Samsonenko, A. D. Bouravleuv, I. P. Soshnikov, et al. Photoluminescence properties of GaAs nanowire ensembles with zincblende and wurtzite crystal structure. physica status solidi (RRL)-Rapid Research Letters, 4(7):175–177, 2010.

[88] Hannah J. Joyce, Jennifer Wong-Leung, Qiang Gao, H. Hoe Tan, and Chennupati Jagadish. Phase perfection in zinc blende and wurtzite III-V nanowires using basic growth parameters. Nano Letters, 10(3):908–915, October 2011.

[89] D. Spirkoska, J. Arbiol, A. Gustafsson, S. Conesa-Boj, F. Glas, I. Zardo, M. Heigoldt, M. H. Gass, A. L. Bleloch, S. Estrade, M. Kaniber, J. Rossler, F. Peiro, J. R. Morante, G. Abstreiter, L. Samuelson, and A. Fontcuberta i Morral. Structural and optical properties of high quality zinc-blende/wurtzite GaAs nanowire heterostructures. Physical Review B, 80(24):245325, December 2009.

[90] Peter Krogstrup, Morten Hannibal Madsen, Wen Hu, Miwa Kozu, Yuka Nakata, Jesper Nygård, Masamitu Takahasi, and Robert FeidenhansŠl. In-situ x-ray characterization of wurtzite formation in GaAs nanowires. Applied Physics Letters, 100(9):093103–093103–4, February 2012.

[91] G. D. Gilliland. Photoluminescence spectroscopy of crystalline semiconductors. Materials Science and Engineering: R: Reports, 18(3-6):99–399, March 1997.

[92] Jacques I. Pankove. Optical Processes in Semiconductors. Courier Dover Publications, 1971.

Bibliography 94

[93] L. V. Titova, Thang B. Hoang, H. E. Jackson, L. M. Smith, J. M. Yarrison-Rice, Y. Kim, H. J. Joyce, H. H. Tan, and C. Jagadish. Temperature dependence of photoluminescence from single core-shell GaAs-AlGaAs nanowires. Applied Physics Letters, 89(17):173126–3, October 2006.

[94] M. R. Brozel and G. E. Stillman. Properties of Gallium Arsenide (3rd Edition). Institution of Engineering and Technology, 1996.

[95] Y. P. Varshni. Temperature dependence of the energy gap in semiconductors. Physica, 34(1): 149–154, 1967.

[96] Diego Olego and Manuel Cardona. Photoluminescence in heavily doped GaAs. I. temperature and hole-concentration dependence. Physical Review B, 22(2):886–893, July 1980.

[97] Q. Huang and R. G. Ulbrich. Impurity-induced resonance Raman scattering at the (e, A0) threshold in lightly carbon-doped p-type GaAs at 2 K. Journal of Luminescence, 99(1):19–28, August 2002.

[98] D. C. Reynolds, D. C. Look, R. Kaspi, and D. N. Talwar. Phonon replicas in the photoluminescence emission of Alx Ga1−x As alloys. Applied Physics Letters, 66(25):3447–3449, June 1995.

[99] Q. Huang and R. G. Ulbrich. Carbon-acceptor-induced cascade scattering by acoustic phonons above the (e, A0) threshold in GaAs. Physical Review B, 64(11):113205, 2001.

[100] G. Borghs, K. Bhattacharyya, K. Denee, P. Van Mieghem, and R. Mertens. Band-gap narrowing in highly doped n- and p-type GaAs studied by photoluminescence spectroscopy. Journal of Applied Physics, 66(9):4381–4386, November 1989.

[101] H. D. Chen, M. S. Feng, P. A. Chen, K. C. Lin, and C. C. Wu. Low-temperature luminescent properties of degenerate p-type GaAs grown by low-pressure metalorganic chemical vapor deposition. Journal of Applied Physics, 75(4):2210–2214, February 1994.

[102] Kazuhiro Kudo, Yunosuke Makita, Ichiro Takayasu, Toshio Nomura, Toshihiko Kobayashi, Tomio Izumi, and Tokue Matsumori. Photoluminescence spectra of undoped GaAs grown by molecular-beam epitaxy at very high and low substrate temperatures. Journal of Applied Physics, 59(3):888–891, February 1986.

[103] Elias Burstein. Anomalous optical absorption limit in InSb. Physical Review, 93(3):632–633, February 1954.

[104] T. S. Moss. Theory of the spectral distribution of recombination radiation from InSb. Proceedings of the Physical Society. Section B, 70(2):247–250, February 1957.

