understanding the growth of iii-v nanowires incorporated ...686916/s4315954_final... · iii-v...

189
Understanding the Growth of III-V Nanowires Incorporated with In by Molecular Beam Epitaxy Chen Zhou Bachelor of Engineering A thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2017 School of Mechanical and Mining Engineering

Upload: hathuy

Post on 29-May-2018

215 views

Category:

Documents


0 download

TRANSCRIPT

Understanding the Growth of III-V Nanowires Incorporated

with In by Molecular Beam Epitaxy

Chen Zhou

Bachelor of Engineering

A thesis submitted for the degree of Doctor of Philosophy at

The University of Queensland in 2017

School of Mechanical and Mining Engineering

I

Abstract

III-V semiconductor nanowires have attracted a great attention due to their peculiar

electronic and optical properties. Among them, GaAs semiconductor nanowires, with direct

band gap, high carrier mobilities (than Si) and the heterojunction formation with other III-V

semiconductor nanowires, such as InAs and AlGaAs, have potential for application in

future nanoscale devices. To realize such optoelectronic devices fabrication, the III-V

nanowires growth must be under control to enable their effective integration into devices,

which can be achieved by Au catalyzed epitaxial nanowires growth under the vapor-liquid-

solid (VLS) mechanism.

Compared with binary III-V nanowires, ternary nanowires have been attracting an

increasing interest as they allow the tunability of desired band gap for different applications

by modulating the composition fraction of their alloys. In particular, InGaAs nanowires

have been demonstrated to be technologically important for a wide range of applications,

attributed to their tunable band gap ranging from near-infrared to infrared regions by tuning

the InGaAs alloy composition, which makes them constitute the ideal material system for

numerous optoelectronic devices, such as light-emitting diodes and nanolasers.

Accordingly, the motivation to investigate ternary InGaAs nanowires system and its related

binary GaAs nanowires system is triggered for their optoelectronic applications in the

future.

In this thesis, binary GaAs nanowires system was first studied. With the help of Au

nanoparticles, GaAs nanowires were grown on GaAs {111}B substrates. By modulating the

growth parameters, such as growth temperature, V/III ratio and growth duration, the

morphological and structural characteristics of grown GaAs nanowires were investigated.

Through extensive electron microscopy investigations, we found that by lowering the

group-V flux to the low V/III ratio, nanowire growth is As-limited and thermodynamically

controlled, leading to the slow nanowire growth and defect-free wurtzite structured grown

nanowires. On the other hand, we found that by prolonging the growth duration, the

structural quality of GaAs nanowires was significantly enhanced from defected to defect-

free wurtzite structure.

Moreover, the growth behavior and compositional characteristics of ternary InxGa1-xAs

nanowires were studied by tuning the In concentration from 50 to 85 at.%. With extensive

cross-sectional study of the grown nanowires by advanced electron microscopy, we found

II

that when x = 0.5, core-multishell structure spontaneously formed at the nanowire bottoms

and Ga concentration in nanowire cores increases towards the nanowire tops; and when x

= 0.85, InGaAs nanowires formed the core-shell structure, with the In-rich core and the

Ga-enriched shell, in which the composition of the cores and shells were varied at varied

nanowire regions. The fundamental reasons behind these new phenomena were

investigated and the corresponding growth mechanism was unveiled.

III

Declaration by author

This thesis is composed of my original work, and contains no material previously published

or written by another person except where due reference has been made in the text. I

have clearly stated the contribution by others to jointly-authored works that I have included

in my thesis.

I have clearly stated the contribution of others to my thesis as a whole, including statistical

assistance, survey design, data analysis, significant technical procedures, professional

editorial advice, and any other original research work used or reported in my thesis. The

content of my thesis is the result of work I have carried out since the commencement of

my research higher degree candidature and does not include a substantial part of work

that has been submitted to qualify for the award of any other degree or diploma in any

university or other tertiary institution. I have clearly stated which parts of my thesis, if any,

have been submitted to qualify for another award.

I acknowledge that an electronic copy of my thesis must be lodged with the University

Library and, subject to the policy and procedures of The University of Queensland, the

thesis be made available for research and study in accordance with the Copyright Act

1968 unless a period of embargo has been approved by the Dean of the Graduate School.

I acknowledge that copyright of all material contained in my thesis resides with the

copyright holder(s) of that material. Where appropriate I have obtained copyright

permission from the copyright holder to reproduce material in this thesis.

IV

Publications during candidature

Peer-viewed papers:

1. Zhou, C.; Zheng, K.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Phase Purification of GaAs

Nanowires by Prolonging the Growth Duration in MBE. J. Mater. Chem. C 2017, 5,

5257-5262.

2. Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting as Flux in Molecular Beam

Epitaxy. J. Phys. Chem. C 2015, 119, 20721-20727.

3. Wei, D.-D.; Shi, S.-X.; Zhou, C.; Zhang, X.-T.; Chen, P.-P.; Xie, J.-T.; Tian, F.; Zou,

J. Formation of GaAs/GaSb Core-Shell Heterostructured Nanowires Grown by

Molecular-Beam Epitaxy. Crystals 2017, 7, 94. 4. Lu, Z.; Zhang, Z.; Chen, P.; Shi, S.; Yao, L.; Zhou, C.; Zhou, X.; Zou, J.; Lu, W.;

Bismuth-induced Phase Control of GaAs Nanowires Grown by Molecular Beam

Epitaxy. Appl. Phys. Lett. 2014, 105, 162102.

5. Shi, S.; Lu, Z.; Zhang, Z.; Zhou C.; Chen, P.; Zou, J.; Lu W.; Morphology and

Microstructure of InAs Nanowires on GaAs Substrate Grown by Molecular Beam

Epitaxy. Chin. Phys. Lett. 2014, 31 (9): 098101.

V

Publications included in this thesis

Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality Control

of GaAs Nanowire Structures by Limiting as Flux in Molecular Beam Epitaxy. J. Phys.

Chem. C 2015, 119, 20721-20727. – incorporated as Chapter 4.

Contributor Statement of contribution

Chen Zhou (Candidate) Carried out characterization (80%)

Carried out data analyses (60%)

Wrote the paper (60%)

Kun Zheng Wrote and edited paper (10%)

Supervised the project (20%)

Zhengyu Lu Carried out sample synthesis (50%)

Zhi Zhang Carried out data analyses (10%)

Zhiming Liao Carried out characterization (20%)

Pingping Chen Carried out sample synthesis (50%)

Designed experiments (40%)

Supervised the project (10%)

Wei Lu Designed experiments (20%)

Supervised the project (10%)

Jin Zou Designed experiments (40%)

Supervised the project (60%)

Carried out data analyses (30%)

Wrote and edited paper (30%)

VI

Zhou, C.; Zheng, K.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Phase Purification of GaAs

Nanowires by Prolonging the Growth Duration in MBE. J. Mater. Chem. C 2017, 5, 5257-

5262. – incorporated as Chapter 5.

Contributor Statement of contribution

Chen Zhou (Candidate) Designed experiments (20%)

Carried out characterization (80%)

Carried out data analyses (60%)

Wrote the paper (60%)

Kun Zheng Carried out data analyses (10%)

Wrote and edited paper (10%)

Supervised the project (20%)

Zhiming Liao Carried out characterization (20%)

Pingping Chen Carried out sample synthesis (80%)

Designed experiments (40%)

Supervised the project (10%)

Wei Lu Designed experiments (10%)

Carried out sample synthesis (20%)

Supervised the project (10%)

Jin Zou Designed experiments (30%)

Supervised the project (60%)

Carried out data analyses (30%)

Wrote and edited paper (30%)

VII

Contributions by others to the thesis

No contributions by others.

Statement of parts of the thesis submitted to qualify for the award of another degree

None

VIII

Acknowledgements

First, I would like to express my sincere gratitude for my principle supervisor Prof. Jin

Zou at The University of Queensland, who gave me the chance to carry out this PhD

research project as an undergraduate student then. He always passes positive attitude to

us PhD students, no matter what problems we faced during the research. Apart from the

strong research skills, the most important thing is that he trained us to build a strong mind,

so that we can have inner peace in facing any difficulties in life. Besides, I would also like

to thank my co-supervisor, Prof. Kun Zheng, a visiting professor from Beijing University of

Technology. He has a strong logic thinking ability so that he can cut to the chase very

quickly, which helps me a lot when I was struggling in experimental data analyses. My

other co-supervisors, Prof. Ping-Ping Chen and Prof. Wei Lu at Shanghai Institute of

Technical Physics, also gave me a lot of support in designing experiment and carrying out

sample syntheses. Without their help and support, this PhD research project cannot be

fulfilled.

Second, I would like to acknowledge all my loverly colleagues: Dr. Zhi Zhang, Dr Hongyi

(Justin) Xu, Dr Guang Han, Dr Wen Sun, Dr Zhi-Gang Chen, Dr Lihua Wang, Dr Lei Yang,

Dr Min Hong, Dr Yichao Zou, Ms Mun Teng (Abby) Soo, Mr Zhiming Liao, Mr Liqing

Huang, Mr Han Gao, Mr Xiang Sun, Mr Van Nguyen, Mr Xiaolei Shi, Mr Weidi Liu, Ms

Yuzhe Yang and Mr Haining Luo in our research group at The University of Queensland.

Thank you all for your kind share of your research experience and skills.

Besides, thanks the Centre of Microscopy and Microanalysis (CMM) at The University of

Queensland for providing those fabulous facilities to support our experiments. Thanks all

the CMM staff for your patient training and technical support.

Last but not least, thanks all my dear families and friends who are always there for me

through thick and thin. Your company and selfless love are the power that keeps me

moving forward.

IX

Keywords

GaAs nanowires, InGaAs nanowires, vapor-liquid-solid mechanism, vapor-solid

mechanism, growth duration, structural quality, phase segregation, core-shell, MBE

Australian and New Zealand Standard Research Classifications (ANZSRC)

ANZSRC code: 100712, Nanoscale Characterisation, 60%

ANZSRC code: 100706, Nanofabrication, Growth and Self Assembly, 30%

ANZSRC code: 091203, Compound Semiconductors, 10%

Fields of Research (FoR) Classification

FoR code: 0912, Materials Engineering, 50%

FoR code: 1007, Nanotechnology, 50%

X

Table of Contents

Chapter 1 Introduction ......................................................................................................... 1

1.1 Background ................................................................................................................ 1

1.2 Objectives and scope ................................................................................................. 2

1.3 Thesis outline ............................................................................................................. 3

Chapter 2 Literature Review ................................................................................................ 5

2.1 Introduction ................................................................................................................. 5

2.2 Synthesis of nanowires ............................................................................................... 6

2.2.1 Vapor-liquid-solid mechanism .............................................................................. 7

2.2.2 Vapor-solid-solid mechanism ............................................................................. 13

2.2.3 Vapor-solid mechanism ...................................................................................... 15

2.3 Morphology of III-V nanowires .................................................................................. 16

2.3.2 Effect of catalysts ............................................................................................... 17

2.3.2 Effect of nanowire density .................................................................................. 21

2.3.2 Effect of growth temperature .............................................................................. 23

2.3.3 Effect of V/III ratio ............................................................................................... 26

2.3.4 Effect of growth duration .................................................................................... 31

2.4 Structural characteristics of III-V nanowires.............................................................. 31

2.4.1 Crystal structure of III-V nanowires .................................................................... 31

2.4.2 Side facets of III-V nanowires ............................................................................. 33

2.4.3 Effect of catalysts diameter ................................................................................ 34

2.4.4 Effect of growth direction .................................................................................... 37

2.4.5 Effect of growth parameters ............................................................................... 39

2.5 Ternary III-V nanowires ............................................................................................ 43

2.5.1 Growth behavior ................................................................................................. 44

2.5.2 Compositional characteristics ............................................................................. 45

2.6 Unclear issues .......................................................................................................... 48

2.6.1 About GaAs nanowires ....................................................................................... 48

2.6.2 About InGaAs nanowires .................................................................................... 48

Chapter 3 Methodology...................................................................................................... 65

3.1 Introduction ............................................................................................................... 65

3.2 Epitaxial growth ........................................................................................................ 66

XI

3.2.1 Molecular beam epitaxy ..................................................................................... 66

3.2.2 Growth procedures ............................................................................................. 69

3.3 Scanning electron microscopy .................................................................................. 70

3.4 Transmission electron microscopy ........................................................................... 76

3.4.1 Imaging mode ..................................................................................................... 77

3.4.2 Diffraction mode ................................................................................................. 80

3.4.3 Scanning mode .................................................................................................. 81

3.4.4 Analytical mode .................................................................................................. 82

3.5 Sample preparation .................................................................................................. 84

3.5.1 SEM sample preparation .................................................................................... 84

3.5.2 TEM sample preparation .................................................................................... 85

3.5.3 Cross-sectional sample preparation ................................................................... 85

Chapter 4 Effect of V/III ratio on binary GaAs nanowires ................................................... 91

4.1 Introduction ............................................................................................................... 91

4.2 Quality control of GaAs nanowire structures by limiting As flux in MBE ................... 92

Chapter 5 Effect of growth time on binary GaAs nanowires ............................................. 109

5.1 Introduction ............................................................................................................. 109

5.2 Phase purification of GaAs nanowires by prolonging growth duration in MBE ....... 110

Chapter 6 Effect of growth time on ternary InGaAs nanowires ........................................ 129

6.1 Introduction ............................................................................................................. 129

6.2 Unexpected formation of hierarchical structure in ternary InGaAs nanowires ........ 130

Chapter 7 Composition characteristics of ternary InGaAs nanowires .............................. 150

7.1 Introduction ............................................................................................................. 150

7.2 Composition gradient ternary InGaAs nanowires caused by strain relaxation ........ 151

Chapter 8 Conclusions and recommendations ................................................................ 168

8.1 Conclusions ............................................................................................................ 168

8.2 Recommendations .................................................................................................. 170

XII

List of Figures

Figure 2.1 (a) Schematic phase diagram is between the catalyst material (Au) and growth

material (Si) with red arrows showing the growth process. (b) Schematic illustration of VLS

growth mechanism of Si whisker from reaction of SiCl4 and H2 vapor phases. Black arrows

indicate the impingement of the vapor Si atoms to the catalyst, nanowires sidewall and

substrate and the diffusion of the atoms.

Figure 2.2 (a-c) TEM images show the top regions of the three grown GaAs nanowires

samples. (d-f) Corresponding electron diffraction patterns taken from the nanowires-

catalysts interfaces marked in (a-c).

Figure 2.3 Au-Ga phase diagram.

Figure 2.4 The model of novel VLS growth of a real nanowire with an elongated droplet on

the top.

Figure 2.5 Environmental TEM images (a) showing the Si nanowires growth with

increased growth duration, in which the corresponding ledge flow and growth direction are

illustrated in (b-d). (e) Plot of step edge position versus growth time.

Figure 2.6 Schematic of VSS nanowires growth. (a) Transport of the gas source to the

nanowire surface. (b) Atoms are delivered to the growth interface by surface or bulk

diffusion processes. (c) Atoms are incorporated at the growth interface leading to the

growth.

Figure 2.7 (a) Top-view SEM image of InAs nanowires arrays on a patterned substrate.

(b-c) Zoom-in 45°-tilted SEM images of grown nanowires.

Figure 2.8 (a,c,e,g) High-resolution TEM images of GaAs NWs tips grown with increased

As/Ga ratio. On the left, relevant diffractogram are shown, indicating the local crystal

structure. (b,d,f,h) Typical scanning TEM images from the nanowires tips (in false color to

enhance the detail visibility).

Figure 2.9 Schematic representations and the corresponding observed nanowires

(indicated by arrows) grown by the VLS mechanism with Ga catalysts at the tip of the

nanowires.

Figure 2.10 30°-tilted SEM images of Au-catalyzed GaAs nanowires with Au film thickness

of 0.5, 1, 2, 3, 5 nm, respectively. The inserts are top-view SEM images of corresponding

Au film after annealing.

XIII

Figure 2.11 (a) Radius dependencies of normalized growth rate of nanowires at different

growth temperature with fixed flux rate. (b) Diameter-dependent growth rates of GaAs

nanowires at different V/III ratios.

Figure 2.12 (a) 30°-tilted SEM images of nanowires grown at (a) 500 °C and V/III ratio of

24; (f) 580°C and V/III ratio of 1.2; and (k) 590°C and V/III ratio of 0.8. (b,g,l)

Corresponding TEM images taken from nanowires tops. Corresponding (c,h,m) HRTEM

images, (d,i,h) enlarged TEM images taken form the catalyst-nanowire interfaces, (e,j)

selective area electron diffraction (SAED) patterns and (o) Fast Fourier Transform (marked

in (b,g,l)).

Figure 2.13 (a) Overview and (b) high-resolution TEM images of Zn-doped InP nanowires.

(c) Schematic growth model for periodic twinning in nanowires (the contract angle between

droplet and nanowires is marked).

Figure 2.14 (a) 30°-tilted and (b) 45°-tilted SEM images of GaP nanotrees with 8.9

branches per tree on average. Top-view SEM images of (c) three branches directions and

(d) six branches directions.

Figure 2.15 SEM images of nanowires with their spacing varying from (a) 400 nm, to (b)

500 nm and (c) 1000 nm.

Figure 2.16 Schematic model of nanowires growth in different regimes. (a) Competitive

regime, d < λs, where the distance of neighbouring nanowires d is within the diffusion

length along the substrate λs. (b) Synergetic regime, λs < d < λg (diffusion length of Ga in

vapour), where the substrate collection areas become fully independent and wire-to-wire

interaction in gas phase occurs. (c) Independent regime, d > λg > λs.

Figure 2.17 45°-tilted SEM images of CBE-grown GaAs nanowires at different

temperatures of (a) 480°C, (b) 500°C, (c) 515°C, (d) 535°C and (e) 560°C. (f) Schematic

growth model illustrating nanowires axial growth and radial growth.

Figure 2.18 30°-tilted SEM images of GaAs grown by MOCVD at the varied growth

temperatures. The V/III ratio was 46 for all samples. Scale bars are 1 μm.

Figure 2.19 Surface density and growth rate of self-catalyzed GaAs nanowires as a

function of Ga flux (left panel) and As flux (right panel).

Figure 2.20 Top-view and side-view SEM images of GaAs nanowires grown under varied

V/III ratios (as described by Table 2.1).

Figure 2.21 Growth rates of InAs nanowires versus (a) In-flux for a fixed As-flux, (b) As-

flux with a fixed In-flux.

XIV

Figure 2.22 (a) Axial and radial nanowires growth rate and (b) tapering parameters of

GaAs nanowires with varied V/III ratio or III flow. (c) 40°-tilted SEM images of InAs

nanowires grown under varied V/III ratios. Scale bars are 1 µm.

Figure 2.23 45°-tilted SEM images of GaAs nanowires grown on GaAs {111}B substrates

for (a) 3 min, (b) 10 min and (c) 30 min at a growth temperature of 525°C and a V/III ratio

of ~1.5 in MBE.

Figure 2.24 Crystal structures of GaAs (Ga in purple and As in green). (a) Crystal

structure of zinc blende phase with unit cell on top and projection along <110> on bottom.

(b) Crystal structure of wurtzite phase with 2 2 2 cells on top and projection along

<110> on bottom.

Figure 2.25 SEM images and schematic diagram of a typical GaAs nanowire with rotated

truncated triangular cross section.

Figure 2.26 TEM images viewed along the <110> zone axis of InAs nanowires with

different diameters grown at 420°C. The fraction of wurtzite phase decreases as the

diameter increases: (a) 100%, (b) 97%, (c) 86% and (d) 15% wurtzite. The arrows mark

the planar defects of the grown nanowires.

Figure 2.27 SEM images of InAs nanowires grown at (a) a high V/III ratio of 54, tilled-view

and (b) a low V/III ratio of 15, top-view. (c) 35°-tilted SEM images of inclined nanowires in

(b). (d) Corresponding sketch of inclined-nanowires side-facets. TEM images of (a) a

typical inclined nanowire, and the nanowire top regions viewed along (f) [ ] and (g) [ 0]

zone-axis. (h) EDS spectra taken from the catalyst and nanowire. (i,j) Corresponding

SAED patterns. (k) High-resolution TEM images taken from the nanowire middle region.

Figure 2.28 TEM images of InAs nanowires grown along < >B or < > direction. (a-

d) High magnified TEM images of nanowires grown at varied temperatures and V/III ratios.

The insets are corresponding High-resolution TEM images. (e-h) Corresponding SAED

patterns.

Figure 2.29 Map of crystal structure of InP nanowires in the range of V/III ratio 44 to 700

and growth temperature of 400 to 510°C.

Figure 2.30 TEM images of a GaAs nanowire (a) at initial growth and (b) after nanowire

cooling. The inset in (a) shows the interface between two phases. (c) TEM images of InAs

nanowire superlattices, defined by 60 periods of alternating zinc blende and wurtzite

structure, viewed along the <110> zone axis.

XV

Figure 2.31 Band gaps and corresponding light wavelength λ for selected binary

semiconductors plotted as a function of lattice parameter.

Figure 2.32 45°-tilted SEM images of InGaAs nanowires grown under varied temperature

and V/III ratios, and corresponding STEM HADDF images of cross sections and their EDS

elements maps. Scale bars are 500 nm.

Figure 2.33 (a,d) TEM images of InGaAs nanowires grown at the same V/III and In/Ga

ratio but varied In flux. (b,e) EDS spectra taken from nanowires and catalysts as marked in

(a,d). (c,f) Corresponding TEM images taken from nanowires bottoms and schematic

illustration of nanowires structural and cross-sectional models.

Figure 3.1 Schematic illustration of a top-view of a typical growth chamber in the MBE

system, which is equipped with the RHEED monitor.

Figure 3.2 RHEED patterns (<110> azimuth) and SEM images of (100)-oriented GaP

substrates at (a) 350°C, 550°C and 600°C, respectively.

Figure 3.3 Signals generated when a high-energy beam of electrons interacts with a thin

sample.

Figure 3.4 Schematic illustrations of SEM layout and function (http://www.ammrf.org.

au/myscope/sem/practice/principles/layout.php).

Figure 3.5 Schematic illustration of interaction volume created by electron beam on the

sample surface with electron singles generated from different depths.

Figure 3.6 Schematic illustration of the “edge effect” on the SE yields.

(© http://laser.phys.ualberta.ca/~egerton/SEM/sem.htm)

Figure 3.7 (a) BSE and (b) SE images of a grain of sand that contains Si and Ti.

(©http://www.ammrf.org.au/myscope/sem/practice/principles/imagegeneration.php)

Figure 3.8 Ray diagram of a TEM in (a) imaging mode, creating bright field TEM images

on the viewing screen and (b) diffraction mode, the diffraction pattern from the sample

captured on the viewing screen.

Figure 3.9 (a) BF and (b) DF TEM image of a polycrystalline thin film of Bi. The insets are

the diffraction pattern taken from the one of single crystalline Bi. (c) Schematic illustration

of atomic planes bending at the edge dislocations. At position P, the angle equals to the

Bragg angle of the incidence electron beam. (d) A BF TEM image of cobalt mental.

Figure 3.10 Two-beam imaging condition and many-beam imaging condition and the

corresponding TEM images.

Figure 3.11 Schematic illustration of the interference of the electron beam waves and the

principles of the electron diffraction pattern.

XVI

(http://www.ammrf.org.au/myscope/tem/background/concepts/imagegeneration/diffraction/

beam/)

Figure 3.12 Schematic illustrations of (a) Z-contrast technique in a STEM and (b) the

HAADF detector setting for the Z-contrast imaging.

Figure 3.13 (a) HAADF STEM image of the interface between GaAs and Bi2Te3 viewed

from the respective <110> and <210> zone-axis. (b) Intensity profile of the atom column

marked in (a).

Figure 3.14 (a) Atomic resolution Z-contrast STEM image of Ba1.7Ca2.4Y0.9Fe5O13 film and

corresponding (b) Ba, (c) Ca, (d) Y, (e) Fe and (f) their overlap EDS maps.

Figure 3.15 (a,b) Images of the resin block after baking. The nanowires sample is at the

container bottom. (c) Image of the nanowires sample after the neighbouring resin being

cleaned. (d) Image of the resin after the substrate being removed. Red arrows mark the

sample.

Figure 3.16 Schematic illustration of nanowires cross sections preparation.

Figure 3.17 (a) Image of the prepared “sandwich” sample with the real sample in the

middle. (b) Schematic illustration of the sample adherence to the holder. (c) Sample

loading on the tripot. (d) The sample interface after polishing and (e) attached to the Cu

grid.

XVII

List of Tables

Table 2.1 Growth conditions for GaAs nanowires grown by MBE.

Table 2.2 Common side-facets in wurtzite and zinc blende structured nanowires.

Table 2.3 Surface energy of sidewall facets in both zinc blende and wurtzite phases.

Table 3.1 Comparison of MBE and MOCVD

XVIII

List of Abbreviations

VLS: vapor-liquid-solid

VS: vapor-solid

VSS: vapor-solid-solid

MBE: molecular beam epitaxy

CBE: chemical beam epitaxy

CVD: chemical vapor deposition

MOVCD: metal-organic chemical vapor deposition

SEM: scanning electron microscopy

FE-SEM: field-emission scanning electron microscopy

SE: secondary electrons

BSE: backscattered electrons

TEM: transmission electron microscopy

SAED: selected area electron diffraction

HRTEM: high-resolution transmission electron microscopy

BF: bright-field

DF: dark-field

STEM: scanning transmission electron microscopy

HADDF: high-angular annular dark field

EDS: energy dispersive X-ray spectroscopy

CCD: charge-coupled device

RHEED: reflection high energy electron diffraction

FIB: focused ion beam

1

Chapter 1

Introduction

1.1 Background

With an increasing demand of nanoscale electronic, optoelectronic and electromechanical

devices, semiconductor nanowire has become one of the most potential and promising

materials due to its peculiar electronic and optical properties.1,2 Among them, III-V

semiconductor nanowires, with direct band gap, high carrier mobility and high degree of

2

stoichiometry, have received great attention.3-5 To explore and exploit the properties in

nanowires more effectively and efficiently, realizing the controllable growth of the

nanowires and understanding their structural characteristics have become a rapid growing

trend of research.6-8

GaAs semiconductors, with wide direct band gap (~1.4 eV), high electron mobility (~

8500 cm-2V-1s-1) and large absorption coefficient, have been extensively studied during the

last decade.9,10 Many devices have been demonstrated using these semiconductor

nanowires, such as field-effect transistors,3 photodetectors11 and solar cells.12 One of the

most popular techniques for nanowires growth is the vapour-liquid-solid (VLS) nanowires

growth catalyzed by Au nanoparticles in molecular beam epitaxy (MBE).13-17 However,

compared with other growth technique like metal-organic chemical vapour deposition

(MOCVD), the growth behaviour and structural quality of GaAs nanowires in MBE has not

been comprehensively understood and precisely controlled.

Compared with binary III-V nanowires, ternary III-V nanowires have been attracting

an increasing interest as they allow the tunability of desired band gap for varied

applications by modulating the composition fraction of their alloys.18-19 For example,

InGaAs nanowires have demonstrated to be technologically important for a wide range of

applications, attributed to their tunable band gap ranging from the near-infrared region to

the infrared region by tuning the InGaAs alloy composition,20 which makes it constitute the

ideal material system for numerous optoelectronic devices, such as light-emitting diodes21

and nanolasers.22 However, with spontaneous nanowire lateral growth on the sidewall via

the vapor-solid mechanism, the composition of the InGaAs nanowires can be varied due to

the phase segregation.23 Therefore, understanding the chemical characteristics of the

InGaAs nanowires is critical for achieving uniform nanowires.

1.2 Objectives and scope

This thesis aims to study the growth of III-V semiconductor nanowires (mainly GaAs)

incorporated with In grown on GaAs {111}B substrate MBE. By performing advanced

electron microscopy investigations on grown nanowires, their physical and chemical

characteristics would be revealed.

To have a better understanding of multinary epitaxial nanowires growth, especially

the GaAs/InGaAs nanowires system, binary GaAs nanowires are first studied. By tuning

the growth parameters, such as growth temperature, V/III ratio and growth time, the growth

3

behaviour and mechanism of Au-assisted GaAs nanowires will be carefully investigated

and the nanowires quality should be under control.

Based on the study of GaAs nanowires, the growth behaviour of hetero-epitaxial

InGaAs nanowires growth on GaAs {111}B substrates should be better understood. In

particular, to obtain the uniform ternary nanowires, the chemical characteristics of grown

III-V nanowires will be investigated with varied In/Ga ratios. In particular, InGaAs

nanowires growth with high In/Ga ratios of 50 at.% and 85 at.% were investigated, which

have narrower band gaps and it is possible to extend the cut-off wavelength up to about

2.6 µm.

1.3 Thesis outline

This thesis includes eight chapters in all.

Chapter 1 presents the introduction of this thesis.

Chapter 2 is the critical literature review related to the main focus of this thesis, which

includes the review of (1) semiconductor nanowire growth mechanisms, such as VLS,

vapor-solid-solid (VSS) and vapor-solid (VS) mechanism; influence of growth parameters

such as growth temperature, V/III ratio and growth time on (2) binary nanowires

morphology and (3) their structural characteristics; (4) ternary nanowires growth, such as

III-III-V (e.g. AlGaAs, InGaAs) and III-V-V (e.g. GaAsP) nanowires, and their compositional

characteristics.

In chapter 3, methodology involved in this project will be introduced. First, the MBE

system will be described in details theoretically, which is the growth system for all

nanowires samples in this thesis. Then, analyses techniques such as scanning electron

microscopy (SEM), transmission electron microscopy (TEM) and energy-dispersive X-ray

spectroscopy (EDS) will be described.

Chapters 4 and 5 study the growth behavior and structural quality of the binary GaAs

nanowires. Chapter 4 depicts the quality control of GaAs nanowires by tuning the V/III ratio

and Chapter 5 shows the growth behavior and quality purification of GaAs nanowires by

prolonging the growth duration.

Based on the study of binary GaAs nanowires, the related ternary InGaAs nanowires

growth with varied In concentrations (In/(In+Ga) ratios) are studied and correspondingly

results are shown in chapters 6 and 7.

4

Chapter 6 shows the growth of hierarchical structure in InGaAs nanowires with the

increased growth duration. The nominal In concentration for InGaAs nanowires growth is

50 at.%.

Chapter 7 demonstrates the composition gradient InGaAs nanowires induced by

strain relaxation with nominal In concentration of 85 at.%.

Finally, chapter 8 summarizes the conclusions of this PhD thesis and the contributions

for future research work based on this thesis.

5

Chapter 2

Literature Review

2.1 Introduction

One-dimensional semiconductor nanostructures, such as nanowires, nanotube, nanobelts

and nanorods, whose diameter normally less than 100 nm, can exhibit extraordinary

properties compared with their counterpart bulk materials. For example, with the size

shrinking below a critical value, quantum size effect and quantum tunnelling effect can be

6

significant, which is the foundation of future microelectronic devices, such as solar cells,24

light-emitting diodes25 and transistors.26

Among various semiconductor materials, III-V semiconductor materials show the

advantages of narrow and direct band-gaps, high electron mobility and possibilities to form

heterostructures with other III-V materials. However, the lattice mismatch between two

different III-V materials can induce strain into the heterostructures, which will be

detrimental to the crystal quality of the heterostructures and their corresponding optical

properties.27 Nevertheless, this misfit strain can be largely released in the one-dimensional

III-V nanowires heterostructures that have large aspect ratios, which enable the high

crystallographic and optical qualities of the heterostructures.28 Therefore, the III-V

semiconductor nanowires and their heterostructures have attracted an increasing

attention. Moreover, based on the study of binary nanowires, the physical properties of

ternary nanowires and their growth behaviour have been further investigated, which broad

the applications of the III-V semiconductor nanowires.

In this chapter, the strategies of fabricating III-V semiconductor nanowires will be first

illustrated. The crystalline nanowires growth principles and their varied growth behaviour

(such as morphology and growth rate) under varied growth parameters (growth

temperature, V/III ratios and growth duration) will be described. Moreover, the research

status of the specific materials system (GaAs, InGaAs) will be depicted, in which the

objectives of this PhD project will be secured.

2.2 Synthesis of nanowires

The approaches of fabricating III-V nanowires can be divided into two categories: top-

down and bottom-up.29 Normally, top-down methods involve lithography, such as electron

beam lithography or focused ion beam milling and etching techniques, which are not likely

to form nanostructure with smaller scale size and high quality.30 Instead, bottom-up

methods can offer opportunities to eliminate such limitations. It consists of two main sub-

categories: template-directed and free-standing.

For template-directed growth, the success of it is to fabricate a suitable template, for

example V-grooves, step edges and mental membrane.31 This method can offer a great

chance to form nanostructure in uniform length and diameter. On the other hand, the over

dependent on the template can cause some problems, such as structural impurity and lack

of effective templates.

7

Free-standing growth, the most popular way to synthesize nanowires, relies most on

the anisotropy of growth rate. After nucleation at a certain point, nanowires elongate at one

direction with the highest growth rate, resulting in the one-dimensional shape. Within this

method, several growth mechanisms can be associated such as VLS growth, VSS growth

and VS growth.

2.2.1 Vapor-liquid-solid mechanism

One of the most popular mechanisms for growing nanoscale one-dimensional structures is

the VLS mechanism, which was first proposed by Wagner and Ellis in 1964 as an

explanation for silicon whiskers growth in vapor ambient assisted by gold liquid droplets on

a silicon substrate.32 It can be described in three stages as shown in Figure 2.1.33 At stage

I, nano-sized Au particles are formed as solid by depositing a thin film of Au on the

substrate or using aerosol Au nanoparticles directly. After the growth temperature is

achieved at around 800K, above the eutectic point, the Au catalyst begins to absorb vapor

Si atoms from the ambient to form liquid alloy. When the Si concentration is high enough in

the Au-Si alloy to cross the second liquidus line, Si starts to nucleate and form the liquid-

solid interface. Until the liquid droplet reaches the supersaturation level, it starts to

precipitate out Si and nuclei of Si grow in the form of nanostructure while the catalyst still

keeps liquid. As illustrated from this process, the VLS growth mechanism means that three

phases are coexisting during the nanostructure growth: the vapor phase of SiCl4 which

ends up with Si, the liquid phase referring to the Au-Si alloy at the top and the solid phase

of the grown nanostructure between the interface of the droplet and the substrate. This

theory was supported by Wu and Yang,17 who reported the VLS growth of Ge nanowires

with real time monitoring and identified these three stages by using in situ transmission

electron microscopy (in situ TEM).

8

Figure 2.1 (a) Schematic phase diagram is between the catalyst material (Au) and growth

material (Si) with red arrows showing the growth process. (b) Schematic illustration of VLS

growth mechanism of Si whisker from reaction of SiCl4 and H2 vapor phases. Black arrows

indicate the impingement of the vapor Si atoms to the catalyst, nanowires sidewall and

substrate and the diffusion of the atoms.33

During the last decade, VLS growth mechanism has been widely applied on the

growth of III-V semiconductor nanowires. For example, the study of Au-assisted GaAs

nanowire on (111)B substrate via VLS growth was conducted by Harmand et al34 in 2005.

Three samples were grown at the same conditions except for the cooling down process,

which ended up with three different phases of post-catalysts at room temperature: the

hexagonal Au7Ga2 phase (Figure 2.2d), the nearly pure Au of face-centred-cubic phase

(Figure 2.2e) and the orthorhombic AuGa phase (Figure 2.2f) (it is noted that the catalysts

9

size in three samples varies, which may cause the varied phases of the catalysts).

According to the Au-Ga phase diagram (Figure 2.3)35 and the growth temperature of

590°C that is above the eutectic point (~339°C), Au catalysts shall be at liquid state during

the nanowires growth. Besides, the VLS mechanism also was adopted to promote self-

catalysed or group III-assisted III-V nanowire growth, during which the presence of group-

III droplets (either Ga or In) can be witnessed on the top of corresponding nanowires.36,37

Figure 2.2 (a-c) TEM images show the top regions of the three grown GaAs nanowires

samples. (d-f) Corresponding electron diffraction patterns taken from the nanowires-

catalysts interfaces marked in (a-c).34

10

Figure 2.3 Au-Ga phase diagram.35

In the VLS growth process, catalyst plays an important role not only in lowering the

reaction energy but also inducing the growth of nanowires at the desired shape and size.

Therefore, the requirements for the catalyst particles can be critical and crucial.

To begin with, it is necessary for the catalyst to form a liquid alloy with the crystal

material at the growth temperature. In VLS growth, the most widely used catalyst particle

is Au, since it can form eutectic alloy with most of semiconductor materials at low

temperatures, such as Au/Ga: 339°C,38 Au/Sb: 360°C,39 Au/Bi: 241°C, Au/In: 454°C.35

11

However, no trace of As can be found in the Au catalyst.40,41 It is of great importance to

analyse the composition concentration in the catalyst, since nanowires are grown via the

nanowires-catalysts interface, the catalysts composition can provide tracks of the specific

catalysts and nanowires status during the growth. Normally, to form a eutectic alloy, the

growth temperature should be above the eutectic point to meet the thermodynamic

conditions for crystallization. However, Kodambaka’s group42 realized Au-assisted Ge

nanowires growth below the eutectic temperature. By using in situ microscopy, they found

the Au-Ge alloy can stay liquid with the dependence on Ge2H6 pressure as well, which can

be ascribed to the kinetic enrichment of the alloy particle.

In addition, the interface energies of vapor-liquid, vapor-solid and liquid-solid should

be well examined before choosing catalyst as they play a key part in determining the

catalyst shape.43 Recently, a new mode of VLS nanowire growth was reported by

Dubrovskii et al,14 who theoretically studied nanowires growth with catalysts in a varied

wetting case. Unlike the standard VLS growth, the non-spherical droplet sits on the corner

of the sidewall and buries the growth interface as shown in Figure 2.4. This elongated

droplet can be observed during self-catalyzed GaAs nanowires growth. As a consequence,

the nucleation on the triple phase line can be supressed and the structural purity would be

affected.

Figure 2.4 The model of novel VLS growth of a real nanowire with an elongated droplet on

the top.14

12

The VLS growth mechanism can be applied with many growth techniques especially

for epitaxial growth, such as chemical vapour deposition (CVD),44,45 MOCVD,46,47 chemical

beam epitaxy (CBE)48 and MBE.49-51 Among these technique, MBE is one of the most

promising and popular technique for III-V nanowire growth as it can guarantee high purity

and atomic accuracy of the material structure. For VLS-grown nanowires by MBE,

essential prerequisites should be satisfied.

From the thermodynamics point of view, the driving force for crystals growth is the

chemical potential difference (supersaturation) between the depositing materials in the

growing crystal and in vapor phase, which can be defined as under

the same temperature.52 Under certain conditions, the supersaturation level is catalyst size

dependent as reported to grow MgO nanowire by Takeshi et al, in which the VLS growth is

no longer available when the catalyst size is above a critical value.53 Additionally, based on

the Gibbs-Thomson effect, the supersaturation can be expressed as a function of catalyst

diameter :

0 4 / d (2.1)

Where is the effective difference between the chemical potential of depositing material

in the nanowire and in the vapour phase, is the same difference at a plane boundary

, is the specific free energy of the nanowire surface, and is the atomic volume of

the depositing material.52

From the kinetics point of view, two aspects should be under concern for crystalline

nanowires growth: (1) the presence of energy barrier for growing crystal and (2) the

identification of growth rate-determining steps. For the first part, the energy barrier for

growth process is more significant at relatively low temperature. However, with the help of

catalysts, the activation barrier can be lowered thus the growth rate would be enhanced.

For the second part, unlike MOCVD, the absorption of vapour source does not need the

thermal decomposition of precursors in MBE system. Thus, several steps of the growth

process can be drawn: (1) mass transport of growing material in the vapour phase; (2)

atom diffusion through the liquid droplet to the growth interface; (3) formation of the crystal

lattice at the liquid-solid interface; (4) atom diffusion along sidewall and substrate to the

nanowire-catalyst interface.52,54 Theoretically, the growth rate of nanowire is reported to

be dependent on supersaturation as (where refers to the kinetic

coefficient) and its diameter. Besides, Dubrovskii et al55 promoted a theoretical model

describing the dependence of the nanowire growth rate on its lateral dimension. It has

13

been demonstrated that the nanowire growth rate increases with the enlarged nanowire

diameter under higher growth temperature due to the Gibbs-Thomson effect, in which

nanowires growth rate is determined by the direct impingement of growth species to the

droplets and nanowires sidewalls. However, with the decreased temperature, the

contribution of atom diffusion becomes a determining factor for nanowire growth, which

minimizes this tendency.

2.2.2 Vapor-solid-solid mechanism

Since the VLS nanowires growth happens when the growth temperature is higher than the

eutectic point, the catalysts are in the liquid state. However, when the growth temperature

is lower or similar with the eutectic temperature, the catalysts are solid or quasi-solid,

which induce nanowires growth under the VSS growth mechanism. This mechanism was

first reported by Kamins et al in 2001,45 when they grew Ti-catalyzed Si nanowires by CVD

and claimed that the growth process is analogous to the traditional VLS growth

mechanism except for the solid catalysts. Hofmann et al56 monitored the VSS growth of Si

nanowires using Pd as catalysts with in situ observation for the first time, in which

nanowires growth was induced by the ledge propagation. As noted from Figure 2.5, the

growth-recorded TEM images show the ledge formation and layer by layer growth during

Si nanowires growth. Since the eutectic temperature Pd2Si alloy is 810°C that is higher

than the growth temperature, catalysts remain solid during nanowires growth, which shows

a faceted cuboid shape rather than the hemisphere shape, suggesting that nanowires

growth is under the VSS mechanism rather than the VLS mechanism. In addition, mental

catalysts like Al,57 Cu58 and Ni59 have been reported to promote VSS growth of Si and Ge

nanowires effectively as well. For III-V nanowire system, the VSS growth mechanism was

proposed by Persson et al48 in growing Au-catalyzed GaAs nanowires by CBE. By

combining in situ heating experiment and EDS measurement, the Au catalysts were

proved in solid state during nanowires growth. Indeed, the in situ observation of nanowires

growth is a powerful tool to confirm the nanowires growth and catalysts status. On the

other hand, the alloy phase diagram is another possible approach for ensure the catalysts

status during nanowires growth, although it is theoretically estimated under the

thermodynamics equilibrium conditions, in which the kinetics factors such as the growth

species flux and pressure should also be taken into consideration.

14

Figure 2.5 Environmental TEM images (a) showing the Si nanowires growth with

increased growth duration, in which the corresponding ledge flow and growth direction are

illustrated in (b-d). (e) Plot of step edge position versus growth time.56

Figure 2.6 Schematic of VSS nanowires growth. (a) Transport of the gas source to the

nanowire surface. (b) Atoms are delivered to the growth interface by surface or bulk

diffusion processes. (c) Atoms are incorporated at the growth interface leading to the

growth.60

15

The VSS nanowires growth process can be divided into four steps as illustrated in

Figure 2.6:60 (i) transport of growing material in vapour phase; (ii) gas atoms stick and

deposit on the catalyst surface and nanowires sidewalls; (iii) diffusion in the solid catalyst

particle; (iv) incorporation of depositing material into the growing crystal via the growth

interface (nanowire-catalyst interface). As can be noticed, the difference in mass transport

from VLS is that atoms have to go through solid catalyst to the growth interface. However,

this step cannot be distinguished as a growth-rate determination in all cases, because it

still depends on the diffusivity of atoms in the catalysts.

2.2.3 Vapor-solid mechanism

For III-V nanowires growing on Si substrates, the use of mental catalysts such as Au can

raise concerns about contaminations during the VLS growth. For instance, it has been

demonstrated that Au can induce deep level traps in Si which can affect the performance

of the nanowire-based electronics, by potentially reducing the minority carrier life time and

increasing the generation-recombination of electron-hole pairs.61 On this basis, catalyst-

free nanowires growth was proposed, which can be realized by patterned growth or

selective area growth.62 It has been demonstrated during InAs nanowires growth on Si

(111) substrate, patterned regions ( , Figure 2.7a) were prepared by electron-

beam lithography and wet chemical etching, on which a SiO2 film was uses as masking

materials.63 The patterns opening diameters were set as 60 nm and the pitch ranges from

400 nm to 800 nm, in which the InAs nanowires were grown on these opening with defined

diameter (60 nm) and controlled positions (Figure 2.7b,c). Such growth process not

involved any seed particles during growth is under the VS mechanism, which has been

demonstrated in a wide materials systems, ranging from oxides nanowires such as ZnO, to

semiconductor compound nanowires such as GaN.64,65 In addition, Ambrosini et al

demonstrated that in Ga-catalyzed GaAs nanowires growth in MBE, nanowires growth can

be modulated from the VLS mode to the VS mode.66 By tuning the As/Ga ratio, they

observed that the Ga droplet was consumed with the increased As/Ga ratio and nanowires

axial growth continues through the VS mechanism (Figure 2.8), in which nanowires tips

have varied morphology and crystal structures. This phenomenon can also be found

during InAs nanowires growth on bare Si (111) substrate.67

16

Figure 2.7 (a) Top-view SEM image of InAs nanowires arrays on a patterned substrate.

(b-c) Zoom-in 45°-tilted SEM images of grown nanowires.63

Figure 2.8 (a,c,e,g) High-resolution TEM images of GaAs NWs tips grown with increased

As/Ga ratio. On the left, relevant diffractogram are shown, indicating the local crystal

structure. (b,d,f,h) Typical scanning TEM images from the nanowires tips (in false color to

enhance the detail visibility).66

2.3 Morphology of III-V nanowires

It has been demonstrated that during nanowires growth, nanowires axial growth is

governed by the catalysts under the VLS growth, in which nanowires diameter depends on

the catalysts size. While, epitaxial nanowire shell growth can happen on the nanowires

sidewall, which is under the VS growth. Therefore, the nanowires morphology is

determined by the combination of nanowire axial and later growth. It has been

demonstrated that the nanowires morphology is closely linked to their possible

applications.68 For example, it has been established that nanoneddle structures with

17

ultrasharp tips are of great importance for the strong electric field improvement, which

points the way to the sharper sensors.69 On the other hand, nanowires are expected to

form uniform and controllable size, and the precise controlled growth sites for integration

into the devices. Accordingly, template directed methods have been taken advantage of.

For instance, Si nanowires70 and uniformly aligned ZnO nanowires71 in a large scale are

synthesized in a mental membrane with the help of Au. However, the entire dependence of

an available template can also act as limitation and hindrance.

2.3.2 Effect of catalysts

Catalysts size Since the nanowires diameter is determined by the catalysts size, the

selective of catalysts material can be a key factor for the nanowires morphology. It has

been demonstrated that during the Ga-catalyzed GaN nanowires growth, the Ga droplets

varies in lateral dimensions with varied growth condition, resulted in varied nanowires

morphology (Figure 2.9).72 Besides, it has also been demonstrated that by controlling the

Au film thickness, the Au catalysts size can be tuned after annealing, which results in

grown nanowires with varied diameters.73 Through quantitative study of the effect of Au

film thickness on the Au catalysts diameter and the grown nanowires, it was found that

with thicker the Au film, Au catalyst are bigger, which promote nanowires growth with

larger diameter (see Figure 2.10). This suggests that nanowires pre-growth process can

play an important role in modulating nanowires morphology.