[105] Franz Urbach. The long-wavelength edge of photographic sensitivity and of the electronic absorption of solids. Physical Review, 92(5):1324–1324, December 1953.

[106] F. Moser and F. Urbach. Optical absorption of pure silver halides. Physical Review, 102(6): 1519–1523, June 1956.

Bibliography 95

[107] Piet Van Mieghem. Theory of band tails in heavily doped semiconductors. Reviews of Modern Physics, 64(3):755–793, July 1992.

[108] P. J. Dean. Inter-impurity recombinations in semiconductors. Progress in Solid State Chemistry, 8:1–126, 1973.

[109] J. J. Hopeld, D. G. Thomas, and M. Gershenzon. Pair spectra in GaP. Physical Review Letters, 10(5):162, March 1963.

[110] Max Born and Kun Huang. Dynamical Theory of Crystal Lattices. Oxford University Press, Oxford, 1954.

[111] W. Y. Han, Y. Lu, H. S. Lee, M. W. Cole, S. N. Schauer, R. P. Moerkirk, K. A. Jones, and L. W. Yang. Annealing eects on heavily carbon-doped GaAs. Applied Physics Letters, 61(1):87–89, July 1992.

[112] D. M. Kozuch, Michael Stavola, S. J. Pearton, C. R. Abernathy, and J. Lopata. Hydrogen in carbon-doped GaAs grown by metalorganic molecular beam epitaxy. Applied Physics Letters, 57 (24):2561–2563, December 1990.

[113] J. Wagner, M. Maier, Th. Lauterbach, K. H. Bachem, M. Ashwin, R. C. Newman, K. Woodhouse, R. Nicklin, and R. R. Bradley. Raman spectroscopic assessment of carbon-hydrogen pairs in carbon-doped GaAs layers. Applied Physics Letters, 60(20):2546–2548, May 1992.

[114] Q. Liu, A. Brennemann, H. Hardtdegen, A. Lindner, W. Prost, and F.-J. Tegude. Characterization of hydrogen passivation and carbon self-compensation of highly C-doped GaAs by means of x-ray diraction. Journal of Applied Physics, 79(2):710–716, January 1996.

[115] S. J. Breuer, R. Jones, P. R. Briddon, and S. Öberg. Interstitial hydrogen and enhanced dissociation of C-H complexes in GaAs. Physical Review B, 53(24):16289–16296, June 1996.

[116] K. Woodhouse, R. C. Newman, R. Nicklin, R. R. Bradley, and M. J. L. Sangster. The lattice sites of carbon and hydrogen incorporated in GaAs grown by MOVPE revealed by infrared spectroscopy. Journal of Crystal Growth, 120(1-4):323–327, May 1992.

[117] Hiroshi Fushimi and Kazumi Wada. The presence of isolated hydrogen donors in heavily carbon-doped GaAs. Journal of Crystal Growth, 145(1-4):420–426, December 1994.

[118] Hiroshi Fushimi and Kazumi Wada. Carbon-related defects in carbon-doped GaAs by high-temperature annealing. Journal of Applied Physics, 82(3):1208–1213, August 1997.

[119] J. Mimila-Arroyo and S. W. Bland. Acceptor reactivation kinetics in heavily carbon-doped GaAs epitaxial layers. Applied Physics Letters, 77(8):1164–1166, August 2000.

[120] J. Mimila-Arroyo, A. Lusson, J. Chevallier, M. Barbé, B. Theys, F. Jomard, and S. W. Bland. Carbon acceptor doping eciency in GaAs grown by metalorganic chemical vapor deposition. Applied Physics Letters, 79(19):3095–3097, November 2001.

Bibliography 96

[121] Kazuo Watanabe and Hajime Yamazaki. Annealing eect on the electrical properties of heavily C-doped p+GaAs. Applied Physics Letters, 59(4):434–436, July 1991.

[122] N. M. Johnson, R. D. Burnham, R. A. Street, and R. L. Thornton. Hydrogen passivation of shallow-acceptor impurities in p-type GaAs. Physical Review B, 33(2):1102–1105, January 1986.

[123] Yunosuke Makita. Novel features of photoluminescence spectra from acceptor-doped GaAs: formation of acceptor-acceptor pair emissions and optical compensation eect. Materials Science and Engineering: R: Reports, 16(6-8):265–398, June 1996.