Figure 2.9 Schematic representations and the corresponding observed nanowires

(indicated by arrows) grown by the VLS mechanism with Ga catalysts at the tip of the

nanowires.72

18

Figure 2.10 30°-tilted SEM images of Au-catalyzed GaAs nanowires with Au film thickness

of 0.5, 1, 2, 3, 5 nm, respectively. The inserts are top-view SEM images of corresponding

Au film after annealing.73

As mentioned previously, the diameter of nanowire can play an important role in

determining the supersaturation in catalyst, due to the classical Gibbs-Thompson effect.

With the increased nanowire diameter, the value of chemical potential difference

(supersaturation) increases, resulting in higher the nanowires growth rate. However, this

only happens within a certain range of the nanowires diameter. With further increasing the

nanowires diameter, the nanowires growth rate decreases, due to the diminished Gibbs-

Thompson effect. Additionally, As reported by Dubrovskii et al,74 there exists a critical

range of diameter for nanowires growth with a certain window of growth temperature,

below which nanowire cannot be promoted. On the other hand, the dependence of

nanowire growth rate on its diameter can be largely affected by growth temperature and

V/III ratio as plotted by Figure 2.11.74--76

Figure 2.11 (a) Radius dependencies of normalized growth rate of nanowires at different

growth temperature with fixed flux rate. (b) Diameter-dependent growth rates of GaAs

nanowires at different V/III ratios.74-76

19

Catalysts materials Additionally, it has been demonstrated that during the Pd-

catalyzed GaAs nanowires growth, the varied Ga concentration in the growth environment

leads to the varied Ga concentration in the Pd catalysts.77 As a consequence, the

nanowire-catalyst interfacial energy can be varied, causing the varied nanowire-catalyst

interface and thus the varied growth orientations and morphology (see Figure. 2.12a). This

nanowires morphology dependence of the catalyst-nanowire interfacial energy was applied

to modulate nanowires morphology. It has been demonstrated that in InP nanowires

growth doped with Zn, twinned zinc-blende superlattice structure was formed with zigzag

nanowires sidewall (see the high-resolution TEM (HRTEM) image in Figure 2.13c), which

can be resulted from the liquid-solid contact angle caused by the Zn dopants.78

Consequently, the solid-liquid and liquid-vapor surface energy were altered. Without Zn,

such periodic structure was disappeared.

Figure 2.12 (a) 30°-tilted SEM images of nanowires grown at (a) 500 °C and V/III ratio of

24; (f) 580°C and V/III ratio of 1.2; and (k) 590°C and V/III ratio of 0.8. (b,g,l)

Corresponding TEM images taken from nanowires tops. Corresponding (c,h,m) HRTEM

images, (d,i,h) enlarged TEM images taken form the catalyst-nanowire interfaces, (e,j)

selective area electron diffraction (SAED) patterns and (o) Fast Fourier Transform (marked

in (b,g,l)).77

20

Figure 2.13 (a) Overview and (b) high-resolution TEM images of Zn-doped InP nanowires.

(c) Schematic growth model for periodic twinning in nanowires (the contract angle between

droplet and nanowires is marked).78

Catalysts position On the other hand, by controlling the dispersion sites of the

catalysts, the nanowires morphology can be controlled. Dick et al successfully grew the

branched GaP and InP nanotrees in MOCVD.79 First, the Au-assisted nanowires was

initially grown under the VLS growth as a trunk. By spreading Au droplets on the grown

nanowire trunk though aerosol deposition method, the nano-branches are grown.79 The

diameter and number of these nano-branches can be controlled by tuning the

corresponding diameter and density of Au particles (Figure 2.14). This controllable

branched nanostructure may offer possibility in mimicking photosynthesis nanotrees.

Besides, although the unexpected nanowires kinking morphology can be observed in

III-V nanowires growth, particularly in the axial hetero-structured nanowires, due to the

varied interfacial energy of two different materials with catalysts, the control of such

morphology can also enable the possible nanowire-based applications.80

21

Figure 2.14 (a) 30°-tilted and (b) 45°-tilted SEM images of GaP nanotrees with 8.9

branches per tree on average. Top-view SEM images of (c) three branches directions and

(d) six branches directions.79

2.3.2 Effect of nanowire density

It has been demonstrated that nanowires morphology can be largely affected by the

nanowires density, i.e. the wire-to-wire spacing.22 As noted in Figure 2.15, with increasing

the nanowires spacing, nanowires grow shorter, suggesting that nanowires growth rate

decreases with the decreased nanowires density. Accordingly, a growth model was

proposed, in which, when nanowires spacing is small, nanowires compete with their

neighboured nanowires for the growth materials (see Figure 2.16a), which is under the

competitive growth mode. In this case, the nanowires growth rate is dependent on their

ability to absorb growth materials, which can be determined by the catalysts size. While, at

a relative large nanowires spacing, the material collection area on the substrate is not

overlapped but nanowires are still close, so that nanowires growth rate can be increased

by the interaction with neighboured nanowires via atoms diffusion (Figure 2.16b). On the

22

other hand, with further increasing the nanowires spacing, nanowires growth is isolated,

which is under the independent growth regime (Figure 2.16c).81 Overall, it is noted that the

nanowires grow faster with relatively small spacing, which can be ascribed to the catalytic

growth precursor decomposition in MOVCD. This process is accelerated with the

increased catalysts number per unit area. Therefore, nanowires growth with narrow

spacing can have more growth species, which leads to faster nanowires growth. However,

this tendency has been demonstrated to be related to the nanowires growth temperature

by Joyce et al.82 At higher temperature, nanowires diffusion length is longer, in which the

density dependence of nanowires growth is more significant: strong competition occurs in

high nanowires density, leading to the short nanowires, while, nanowires height increases

in the low-density nanowires growth. On the other hand, when the growth temperature is

low, atoms diffusivity becomes low, leading to the decreased competition between

neighboured nanowires, so that the nanowires length is similar in both high-density and

low-density growth. This dependence on nanowires density is largely dependent on the

atoms diffusion length on {111}B substrates.82

On the other hand, the synergetic effect of narrow nanowires spacing on the

nanowires growth is varied with nanowires patterned growth by CBE, in which nanowires

grow shorter under the small nanowires spacing that leads to the growth materials

competition.83

Figure 2.15 SEM images of nanowires with their spacing varying from (a) 400 nm, to (b)

500 nm and (c) 1000 nm.82

23

Figure 2.16 Schematic model of nanowires growth in different regimes. (a) Competitive

regime, d < λs, where the distance of neighbouring nanowires d is within the diffusion

length along the substrate λs. (b) Synergetic regime, λs < d < λg (diffusion length of Ga in

vapour), where the substrate collection areas become fully independent and wire-to-wire

interaction in gas phase occurs. (c) Independent regime, d > λg > λs.82

2.3.2 Effect of growth temperature

It has been demonstrated that nanowires can only successfully grow within a certain range

of temperature. This temperature growth window depends on varied III-V materials and the

other growth parameters, mainly V/III ratios.84 It has been established that for Au-assisted

GaAs nanowires grown by MBE, when the growth temperature is below 339°C (Figure

2.3), the lowest eutectic point of the possible Au-Ga alloy, the catalyst is solidified and

nanowires fail to grow.85-87 As noted from Figure 2.17, GaAs nanowires growth

morphology is varied with varied growth temperature: at low growth temperature,

24

nanowires show tapering morphology, while, with increasing the growth temperature, the

tapering decreases and nanowires have uniform diameters along the growth directions.

The tapering degree can be evaluated by the ratio of nanowires axial growth to their lateral

growth (see Figure 2.17f). Besides, nanowires growth rate also shows dependence on the

growth temperature: with a certain range of the growth temperature, nanowires grow faster

with the increased temperature; while, with further increased temperature, nanowires grow

shorter and two-dimensional layer growth on the substrates becomes significant (Figure

2.17d,e).48

In addition, it is noted from Figure 2.17b, nanowires have the pencil shape

morphology, which can be particularly observed in GaAs nanowires growth in MBE87 and

other III-V nanowires systems, such as InAs88 and GaN.89 This interesting morphology

endows nanowires with straight sidewalls rather than the inclined high-index planes.

However, the growth mechanism of the pencil-shape morphology is still controversial,

which has been demonstrated as layer-by-layer growth, similar with the epitaxial thin film

growth, or the step-by-step growth.90 On the other hand, the tapering morphology is

commonly observed during III-V nanowires growth.91,92 The growth mechanism of this

morphology is established mostly depends on the theoretical mass diffusion model,93 in

which growth temperature plays an important role in affecting the atoms diffusivity. For

short nanowires, when atoms diffusion length is larger than the nanowire length, atoms

can diffuse to the catalysts along nanowire sidewall and participate in nanowires axial

growth. When the sidewall instead of nanowire length exceeds the diffusion length, atoms

tend to stay at incorporating into the catalyst. Consequently, nanowires lateral growth

increases, causing nanowire tapering.

25

Figure 2.17 45°-tilted SEM images of CBE-grown GaAs nanowires at different

temperatures of (a) 480°C, (b) 500°C, (c) 515°C, (d) 535°C and (e) 560°C. (f) Schematic

growth model illustrating nanowires axial growth and radial growth.48

On the other hand, in III-V nanowires growth by MOCVD, such as GaAs or InAs, the

effect of growth temperature on nanowires growth is varied. As demonstrated by Joyce et

al,94 nanowires grown above 400°C epitaxial and straight aligned, while, nanowires grown

at a lower temperature show kinking and other irregular morphology (see Figure 2.18).

Since nanowires axial growth is driven by the catalysts under the VLS growth mechanism,

it was thought that under the relative low growth temperatures, Au catalysts are not

completely in the liquid state, so that the catalysts cannot well induce perfect crystal

growth. Besides, it can be seen that above 450°C, nanowires axial growth decreases,

which can be results from the onsets of nanowires lateral growth and two-dimensional

growth on the substrate under the high temperature. These two-dimensional growths

compete with the one-dimensional nanowires axial growth, leading to the mass transport

limitation for the nanowire axial growth and hence the short grown nanowires. In terms of

nanowires morphology, nanowires tapering is minimized at the relatively low growth

temperature, in which nanowires lateral growth is not significant.

26

Figure 2.18 30°-tilted SEM images of GaAs grown by MOCVD at the varied growth

temperatures. The V/III ratio was 46 for all samples. Scale bars are 1 μm.94

2.3.3 Effect of V/III ratio

Another key growth parameter for III-V nanowires growth is V/III ratio, the beam equivalent

ratio of group-III to group-V flux rate in MBE, which is a scale of quantity of gas atoms per

unit time. The V/III ratio, together with respective III and V flux rate can exert a great

impact on the nanowire growth rate and morphology.

It has been demonstrated in self-catalyzed GaAs nanowires growth by MBE,

nanowires growth rate mainly depends on the As flux rate rather than the Ga flux (Figure

2.19).95 This is because the group-III Ga is the catalysts material, nanowires growth

depends on the rate of the combination rate of Ga and As and their incorporation to the

lattice.96,97 On the other hand, for Au-catalyzed GaAs nanowires grown under varied V/III

ratios (see Table 2.1), with increasing the V/III ratio by increasing the V flux, nanowires

grow longer (Figure 2.20a-c).98 Besides, nanowires have tapering morphology under the

low V/III ratio, while nanowires have uniform diameter along their whole growth direction

under the high V/III ratio. However, it is noted that under the low two-dimensional growth

rate (samples H and J in Figure 2.20), increasing the V/III ratio have a minor impact on

nanowires growth rate and results in similar nanowires length.

27

Figure 2.19 Surface density and growth rate of self-catalyzed GaAs nanowires as a

function of Ga flux (left panel) and As flux (right panel).95

Table 2.1 Growth conditions for GaAs nanowires grown by MBE.98

Sample Temperature (°C) V/III ratio Two-dimensional growth rate (nm s-1)

E 600 1.1 0.28

F 600 1.7 0.28

G 600 2.3 0.28

H 600 2.3 0.14

I 600 2.3 0.07

J 600 4.6 0.14

28

Figure 2.20 Top-view and side-view SEM images of GaAs nanowires grown under varied

V/III ratios (as described by Table 2.1).98

Besides, for Au-catalyzed InAs nanowires grown by MBE, by increasing In flux with a

constant As flux, nanowires grow faster due to the increase of In atoms supply for

nanowires axial growth. However, with further increased In flux, nanowires growth

reduced, possibly owing to the decreased V/III ratio, in which As is relatively insufficient for

nanowires growth, leading to the desorption of III atoms (Figure 2.21a). By estimating the

corresponding V/III ratio, the peak of nanowire growth rate occurs at the V/III ratio of ~15,

and nanowire growth rate decreases with reduced V/III ratio. On the other hand, by

increasing As flux with a constant In flux, nanowires growth rate keeps increasing (Figure

2.21b). It is noted that this trend is obtained at the V/III ratio ranging from 357 to 21, which

is consistent with that in Figure 2.21a.99

29

Figure 2.21 Growth rates of InAs nanowires versus (a) In-flux for a fixed As-flux, (b) As-

flux with a fixed In-flux.99

In MOCVD growth, the V/III also has a distinct effect on the growth rate and

morphology of III-V nanowires (Figure 2.22). It can be seen that by increasing the V/III

ratio via increasing the V flux, nanowires axial growth slightly increases while their lateral

growth increases more significantly. On the other hand, by keeping the same V/III ratio

while increasing the III and V flux, both nanowires axial and lateral growth increase (Figure

2.22a).100 This trend suggests that nanowires axial growth is governed by both mass

transport and growth kinetics, while nanowires lateral growth is mostly kinetics limited.

Besides, under a constant V/III ratio, with increasing the III and V flux, nanowires tapering

30

reduces, which can be ascribed to the increased III atoms for nanowires axial growth, so

that the nanowires aspect ratio increases (Figure 2.22b). In addition, similar dependence

of InAs nanowires growth on the V/III ratio has been demonstrated. From Figure 2.22c,

with increasing the V/III ratio, nanowires growth in both axial and lateral directions

increases, owing to the minimized activation energy for nanowires growth at relatively high

V/III ratios. The nanowires axial growth rate reaches highest at a critical V/III ratio. When

the V/III ratio exceeds this value, the axial rate drops dramatically and the radial growth

dominates, resulting in severe nanowires tapering. With even higher V/III ratio as 93 and

above, nanostructure shows island-like and irregular shape. This may be due to the high

As pressure, the diffusion of group-III can be diminished, which promotes three

dimensional growth.101 This impact of V/III ration on nanowires growth rate and

morphology is also consistent with that during InP102 and InAs94 nanowires growth.

Figure 2.22 (a) Axial and radial nanowires growth rate and (b) tapering parameters of

GaAs nanowires with varied V/III ratio or III flow. (c) 40°-tilted SEM images of InAs

nanowires grown under varied V/III ratios. Scale bars are 1 µm.100,101

31

2.3.4 Effect of growth duration

Growth duration is a basic growth parameter for nanowires growth. By examining

nanowires growth at different stages of the growth, nanowires growth status and the

related growth mechanism can be unveiled. It has been demonstrated in GaAs nanowires

growth by MBE, with increased growth duration, nanowires morphology transits from the

rod-shape to the pencil-shape (Figure 2.23), which can be ascribed to the increased

nanowires later growth.92

On the other hand, it has been demonstrated during the MOVCD growth of GaAs

nanowires, the nanowires length can be controlled by the growth duration. It was found

that nanowires length is proportional to growth duration, which shows a linear relationship.

This tendency was also observed under the varied growth temperatures. However, it

should be noted that the maximum experimental data is obtained under the growth time of

120s, which is short compared with normal nanowire growth durations.75

Figure 2.23 45°-tilted SEM images of GaAs nanowires grown on GaAs {111}B substrates

for (a) 3 min, (b) 10 min and (c) 30 min at a growth temperature of 525°C and a V/III ratio

of ~1.5 in MBE.92

2.4 Structural characteristics of III-V nanowires

2.4.1 Crystal structure of III-V nanowires

Most III-V semiconductors, including the III-As, normally adopt cubic zinc blende crystal

structure. However, for III-V nanowires, hexagonal wurtzite structure can also be found,

32

ranging from GaAs nanowires by MBE to GaP nanowires by MOCVD.103 The only

difference of these two structures is the stacking sequence in the close-packed directions.

As illustrated in Figure 2.24, for zinc blende phase, the stacking sequence is described as

ABCABC… along the <111> direction, compared with ABAB…along <0001> direction for

wurtzite phase. Furthermore, zinc blende structure is polar in nature owning two different

<111> directions, i.e. <111>A and <111>B. Among them, <111>B is the preferential growth

direction in nanowire growth, due to the lower surface energy.104 The interruption of the

stacking sequence can lead to defects or different phase. Often in zinc blende structure,

the stacking sequence of ABCACB... can be found, where the plane A is called a twin

plane. The polytypism of wurtzite-zinc blende phase can be normally found in

semiconductor nanowire with varied growth conditions. Besides, the 4H and 6H polytype

have been recently observed in a few cases in III-V nanowires.105,106 The stacking

sequences observed from TEM are noted as ABCBABCB…and ABCBACABCBAC…

respectively.

Figure 2.24 Crystal structures of GaAs (Ga in purple and As in green). (a) Crystal

structure of zinc blende phase with unit cell on top and projection along <110> on bottom.

(b) Crystal structure of wurtzite phase with 2 2 2 cells on top and projection along

<110> on bottom.

33

2.4.2 Side facets of III-V nanowires

From SEM and HRTEM study of III-V nanowires, Dick et al103 reported that {1120} and

{1100} facets can be normally found in wurtzite structure, while, {112} and {110} facets are

common in zinc blende phase. As described in Table 2.2,107 the atomic planes are

projected along the growth directions, normally [111]B direction in zinc blende structure and

[0001] direction in wurtzite structure. Both of them have hexagonal shape with symmetry in

cross section. Two types of nanowire side facets can be noticed: {110}-type parallel to the

cleavage plane in III-V compounds and {112}-type with a rotation of 30° to (110) planes. It

is noteworthy that only {112} facets in zinc blende is polar and nano-faceting, which can be

divided into {112}A and {112}B planes. A plane is defined as terminating with the group V

atoms which can rescue three dangling bonds. While, B plane ends with group III atoms

that can only rescue one or two broken bonds. Such polarity can result in different

chemical and physical behaviour of non-equivalent {112} sidewalls.108 With {112}A growing

faster, the asymmetric truncated hexagonal cross section can be observed in

nanowires.109 As can be seen from Figure 2.25, a number of overlapping truncated

triangles are organized in orientations rotated with 60°. Theoretically, nanowires tend to

keep their side facets with lowest energy as shown in Table 2.3.110-112 However, it is also

related to the growth parameters. As reported by Plante et al,113 the side facets of GaAs

nanowires are intermixed by both type {110} and {112} planes and even nanowires grow

with no side facets when the growth temperature and V/III increase.

Table 2.2 Common side-facets in wurtzite and zinc blende structured nanowires.107

34

Figure 2.25 SEM images and schematic diagram of a typical GaAs nanowire with rotated

truncated triangular cross section.109

Table 2.3 Surface energy of sidewall facets in both zinc blende and wurtzite phases.

Facet type Surface energy (J/m2)110,111 Surface energy (meV/Å2)112

(112) 1.73 1.04 25/23/21

(110) 1.50 0.86 21/19/16

( ) 1.30 0.69 30/28/29

) 1.50 0.73 22/20/17

2.4.3 Effect of catalysts diameter

Catalysts diameter is a key factor that not only impact nanowires growth rate and

morphology but also the crystal structure of nanowires. It has been well demonstrated that

small nanowires diameter favours wurtzite structure while large sized nanowires tend to

form zinc blende structure. For instance, Johansson et al114 have demonstrated that by

controlling the diameter of nanowire, the phase purity of InAs nanowires can be

modulated. As described in Figure 2.26, nanowires with small diameters have the pure

wurtzite structure, while, with the increasing of nanowire diameter, the nanowires sections

of pure wurtzite phase decreases and the density of stacking faults increases. When the

diameter becomes large, there is a gradual change from pure wurtzite phase to a mixed

twining zinc blende phase. This catalysts diameter dependence of crystal structure has

been confirmed by nanowires growth under varied growth temperature. In addition,

Shtrikman et al115,116 have observed pure wurtzite phase in ultrathin GaAs and InAs

nanowires (~10 nm).

35

Figure 2.26 TEM images viewed along the ⟨110⟩ zone axis of InAs nanowires with

different diameters grown at 420 °C. The fraction of wurtzite phase decreases as the

diameter increases: (a) 100%, (b) 97%, (c) 86% and (d) 15% wurtzite. The arrows mark

the planar defects of the grown nanowires.114

Besides, as Hoang et al reported,117 MBE grown GaAs nanowires with small diameters

(30-40 nm) showed wurtzite phase with high density of stacking faults, whereas pure zinc

blende structure was obtained with diameter at around 60 nm.

The purity of nanowires phase with varied diameters can be ascribed to the nanowires

nucleation status at the nanowire-catalyst interface. It has been demonstrated that the

Gibbs free energy for nanowires nucleation is reversely related to the supersaturation of

the catalysts during nanowires growth ( ), which is the chemical potential of the Au-In

particle and the nanowire (take Au-catalyzed InAs nanowires growth for an

example). The supersaturation of the catalysts is given by114

(2.2)

where is Boltzmann’s constant, is the absolute temperature, is the atomic

36

fraction of In in Au-In alloy particle with radius and the solubility of In, is the surface

energy of the particle, and is the atomic volume of indium in the particle. According to

the Equation 2.2, the catalysts supersaturation increases with the decreased nanowire

diameter. Based on the theoretical modelling and calculation of the nucleation energy of

the wurtzite and zinc blende phases, high catalysts supersaturation growth favours

wurtzite structured nanowires, while low supersaturation growth favours zinc blende

structured nanowires, which is in good agreement with the experimental observations that

small catalysts induce pure nanowires structure. As demonstrated by Zhang et al,118

during InAs nanowires growth in MBE, small Au catalysts absorb more In during nanowires

growth, due to the high droplet surface tension, and the grown nanowires structure is

defect-free wurtzite structure. While, large Au catalysts have low In concentration during

the growth, which results in defected wurtzite structure. This founding shows a good

agreement with the theoretical predictions, verifying the strong dependence of catalysts

diameter on the III-V nanowires crystal structures and their quality. Besides, it has been

demonstrated that for a given supersaturation, the diameter crossover of wurtzite and zinc

blende structure is smaller, below which no nanowires can grow.

On the other hand, it has been established that the side-facets or sidewall of

nanowires grown on (111)B substrate can be related to their diameter. As reported by

Chen et al,119 during GaAsP/GaAs heterostructured nanowires growth, nanowires sidewall

varies with varied nanowires diameters. They found that when nanowires diameter is

smaller than 200 nm, {110} nanowires side facets formed, which have low surface energy.

While, wider nanowires form {112} side facets. Since thin nanowires have large surface to

volume ratio, they prefer to adopt low surface energy-surfaces.

Besides, Li et al have found periodic faceting in <111> grown Si nanowires, which

have six {112} sidewalls.120 This facet period was demonstrated to be dependent on the

nanowires diameter,121 in which the facet size decreases with the increased nanowires

diameter, suggesting that under a constant length, wider nanowires have more facets

change. The corresponding theoretical estimation indicates that the nanowires sidewall

surface energy increases with the additional formation energy of edges and apexes at

nanowires sidewall. Therefore, for thin nanowires preferring to keep low-energy surfaces, it

is energy favourable for them to form a small number of faceted surfaces on the sidewall.

37

2.4.4 Effect of growth direction

In general, III-V nanowires tend to grow in preferential crystal directions such as <111>,

<110>, <001> and <112>, which have relatively low surface energy. With varied growth

directions, nanowires structural quality and facets formation would be varied. It has been

well demonstrated that III-V nanowires tend to grow in <111> directions regardless of

substrate orientations, owing to the lower surface energy.104 For example, <111>B growth

direction has been found for InP nanowires grown on (110), (111)A and (111)B

substrates,122 and GaAs nanowires grown on (111)B and (100) substrates.123 However,

this growth behaviour can cause some drawbacks, for example, the stacking faults

formation. Krishnamachari et al found that, in InP nanowires growth on InP(001) substrate,

both <001>- and <111>B-grown nanowires were observed, in which nanowires grown

along <001> direction were perfect zinc blende structured, whereas the <111>B-grown

nanowires have planar defects along the whole nanowire.124 This is because the formation

energy difference between wurtzite and zinc blende structures along <111>B direction is

small, so that the mix phase of these two different structures are easy to form. Besides,

they also found that the nanowires side-facets are varied with varied growth orientations.

The <001>-grown nanowires have rectangular shaped cross sections, indicating the four

{110} side facets, which is consistent with the <001>-grown InAs nanowires by MBE.

While, the <111>B grown nanowires normally adopt hexagon shaped cross sections, in

which the side facets depends on their crystal structure and growth conditions.

On the other hand, for Au-catalyzed III-V nanowires grown along <111>B diretcions,

their crystal structure are mainly wurtzite structure, which is varied from those grown in

MOCVD. However, zinc-blende structured III-V nanowires can be obtained in non-<111>B

nanowires growth by using non-{111} substrates. For example, Shtrikman observed that

<001>-GaAs nanowires were grown successfully on the (311)B substrates, which have

pure zinc-blende structure. In addition, Zhang in our group recently reported his work of

InAs nanowires growth on GaAs {111}B substrate by MBE.125 In his work, nanowires were

grown at varied V/III ratios, in which both vertical and inclined nanowires were observed

under the low V/III ratio growth. Through SEM investigations, two types of nanowires

directions were confirmed as <111> and <110> directions (Figure 2.27). The <110> grown

nanowires has an elongated hexagonal cross section with four {111} and two {110} side-

facets. Further TEM investigations show that the <110>-grown InAs nanowires have pure

38

Figure 2.27 SEM images of InAs nanowires grown at (a) a high V/III ratio of 54, tilled-view

and (b) a low V/III ratio of 15, top-view. (c) 35°-tilted SEM images of inclined nanowires in

(b). (d) Corresponding sketch of inclined-nanowires side-facets. TEM images of (a) a

typical inclined nanowire, and the nanowire top regions viewed along (f) [ ] and (g) [ ]

zone-axis. (h) EDS spectra taken from the catalyst and nanowire. (i,j) Corresponding

SAED patterns. (k) High-resolution TEM images taken from the nanowire middle region.125

zinc-blende structure along their growth direction, and the the nanowire-catalyst interface

is (111)B surface, which is energy favourable to keep the low interfacial energy of the

nanowire-catalyst interface. While, the <111>B grown InAs nanowires have the wurtzite

structure. The crystal structure difference in the InAs nanowires growth along varied

directions was clarified by the nucleation energy difference at the nanowire-catalysts

interface for nanowires growth. It was descried that the nucleation barrier is droplet surface

energy dependent, which can be determined by the contact angle between droplet and

nanowire side facet. This contact angle is small in the inclined nanowires growth, which

gives rise to a lower nucleation energy, suggesting that the zinc-blended structured <110>-

nanowires are energy favourable to grow.

39

2.4.5 Effect of growth parameters

Faced with the fact that the dominated <111> nanowires growth could cause planar

defects formation, researchers are searching other ways to achieve high purity III-V

nanowires. From the theoretical point of view, Glas et al108 built a growth model about the

preference of wurtzite and zinc blende structure formation in III-V nanowires.94 They

pointed out that the fundamental reasons behind the phase difference of nanowires is the

supersaturation in the catalyst droplets during the growth, which would affect the

nucleation of both wurtzite and zinc-blende structure. This theory has laid a solid

foundation for later research study on the structure perfection of III-V nanowires. So far, it

has been well demonstrated that by tuning the growth parameters, such as growth

temperature, V/III ratio, group III or group V flux during III-V nanowires growth, nanowires

structural quality can be under control.

In the MOCVD growth of III-V nanowires, Joyce et al systematically investigated the

impacts of growth temperature and V/III ratio on nanowires quality and obtained pure

wurtzite and zinc blende crystal structured nanowires.126 It has been demonstrated during

the growth of InAs nanowires, by tuning the growth temperature and V/III ratio, nanowires

structure was tuned from pure wurtzite structure to pure zinc blende structure (Figure

2.28). Under the growth with low temperature and high V/III ratio, pure wurtzite structured

InAs nanowires formed. While, under the growth of high temperature and low V/III ratio,

zinc blende structured nanowires were grown. Between these two ends, nanowires show a

trend of phase transition with formation of stacking faults or twin defects. The structure

dependence of the catalyst supersaturation can be impacted by the growth temperature

and V/III ratio. According to the nanowire nucleation model, high growth temperature and

high III concentration in the catalysts give rise to the high catalysts supersaturation, which

leads to the wurtzite structure. While, under a low temperature or high V/III ratio, the

nucleation and growth of zinc blende nanowires are more energetically favourable, due to

the low-energy surface construction of zinc blende surface.

Similar phenomena were also observed when they grow GaAs nanowires under the

varied temperature and V/III ratio.127 Besides, it was found that GaAs nanowires grown

under a relatively low temperature can secure high quality, in which however nanowires

morphology may be irregular and their growth direction can be inclined. To solve this

problem, a two-temperature growth strategy was proposed, in which nanowires nucleation

40

Figure 2.28 TEM images of InAs nanowires grown along < >B or < > direction. (a-

d) High magnified TEM images of nanowires grown at varied temperatures and V/III ratios.

The insets are corresponding High-resolution TEM images. (e-h) Corresponding SAED

patterns.126

at a high temperature while growth at a low temperature. By comparing with the single-

temperature, this strategy works effectively in achieving the high-quality GaAs nanowires.

In addition, the impacts of temperature and V/III ratio on the nanowire structure have

also been demonstrated during the InP nanowires growth by MOCVD.103,128 As depicted

by Paiman et al,129 the InP nanowires favour zinc blende crystal structure under low

growth temperature and a wide range of V/III ratios from 44 to 350. While, wurtzite

structure formed at the high temperature above 485°C and the V/III ratio between 44 and

700 (see Figure 2.29). These results show that high V flux and high temperature promotes

the wurtzite structured nanowires growth, due to the high catalysts supersaturation, which

is consistent with previous study of III-V nanowires growth in MOCVD.

On the other hand, during the MBE growth of III-V nanowires, such as GaAs and

InAs, the influence of growth parameters on the nanowires structure is varied. In the self-

catalyzed GaAs nanowires growth, nanowires form the pure wurtzite structure with the

decreased V/III ratio. While, under the high V/III ratio growth, nanowires adopt zinc blende

41

Figure 2.29 Map of crystal structure of InP nanowires in the range of V/III ratio 44 to 700

and growth temperature of 400 to 510°C.129

structure, in which the twin defects decreases with the increased As flux.130,131 Besides, it

has been demonstrated in the InAs nanowires growth by MBE, nanowires grown at a low

V/III ratio have the defect-free wurtzite structure, while stacking faults were formed in

nanowires with the increased V/III ratio.125 This can be ascribed to the catalysts

supersaturation, which increases with the increased III in the catalysts during nanowires

growth. Additionally, Dubrovskii et al presented an interesting phenomenon in the Au-

catalyzed GaAs nanowires growth, by applying a two-temperature growth.132 They found

that nanowires grown at low temperature is pure wurtzite structure, while with the

increased growth temperature, nanowires form the zinc blende structured sections.

However, instead of a sharp interface between them, a faulted wurtzite/zinc-blende mix

structured nanowire section of around 30 nm can be observed. This observation is varied

from those in the MOCVD growth, which can be ascribed to the small nanowires diameters

and the impact of surface energy.

In Au-assisted III-V nanowire growth by MBE, few work has reported pure zinc

blende structured nanowires grown along <111> directions.133 It is common for zinc blende

phase being observed in the nanowire at very beginning and very end of the nanowires

growth, where the low III concentration leads to the low catalysts supersaturation which

favours zinc blende structure formation.108,115 As shown in Figure 2.30a, at the initial stage

of nanowire growth, the flux of Ga and As started to be pumped in the growth environment.

42

Figure 2.30 TEM images of a GaAs nanowire (a) at initial growth and (b) after nanowire

cooling. The inset in (a) shows the interface between two phases. (c) TEM images of InAs

nanowire superlattices, defined by 60 periods of alternating zinc blende and wurtzite

structure, viewed along the <110> zone axis.108, 134

After nucleation, the insufficient source supply causes low catalysts supersaturation,

leading to zinc blende phase formation. During the cooling down stage (see Fig 2.30b), the

Ga source is usually switched off. The maintenance of As would consume Ga residual in

the chamber and catalysts particles, forming a so-called “neck” region. The structure of

this region normally exhibits zinc blende structure. By utilizing this feature, Dick et al

demonstrated that by applying this flux interruption via the group-V switched on and off,

InAs nanowire superlattices were grown, which has 60 periods of alternating wurtzite and

zinc blende phases.134

In addition, it has been established that the variation in Ga or As flux would result in

different nanowire side-facets. Under the growth of a constant Ga flux, with increasing of

As flux, nanowires side-facets changed from {1100} planes to {2110} planes. While,

nanowires grown under a constant As flux, by changing the Ga flux, their side-facets did

not change and kept the {2110} planes.135 This phenomenon indicates that nanowire side-

facets formation may be more sensitive to the change of As flux. Besides, as presented by

Jiang et al for understanding the true shape of GaAs nanowires grown in MOCVD,136 they

did the cross-sectional study of the grown zinc blende structured GaAs nanowires, and

found that the newly grown nanowires actually have the high-index side-facets, and with

nanowires lateral growth, those high-energy side-facets transit to the low energy {112}

planes. They also found that the nanowires side-facets are temperature dependent. By

increasing the growth temperature during the later stage of the growth, nanowires side-

facets transit from {112} facets to the lower-energy {110} facets.

43

2.5 Ternary III-V nanowires

Compared with binary nanowires systems, ternary III-V semiconductor nanowires provide

large possibilities in composition tunability of the nanowires, which in turn, brings in

possibilities of band structure engineering and the design of novel nanowire-based

devices.137-140 For example, InGaAs nanowires system, as one of important ternary III-V

nanowires systems, has been attract an increasing interest. Comparing with GaAs,

InGaAs compound has higher electron mobility, lattice match with InP which is a

technologically important material for optoelectronic devices.141 The band gap of InGaAs

can be tuned between 0.34 (InAs) and 1.42 eV (GaAs) by tuning the chemical composition

of InGaAs compound alloy (see Figure 2.31).142-147 This band-gap range is between the

infrared and near-infrared regions, making ternary InGaAs nanowires as building blocks for

possible optoelectronic devices, such as light-emitting diodes,148 nanolasers149 and

photodetector.150

Figure 2.31 Band gaps and corresponding light wavelength λ for selected binary

semiconductors plotted as a function of lattice parameter.142-147

Basically, there are two types of ternary III-V nanowires: III-III-V nanowire, such as

AlGaAs139,151 and InGaAs,152-154 and III-V-V nanowires, such as GaAsP.155-157 In particular,

for Au-catalyzed III-III-V nanowires, both group-III elements can be alloyed with Au

catalysts, and incorporate to the lattice by alloyed with group-V at the catalyst-nanowire

44

interface, which is under the VLS growth mode.118,158 Therefore, the nanowires

composition depends on the rate of III atoms bonded with V or the interfacial energy of III-

V at the catalyst-nanowire interface. On the other hand, nanowires lateral growth can

simultaneously take place at nanowires sidewall, which is under the VS growth mode. This

can cause phase segregation in the ternary III-V nanowires, which form the core-shell

structures and leads to the complicated composition distribution in the nanowires.159,160

The composition of grown nanowires can be varied from the nominal III/III ratio, which

makes the composition of ternary III-V nanowires difficult to control.

2.5.1 Growth behavior

Similar with binary nanowires morphology, cylinder shape and tapered morphology can be

observed in ternary III-V nanowires, which are determined by nanowires lateral growth.132

Additionally, nanowires morphology is also catalyst-size dependent. For example, in

MOCVD growth of InGaAs nanowires, nanowires catalyzed by small droplets have uniform

diameters, while, nanowires have tapering morphology when their catalysts are larger.

Besides, Kim et al demonstrated that during InGaAs nanowires growth, nanowires tapering

degree, the ratio of nanowires axial growth to lateral growth, increases with the increased

nominal In concentration.22,152

On the other hand, the growth rate of ternary III-V nanowires can also be impacted by

the growth parameters, such as growth temperature and V/III ratio, catalysts diameters

and nanowires density. It has been demonstrated that, during InGaAs ternary nanowires

growth under a Ga-enriched environment (Ga/In = 7/3), with increasing the V/III to a critical

value below 10, nanowires axial growth increases. While, with further increasing the V/III

ratio beyond the critical value, nanowires axial growth decreases.161 This can be ascribed

to the Gibbs-Thomson effect, in which As can be easily desorbed from the catalysts due to

the large curvature of small catalysts. Therefore, increasing the V/III ratio enable more As

bonding with group-III atoms, leading to the increased nanowires growth. With the further

increased V/III ratio, nanowires growth is III-controlled, in which high V/III ratio causes less

III available for nanowires growth. Additionally, a relatively low temperature can reduce

InGaAs nanowires tapering, which is consistent with that in binary GaAs and InAs

nanowires growth. This can be attributed to the reduced possibility of overcoming the

kinetic barrier for nanowires lateral growth, or the reduced atoms diffusivity, so that the

atom incorporation to the nanowires sidewalls is decreased.

45

Besides, in the MBE growth of InGaP nanowires, it was found that nanowires with

significant later growth have pencil-shape morphology, while, nanowires with less lateral

growth have tapering morphology. The nanowires later growth was found to increase with

decreased group-III flux. Additionally, nanowires axial growth rate increases with the

increased III flux.160

On the other hand, axial heterostructured nanowires containing ternary III-V

nanowires, such as InGaP with InAs,162 InGaP with InP,163 GaInP with GaP164 and

GaAsSb with GaAs,105 were grown straight and had sharp interfaces. However, the sharp

interfaces and straight growth directions of heterostructured nanowires with ternary III-V

nanowires can be only in one direction, such InGaAs on top of InAs, not in the growth of

InGaAs on top of the GaAs.165 This fact can be ascribed to the growth thermodynamics. If

the interface energy between Au catalysts and nanowire A is lower than nanowire B,

straight nanowires can be grown for A growing on top of B, but nanowires can be inclined

or kinked if B grows on top of A to minimize the system energy.166-168

2.5.2 Compositional characteristics

Impact of nanowires diameter It has been demonstrated that for ternary III-V nanowires,

their composition can be morphology dependent. As reported by Hou et al, InGaAs

nanowires with uniform diameters along the nanowire length have uniform composition

distributions. However, the nanowires composition can be deviated from the nominal

composition.20 Moreover, during the ternary InGaP nanowires growth with varied catalysts

diameters, nanowires grown by small catalysts have lower In concentration, which may be

due to the Gibbs-Thompson effect. This In incorporation limitation was eliminated with

increasing the catalysts diameter or the group-III flux, in which In supersaturation in the

catalysts can be increased, leading to the increased In incorporation in the catalysts.160

Impact of growth parameters Similar with binary nanowires, ternary nanowires

morphology can be also determined by growth parameters. During the ternary InGaAs

nanowires grown at high V/III ratio by MOCVD, with increasing the growth temperature,

nanowires lateral growth increases, leading to the increased nanowires tapering.82 While,

under the growth of low V/III ratio, increasing the growth temperature would result in faster

nanowires growth, and further the increased temperature would also ends up with

nanowires tapering. On the other hand, under the growth of low temperature, increasing

46

Figure 2.32 45°-tilted SEM images of InGaAs nanowires grown under varied temperature

and V/III ratios, and corresponding STEM (scanning TEM) HADDF images of cross

sections and their EDS elements maps. Scale bars are 500 nm.168

the V/III ratio would leads to the faster nanowires axial growth; while under the growth of

high temperature, increasing the V/III ratio would leads to the increased nanowires

tapering and decreased nanowires axial growth. Compositional measurements of the

cross sections of the nanowires indicate that the nanowires with uniform diameters along

the nanowires have uniform compositions; while, tapered nanowires grown at low

temperature have core-shell structure, due to phase segregation, but tapered nanowires

grown at high temperature have uniform compositions (Figure 2.32), possibility due to the

increased atoms interdiffusion at higher temperature.169

Apart from the impact of growth temperature and V/III ratio on III-V ternary nanowires

chemical composition, the III/III or V/V ratio can also exert an impact on nanowires growth

and chemical compositions. Guo et al have demonstrated that for Ga-rich InGaAs

nanowires growth under the low In/Ga ratios, nanowires composition can be

modulated.23,170 Nanowires were grown at the same V/III ratio (As/(In+Ga)) of 44 and

In/Ga ratio of 1/19, but the varied In flux rate of 7.13 × 10-7 mol/min for sample A, while

8.20 × 10-6 mol/min for sample B. From EDS analyses on both samples, it was found that

only In was contained in the post-growth catalysts, and In concentrations in the nanowire

bottoms are higher than those in the nanowires tops. Besides, nanowires show Moiré

47

Figure 2.33 (a,d) TEM images of InGaAs nanowires grown at the same V/III and In/Ga

ratio but varied In flux. (b,e) EDS spectra taken from nanowires and catalysts as marked in

(a,d). (c,f) Corresponding TEM images taken from nanowires bottoms and schematic

illustration of nanowires structural and cross-sectional models.23

contrast in the nanowires bottoms, suggesting the formation of core-shell structure, which

was verified by further cross-sectional study of grown nanowires. By comparison of two

samples, it was noted that In concentration in the catalysts was lower in sample A, in

which no In was contained in nanowires cores. Consequently, nanowires in sample A

formed the core-shell structure, with GaAs core and In-enriched InGaAs shell; while

nanowires in sample B formed the core-shell structure with Ga-enriched nanowire cores

and In-enriched shells. These results suggested that there is an energy barrier for In

concentration in the catalysts to incorporate to the nanowires. Based on the first-principle

calculations, it was verified that the thermodynamics affinity of Au-In is higher than that of

Au-Ga. Therefore, Au catalysts prefer to keep In while expel Ga to the nanowires. When

nanowires were grown at the In-limited growth environment, In supersaturation in the

catalysts is low, so that In cannot precipitate out from catalysts and incorporate to the

nanowires. However, with increasing the In flux rate, more In can be absorbed by catalysts,

leading to the higher In supersaturation in the catalysts, and thus In incorporation to the

nanowire cores.

Impact of planar layer growth During the one-dimensional nanowires growth, two-

dimensional planar layer simultaneously occurs on the substrate.171-173 It has been

48

demonstrated that during ternary GaAsP nanowires growth on the GaAs substrates, the

group-V incorporation to the GaAsP planar layer was varied from the nominal As/P

ratio,174,175 due to requirement of misfit strain minimization caused by the lattice mismatch,

which in turn leading to the nanowires composition deviated from the nominal

composition.176-178 Therefore, the co-growth of nanowires and planar layer can result in the

uncontrollable nanowires composition.

2.6 Unclear issues

2.6.1 About GaAs nanowires

In order to have a better understanding the growth of ternary InGaAs nanowires or

InGaAs/GaAs heterostructured nanowires, binary GaAs nanowires should be first

investigated. From current study, the morphology of GaAs nanowires can be under control

by modulating the growth parameters, such as growth temperature and V/III ratio.

However, there are still a few issues need to be clarified.

Zinc-blended structure has not been achieved in GaAs nanowires grown along

<111>B directions in MBE.

A more comprehensive study of GaAs nanowires growth by tuning the respective

group-III and group-V flux.

The growth mechanism of pencil-shape GaAs nanowires.

2.6.2 About InGaAs nanowires

So far, although ternary InGaAs nanowires were successfully grown on the GaAs

substrates, the nominal In concentration was low, possibility to achieve the lattice match

between nanowires and substrate to ensure the straight grown nanowires. Therefore, the

growth of In-rich InGaAs nanowires on GaAs substrates needs to be achieved.

Besides, since ternary nanowires can also have lateral growth, core-shell structures

can spontaneously form in the nanowires, nanowires with uniform composition are hard to

achieve. Apart from that, planar layer growth also simultaneously takes place on the

substrate, in which the composition can be varied from the nominal composition. This in

turn, can lead to the even more complex nanowires compositional characteristics. In this

49

regard, the investigation of InGaAs nanowires chemical compositions and growth behavior

under varied growth parameters, particularly the In/Ga ratio, is necessary to achieve the

uniform ternary InGaAs nanowires.

References

1. Gudiksen, M. S.; Lauhon, L. J.; Wang, J.; Smith, D. C.; Lieber, C. M., Growth of

Nanowire Superlattice Structures for Nanoscale Photonics and Electronics. Nature 2002,

415, 617-620.

2. Robinson, R. D.; Sadtler, B.; Demchenko, D. O.; Erdonmez, C. K.; Wang, L.-W.;

Alivisatos, A. P., Spontaneous Superlattice Formation in Nanorods through Partial Cation

Exchange. Science 2007, 317, 355-358.

3. Han, N.; Wang, F.; Hou, J. J.; Xiu, F.; Yip, S.; Hui, A. T.; Hung, T.; Ho, J. C.,

Controllable P–N Switching Behaviors of GaAs Nanowires via an Interface Effect. ACS

Nano 2012, 6, 4428-4433.

4. Miao, J.; Hu, W.; Guo, N.; Lu, Z.; Zou, X.; Liao, L.; Shi, S.; Chen, P.; Fan, Z.; Ho, J. C.;

Li, T.-X.; Chen, X. S.; Lu, W., Single InAs Nanowire Room-Temperature Near-Infrared

Photodetectors. ACS Nano 2014, 8, 3628-3635.

5. Zhou, X.; Lu, M.-Y.; Lu, Y.-J.; Jones, E. J.; Gwo, S.; Gradečak, S., Nanoscale Optical

Properties of Indium Gallium Nitride/Gallium Nitride Nanodisk-in-Rod Heterostructures.

ACS Nano 2015, 9, 2868-2875.

6. Chiaramonte, T.; Tizei, L. H. G.; Ugarte, D.; Cotta, M. N. A., Kinetic Effects in Inp

Nanowire Growth and Stacking Fault Formation: The Role of Interface Roughening. Nano

Lett. 2011, 11, 1934-1940.

7. Chen, Y.; Burgess, T.; An, X.; Mai, Y.-W.; Tan, H. H.; Zou, J.; Ringer, S. P.; Jagadish,

C.; Liao, X., Effect of a High Density of Stacking Faults on the Young’s Modulus of Gaas

Nanowires. Nano Lett. 2016, 16, 1911-1916.

8. Conesa-Boj, S.; Kriegner, D.; Han, X.-L.; Plissard, S.; Wallart, X.; Stangl, J.;

Fontcuberta i Morral, A.; Caroff, P., Gold-Free Ternary III–V Antimonide Nanowire Arrays

on Silicon: Twin-Free Down to the First Bilayer. Nano Lett. 2013, 14, 326-332.