[124] Hannah J. Joyce, Qiang Gao, H. Hoe Tan, Chennupati Jagadish, Yong Kim, Xin Zhang, Yanan Guo, and Jin Zou. Twin-free uniform epitaxial GaAs nanowires grown by a two-temperature process. Nano Letters, 7(4):921–926, 2007.

[125] J. H. Kang, Q. Gao, P. Parkinson, H. J. Joyce, H. H. Tan, Y. Kim, Y. Guo, H. Xu, J. Zou, and C. Jagadish. Precursor ow rate manipulation for the controlled fabrication of twin-free GaAs nanowires on silicon substrates. Nanotechnology, 23(41):415702, October 2012.

[126] Anders Gustafsson, Karla Hillerich, Maria E. Messing, Kristian Storm, Kimberly A. Dick, Knut Deppert, and Jessica Bolinsson. A cathodoluminescence study of the inuence of the seed particle preparation method on the optical properties of GaAs nanowires. Nanotechnology, 23 (26):265704, July 2012.

[127] Veer Dhaka, Tuomas Haggren, Henri Jussila, Hua Jiang, Esko Kauppinen, Teppo Huhtio, Markku Sopanen, and Harri Lipsanen. High quality GaAs nanowires grown on glass substrates. Nano Letters, 12(4):1912–1918, 2012.

[128] Xin-Yu Bao, Cesare Soci, Darija Susac, Jon Bratvold, David P. R. Aplin, Wei Wei, Ching-Yang Chen, Shadi A. Dayeh, Karen L. Kavanagh, and Deli Wang. Heteroepitaxial growth of vertical GaAs nanowires on Si (111) substrates by metal-organic chemical vapor deposition. Nano Letters, 8(11):3755–3760, November 2008.

[129] I. Zardo, S. Conesa-Boj, F. Peiro, J. R. Morante, J. Arbiol, E. Uccelli, G. Abstreiter, and A. Fontcuberta i Morral. Raman spectroscopy of wurtzite and zinc-blende GaAs nanowires: Polarization dependence, selection rules, and strain eects. Physical Review B, 80(24):245324, December 2009.

[130] M. S. Skolnick, C. W. Tu, and T. D. Harris. High-resolution spectroscopy of defect-bound excitons and acceptors in GaAs grown by molecular-beam epitaxy. Physical Review B, 33(12): 8468–8474, June 1986.

[131] R. Loudon. Theory of the rst-order Raman eect in crystals. Proceedings of the Royal Society of London. Series A, Mathematical and Physical Sciences, 275(1361):218–232, September 1963.

[132] R. Loudon. The Raman eect in crystals. Advances in Physics, 13(52):423, 1964.

[133] R. P. Feynman. The theory of positrons. Physical Review, 76(6):749–759, September 1949.

Bibliography 97

[134] V. I. Belitsky, M. Cardona, I. G. Lang, and S. T. Pavlov. Spatial correlation of electrons and holes in multiphonon resonant Raman scattering in a high magnetic eld. Physical Review B, 46(24):15767–15788, December 1992.

[135] Diego Olego, A. Pinczuk, A. C. Gossard, and W. Wiegmann. Plasma dispersion in a layered electron gas: A determination in GaAs-(AlGa)As heterostructures. Physical Review B, 25(12): 7867, June 1982.

[136] E. Burstein, A. Pinczuk, and D. L. Mills. Inelastic light scattering by charge carrier excitations in two-dimensional plasmas: Theoretical considerations. Surface Science, 98(1-3):451, August 1980.

[137] J. Menéndez, A. Pinczuk, D. J. Werder, A. C. Gossard, and J. H. English. Light scattering determination of band osets in GaAs-Alx Ga1−x As quantum wells. Physical Review B, 33(12): 8863, June 1986.

[138] C. V. Raman and K. S. Krishnan. A new type of secondary radiation. Nature, 121(3048): 501–502, March 1928.

[139] A. K. Sood, W. Kauschke, J. Menéndez, and M. Cardona. Resonance Raman scattering by optical phonons in GaAs near the E0 band gap. Physical Review B, 35(6):2886, February 1987. [140] J. E. Zucker, A. Pinczuk, and D. S. Chemla. Relation of resonant Raman line shape to electronic structure in quantum wells. Physical Review B, 38(6):4287, 1988.

[141] Wolfgang Kauschke and Manuel Cardona. Resonant Raman scattering and interference eects of LO phonons at the E0 + ∆0 gap of InP. Physical Review B, 33(8):5473–5481, April 1986. [142] Patryk Kusch, Steen Breuer, Manfred Ramsteiner, Lutz Geelhaar, Henning Riechert, and Stephanie Reich. Band gap of wurtzite GaAs: a resonant Raman study. Physical Review B, 86(7): 075317, August 2012.