9. Han, N.; Hou, J. J.; Wang, F.; Yip, S.; Yen, Y.-T.; Yang, Z.-X.; Dong, G.; Hung, T.;

Chueh, Y.-L.; Ho, J. C., GaAs Nanowires: From Manipulation of Defect Formation to

Controllable Electronic Transport Properties. ACS Nano 2013, 7, 9138-9146.

50

10. Xia, H.; Lu, Z.-Y.; Li, T.-X.; Parkinson, P.; Liao, Z.-M.; Liu, F.-H.; Lu, W.; Hu, W.-D.;

Chen, P.-P.; Xu, H.-Y.; Zou, J.; Jagadish, C., Distinct Photocurrent Response of Individual

GaAs Nanowires Induced by N-Type Doping. ACS Nano 2012, 6, 6005-6013.

11. Prechtel, L.; Padilla, M.; Erhard, N.; Karl, H.; Abstreiter, G.; Fontcuberta I Morral, A.;

Holleitner, A. W., Time-Resolved Photoinduced Thermoelectric and Transport Currents in

GaAs Nanowires. Nano Lett. 2012, 12, 2337-2341.

12. Hu, S.; Chi, C.-Y.; Fountaine, K. T.; Yao, M.; Atwater, H. A.; Dapkus, P. D.; Lewis, N.

S.; Zhou, C., Optical, Electrical, and Solar Energy-Conversion Properties of Gallium

Arsenide Nanowire-Array Photoanodes. Energy Environ. Sci. 2013, 6, 1879-1890.

13. Jeon, N.; Dayeh, S. A.; Lauhon, L. J., Origin of Polytype Formation in VLS-Grown Ge

Nanowires through Defect Generation and Nanowire Kinking. Nano Lett. 2013, 13, 3947-

3952.

14. Dubrovskii, V. G.; Cirlin, G. E.; Sibirev, N. V.; Jabeen, F.; Harmand, J. C.; Werner, P.,

New Mode of Vapor−Liquid−Solid Nanowire Growth. Nano Lett. 2011, 11, 1247-1253.

15. Kolasinski, K. W., Catalytic Growth of Nanowires: Vapor–Liquid–Solid, Vapor–Solid–

Solid, Solution–Liquid–Solid and Solid–Liquid–Solid Growth. Curr. Opin. Solid State

Mater.Sci. 2006, 10, 182-191.

16. Glas, F., Chemical Potentials for Au-assisted Vapor-Liquid-Solid Growth of III-V

Nanowires. J. Appl. Phys. 2010, 108, 073506.

17. Wu, Y.; Yang, P., Direct Observation of Vapor-Liquid-Solid Nanowire Growth. J. Am.

Chem. Soc. 2001, 123, 3165-3166.

18. Kuykendall, T.; Ulrich, P.; Aloni, S.; Yang, P., Complete Composition Tunability of

InGaN Nanowires Using a Combinatorial Approach. Nat. Mater. 2007, 6, 951-956.

19. Chen, H.; Feenstra, R. M.; Northrup, J. E.; Zywietz, T.; Neugebauer, J., Spontaneous

Formation of Indium-Rich Nanostructures on InGaN(0001) Surfaces. Phys. Rev. Lett. 2000,

85, 1902-1905.

20. Hou, J. J.; Wang, F.; Han, N.; Xiu, F.; Yip, S.; Fang, M.; Lin, H.; Hung, T. F.; Ho, J. C.,

Stoichiometric Effect on Electrical, Optical, and Structural Properties of Composition-

Tunable InxGa1–X As Nanowires. ACS Nano 2012, 6, 9320-9325.

21. Boroditsky, M.; Krauss, T. F.; Coccioli, R.; Vrijen, R.; Bhat, R.; Yablonovitch, E., Light

Extraction from Optically Pumped Light-Emitting Diode by Thin-Slab Photonic Crystals.

Appl. Phys. Lett. 1999, 75, 1036-1038.

51

22. Kim, Y.; Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Paladugu, M.; Zou, J.;

Suvorova, A. A., Influence of Nanowire Density on the Shape and Optical Properties of

Ternary InGaAs Nanowires. Nano Lett. 2006, 6, 599-604.

23. Guo, Y.-N.; Xu, H.-Y.; Auchterlonie, G. J.; Burgess, T.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C.; Shu, H.-B.; Chen, X.-S.; Lu, W.; Kim, Y.; Zou, J., Phase Separation

Induced by Au Catalysts in Ternary InGaAs Nanowires. Nano Lett. 2013, 13, 643-650.

24. Law, M.; Greene, L. E.; Johnson, J. C.; Saykally, R.; Yang, P., Nanowire Dye-

Sensitized Solar Cells. Nat. Mater. 2005, 4, 455-459.

25. Qian, F.; Gradečak, S.; Li, Y.; Wen, C.-Y.; Lieber, C. M., Core/Multishell Nanowire

Heterostructures as Multicolor, High-Efficiency Light-Emitting Diodes. Nano Lett. 2005, 5,

2287-2291.

26. Thelander, C.; Mårtensson, T.; Björk, M. T.; Ohlsson, B. J.; Larsson, M. W.;

Wallenberg, L. R.; Samuelson, L., Single-Electron Transistors in Heterostructure

Nanowires. Appl. Phys. Lett. 2003, 83, 2052-2054.

27. Brandt, O.; Tapfer, L.; Ploog, K.; Bierwolf, R.; Hohenstein, M., Effect of in Segregation

on the Structural and Optical Properties of Ultrathin InAs Films in GaAs. Appl. Phys. Lett.

1992, 61, 2814-2816.

28. Dick, K. A.; Kodambaka, S.; Reuter, M. C.; Deppert, K.; Samuelson, L.; Seifert, W.;

Wallenberg, L. R.; Ross, F. M., The Morphology of Axial and Branched Nanowire

Heterostructures. Nano Lett. 2007, 7, 1817-1822.

29. Biswas, A.; Bayer, I. S.; Biris, A. S.; Wang, T.; Dervishi, E.; Faupel, F., Advances in

Top–down and Bottom–up Surface Nanofabrication: Techniques, Applications &Amp;

Future Prospects. Adv. Colloid Interface 2012, 170, 2-27.

30. Tseng, A. A.; Kuan, C.; Chen, C. D.; Ma, K. J., Electron Beam Lithography in

Nanoscale Fabrication: Recent Development. IEEE Trans. Electron. Packag. Manuf. 2003,

26, 141-149.

31. Dick, K. A., 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 2008, 54, 138-173.

32. Wagner, R. S.; Ellis, W. C., Vapor‐Liquid‐Solid Mechanism of Single Crystal Growth.

Appl. Phys. Lett. 1964, 4, 89-90.

33. Ross, F. M., Controlling Nanowire Structures through Real Time Growth Studies. Rep.

Prog. Phys. 2010, 73, 114501.

52

34. Harmand, J. C.; Patriarche, G.; Ṕŕ-Laperne, N.; Ḿrat-Combes, M. N.; Travers, L.; Glas,

F., Analysis of Vapor-Liquid-Solid Mechanism in Au-assisted GaAs Nanowire Growth. Appl.

Phys. Lett. 2005, 87, 1-3.

35. Okamoto, H., Desk Handbook: Phase Diagram for Bianry Alloys, 2nd ed.: ASM

International, Materials Park, OH, USA, 2000.

36. Fontcuberta i Morral, A.; Colombo, C.; Abstreiter, G.; Arbiol, J.; Morante, J. R.,

Nucleation Mechanism of Gallium-assisted Molecular Beam Epitaxy Growth of Gallium

Arsenide Nanowires. Appl. Phys. Lett. 2008, 92, 063112.

37. Mattila, M.; Hakkarainen, T.; Lipsanen, H.; Jiang, H.; Kauppinen, E. I., Catalyst-Free

Growth of In(As)P Nanowires on Silicon. Appl. Phys. Lett. 2006, 89, 063119.

38. Cooke, C. J.; Hume-Rothery, W., The Equilibrium Diagram of the System Gold-

Gallium. Journal of the Less Common Metals 1966, 10, 42-51.

39. Kim, J.; Jeong, S.; Lee, H., A Thermodynamic Study of Phase Equilibria in the Au-Sb-

Sn Solder System. Journal of Elec Materi 2002, 31, 557-563.

40. Ann I. Persson; Magnus W. Lsrsson; Stig Stenstom; B. Jonas Ohlsson; Lars

Samulson; Wallenberg, L. R., Solid-Phase Diffusion Mechanism for GaAs Nanowire

Growth. Nat.Mater. 2004, 3, 677-681.

41. Dick, K. A.; Deppert, K.; Mårtensson, T.; Mandl, B.; Samuelson, L.; Seifert, W., Failure

of the Vapor−Liquid−Solid Mechanism in Au-Assisted Movpe Growth of InAs Nanowires.

Nano Lett. 2005, 5, 761-764.

42. Kodambaka, S.; Tersoff, J.; Reuter, M. C.; Ross, F. M., Germanium Nanowire Growth

Below the Eutectic Temperature. Science 2007, 316, 729-732.

43. Algra, R. E.; Verheijen, M. A.; Feiner, L.-F.; Immink, G. G. W.; Enckevort, W. J. P. v.;

Vlieg, E.; Bakkers, E. P. A. M., The Role of Surface Energies and Chemical Potential

During Nanowire Growth. Nano Lett. 2011, 11, 1259-1264.

44. Chang, P.-C.; Fan, Z.; Wang, D.; Tseng, W.-Y.; Chiou, W.-A.; Hong, J.; Lu, J. G., Zno

Nanowires Synthesized by Vapor Trapping CVD Method. Chem. Mater. 2004, 16, 5133-

5137.

45. Kamins, T. I.; Stanley Williams, R.; Basile, D. P.; Hesjedal, T.; Harris, J. S., Ti-

Catalyzed Si Nanowires by Chemical Vapor Deposition: Microscopy and Growth

Mechanisms. J. Appl. Phys. 2001, 89, 1008-1016.

46. Seifert, W.; Borgström, M.; Deppert, K.; Dick, K. A.; Johansson, J.; Larsson, M. W.;

Mårtensson, T.; Sköld, N.; Patrik T. Svensson, C.; Wacaser, B. A.; Reine Wallenberg, L.;

53

Samuelson, L., Growth of One-Dimensional Nanostructures in MOVPE. J. Cryst. Growth

2004, 272, 211-220.

47. Murray, S. L.; Newman, F. D.; Murray, C. S.; DavidMWilt; MarkWWanlass; Ahrenkiel,

P.; Messhaw, R.; Siergiej, R. R., MOCVD Growth of Lattice-Matched and Mismatched

InGaAs Materials for Thermophotovoltaic Energy Conversion. Semicond. Sci. Technol.

2003, 18, 202-208.

48. Persson, A. I.; Ohlsson, B. J.; Jeppesen, S.; Samuelson, L., Growth Mechanisms for

GaAs Nanowires Grown in CBE. J. Cryst. Growth 2004, 272, 167-174.

49. Paek, J. H.; Nishiwaki, T.; Yamaguchi, M.; Sawaki, N., Catalyst Free MBE-VLS

Growth of GaAs Nanowires on (111)Si Substrate. Phys. Status Solidi 2009, 6, 1436-1440.

50. Haapamaki, C. M.; LaPierre, R. R., Mechanisms of Molecular Beam Epitaxy Growth in

InAs/InP Nanowire Heterostructures. Nanotechnology 2011, 22, 335602.

51. Cirlin, G. E.; Dubrovskii, V. G.; Soshnikov, I. P.; Sibirev, N. V.; Samsonenko, Y. B.;

Bouravleuv, A. D.; Harmand, J. C.; Glas, F., Critical Diameters and Temperature Domains

for MBE Growth of III–V Nanowires on Lattice Mismatched Substrates. Phys. Status Solidi

2009, 3, 112-114.

52. Givargizov, E. I., Fundamental Aspects of VLS Growth. J. Cryst. Growth 1975, 31, 20-

30.

53. Yanagida, T.; Nagashima, K.; Tanaka, H.; Kawai, T., Mechanism of Critical Catalyst

Size Effect on MgO Nanowire Growth by Pulsed Laser Deposition. J. Appl. Phys. 2008,

104.

54. Verheijen, M. A.; Immink, G.; de Smet, T.; Borgström, M. T.; Bakkers, E. P. A. M.,

Growth Kinetics of Heterostructured GaP−GaAs Nanowires. J. Am. Chem. Soc. 2006, 128,

1353-1359.

55. Dubrovskii, V. G.; Sibirev, N. V., General Form of the Dependences of Nanowire

Growth Rate on the Nanowire Radius. J. Cryst. Growth 2007, 304, 504-513.

56. Hofmann, S.; Sharma, R.; Wirth, C. T.; Cervantes-Sodi, F.; Ducati, C.; Kasama, T.;

Dunin-Borkowski, R. E.; Drucker, J.; Bennett, P.; Robertson, J., Ledge-Flow-Controlled

Catalyst Interface Dynamics During Si Nanowire Growth. Nat. Mater. 2008, 7, 372-375.

57. Wang, Y.; Schmidt, V.; Senz, S.; Gosele, U., Epitaxial Growth of Silicon Nanowires

Using an Aluminium Catalyst. Nat. Nanotechnol. 2006, 1, 186-189.

58. Arbiol, J.; Kalache, B.; Cabarrocas, P. R. i.; Morante, J. R.; Morral, A. F. I., Influence

of Cu as a Catalyst on the Properties of Silicon Nanowires Synthesized by the Vapour–

Solid–Solid Mechanism. Nanotechnology 2007, 18, 305606.

54

59. Kang, K.; Gu, G. H.; Kim, D. A.; Park, C. G.; Jo, M.-H., Self-Organized Growth of Ge

Nanowires from Ni−Cu Bulk Alloys. Chem. Mater. 2008, 20, 6577-6579.

60. Lensch-Falk, J. L.; Hemesath, E. R.; Perea, D. E.; Lauhon, L. J., Alternative Catalysts

for VSS Growth of Silicon and Germanium Nanowires. J. Mater. Chem. 2009, 19, 849-857.

61. Allen, J. E.; Hemesath, E. R.; Perea, D. E.; Lensch-Falk, J. L.; LiZ.Y; Yin, F.; Gass, M.

H.; Wang, P.; Bleloch, A. L.; Palmer, R. E.; Lauhon, L. J., High-Resolution Detection of Au

Catalyst Atoms in Si Nanowires. Nat. Nanotechnol. 2008, 3, 168-173.

62. Sato, T.; Hiruma, K.; Shirai, M.; Tominaga, K.; Haraguchi, K.; Katsuyama, T.; Shimada,

T., Site‐Controlled Growth of Nanowhiskers. Appl. Phys. Lett. 1995, 66, 159-161.

63. Tomioka, K.; Motohisa, J.; Hara, S.; Fukui, T., Control of InAs Nanowire Growth

Directions on Si. Nano Lett. 2008, 8, 3475-3480.

64. Fan, H. J.; Bertram, F.; Dadgar, A.; Christen, J.; Krost, A.; Zacharias, M., Self-

Assembly of ZnO Nanowires and the Spatial Resolved Characterization of Their

Luminescence. Nanotechnology 2004, 15, 1401–1404.

65. Zhou, S. M.; Feng, Y. S.; Zhang, L. D., A Physical Evaporation Synthetic Route to

Large-Scale GaN Nanowires and Their Dielectric Properties. Chem. Phys. Lett. 2003, 369,

610-614.

66. Ambrosini, S.; Fanetti, M.; Grillo, V.; Franciosi, A.; Rubini, S., Vapor-Liquid-Solid and

Vapor-Solid Growth of Self-Catalyzed GaAs Nanowires. AIP Adv. 2011, 1, 042142.

. Dimakis, E.; L hnemann, J.; Jahn, .; Breuer, S.; Hilse, M.; Geelhaar, L.; iechert, H.,

Self-Assisted Nucleation and Vapor–Solid Growth of InAs Nanowires on Bare Si(111).

Cryst. Growth Des. 2011, 11, 4001-4008.

68. Moewe, M.; Chuang, L. C.; Crankshaw, S.; Chase, C.; Chang-Hasnain, C., Atomically

Sharp Catalyst-Free Wurtzite GaAs∕AlGaAs Nanoneedles Grown on Silicon. Appl. Phys.

Lett. 2008, 93, 023116.

69. Tang, Y. B.; Cong, H. T.; Wang, Z. M.; Cheng, H.-M., Catalyst-Seeded Synthesis and

Field Emission Properties of Flowerlike Si-Doped AlN Nanoneedle Array. Appl. Phys. Lett.

2006, 89, 253112.

70. Lew, K.-K.; Reuther, C.; Carim, A. H.; Redwing, J. M.; Martin, B. R., Template-

Directed Vapor-Liquid-Solid Growth of Silicon Nanowires. J. Vac. Sci. Technol., B 2002, 20,

389-392.

71. Fan, H. J.; Lee, W.; Scholz, R.; Dadgar, A.; Krost, A.; Nielsch, K.; Zacharias, M.,

Arrays of Vertically Aligned and Hexagonally Arranged ZnO Nanowires: A New Template-

Directed Approach. Nanotechnology 2005, 16, 913–917.

55

72. Sahoo, P.; Dhara, S.; Amirthapandian, S.; Kamruddin, M., Evolution of GaN Nanowire

Morphology During Catalyst-Induced Growth Process. J. Mater. Chem. C 2013, 1, 7237-

7245.

73. Xu, H.-Y.; Guo, Y.-n.; Sun, W.; Liao, Z.-m.; Burgess, T.; Lu, H.-f.; Gao, Q.; Tan, H. H.;

Jagadish, C.; Zou, J., Quantitative Study of GaAs Nanowires Catalyzed by Au Film of

Different Thicknesses. Nanoscale Res. Lett. 2012, 7, 1-6.

74. Dubrovskii, V. G.; Sibirev, N. V.; Harmand, J. C.; Glas, F., Growth Kinetics and Crystal

Structure of Semiconductor Nanowires. Phys. Rev. B 2008, 78, 235301.

75. Soci, C.; Bao, X.-Y.; Aplin, D. P. R.; Wang, D., A Systematic Study on the Growth of

GaAs Nanowires by Metal−Organic Chemical Vapor Deposition. Nano Lett. 2008, 8, 4275-

4282.

76. Bauer, J.; Gottschalch, V.; Paetzelt, H.; Wagner, G.; Fuhrmann, B.; Leipner, H. S.,

MOVPE Growth and Real Structure of Vertical-Aligned GaAs Nanowires. J. Cryst. Growth

2007, 298, 625-630.

77. Hallberg, R. T.; Lehmann, S.; Messing, M. E.; Dick, K. A., Palladium Seeded GaAs

Nanowires. J. Mater. Res. 2016, 31, 175-185.

78. Algra, R. E.; Verheijen, M. A.; Borgstrom, M. T.; Feiner, L.-F.; Immink, G.; van

Enckevort, W. J. P.; Vlieg, E.; Bakkers, E. P. A. M., Twinning Superlattices in Indium

Phosphide Nanowires. Nature 2008, 456, 369-372.

79. Dick, K. A.; Deppert, K.; Larsson, M. W.; Martensson, T.; Seifert, W.; Wallenberg, L.

R.; Samuelson, L., Synthesis of Branched 'Nanotrees' by Controlled Seeding of Multiple

Branching Events. Nat. Mater. 2004, 3, 380-384.

80. Tian, B.; Xie, P.; Kempa, T. J.; Bell, D. C.; Lieber, C. M., Single-Crystalline Kinked

Semiconductor Nanowire Superstructures. Nat. Nanotechnol. 2009, 4, 824-829.

81. Borgstrom, M. T.; Immink, G.; Ketelaars, B.; Algra, R.; Bakkers, E. P. A. M.,

Synergetic Nanowire Growth. Nat. Nanotechnol. 2007, 2, 541-544.

82. Joyce, H. J.; Qiang, G.; Wong-Leung, J.; Kim, Y.; Tan, H. H.; Jagadish, C., Tailoring

GaAs, InAs, and InGaAs Nanowires for Optoelectronic Device Applications. IEEE J. Sel.

Top. Quantum Electron. 2011, 17, 766-778.

83. Jensen, L. E.; Björk, M. T.; Jeppesen, S.; Persson, A. I.; Ohlsson, B. J.; Samuelson, L.,

Role of Surface Diffusion in Chemical Beam Epitaxy of InAs Nanowires. Nano Lett. 2004,

4, 1961-1964.

84. Dayeh, S. A.; Yu, E. T.; Wang, D., Iii-V Nanowire Growth Mechanism: V/III Ratio and

Temperature Effects. Nano Letters 2007, 7, 2486-2490.

56

85. Harmand, J. C.; Tchernycheva, M.; Patriarche, G.; Travers, L.; Glas, F.; Cirlin, G.,

Gaas Nanowires Formed by Au-assisted Molecular Beam Epitaxy: Effect of Growth

Temperature. J. Cryst. Growth 2007, 301-302, 853-856.

86. Tchernycheva, M.; Harmand, J. C.; Patriarche, G.; Travers, L.; Cirlin, G. E.,

Temperature Conditions for GaAs Nanowire Formation by Au-assisted Molecular Beam

Epitaxy. Nanotechnology 2006, 17, 4025-4030.

87. Lu, Z. Y.; Chen, P. P.; Liao, Z. M.; Shi, S. X.; Sun, Y.; Li, T. X.; Zhang, Y. H.; Zou, J.;

Lu, W., Impact of Growth Parameters on the Morphology and Microstructure of Epitaxial

GaAs Nanowires Grown by Molecular Beam Epitaxy. J. Alloys Compd. 2013, 580, 82-87.

88. Dayeh, S. A.; Yu, E. T.; Wang, D., III−V Nanowire Growth Mechanism:  V/III Ratio and

Temperature Effects. Nano Lett. 2007, 7, 2486-2490.

89. Chen, X. J.; Gayral, B.; Sam-Giao, D.; Bougerol, C.; Durand, C.; Eymery, J., Catalyst-

Free Growth of High-Optical Quality GaN Nanowires by Metal-Organic Vapor Phase

Epitaxy. Appl. Phys. Lett. 2011, 99, 251910.

90. Chen, C.; Plante, M. C.; Fradin, C.; LaPierre, R. R., Layer-by-Layer and Step-Flow

Growth Mechanisms in GaAsP/GaP Nanowire Heterostructures. J. Mater. Res. 2006, 21,

2801-2809.

91. Dick, K. A.; Deppert, K.; Samuelson, L.; Seifert, W., InAs Nanowires Grown by

MOVPE. J. Cryst. Growth 2007, 298, 631-634.

92. Plante, M. C.; LaPierre, R. R., Au-assisted Growth of GaAs Nanowires by Gas Source

Molecular Beam Epitaxy: Tapering, Sidewall Faceting and Crystal Structure. J. Cryst.

Growth 2008, 310, 356-363.

93. Rueda-Fonseca, P.; Bellet-Amalric, E.; Vigliaturo, R.; den Hertog, M.; Genuist, Y.;

André, R.; Robin, E.; Artioli, A.; Stepanov, P.; Ferrand, D.; Kheng, K.; Tatarenko, S.; Cibert,

J., Structure and Morphology in Diffusion-Driven Growth of Nanowires: The Case of ZnTe.

Nano Lett. 2014, 14, 1877-1883.

94. Joyce, H. J.; Gao, Q.; Hoe Tan, H.; Jagadish, C.; Kim, Y.; Zou, J.; Smith, L. M.;

Jackson, H. E.; Yarrison-Rice, J. M.; Parkinson, P.; Johnston, M. B., III–V Semiconductor

Nanowires for Optoelectronic Device Applications. Prog. Quant. Electron. 2011, 35, 23-75.

95. Yu, X.; Wang, H.; Lu, J.; Zhao, J.; Misuraca, J.; Xiong, P.; von Molnár, S., Evidence

for Structural Phase Transitions Induced by the Triple Phase Line Shift in Self-Catalyzed

GaAs Nanowires. Nano Lett. 2012, 12, 5436-5442.

96. Glas, F.; Ramdani, M. R.; Patriarche, G.; Harmand, J. C., Predictive Modeling of Self-

Catalyzed III-V Nanowire Growth. Phys. Rev. B: Condens. Matter 2013, 88, 195304.

57

97. Colombo, C.; Spirkoska, D.; Frimmer, M.; Abstreiter, G.; Fontcuberta i Morral, A., Ga-

Assisted Catalyst-Free Growth Mechanism of GaAs Nanowires by Molecular Beam

Epitaxy. Phys. Rev. B 2008, 77, 155326.

98. Plante, M. C.; LaPierre, R. R., Control of GaAs Nanowire Morphology and Crystal

Structure. Nanotechnology 2008, 19, 495603.

99. Bubesh Babu, J.; Yoh, K., Growth Rate Enhancement of InAs Nanowire by Molecular

Beam Epitaxy. J. Cryst. Growth 2011, 322, 10-14.

100. Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Kim, Y.; Fickenscher, M. A.; Perera,

S.; Hoang, T. B.; Smith, L. M.; Jackson, H. E.; Yarrison-Rice, J. M.; Zhang, X.; Zou, J.,

Unexpected Benefits of Rapid Growth Rate for III−V Nanowires. Nano Lett. 2008, 9, 695-

701.

101. Zhang, Z. H.; Pickrell, G. W.; Chang, K. L.; Lin, H. C.; Hsieh, K. C.; Cheng, K. Y.,

Surface Morphology Control of Inas Nanostructures Grown on InGaAs/InP. Appl. Phys.

Lett. 2003, 82, 4555-4557.

102. Paiman, S.; Gao, Q.; Tan, H. H.; Jagadish, C.; Pemasiri, K.; Montazeri, M.; Jackson,

H. E.; Smith, L. M.; Yarrison-Rice, J. M.; Zhang, X.; Zou, J., The Effect of V/III Ratio and

Catalyst Particle Size on the Crystal Structure and Optical Properties of InP Nanowires.

Nanotechnology 2009, 20, 225606.

103. Dick, K. A.; Caroff, P.; Bolinsson, J.; Messing, M. E.; Johansson, J.; Deppert, K.;

Wallenberg, L. R.; Samuelson, L., Control of III–V Nanowire Crystal Structure by Growth

Parameter Tuning. Semicond. Sci. Technol. 2010, 25, 024009.

104. Braun, W.; Kaganer, V. M.; Trampert, A.; Schönherr, H.-P.; Gong, Q.; Nötzel, R.;

Däweritz, L.; Ploog, K. H., Diffusion and Incorporation: Shape Evolution During

Overgrowth on Structured Substrates. J. Cryst. Growth 2001, 227–228, 51-55.

105. Dheeraj, D. L.; Patriarche, G.; Zhou, H. L.; Hoang, T. B.; Moses, A. F.; Gronsberg, S.;

van Helvoort, A. T. J.; Fimland, B. O.; Weman, H., Growth and Characterization of

Wurtzite GaAs Nanowires with Defect-Free Zinc Blende GaAsSb Inserts. Nano Lett. 2008,

8, 4459-4463.

106. Soshnikov, I. P.; Cirlin, G. É.; Tonkikh, A. A.; Samsonenko, Y. B.; Dubovskii, V. G.;

Ustinov, V. M.; Gorbenko, O. M.; Litvinov, D.; Gerthsen, D., Atomic Structure of MBE-

Grown GaAs Nanowhiskers. Phys. Solid State 2005, 47, 2213-2218.

107. Breuer, S.; Feiner, L.-F.; Geelhaar, L., Droplet Bulge Effect on the Formation of

Nanowire Side Facets. Cryst. Growth Des. 2013, 13, 2749-2755.

58

108. Glas, F.; Harmand, J. C.; Patriarche, G., Why Does Wurtzite Form in Nanowires of

III-V Zinc Blende Semiconductors? Phys. Rev.Lett. 2007, 99, 146101.

109. Zou, J.; Paladugu, M.; Wang, H.; Auchterlonie, G. J.; Guo, Y.-N.; Kim, Y.; Gao, Q.;

Joyce, H. J.; Tan, H. H.; Jagadish, C., Growth Mechanism of Truncated Triangular III–V

Nanowires. Small 2007, 3, 389-393.

110. Ren, X.; Huang, H.; Dubrovskii, V. G.; Sibirev, N. V.; Nazarenko, M. V.; Bolshakov, A.

D.; X Ye, Q. W.; Huang, Y.; Zhang, X.; Guo, J.; Liu, X., Experimental and Theoretical

Investigations on the Phase Purity of GaAs Zincblende Nanowires. Semicond. Sci.

Technol. 2011, 26, 014034.

111. Messmer, C.; Bilello, J. C., The Surface Energy of Si, GaAs, and GaP. J. Appl. Phys.

1981, 52, 4623-4629.

112. Leitsmann, R.; Bechstedt, F., Surface Influence on Stability and Structure of

Hexagon-Shaped III-V Semiconductor Nanorods. J. Appl. Phys. 2007, 102, 063528.

113. M.C. Plante, R. R. L., Growth Mechanisms of GaAs Nanowires by Gas Source

Molecular Beam Epitaxy. J. Cryst. Growth 2006, 286, 394-399.

114. Johansson, J.; Dick, K. A.; Caroff, P.; Messing, M. E.; Bolinsson, J.; Deppert, K.;

Samuelson, L., Diameter Dependence of the Wurtzite−Zinc Blende Transition in InAs

Nanowires. J. Phys. Chem. C 2010, 114, 3837-3842.

115. Shtrikman, H.; Popovitz-Biro, R.; Kretinin, A.; Heiblum, M., Stacking-Faults-Free Zinc

Blende GaAs Nanowires. Nano Lett. 2008, 9, 215-219.

116. Shtrikman, H.; Popovitz-Biro, .; Kretinin, A.; Houben, L.; Heiblum, M.; Bukała, M.;

Galicka, M.; Buczko, R.; Kacman, P., Method for Suppression of Stacking Faults in

Wurtzite III−V Nanowires. Nano Lett. 2009, 9, 1506-1510.

117. Hoang, T. B.; Zhou, H. L.; Moses, A. F.; Dheeraj, D. L.; Van Helvoort, A. T. J.;

Fimland, B. O.; Weman, H., Optical Properties of Single Wurtzite GaAs Nanowires and

GaAs Nanowires with GaAsSb Inserts, Mater. Res. Soc. Symp. Proc. 2009, 61, 7-12.

118. Zhang, Z.; Lu, Z.-Y.; Chen, P.-P.; Xu, H.-Y.; Guo, Y.-N.; Liao, Z.-M.; Shi, S.-X.; Lu,

W.; Zou, J., Quality of Epitaxial InAs Nanowires Controlled by Catalyst Size in Molecular

Beam Epitaxy. Appl. Phys. Lett. 2013, 103, 073109.

119. Chen, C., Plante, M. C., Fradin, C., LaPierre, R. R., Layer-by-Layer and Step-Flow

Growth Mechanisms in GaAsP/GaP Nanowire Heterostructures. J. Mater. Res. 2006, 21,

2801-2809.

120. Li, F.; Nellist, P. D.; Lang, C.; Cockayne, D. J. H., Dependence of Surface Facet

Period on the Diameter of Nanowires. ACS Nano 2010, 4, 632-636.

59

121. Ross, F. M.; Tersoff, J.; Reuter, M. C., Sawtooth Faceting in Silicon Nanowires. Phys.

Rev. Lett. 2005, 95, 146104.

122. Mattila, M.; Hakkarainen, T.; Jiang, H.; Kauppinen, E. I.; Lipsanen, H., Effect of

Substrate Orientation on the Catalyst-Free Growth of InP Nanowires. Nanotechnology

2007, 18, 155301.

123. Wu, Z. H.; Mei, X.; Kim, D.; Blumin, M.; Ruda, H. E.; Liu, J. Q.; Kavanagh, K. L.,

Growth, Branching, and Kinking of Molecular-Beam Epitaxial〈110〉GaAs Nanowires.

Appl. Phys. Lett. 2003, 83, 3368-3370.

124. Krishnamachari, U.; Borgstrom, M.; Ohlsson, B. J.; Panev, N.; Samuelson, L.; Seifert,

W.; Larsson, M. W.; Wallenberg, L. R., Defect-Free InP Nanowires Grown in [001]

Direction on InP (001). Appl. Phys. Lett. 2004, 85, 2077-2079.

125. Zhang, Z.; Lu, Z.; Xu, H.; Chen, P.; Lu, W.; Zou, J., Structure and Quality Controlled

Growth of InAs Nanowires through Catalyst Engineering. Nano Res. 2014, 7, 1640-1649.

126. Joyce, H. J.; Wong-Leung, J.; Gao, Q.; Tan, H. H.; Jagadish, C., Phase Perfection in

Zinc Blende and Wurtzite III−V Nanowires Using Basic Growth Parameters. Nano Lett.

2010, 10, 908-915.

127. Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Kim, Y.; Zhang, X.; Guo, Y.; Zou, J.,

Twin-Free Uniform Epitaxial GaAs Nanowires Grown by a Two-Temperature Process.

Nano Lett. 2007, 7, 921-926.

128. Mattila, M.; Hakkarainen, T.; Mulot, M.; Lipsanen, H., Crystal-Structure-Dependent

Photoluminescence from InP Nanowires. Nanotechnology 2006, 17, 1580–1583.

129. Paiman, S.; Gao, Q.; Joyce, H. J.; Kim, Y.; Tan, H. H.; Jagadish, C.; Zhang, X.; Guo,

Y.; Zou, J., Growth Temperature and V/III Ratio Effects on the Morphology and Crystal

Structure of InP Nanowires. Nanotechnology 2010, 43, 445402.

130. Ihn, S. G.; Song, J. I.; Kim, Y. H.; Lee, J. Y.; Ahn, I. H., Growth of GaAs Nanowires

on Si Substrates Using a Molecular Beam Epitaxy. IEEE Trans. Nanotechnol. 2007, 6,

384-389.

131. Spirkoska, D.; Arbiol, J.; Gustafsson, A.; Conesa-Boj, S.; Glas, F.; Zardo, I.; Heigoldt,

M.; Gass, M. H.; Bleloch, A. L.; Estrade, S.; Kaniber, M.; Rossler, J.; Peiro, F.; Morante, J.

R.; Abstreiter, G.; Samuelson, L.; Fontcuberta I Morral, A., Structural and Optical

Properties of High Quality Zinc-Blende/Wurtzite GaAs Nanowire Heterostructures. Phys.

Rev. B 2009, 80, 245325.

132. Dubrovskii, V. G.; Sibirev, N. V.; Cirlin, G. E.; Bouravleuv, A. D.; Samsonenko, Y. B.;

Dheeraj, D. L.; Zhou, H. L.; Sartel, C.; Harmand, J. C.; Patriarche, G.; Glas, F., Role of

60

Nonlinear Effects in Nanowire Growth and Crystal Phase. Phys. Rev. B: Condens. Matter

2009, 80, 205305.

133. Dheeraj, D. L.; Munshi, A. M.; Scheffler, M.; Helvoort, A. T. J. v.; Weman, H.;

Fimland, B. O., Controlling Crystal Phases in GaAs Nanowires Grown by Au-Assisted

Molecular Beam Epitaxy. Nanotechnology 2013, 24, 015601.

134. Dick, K. A.; Thelander, C.; Samuelson, L.; Caroff, P., Crystal Phase Engineering in

Single InAs Nanowires. Nano Lett. 2010, 10, 3494-3499.

135. Li, X.; Guo, H.; Yin, Z.; Shi, T.; Wen, L.; Zhao, Z.; Liu, M.; Ma, W.; Wang, Y.,

Morphology and Crystal Structure Control of GaAs Nanowires Grown by Au-assisted MBE

with Solid As4 Source. J. Cryst. Growth 2011, 324, 82-87.

136. Jiang, N.; Wong-Leung, J.; Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.,

Understanding the True Shape of Au-Catalyzed GaAs Nanowires. Nano Lett. 2014, 14,

5865-5872.

137. Hou, J. J.; Han, N.; Wang, F.; Xiu, F.; Yip, S.; Hui, A. T.; Hung, T.; Ho, J. C.,

Synthesis and Characterizations of Ternary InGaAs Nanowires by a Two-Step Growth

Method for High-Performance Electronic Devices. ACS Nano 2012, 6, 3624-3630.

138. Leon, R.; Kim, Y.; Jagadish, C.; Gal, M.; Zou, J.; Cockayne, D. J. H., Effects of

Interdiffusion on the Luminescence of InGaAs/GaAs Quantum Dots. Appl. Phys. Lett. 1996,

69, 1888-1890.

139. Chen, C.; Shehata, S.; Fradin, C.; LaPierre, R.; Couteau, C.; Weihs, G., Self-

Directed Growth of AlGaAs Core-Shell Nanowires for Visible Light Applications. Nano Lett.

2007, 7, 2584-2589.

140. Sun, B. Q.; Gal, M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Puzzer, T.; Ouyang, L.; Zou,

J., Epitaxially Grown GaAsN Random Laser. J. Appl. Phys. 2003, 93, 5855-5858.

141. Zhang, J.; Itzler, M. A.; Zbinden, H.; Pan, J.-W., Advances in InGaAs/InP Single-

Photon Detector Systems for Quantum Communication. Light Sci. Appl 2015, 4, e286.

142. Caroff, P.; Bolinsson, J.; Johansson, J., Crystal Phases in III--V Nanowires: From

Random toward Engineered Polytypism. IEEE J. Sel. Top. Quantum Electron. 2011, 17,

829-846.

143. Mohseni, P. K.; Behnam, A.; Wood, J. D.; English, C. D.; Lyding, J. W.; Pop, E.; Li,

X., InxGa1–XAs Nanowire Growth on Graphene: Van Der Waals Epitaxy Induced Phase

Segregation. Nano Lett. 2013, 13, 1153-1161.

61

144. Heiss, M.; Ketterer, B.; Uccelli, E.; Morante, J. R.; Arbiol, J.; Morral, A. F. I., In(Ga)As

Quantum Dot Formation on Group-III Assisted Catalyst-Free InGaAs Nanowires.

Nanotechnology 2011, 22, 195601.

145. Jabeen, F.; Rubini, S.; Grillo, V.; Felisari, L.; Martelli, F., Room Temperature

Luminescent InGaAs/GaAs Core-Shell Nanowires. Appl. Phys. Lett. 2008, 93, 083117.

146. Ng, K. W.; Tran, T.-T. D.; Ko, W. S.; Chen, R.; Lu, F.; Chang-Hasnain, C. J., Single

Crystalline InGaAs Nanopillar Grown on Polysilicon with Dimensions Beyond the Substrate

Grain Size Limit. Nano Lett. 2013, 13, 5931-5937.

147. Ng, K. W.; Ko, W. S.; Chen, R.; Lu, F.; Tran, T.-T. D.; Li, K.; Chang-Hasnain, C. J.,

Composition Homogeneity in InGaAs/GaAs Core–Shell Nanopillars Monolithically Grown

on Silicon. ACS Appl. Mater. Interfaces 2014, 6, 16706-16711.

148. Dimakis, E.; Jahn, U.; Ramsteiner, M.; Tahraoui, A.; Grandal, J.; Kong, X.; Marquardt,

O.; Trampert, A.; Riechert, H.; Geelhaar, L., Coaxial Multishell (In,Ga)As/GaAs Nanowires

for Near-Infrared Emission on Si Substrates. Nano Lett. 2014, 14, 2604-2609.

149. Chen, R.; Tran, T.-T. D.; Ng, K. W.; Ko, W. S.; Chuang, L. C.; Sedgwick, F. G.;

Chang-Hasnain, C., Nanolasers Grown on Silicon. Nat. Photon. 2011, 5, 170-175.

150. Treu, J.; Stettner, T.; Watzinger, M.; Morkötter, S.; Döblinger, M.; Matich, S.; Saller,

K.; Bichler, M.; Abstreiter, G.; Finley, J. J.; Stangl, J.; Koblmüller, G., Lattice-Matched

InGaAs–InAlAs Core–Shell Nanowires with Improved Luminescence and Photoresponse

Properties. Nano Lett. 2015, 15, 3533-3540.

151. Rudolph, D.; Funk, S.; Döblinger, M.; Morkötter, S.; Hertenberger, S.; Schweickert, L.;

Becker, J.; Matich, S.; Bichler, M.; Spirkoska, D.; Zardo, I.; Finley, J. J.; Abstreiter, G.;

Koblmüller, G., Spontaneous Alloy Composition Ordering in GaAs-AlGaAs Core–Shell

Nanowires. Nano Lett. 2013, 13, 1522-1527.

152. Wu, J.; Borg, B. M.; Jacobsson, D.; Dick, K. A.; Wernersson, L.-E., Control of

Composition and Morphology in InGaAs Nanowires Grown by Metalorganic Vapor Phase

Epitaxy. J. Cryst. Growth 2013, 383, 158-165.

153. Ng, K. W.; Ko, W. S.; Lu, F.; Chang-Hasnain, C. J., Metastable Growth of Pure

Wurtzite InGaAs Microstructures. Nano Lett. 2014, 14, 4757-4762.

154. Ameruddin, A. S.; Caroff, P.; Tan, H. H.; Jagadish, C.; Dubrovskii, V. G.,

Understanding the Growth and Composition Evolution of Gold-Seeded Ternary InGaAs

Nanowires. Nanoscale 2015, 7, 16266-16272.

62

155. Sun, W.; Guo, Y.; Xu, H.; Gao, Q.; Hoe Tan, H.; Jagadish, C.; Zou, J., Polarity Driven

Simultaneous Growth of Free-Standing and Lateral Gaasp Epitaxial Nanowires on Gaas

(001) Substrate. Appl. Phys. Lett. 2013, 103, 223104.

156. Sun, W.; Huang, Y.; Guo, Y.; Liao, Z. M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Liao, X.

Z.; Zou, J., Spontaneous Formation of Core-Shell GaAsP Nanowires and Their Enhanced

Electrical Conductivity. J. Mater. Chem. C 2015, 3, 1745-1750.

157. Zhang, Y.; Wu, J.; Aagesen, M.; Holm, J.; Hatch, S.; Tang, M.; Huo, S.; Liu, H., Self-

Catalyzed Ternary Core–Shell GaAsP Nanowire Arrays Grown on Patterned Si Substrates

by Molecular Beam Epitaxy. Nano Lett. 2014, 14, 4542-4547.

158. Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting As Flux in Molecular Beam Epitaxy. J.

Phys. Chem. C 2015, 119, 20721-20727.

159. Oliveira, D. S.; Tizei, L. H. G.; Li, A.; Vasconcelos, T. L.; Senna, C. A.; Archanjo, B.

S.; Ugarte, D.; Cotta, M. A., Interaction between Lamellar Twinning and Catalyst Dynamics

in Spontaneous Core-Shell InGaP Nanowires. Nanoscale 2015, 7, 12722-12727.

160. Fakhr, A.; Haddara, Y. M.; LaPierre, R. R., Dependence of InGaP Nanowire

Morphology and Structure on Molecular Beam Epitaxy Growth Conditions.

Nanotechnology 2010, 21, 165601.

161. Sato, T.; Motohisa, J.; Noborisaka, J.; Hara, S.; Fukui, T., Growth of InGaAs

Nanowires by Selective-Area Metalorganic Vapor Phase Epitaxy. J. Cryst. Growth 2008,

310, 2359-2364.

162. Persson, A. I.; Björk, M. T.; Jeppesen, S.; Wagner, J. B.; Wallenberg, L. R.;

Samuelson, L., InAs1-XPx Nanowires for Device Engineering. Nano Lett. 2006, 6, 403-407.

163. Tchernycheva, M.; Cirlin, G. E.; Patriarche, G.; Travers, L.; Zwiller, V.; Perinetti, U.;

Harmand, J.-C., Growth and Characterization of InP Nanowires with InAsP Insertions.

Nano Lett. 2007, 7, 1500-1504.

164. Svensson, C. P. T.; Seifert, W.; Larsson, M. W.; Wallenberg, L. R.; Stangl, J.; Bauer,

G.; Samuelson, L., Epitaxially Grown GaP/GaAs1−XPX/GaP Double Heterostructure

Nanowires for Optical Applications. Nanotechnology 2005, 16, 936.

165. Krogstrup, P.; Yamasaki, J.; Sørensen, C. B.; Johnson, E.; Wagner, J. B.;

Pennington, R.; Aagesen, M.; Tanaka, N.; Nygård, J., Junctions in Axial III−V

Heterostructure Nanowires Obtained via an Interchange of Group III Elements. Nano Lett.

2009, 9, 3689-3693.

63

166. Dick, K. A.; Bolinsson, J.; Borg, B. M.; Johansson, J., Controlling the Abruptness of

Axial Heterojunctions in Iii–V Nanowires: Beyond the Reservoir Effect. Nano Lett. 2012, 12,

3200-3206.

167. Messing, M. E.; Wong-Leung, J.; Zanolli, Z.; Joyce, H. J.; Tan, H. H.; Gao, Q.;

Wallenberg, L. R.; Johansson, J.; Jagadish, C., Growth of Straight InAs-on-GaAs

Nanowire Heterostructures. Nano Lett. 2011, 11, 3899-3905.

168. Scarpellini, D.; Somaschini, C.; Fedorov, A.; Bietti, S.; Frigeri, C.; Grillo, V.; Esposito,

L.; Salvalaglio, M.; Marzegalli, A.; Montalenti, F.; Bonera, E.; Medaglia, P. G.; Sanguinetti,

S., InAs/GaAs Sharply Defined Axial Heterostructures in Self-assisted Nanowires. Nano

Lett. 2015, 15, 3677-3683.

169. Ameruddin, A. S.; Fonseka, H. A.; Caroff, P.; Wong-Leung, J.; Veld, R. L. O. h.;

Boland, J. L.; Johnston, M. B.; Tan, H. H.; Jagadish, C., InXGa1−XAs Nanowires with

Uniform Composition, Pure Wurtzite Crystal Phase and Taper-Free Morphology.

Nanotechnology 2015, 26, 205604.

170. Guo, Y.-N.; Burgess, T.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou, J., Polarity-Driven

Nonuniform Composition in InGaAs Nanowires. Nano Lett. 2013, 13, 5085-5089.

171. Bauer, J.; Gottschalch, V.; Wagner, G., The Influence of the Droplet Composition on

the Vapor-Liquid-Solid Growth of InAs Nanowires on GaAs ( )B by Metal-Organic Vapor

Phase Epitaxy. J. Appl. Phys. 2008, 104, 114315.

172. Liao, Z.-M.; Chen, Z.-G.; Lu, Z.-Y.; Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Zhang, Z.; Yang,

L.; Chen, P.-P.; Lu, W.; Zou, J., Au Impact on GaAs Epitaxial Growth on GaAs (111)B

Substrates in Molecular Beam Epitaxy. Appl. Phys. Lett. 2013, 102, 063106.

173. Nazarenko, M. V.; Sibirev, N. V.; Ng, K. W.; Ren, F.; Ko, W. S.; Dubrovskii, V. G.;

Chang-Hasnain, C., Elastic Energy Relaxation and Critical Thickness for Plastic

Deformation in the Core-Shell InGaAs/GaAs Nanopillars. J. Appl. Phys. 2013, 113, 104311.