[143] P. Y. Yu, Y. R. Shen, Y. Petro, and L. M. Falicov. Resonance Raman scattering at the forbidden yellow exciton in Cu2O. Physical Review Letters, 30(7):283–286, February 1973. [144] Bernt Ketterer, Martin Heiss, Marie J. Livrozet, Andreas Rudolph, Elisabeth Reiger, and Anna Fontcuberta i Morral. Determination of the band gap and the split-o band in wurtzite GaAs using Raman and photoluminescence excitation spectroscopy. Physical Review B, 83(12):125307, March 2011.

[145] Bernt Ketterer, Martin Heiss, Emanuele Uccelli, Jordi Arbiol, and Anna Fontcuberta i Morral. Untangling the electronic band structure of wurtzite GaAs nanowires by resonant Raman spectroscopy. ACS Nano, 5(9):7585–7592, 2011.

[146] J. Wagner, A. Fischer, and K. Ploog. Resonant Raman scattering and photoluminescence at the E0 band gap of carbon-doped AlAs. Applied Physics Letters, 62(26):3482–3484, June 1993. [147] C. Trallero-Giner, A. Cantarero, M. Cardona, and M. Mora. Impurity-induced resonant Raman scattering. Physical Review B, 45(12):6601–6613, March 1992.

Bibliography 98

[148] W. Kauschke, A. K. Sood, M. Cardona, and K. Ploog. Resonance Raman scattering in GaAs-Alx Ga1−x As superlattices: Impurity-induced Fröhlich-interaction scattering. Physical Review B, 36(3):1612–1619, July 1987.

[149] R. C. C. Leite, J. F. Scott, and T. C. Damen. Multiple-phonon resonant Raman scattering in CdS. Physical Review Letters, 22(15):780, April 1969.

[150] Miles V. Klein and S. P. S. Porto. Multiple-phonon-resonance Raman eect in CdS. Physical Review Letters, 22(15):782, April 1969; 22(18):985, May 1969.

[151] Megan Brewster, Oliver Schimek, Stephanie Reich, and Silvija Gradečak. Exciton-phonon coupling in individual GaAs nanowires studied using resonant Raman spectroscopy. Physical Review B, 80(20):201314, November 2009.

[152] R. Zeyher. Theory of multiphonon Raman spectra above the energy gap in semiconductors. Solid State Communications, 16(1):49–52, January 1975.

[153] A. Cantarero, C. Trallero-Giner, and M. Cardona. Excitons in one-phonon resonant Raman scattering: Fröhlich and interference eects. Physical Review B, 40(18):12290–12295, December 1989.

[154] Peter Y. Yu and Y. R. Shen. Resonance Raman studies in Cu2O. I. The phonon-assisted 1s yellow excitonic absorption edge. Physical Review B, 12(4):1377–1394, August 1975.

[155] Miles V. Klein. Equivalence of resonance Raman scattering in solids with absorption followed by luminescence. Physical Review B, 8(2):919–921, July 1973.

[156] Y. R. Shen. Distinction between resonance Raman scattering and hot luminescence. Physical Review B, 9(2):622–626, January 1974.

[157] José Menéndez and Manuel Cardona. Interference eects: A key to understanding forbidden Raman scattering by LO phonons in GaAs. Physical Review B, 31(6):3696–3704, March 1985.

[158] H. Fröhlich. Electrons in lattice elds. Advances in Physics, 3(11):325–361, 1954.

[159] Emmanuel Rosencher and Borge Vinter. Optoelectronics. Cambridge University Press, 2002.

[160] C. Trallero-Giner, A. Cantarero, and M. Cardona. One-phonon resonant Raman scattering: Fröhlich exciton-phonon interaction. Physical Review B, 40(6):4030, 1989.

[161] Jian Wu, Duming Zhang, Qiujie Lu, Humberto R. Gutierrez, and Peter C. Eklund. Polarized Raman scattering from single GaP nanowires. Physical Review B, 81(16):165415, April 2010.

[162] Hiroshi Ajiki and Tsuneya Ando. Aharonov-Bohm eect in carbon nanotubes. Physica B: Condensed Matter, 201:349–352, July 1994.

[163] Jianfang Wang, Mark S. Gudiksen, Xiangfeng Duan, Yi Cui, and Charles M. Lieber. Highly polarized photoluminescence and photodetection from single indium phosphide nanowires. Science, 293(5534):1455–1457, August 2001.