174. Cunningham, J. E.; Santos, M. B.; Goossen, K. W.; Williams, M. D.; Jan, W.,

Dimerization Induced Incorporation Nonlinearities in GaAsP. Appl. Phys. Lett. 1994, 64,

2418-2420.

175. Sun, W.; Guo, Y.-N.; Xu, H.-Y.; Liao, Z.-M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou,

J., Unequal P Distribution in Nanowires and the Planar Layer During GaAsP Growth on

GaAs {111}B by Metal–Organic Chemical Vapor Deposition. J. Phys. Chem. C 2013, 117,

19234-19238.

64

176. Zou, J.; Chou, C. T.; Cockayne, D. J. H.; Sikorski, A.; Vaughan, M. R., Misfit

Dislocations Lying Along <100> in [001] GaAs/In0.25Ga0.75As/GaAs Quantum Well

Heterostructures. Appl. Phys. Lett. 1994, 65, 1647-1649.

177. Zou, J.; Cockayne, D. J. H.; Usher, B. F., Misfit Dislocations and Critical Thickness in

InGaAs/GaAs Heterostructure Systems. J. Appl. Phys. 1993, 73, 619-626.

178. Woicik, J. C.; Miyano, K. E.; Pellegrino, J. G.; Shaw, P. S.; Southworth, S. H.; Karlin,

B. A., Strain and Relaxation in InAs and InGaAs Films Grown on GaAs(001). Appl. Phys.

Lett. 1996, 68, 3010-3012.

65

Chapter 3

Methodology

3.1 Introduction

In this thesis, III-V semiconductor nanowires are first grown by molecular beam epitaxy

and then investigated by advanced electron microscopy, which is the core to this research.

Specifically, the morphology of the grown nanowires is studied by SEM and their structural

and chemical characteristics are investigated by TEM and EDS, respectively. This chapter

66

will give detailed introductions on these important experimental techniques and their

working principles, as well as the related sample preparation methods.

3.2 Epitaxial growth

3.2.1 Molecular beam epitaxy

Molecular beam epitaxy is a physical deposition method for growing single crystals,

especially thin epitaxial film such as oxides, mentals and semiconductors. It was invented

in the late 1960s and first applied to the growth of compound semiconductors. It has been

employed in manufacturing environment since 1984 to fabricate GaAs laser diodes1 and to

make monolithic microwave integrated circuits and high-power hetero-junction bipolar

transistors since 1986.2 Until now, MBE is still widely used in the manufacture of

semiconductor devices, such as lasers and transistors, which can be further used in such

applications as cellular phones, radar system, solar cell and display devices.3-5

MBE essentially involves highly controlled evaporation in an ultrahigh vacuum (~10-8

Torr) system, which enables compound semiconductor materials with great precision and

high purity. For example, in the process of depositing thin film by solid source MBE,

elements such as Ga and As are heated in separated effusion cells until they sublime (as

illustrated in Figure 3.1).6 The reaction of one or more evaporated beams of atoms and

molecules with the single crystal substrate yields the desired epitaxial film. The term

“beam” means that evaporated source do not interact with each other until they arrive at

the substrate. It has also been demonstrated that by applying MBE, it is of great possibility

to grow “superlattice” structures, for example, the alternate thin layers of GaAs and

AlGaAs with abrupt interfaces.7 It has been developed as a popular method to grow

semiconductor nanowires with precise control of their composition during growth.

RHEED Nowadays, MBE systems have been equipped with an electron gun and

fluorescent screen for the display of reflection high energy electron diffraction (RHEED)

patterns, which is often used for monitoring the structure and composition of epitaxial film

during the crystal growth. In RHEED system, an electron gun is set up to generate the

electron beam that strikes the sample surface with a small angle (note that the electron

beam is directly incident on the substrate without any physically interference with the

67

deposition source). The incident electrons interact with the sample and a part of them

diffracted and interfere at specific angles to form diffraction patterns.

During the thin film epitaxial growth, normally both spotted and streaked patterns can

be observed: spots often appear as a consequence of three-dimensional volume

diffraction like islands (Figure 3.2a), while streaked patterns occur when smooth layer

surface formed (Figure 3.2c).8 In nanowires growth, this technique is often used to monitor

the quality of the buffer layer that deposit directly on the substrate and the crystallization of

the grown nanowires.

Figure 3.1 Schematic illustration of a top-view of a typical growth chamber in the MBE

system, which is equipped with the RHEED monitor.6

68

Figure 3.2 RHEED patterns (<110> azimuth) and SEM images of (100)-oriented GaP

substrates at (a) 350°C , 550°C and 600°C, respectively.8

Apart from that, several parameters in MBE can also be independently adjusted and

monitored to enhance the crystal quality, such as:

Flux rate: flux is determined as the flow rate of source per unit area. In MBE, the

flux rate determines the amount of sources reaching the substrate and is

measured as the beam equivalent pressure (Torr). The flux rate of sources such

as Ga and As can be tuned by the their respective source temperature

depending on the heating methods. The respective flux of group-III and group-V

determine the V/III ratio, which is a key factor for nanowire growth. Normally, for

III-V nanowires growth, group-V (normally As) is excessive to prevent deposited

material evaporation and to ensure a stable crystal growth.

Substrate temperature: since the crystal growth directly deposited on the

substrate surface, the substrate temperature approximately equals to the growth

temperature, which determines the thermodynamic conditions for crystal growth

and affect the diffusivity of the impinging atoms.

69

Table 3.1 Comparison of MBE and MOCVD

MBE MOCVD

Vaccum conditions high (10-5 10-8 Torr) low

Growth species elements source metalorganic precursors

Growth temperature low high

Growth rate slow: layer by layer growth quick

In situ monitoring yes: RHEED no

Growth control precise (at atomic level) rough

Growth state physical process, non-equilibrium

not thermodynamically favorable

Chemical process,

thermodynamically favorable

III-V nanowires

product structure mainly wurtzite wurtzite and zinc blende

3.2.2 Growth procedures

All nanowires investigated in this thesis were grown by solid-source Riber 32 MBE system

with a RHEED monitor. For instance, the process of growing GaAs nanowires is described

as below in details.

In the pre-growth stage, the growth chamber in MBE system was cooled down by

liquid nitrogen until the vacuum condition reaches as high as 10-5 Torr. Then, the GaAs

substrate was transferred to the buffer chamber, where the degassing and deoxidization

took place under As flux for about 10 minutes. During this essential process, surface

contaminants can be effectively desorbed. After that, a GaAs buffer layer was deposited

on the substrate to achieved atomic flat surface, which would be secured by the RHEED

monitoring. Then, a thin film of Au was deposited on the smooth GaAs buffer layer.

Next, the substrate with Au film was sent back to the growth chamber, in which

annealing took place at 550°C, above the eutectic point, to form the Au nanoparticles

acting as catalysts. The function of the Au catalysts is not only by minimizing the activation

energy for nanowire growth but also can attract source materials especially the group-III to

realize crystallization at the nanowire-catalyst interface.

After annealing for 5-10 min, the substrate temperature dropped down to the

preferential temperature for nanowire growth (normally higher than the eutectic

70

temperature). Meanwhile, the group-III flux (Ga) was introduced to the growth chamber to

initiate the GaAs nanowires growth. With varied growth recipes, the growth parameters,

such as growth temperature, V/III ratios and growth duration, can be controlled accordingly.

When the GaAs nanowires growth was terminated, the valves of effusion cells were closed

and the grown nanowires sample was cooling down under the As protection until the

temperature decreased to 300°C. After the substrate temperature was dropped close to

the room temperature, the grown nanowire sample was removed from the growth chamber.

3.3 Scanning electron microscopy

An electron microscope uses a beam of accelerated electrons (rather than photons) to

illuminate the sample and generates a highly magnified image. With the wavelength

100,000 times shorter than that of the visible light photons, the electron microscope can

have a greater resolution power than an optical microscope, being able to reveal more

information about smaller objects in details. For example, electron microscopies, such as

SEM and TEM, have become essential and irreplaceable tools to characterize the

morphology, structure and composition of the crystal sample, in particular the

nanostructures.

In electron microscopy, the electron beam interacting with the sample can generate a

number of signals as described in Figure 3.3.9 If the sample is sufficiently thin, the electron

beam can go through it thoroughly without any hindrance and interaction, acting as a

transmitted beam. Most of these singles can be detected in different type of TEM modes,

such as imaging, diffraction and spectroscopy mode. It is important to note that not all the

incident electron beams keep their original directions. Instead, some of them are scattered,

either elastically or inelastically. When the incident electrons are elastically scattered, they

do not lose any energy but only changed their incident orientation. On the contrary, when

the coming electrons are inelastically scattered, they have energy loss which can be

transferred to the atoms on the sample.

71

Figure 3.3 Signals generated when a high-energy beam of electrons interacts with a thin

sample.9

As one of the three types of electron microscopies, SEM uses a focused beam of

electrons to scan the sample surface and create signals for imaging.10 SEMs can magnify

an object from 10 times up to 300,000 times with high resolution better than 1 nm, which is

inaccessible by light microscopy. Sample can be observed in high vacuum, low vacuum or

wet conditions. SEM can detect singles generated from the sample surface, including

secondary electrons, backscattered electrons and characteristic X-rays. As can be seen

from Figure 3.4, different singles come from different depths of the sample. With higher

accelerating voltage, the electron beam penetration is greater and the interaction volume

increases. Electron scattered from the sample surface is actually made of atoms, which

has two major parts resulting in two main singles: secondary electrons (SE) and

backscattered electrons (BSE). Primary beam “knocks out” out-shells electrons from

sample atoms forming low energy secondary electrons. With higher energy, electron beam

accelerates towards the positive nucleus and leaves as backscattered electrons. In terms

of X-ray production, continuum and characteristic X-ray are included. Continuum X-ray is

produced by the inelastic scattering process of the primary electron beam and the sample

atoms. Most of them are low energy X-rays, which are heavily absorbed at the low energy

end of the spectrum. On the other hand, characteristic X-rays are emitted when outer shell

electrons fill in the inner shell vacancy and release the excess energy. The inner shell

ionisation caused by the incident electron beam leaves some empty states. To keep the

system stable, electrons from out shells relax to fill the empty state in the inner shell and

72

release the extra energy in the form of x-ray photons at the same time. Each electron jump

has a unique or characteristic energy that represents the patent atoms. Since different

elements have different electronic structure, these characteristic x-rays can be used to

determine the element composition of the sample.

Figure 3.4 schematically illustrated the working principles of the SEM. After electrons

being accelerated, it is first confined by the condenser lens to travel in a desired direction

and get converged at the same time. Before the beam flares out again, it is converged

back again by the objective lens and down onto the sample. Condenser and objective lens

are both electromagnetic, making electron beam travel spiralling down the column.

Besides, they both have an aperture sitting well below them. The condenser aperture

controls the probe size by demagnifying the electron beam from the gun. While the

objective aperture focuses electron beam into a smaller probe to control the focus of the

final image. By applying magnetic deflection coils, the electron beam is scanned the

Figure 3.4 Schematic illustrations of SEM layout and function. (http://www.ammrf.org.

au/myscope/sem/practice/principles/layout.php)

73

sample surface and generated the electric signals that can be collected by the electron

detectors. These signals are transferred and magnified and images are built up by these

scanned signals reflected on the screen. It should be noted that the specimen can be tilt

by a small angle to observe more detailed information of the sample and obtain much

more to lower the signal/noise ratio and thus secure higher-resolution SEM images.

Unlike in TEM, the electron accelerating voltage in SEM is much lower, ranging from

1 to 50 kV, and the sample is thick enough so that electron beam cannot penetrate

through the sample. The main electron signals generated by the incident electron beam

are secondary electrons (SE) and backscattered electrons (BSE).

SE imaging Secondary electrons are produced by inelastic scattering of the incident

electrons from the sample atom electron clouds, which usually have low energy (0-50 eV).

Most secondary electrons are reabsorbed by the sample. Figure 3.5 shows the schematic

illustration of the interaction between incident electron beam with the sample. As noted,

different signals generated from the different depths from the sample surface, bring in

different information about the sample. The interaction volume in the sample depends on

the energy of the incident electrons (accelerating voltage applied). It is noted that most of

secondary electron signals emit near the sample surface (at a depth from 5 to 50 nm),

bring in surface information or tomography of the sample. The signal intensity mainly

depends on the sample surface topography. As shown in Figure 3.6, the variation of

electron signals brings in contrast in SEM images due to the “edge effect”. Edges are

normally brighter in the SE SEM image because of a larger surface area for electrons to

escape from, which generate more signals that can be collected by the detector. By

utilizing this edge effect, the topography of the sample can be well investigated.

Normally, high accelerating voltage, small spot size and short working distance can

endow SEM images with high resolution. With relatively higher accelerating voltage, more

electrons can be generated out, leading to the enhanced signal/noise ratio and thus the

higher resolution of the image. Besides, by narrowing the spot size, those scattered

electrons can be excluded, in which the noise is decreased and the fine electron beam can

74

Figure 3.5 Schematic illustration of interaction volume created by electron beam on the

sample surface with electron singles generated from different depths.

Figure 3.6 Schematic illustration of the “edge effect” on the SE yields. (©

http://laser.phys.ualberta.ca/~egerton/SEM/sem.htm)

carry on more fine sample details. On the other hand, depth of field also plays an import

role in SEM imaging. Although for SEM image with a small depth of field, the main features

of the samples can be revealed, the whole quality of the SEM images is poor, since the

features beyond the depth of field can be blur. Therefore, extend the depth of the field in

75

the SEM image is essential to reveal the comprehensive information contained in the

sample. The larger can be simply achieved by increasing the working the distance, which

would sacrifice part of the resolution. Therefore, the high resolution SEM image can be

obtained under the balance of the accelerating voltage, spot size and working distance,

which also depends on sensitivity and conductivity of sample itself.

BSE imaging Backscattered electrons (BSEs) are the electrons scattering elastically

from sample nucleus. With large interaction volume from higher energy beam, the BSE

signals can provide chemical composition information that cannot be achieved by SE. The

intensity of signal generated depends on the atomic number and results in contrast in the

BSE images. The higher the average atomic number, more primary electrons are

scattered back out of the sample and generate stronger signals. The intensity of the signal

brings in sample information, particularly the element weight. Therefore, the contrast

shown in the BSE image can tell the rough distribution of heavy or light elements.

Figure 3.7 (a) BSE and (b) SE images of a grain of sand that contains Si and Ti.

(©http://www.ammrf.org.au/myscope/sem/practice/principles/imagegeneration.php)

The difference of SE and BSE images can be determined when both imaging modes

are applied on the same sample. Figure 3.7a shows a SE image of grains of

a mixture mineral sand, which indicate the grains morphology and size, and the smooth

surface of the grains. While, Figure 3.7b is the corresponding BSE image of the sand, in

which strong contrast can be witnessed: the brighter grain suggests the Ti grain while the

dark contrast indicates the Si grain.

76

X-ray microanalysis In SEM, X-rays are generated though the whole interaction

volume (Figure 3.4). Therefore, the typical spatial resolution for X-ray microanalysis in

SEM is up to a few microns. The electron accelerating voltage in SEM is typical 5-30 keV,

way lower than the conventional TEM, which can largely affect the X-ray intensities (X-ray

peaks). Since the sample composition is determined based on the ratio of the X-ray peaks,

the sample must be stable, flat and void free to ensure the accuracy of the X-ray detection.

The low X-ray signals in SEM also brings in several limitations: light elements (Z < 11)

cannot be analysed and the composition of small grains with order of nanometres cannot

be guaranteed.

In this thesis, the morphology of nanowires, such as nanowires length, diameter,

growth direction and side facets, are investigated by the field emission SEM, mainly JEOL

JSM-7001F and JEOL JSM-7800F SEM, which are equipped with field emission gun that

can generate electron beam with narrower diameter and higher current density than

conventional thermionic emitters and they are normally set up with SE, BSE or X-ray

detectors.

3.4 Transmission electron microscopy

As mentioned previously, TEM is one of the most powerful and comprehensive techniques

for characterizations of nanoscale materials and devices. Apart from the TEM imaging

mode, a TEM can appear in other forms and functions, such as HRTEM, STEM, EDS and

electron energy loss spectroscopy.

In conventional TEM, an ultrahigh energy electron beam transmitting through an

ultra-thin sample, normally less than 100 nanometres, creates images in several different

forms. Figure 3.8 illustrates ray diagram of a typical conventional TEM. From the top down,

by connecting the electron gun to a high voltage between 100-300 kV, the gun begins to

emit electrons into the vacuum column. In the column, the electromagnetic lenses help to

shape the electron beam. Specifically, the emitted electron beam is paralleled by the

condenser lens before it penetrates the thin sample and further scattered. The transmitted

and scattered electrons pass through the objective lens and form two planes: the image

plane and the back focal plane, resulting in two types of modes: imaging mode and

diffraction mode. At the bottom of the column, it is a viewing screen with phosphor

coatings to visualize both TEM images and diffraction patterns.11

77

3.4.1 Imaging mode

In the imaging mode, the number of transmitted electrons depends on the mass of atoms

in the sample or the thickness of the sample. With the various intensity of electron beam

reaching the screen, the contrast can be seen in the TEM image. After the electron beam

being focused by the objective lens, the intermediate lens projects the image of sample at

the image plane, which can be seen on the fluorescent viewing screen and recorded by a

charge-coupled device (CCD) camera (Figure 3.8a). By choosing the size of objective

aperture, the contrast of the image can be adjusted.

Diffraction-contrast imaging The contrast witnessed in TEM images is due to the

scattering of electron beam by the thin sample. The electron wave interacted with sample

can change its amplitude and phase, leading to the amplitude contrast and phase contrast

Figure 3.8 Ray diagram of a TEM in (a) imaging mode, creating bright field TEM images

on the viewing screen and (b) diffraction mode, the diffraction pattern from the sample

captured on the viewing screen.11

78

in TEM images, respectively. Diffraction contrast is a special form of amplitude contrast, in

which electron scattering at Bragg angles. Since the crystal defects rotate the defecting

planes near the defects and the intensity of the diffracted beam depends on the deviation

factor, the diffraction contrast imaging is used to determine the crystal defects in the

sample. As noted in the inset of Figure 3.9a, if the aperture selects the centre spot, bright-

field (BF) TEM image is formed, while if it select the diffracted spot (Figure 3.9b inset), a

dark-field (DF) image is generated. The dark contrast in BF image and the bright contrast

in DF image indicate the bend contour areas. This is due to the fact that when the atomic

Figure 3.9 (a) BF and (b) DF TEM image of a polycrystalline thin film of Bi. The insets are

the diffraction pattern taken from the one of single crystalline Bi. (c) Schematic illustration

of atomic planes bending at the edge dislocations. At position P, the angle equals to the

Bragg angle of the incidence electron beam. (d) A BF TEM image of cobalt mental.

planes bend and then cause the electron beam incidence angle approximate to the local

Bragg angle, electron would largely diffracted (see Figure 3.9c) and these diffracted

electrons will be absorbed at the TEM objective aperture, leading to the dark in the TEM

79

image. By applying this characteristic, the dislocations and stacking faults in the sample

can be determined. Figure 3.9d is a BF TEM image of the cobalt mental, in which the

dislocations are witnessed as dark curves while the stacking faults are seen as dark

straight stacking lines.12

Phase-contrast imaging In TEM, phase contrast imaging normally refers to high-

resolution TEM. A phase contrast image aims to see the atomic structure of the sample

and thus requires a very thin sample (normally thinner than 30 nm) and the aperture

selecting more than two beam of the diffraction pattern.13 Figure 3.10 compares the

difference of on-axis TEM images under two-beam imaging condition and many-beam

Figure 3.10 Two-beam imaging condition and many-beam imaging condition and the

corresponding TEM images.

imaging condition respectively, in which two -beam imaging form the image with lattice

fringe only along a certain direction, while a real structure can be reflected under many-

beam imaging condition. By applying the phase contrast imaging technique, crystal

structure and growth orientation of the sample can be determined and their dislocations

and stacking defects can be revealed. Therefore, more beam collected increases the

resolution of the HRTEM image, which is normally obtained at the low-index zone axis.

HRTEM is a common technique used to characterize the structure of nanowires, which is

one of the main techniques applied in this thesis.

80

3.4.2 Diffraction mode

As described in Figure 3.8b, when the electron beam lines parallel to zone axis of the

crystalline sample, the objective lens take the electron beam scattered from the sample

and then disperse them to generate diffraction pattern in the back-focal plane. Readjusting

the intermediate lens for the imaging mode, the back-focal plane acts as an objective

plane this time and the diffraction pattern can be further projected onto the CCD. However,

such a strong beam could saturate the CCD camera, so we normally select a specific area

by inserting an aperture above the sample to get a typical diffraction pattern called SAED

pattern, which can be seen on the viewing screen. The space where the electron

diffraction pattern forms is called “real space” and the diffraction spots form when the

interference of the electron beam scattered beam form two adjacent planes that meets the

Bragg’s law = 2dsin B (λ is the electron beam wavelength, d is the lattice spacing of the

Figure 3.11 Schematic illustration of the interference of the electron beam waves and the

principles of the electron diffraction pattern. (http://www.ammrf.org.au/myscope/tem/

background/concepts/imagegeneration/diffraction/beam/)

81

crystalline sample, 2 B is the beam diffracted angle, Figure 3.11).14 it is noted that for the

unscattered electrons, they directly travel through O and produce a bright centre spot. The

distance between the diffracted spot G and the centre spot O is inversely proportional to

the lattice spacing of the sample.

Therefore, by measuring the distance between the diffraction spots, the crystalline

information of the sample can be studied. However, it is noted that this distance depends

not only to the scattered angle of the electron beam, but also the camera length, which is

the length between the sample and the place that the diffraction pattern is recorded.

Additionally, the diffraction pattern can also be applied to investigate the crystal structure

and quality of the sample.

3.4.3 Scanning mode

It has been demonstrated in early days that STEM can image heavy and light atoms,

which generates signals collected by the high-angular annular dark field (HAADF) that are

sensitive to the electron scattering by one atom or one column of atoms (Figure 3.12a).15

In terms of STEM, it is more like the way that SEM works: by adjusting the scan coil, the

electron beam is scanned on the sample. The signals are generated at a point on the

sample and a corresponding one displayed at an equivalent point on the viewing screen.

The electron detector used in the STEM mode is like an aperture we used in the TEM

mode. So when we insert the BF detector, direct-beam signals are captured and further

processed to become a BF STEM image. If scattered electrons (with scattered angle

larger than 50 mrads, Figure 3.12b) are captured by the detector with shifting to the

diffraction mode, we can end up with the DF STEM images. As a result, HAADF images

can be created, which is used to determine the distribution of elements depending on the

contrast. In HAADF images, material with higher atomic number is usually brighter than

the light one. By controlling the probe size that determines the resolution of STEM image,

high resolution HAADF STEM image can be obtained. Figure 3.13a describes the atomic-

resolution HAADF image of the interface between GaAs and Bi2Te3, in which the brighter

contrast of the right-hand side indicate the heavier element material of Bi2Te3.16 By

correlating the intensity profile of atom columns across the interface, the exact atoms

position can be secured (Figure 3.13b).

82

Figure 3.12 Schematic illustrations of (a) Z-contrast technique in a STEM and (b) the

HAADF detector setting for the Z-contrast imaging.