Bibliography 99

[164] Tsachi Livneh, Jinping Zhang, Guosheng Cheng, and Martin Moskovits. Polarized Raman scattering from single GaN nanowires. Physical Review B, 74(3):035320, July 2006.

[165] L. D. Landau and E. M. Lifshitz. Electrodynamics of Continuous Media. Pergamon Press, Oxford, 1960.

[166] D. E. Aspnes and A. A. Studna. Dielectric functions and optical parameters of Si, Ge, GaP, GaAs, GaSb, InP, InAs, and InSb from 1.5 to 6.0 eV. Physical Review B, 27(2):985–1009, January 1983.

[167] Richard L. Burden and J. Douglas Faires. Numerical Analysis. Brooks Cole, 9th Edition, August 2010.

Bibliography 100

Vita

Oren D. Leaer recieved a B.S. in Engineeering Physics from the University of California at

Berkeley.

Publications

[OL1] O. D. Leaer, S. Gupta, M. W. Barsoum, and J. E. Spanier. On Raman scattering from selected M2AC compounds. Journal of Materials Research, 22(10):2652, 2007. [OL2] M. K Drulis, H. Drulis, A. E. Hackemer, O. Leaer, J. Spanier, S. Amini, M. W. Barsoum, T. Guilbert, and T. El-Raghy. On the heat capacities of Ta2AlC, Ti2SC, and Cr2GeC. Journal of Applied Physics, 104(2):023526, 2008.

[OL3] B. Garipcan, J. Winters, J. S. Atchison, M. D. Cathell, J. D. Schiman, O. D. Leaer, S. S. Nonnenmann, C. L. Schauer, E. Pişkin, B. Nabet, and J. E. Spanier. Controllable formation of nanoscale patterns on TiO2 by conductive-AFM nanolithography. Langmuir, 24(16): 8944, 2008.

[OL4] T. H. Scabarozi, S. Amini, P. Finkel, O. D. Leaer, J. E. Spanier, M. W. Barsoum, M. Drulis, H. Drulis, W. M. Tambussi, J. D. Hettinger, and S. E. Loand. Electrical, thermal, and elastic properties of the MAX-phase Ti2SC. Journal of Applied Physics, 104 (3):033502, 2008.

[OL5] V. R. Binetti, J. D. Schiman, O. D. Leaer, J. E. Spanier, and C. L. Schauer. The natural transparency and piezoelectric response of the Greta oto buttery wing. Integrative Biology, 1(4):324, 2009.

[OL6] B. Manoun, O. D. Leaer, S. Gupta, E. N. Homan, S. K. Saxena, J. E. Spanier, and M. W. Barsoum. On the compression behavior of Ti2InC,(Ti0.5, Zr0.5)2InC, and M2SnC (M = Ti, Nb, Hf) to quasi-hydrostatic pressures up to 50 GPa. Solid State Communications, 149 (43-44):1978, 2009.

[OL7] T. H. Scabarozi, S. Amini, O. Leaer, A. Ganguly, S. Gupta, W. Tambussi, S. Clipper, J. E. Spanier, M. W. Barsoum, J. D. Hettinger, and S. E. Loand. Thermal expansion of select Mn+1AXn (M = early transition metal, A = A group element, X = C or N) phases measured by high temperature x-ray diraction and dilatometry. Journal of Applied Physics, 105(1):013543, 2009. 101

[OL8] Stephen S. Nonnenmann, Oren D. Leaer, Eric M. Gallo, Michael T. Coster, and Jonathan E. Spanier. Finite curvature-mediated ferroelectricity. Nano Letters, 10(2):542, February 2010.

[OL9] Guannan Chen, Eric M. Gallo, Oren D. Leaer, Terrence McGuckin, Paola Prete, Nico Lovergine, and Jonathan E. Spanier. Tunable Hot-Electron transfer within a single Core-Shell nanowire. Physical Review Letters, 107(15):156802, October 2011.

[OL10] Stephanie H. Johnson, Peter Finkel, Oren D. Leaer, Stephen S. Nonnenmann, Konrad Bussmann, and Jonathan E. Spanier. Magneto-elastic tuning of ferroelectricity within a magnetoelectric nanowire. Applied Physics Letters, 99(18):182901, October 2011.

[OL11] Jonathan E. Spanier, Stephen S. Nonnenmann, and Oren David Leaer. Ferroelectric nanoshell devices, 2009. U.S. Patent Application: 2012/0098589.

Vita