Figure 3.13 (a) HAADF STEM image of the interface between GaAs and Bi2Te3 viewed

from the respective <110> and <210> zone-axis. (b) Intensity profile of the atom column

marked in (a).16

3.4.4 Analytical mode

EDS is a quantitative and qualitative analytical technique that can provide information

of chemical characteristics of the sample. Unlike SEM sample, TEM samples are usually

very thin, therefore, there is less electron beam spreading as it passes though the sample

and the spatial resolution of X-ray microanalysis is nanometres order. Two types of X-rays

83

Figure 3.14 (a) Atomic resolution Z-contrast STEM image of Ba1.7Ca2.4Y0.9Fe5O13 film and

corresponding (b) Ba, (c) Ca, (d) Y, (e) Fe and (f) their overlap EDS maps.17

are generated by the interaction of the primary electron beam with nucleus of the sample:

the background X-rays; and Characteristics X-rays. These X-rays are detected by Ener

gy Dispersive detector which transfer the signals and shows the spectrum of intensity

versus X-ray energy under the STEM mode. Specific peaks in the spectrum results from

the Characteristics X-rays that can be used to analyses the sample compositions. In

addition, X-ray line scan profile provides the intensity of elements distributed along a

certain direction. By comparing the intensities of different elements of the sample, the

qualitative composition of the sample can be determined. Moreover, X-ray mapping can

provides image of elements distributions of the whole area of the sample. This technique

also enable atom-by-atom mapping at high resolution HAADDF STEM imaging (Figure

3.14), in which a very small probe size is selected.17

In this report, the grown nanowires are characterized by Philips Tecnai F20 FEG-

S/TEM (operated at 200 kV). A double-tilt holder was used to load the TEM sample, which

allows 0° α-tilt and 30° β-tilt to get the nanowire to the low-index zone-axis <110> to

identify the crystal structure, twin plane and stacking faults of the nanowires. Beside,

selected electron diffraction pattern, HRTEM and EDS techniques are also applied to

confirm the nanowires quality and the chemical characteristics of the nanowires.

84

Main features of Philips Tecnai F20:

Schottky type field emission source

Accelerating voltage 200 kV

Airlock system

Minimum probe size 1 nm

4k 4k One-View complementary metal–oxide–semiconductor camera with in-situ

software

Gatan image filter with a 2k 2k slow scan CCD

BF/DF TEM; BF/DF/HAADF STEM

Twin-lens objective pole-piece

Single-tilt holder, double-tilt holder, cryo-sample holders (-183°C), double-tilt (±30°)

low background (Be) EDS holder, high temperature double-tilt (±30°) holder

3.5 Sample preparation

3.5.1 SEM sample preparation

For non-conductive samples, a coating of carbon or gold or platinum is essential for SEM

characterizations to prevent sample charging. The coating materials is based on what

technique to use, for example, metal coating may affect the X-ray signals with the signals

generated by the sample itself. Besides, depending on the varied SEM setting, all samples

should be cut in an appropriate size to fill in the sample chamber and mounted on the

sample stub.

As III-V nanowires are semiconductor, the conductive adhesive is usually used in

case of sample damages by the confinement of electrons. However, the conductive

adhesive contains carbon and sample also may have been contaminated. Therefore, it is

necessary to put the sample into the vacuum oven baking for 7-8 hours at 70°C. After that,

all samples are sent to plasma cleaning for about 10 minutes to further remove the

contaminants. In this thesis, since all the nanowires were grown on the GaAs {111}B

substrates that have {110} cleavage plane, cutting the sample along the cleavage plan is

one of the methods to determine the side-facets of the grown nanowires during SEM

investigation. When the SEM session is finished, samples should be kept in the vacuum

container.

85

3.5.2 TEM sample preparation

For individual nanowires being investigated by TEM, a small piece of nanowire sample (a

few millimetre wide) is cut and put in a small container, in which a few ml ethanol is filled to

cover the sample and then the container is sealed. After that, the container with nanowire

sample is put into the ultrasonicator to take a bath for around 20 minutes to shake the

nanowires off the substrate. Then, a plastic pasteur pipette is used to disperse the ethanol

onto a 3mm 3mm holy carbon film coated copper grid, which was put on a filter paper on

a glass board to ensure the nanowires are well dispersed on the Cu grid.

Apart from the individual TEM sample, the nanowires cross sections and the cross

section of nanowires/substrate are also prepared for TEM investigations in this thesis,

which is detailedly described as follows.

3.5.3 Cross-sectional sample preparation

Nanowires cross-sections To investigate the side-facets of nanowires or chemical

characteristics of the nanowires, particular for the core-shell nanowires, the cross-sectional

sample of individual nanowires shall be prepared. In this thesis, we use the Leica Ultracut

UC6 to section our grown nanowires. This state-of-the-art ultramicrotone is utilized for

cutting ultrathin sections of both biological and physical sciences samples.

Before the sectioning, a small piece of substrate with nanowire grown is cut and put

into a cylinder or cuboid container, in which the surface with nanowires are faced up. Then,

liquid epoxy resin is filled into the container until full and the container is transferred into

the oven and baked at 60° for more than 24 h. As can be seen in Figure 3.14a,b, the resin

block is solidified after baking and the nanowires sample keeps at the container bottom.

After that, the container is removed and the resin block is well located on the holder to

make sure it is locked and stable. Then, a razor blade is used to trim the resin block and

remove the resin near the sample (Figure 3.15c). When the sample is exposed and

isolated, the substrate can be easily removed and leave the nanowires in the resin.

86

Figure 3.15 (a,b) Images of the resin block after baking. The nanowires sample is at the

container bottom. (c) Image of the nanowires sample after the neighbouring resin being

cleaned. (d) Image of the resin after the substrate being removed. Red arrows mark the

sample.

A bright and smooth surface is usually obtained after this process (refer Figure 3.15d).

After that, the exposed surface is further trimmed to a square or rectangle shape and

sectioned by the diamond knife. As schematically illustrated in Figure 3.16, after the resin

block is trimmed well, the sample holder is connected into the microtome, in which the

holder can be both manually and electrically driven to move up and down. Meanwhile, a

small “boat” filled with distilled water and with a diamond knife at the front is set up as

close to the resin surface. By setting the boat advance distance, the thin slices of the resin

are continuously cut, in which the thickness of the slices depends on the knife advance

distance (the distance is set 30-50 nm in this thesis) and the nanowires cross sections are

contained in the resin slice (step 3). These cut resin slices are floating on the water and a

Cu grid was used as a carrier to collect these slices for TEM investigations.

87

Figure 3.16 Schematic illustration of nanowires cross sections preparation.

Nanowire/substrate cross sections Investigating the interface between the grown crystal

and the substrate is a common technique in two-dimensional thin film or zero-dimensional

quantum dots growth.18,19 Similarly, this technique can also be applied in one-dimensional

nanowires growth.20 It has been well demonstrated that the nanowires/substrate cross

section can give a strong proof for investigate the nanowires initial growth behaviour on

the substrate, particularly for the hetero-epitaxial growth. Figure 3.17 shows a rough

process for preparing the nanowires/substrate cross sections for TEM investigations. To

begin with, a nanowires sample is cut into a 1cm 3mm rectangle and two similar-size Si

plates are also prepared. These three plates are glued together as a “sandwich” with the

middle plate being the sample (Figure 3.17a). This sandwich plate is further glued

vertically and tight on the sample holder (Figure 3.17b). After baking the sandwich plate

and ensure its stable on the the holder, the holder is transferred and fixed on the tripot and

tilted to a small angle with respect to the horizontal (Figure 3.17c). By manually polishing

88

Figure 3.17 (a) Image of the prepared “sandwich” sample with the real sample in the

middle. (b) Schematic illustration of the sample adherence to the holder. (c) Sample

loading on the tripot. (d) The sample interface after polishing and (e) attached to the Cu

grid.

the sample until the interface is thin though and then attached the sample to the Cu grid

for TEM invetigations (Figure 3.17d,e).

Another popular technique to prepare nanowire/substrate cross section sample for

TEM investigations is focus ion beam (FIB).21,22 The main benefits of FIB is site selective,

due to its ability to image the sample surface by using an integrated SEM, and it can be

applied in a wide range of materials, such as semiconductors, metals, ceramics and

polymers.23 Similar with electron beam for SEM, ion beam was used in FIB, in which

secondary electrons can also be detected to generate the image of the sample. Normally,

FIB uses Ga as an ion source, which has a low melting point and can be focused into a

very narrow probe (<10 nm in diameter). The primary ion penetration depth is ~20 nm for

25 keV Ga+.

The nanowires sample’s cross section was prepared as follows: the sample was first

cut along the cleavage plane, and took an ultrasonic bath in ethanol to shake off

nanowires. Then the cleaned sample was mounted horizontally in the chamber, in which

the electron beam was used to deposit a ~200 nm thick and 10-20 μm long of Pt layer on

the interest area of the sample surface, which is parallel to one of the samples cleavage

plane. Large trenches are sputtered on either side of interest area using a high Ga beam

89

current, so that the selected area was isolated, and further milling the Pt layer down to the

100 nm thick. This thin lamellar sample was lifted out and attached on the Cu grid for TEM

investigations.

References

1. Mataki, H.; Tanaka, H., Mass Production of Laser Diodes by MBE. Microelectronics

Journal 1994, 25, 619-630.

2. Izumi, S.; Kouji, Y.; Hayafuji, N., Multiwafer Gas Source Molecular Beam Epitaxial

System for Production Technology. J. Vac. Sci. Technol., B 1999, 17, 1011-1016.

3. Tsang, W. T., Extremely Low Threshold (AlGa)As Modified Multiquantum Well

Heterostructure Lasers Grown by Molecular‐Beam Epitaxy. Appl. Phys. Lett. 1981, 39,

786-788.

4. Patton, G. L.; Iyer, S. S.; Delage, S. L.; Tiwari, S.; Stork, J. M. C., Silicon-Germanium

Base Heterojunction Bipolar Transistors by Molecular Beam Epitaxy. IEEE Electron Device

Lett. 1988, 9, 165-167.

5. Jackrel, D. B.; Bank, S. R.; Yuen, H. B.; Wistey, M. A.; Jr., J. S. H.; Ptak, A. J.;

Johnston, S. W.; Friedman, D. J.; Kurtz, S. R., Dilute Nitride GaInNAs and GaInNAsSb

Solar Cells by Molecular Beam Epitaxy. J. Appl. Phys. 2007, 101, 114916.

6. Arthur, J. R., Molecular Beam Epitaxy. Surface Science 2002, 500, 189-217.

7. Cho, A. Y., Growth of Periodic Structures by the Molecular‐Beam Method. Appl. Phys.

Lett. 1971, 19, 467-468.

8. Cho, A. Y.; Arthur, J. R., Molecular Beam Epitaxy. Prog. Solid State Chem. 1975, 10,

157-191.

9. Williams, D. B.; Carter, C. B., Transmission Electron Microscopy: A Textbook for

Materials Science. Springer 2009.

10. Joy, D. C., Scanning Electron Microscopy. In Mater Sci Tech-Lond, Wiley-VCH Verlag

GmbH & Co. KGaA: 2006.

11. Fultz, B.; Howe, J., Transmission Electron Microscopy and Diffractometry of Materials.

Springer 2005.

12. Perovic, D. D.; Rossouw, C. J.; Howie, A., Imaging Elastic Strains in High-Angle

Annular Dark Field Scanning Transmission Electron Microscopy. Ultramicroscopy 1993, 52,

353-359.

90

13. Lentzen, M.; Jahnen, B.; Jia, C. L.; Thust, A.; Tillmann, K.; Urban, K., High-Resolution

Imaging with an Aberration-Corrected Transmission Electron Microscope. Ultramicroscopy

2002, 92, 233-242.

14. Kacher, J.; Landon, C.; Adams, B. L.; Fullwood, D., Bragg's Law Diffraction

Simulations for Electron Backscatter Diffraction Analysis. Ultramicroscopy 2009, 109,

1148-1156.

15. Krivanek, O. L.; Chisholm, M. F.; Nicolosi, V.; Pennycook, T. J.; Corbin, G. J.; Dellby,

N.; Murfitt, M. F.; Own, C. S.; Szilagyi, Z. S.; Oxley, M. P.; Pantelides, S. T.; Pennycook, S.

J., Atom-by-Atom Structural and Chemical Analysis by Annular Dark-Field Electron

Microscopy. Nature 2010, 464, 571-574.

16. Dycus, J. H.; White, R. M.; Pierce, J. M.; Venkatasubramanian, R.; LeBeau, J. M.,

Atomic Scale Structure and Chemistry of Bi2Te3/GaAs Interfaces Grown by Metallorganic

Van Der Waals Epitaxy. Appl. Phys. Lett. 2013, 102, 081601.

17. Sayers, R.; Flack, N. L. O.; Alaria, J.; Chater, P. A.; Palgrave, R. G.; McMitchell, S. R.

C.; Romani, S.; Ramasse, Q. M.; Pennycook, T. J.; Rosseinsky, M. J., Epitaxial Growth

and Enhanced Conductivity of an It-Sofc Cathode Based on a Complex Perovskite

Superstructure with Six Distinct Cation Sites. Chem. Sci. 2013, 4, 2403-2412.

18. Liao, X.; Zou, J.; Cockayne, D.; Wan, J.; Jiang, Z.; Jin, G.; Wang, K., Alloying,

Elemental Enrichment, and Interdiffusion During the Growth of Ge(Si)/Si(001) Quantum

Dots. Phys. Rev. B 2002, 65, 153306.

19. Liao, X.; Zou, J.; Cockayne, D.; Leon, R.; Lobo, C., Indium Segregation and

Enrichment in Coherent InxGa1-XAs/GaAs Quantum Dots. Phys. Rev. Lett. 1999, 82, 5148-

5151.

20. Liao, Z.-M.; Chen, Z.-G.; Lu, Z.-Y.; Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Zhang, Z.; Yang, L.;

Chen, P.-P.; Lu, W.; Zou, J., Au Impact on GaAs Epitaxial Growth on GaAs (111)B

Substrates in Molecular Beam Epitaxy. Appl. Phys. Lett. 2013, 102, 063106.

21. Langford, R. M.; Petford-Long, A. K., Preparation of Transmission Electron

Microscopy Cross-Section Specimens Using Focused Ion Beam Milling. J. Vac. Sci.

Technol., A 2001, 19, 2186-2193.

22. Phaneuf, M. W., Applications of Focused Ion Beam Microscopy to Materials Science

Specimens. Micron 1999, 30, 277-288.

23. Giannuzzi, L. A.; Stevie, F. A., A Review of Focused Ion Beam Milling Techniques for

TEM Specimen Preparation. Micron 1999, 30, 197-204.

91

Chapter 4

Effect of V/III ratio on binary GaAs nanowires

4.1 Introduction

In this chapter, the influence of V/III ratio on the growth behavior and structural quality of

GaAs nanowires is investigated. In particular, by tuning the As flux in MBE, GaAs

nanowires were grown at varied V/III ratios. Correponsing results were published on The

Journal of Physical Chemistry C (http://pubs.acs.org/doi/abs/10.1021/acs.jpcc.5b05606)

and intergrated in the Section 4.2 of this chapter.

92

4.2 Quality control of GaAs nanowire structures by limiting As flux in

MBE

In this study, we demonstrate that by merely limiting the As flux, the growth behavior and

structural quality of Au-catalyzed GaAs nanowires can be modulated in molecular beam

epitaxy. With decreasing the As flux through lowering the V/III ratio, GaAs nanowire growth

is found to be slow and defect-free wurtzite structured GaAs nanowires can be obtained

regardless of catalyst sizes. While, in the As-enrich environment (such as at relatively high

V/III ratio), thinner nanowires can grow longer with fewer planar defects. Based on our

extensive morphological, structural and compositional investigations, it is found that GaAs

nanowires grown under an As-limited status can lead to a thermodynamically controlled

growth process, while, when the nanowires grown under a relative high V/III ratio, a typical

kinetically dominated process is observed. This study provides a new insight for controlling

the structural quality of III-V nanowires by tuning the group-V flux.

Introduction

Development of III-V semiconductor nanowires has attracted great attention in the past

decade due to their intrinsic and superior properties,1,2 which makes them of interest for

applications in electronic and optoelectonic devices, such as solar cells,3 light-emitting

diodes4 and field-effect transistors.5 With higher electron mobility than Si,6 wide direct

bandgap (~1.4 eV)7 and possibilities to fabricate hetero-structured nanowires for band

structure engineering, such as GaAs/AlGaAs,8 and GaAs/GaAs(Sb),9 GaAs nanowire is

regarded as one of the essential components for a vast variety of future applications,

particularly in the near infrared region.7

To ensure the premium properties of III-V nanowire based devices, realization of

precise control in nanowire morphology, uniformity and structural quality becomes priority.

Normally, III-V nanowires are grown with metal catalysts (such as Au), which follows the v

VLS growth mechanism.10 MBE11 is one of the most effective techniques for growing

epitaxial III-V nanowires, during which growth parameters such as temperature,12 V/III

ratio,13-15 the absolute pressure of the group-III and group-V species15 can be well

controlled for varying the crystal structures and their structural qualities. Specifically, GaAs

nanowires have been grown by MBE in the past decade,16-18 with focus on optimizing the

93

nanowire structures. For example, Bauer et al.19 have studied the structural quality of self-

catalyzed GaAs nanowires. Also, several research groups have been dedicated to achieve

Au-catalyzed GaAs nanowires with defect-free structure by using different strategies, such

as lowering the As flux rate,20 controlling two-dimensional growth rate by tuning Ga flux

rate21 and lowering the growth rate of nanowires.22 It is well known that Au-catalyzed GaAs

nanowire growth by MBE depends on the different kinetics of Ga and As.23 It has been well

documented that increasing the flux rates of Ga or As can both lead to higher growth

rates.24,25 Also, the preferred formation of crystalline structure is attributed profoundly to

the supersaturation level in the catalysts,26 which is closely related to partial pressure of

Ga and As.27,28 However, few experimental studies in MBE have thoroughly explored the

impact of the group-V flux on the structural quality of GaAs nanowires.12,15,20 Therefore,

the role of As played in nanowire growth and structure control is still unclear in MBE, which

needs further clarification.

In this study, we explore the impact of the As flux on the growth behavior and

structural quality of Au-catalyzed GaAs nanowires grown by MBE. By tuning the absolute

As flux rate while keeping the Ga flux as a constant, we designed two nanowire growth

with different V/III ratios. Through carefully investigating the morphological and structural

features of grown nanowires, we found that nanowires grow slowly in the As-limited

condition (low V/III ratio), and generate low energy {100} facets with defect-free wurtzite

structure. However, in the As-enrich environment, nanowires show entirely different growth

behavior and structural characteristics. The fundamental reasons for such difference are

investigated and two growth mechanisms are proposed.

Experimental section

GaAs nanowires were grown on GaAs {111}B substrates using Au catalysts in a Riber 32

MBE system. Before nanowire growth, the GaAs {111}B substrates were degassed and

deoxidized to remove any contaminants in the growth chamber. In a typical growth, a

GaAs buffer layers with a thickness of 200 nm was grown on a pre-cleaned GaAs

substrate at 580°C to ensure that nanowires can grow on the atomically flat surface. The

substrate was then transferred to the preparation chamber and a ~1 nm thick Au thin film

was deposited directly on the buffer layer surface. After that, the Au coated substrate was

transferred back to the growth chamber. With annealing at 550°C for 5 min under the As

source protection, solid Au nanoparticles were generated on the GaAs substrate acting as

94

catalysts.18,29,30 The substrate temperature was then dropped to 500°C, where the growth

of GaAs nanowires was initiated by introducing Ga source to the growth chamber with a

flux rate of 1.2 10-7 Torr and adjusting flux rate of As source to be around1.2 10-6 Torr

and 2.4 10-6 Torr respectively, to achieve the V/III ratios of 10 and 20. The V/III ratio was

calculated by the As/Ga beam equivalent pressure ratio. After 60 min of nanowire growth,

the growth was terminated by switching off the Ga source while keeping the As source

until the system was cooled naturally to the room temperature.

The morphological characteristics of grown GaAs nanowires were investigated SEM

(JEOL 7800F, operated at an accelerating voltage of 5 kV) and their structural and

chemical characteristics were investigated by TEM (Philips Tecnai F20 equipped with EDS

for compositional analysis). SEM specimen for side-view investigations were cleaved

along the {110} cleavage plane31 of GaAs {111}B substrates. Individual nanowires for TEM

investigation were prepared through ultrasonicating the nanowire samples in ethanol and

dispersion individual nanowires onto the Cu grids coated with the holey carbon films.

Results and discussion

Figure 1a,b is typical plane-view SEM images taken from both samples, in which “white”

dots with similar densities are seen in all two samples, suggesting that nanowires might

grow vertically on the substrates. The insets of each SEM image reveal the hexagonal

faceted features of these “white dots”, indicating faceted side-walls of grown nanowires. It

is of interest to note that the sizes of “white dots” are varied in both cases, which should be

attributed to the annealing process. During which, the Au thin film breaks up and

aggregates into nanoparticles with different sizes, with small particles dissolving and

redepositing onto large ones due to the Ostwald ripening,32 leading to a significant large of

size distribution. In fact, such a phenomenon has been also observed by other groups

studying Au-catalyzed GaAs nanowires.15,33 To understand the growth status of these

white dots, side-view SEM investigations were performed and the results are shown in

Figure 1c,d correspondingly, in which nanowires were vertically grown on the substrates,

and the lengths of grown nanowires can be statistically determined. As can be seen,

grown nanowires in the sample with the V/III ratio of 10 are relatively short (less than 1 µm

95

Figure 1 SEM images of GaAs nanowires grown on GaAs {111}B substrate under different

V/III ratios. (a,b) Top-view SEM images of GaAs nanowires with the V/III ratios of 10 and

20, respectively. The inserts are high-magnified SEM images showing the side-facets of

nanowires. (c,d) Corresponding side-view SEM images with the insert in (d) being a

slightly tilted SEM image. All scale bars are 200 nm.

), and they have almost the same height. In contrast, in the sample with the V/III ratio of

20, grown nanowires have diverse heights with a tendency of thinner nanowires growing

longer, as indicated in the inset of Figure 1d. In both cases, no significant nanowire

tapering was observed, except some tapering were found at the bottom of the nanowires

caused by the lateral growth.34 It should be noted that the lateral dimensions of grown

nanowires vary and the observed variations fit well with the size distribution estimated from

the plan-view SEM images.

To understand the structural characteristics of grown nanowires, TEM investigations

were employed. Figure 2a,b is BF TEM images of a typical thin nanowire (~25 nm in

diameter) and a thick one (~65 nm in diameter) both taken from the sample with the V/III

ratio of 10. Figure 2c,d is the corresponding HRTEM images showing that nanowires have

the wurtzite structure, which were taken along the <110> zone axis. Interestingly, these

nanowires are defect-free along their entire lengths, irrespective of their catalysts’ sizes.

96

Figure 2 (a,b) BF TEM images of typical thin and thick nanowires from the sample grown

with the V/III ratio of 10. (c,d) Corresponding HRTEM images from both nanowires. (e,f)

BF TEM images of typical thin and thick nanowires from the sample grown with the V/III

ratio of 20. (g h) Corresponding HRTEM images.

This conclusion was further confirmed by our extensive investigations over a half dozen of

thin and thick nanowires. Similar TEM investigations were carried out for the sample grown

under the V/III ratio of 20, and the typical results are shown in Figure 2e–h. As presented

in Figure 2e,f, both thin and thick nanowires contain planar defects along their growth

directions. Figure 2g,h is the corresponding HRTEM images, from which planar defects

can be confirmed at the atomic level. Furthermore, our extensive HRTEM investigations

indicate that, in the sample grown under the V/III ratio of 20, the thicker the nanowires, the

higher the density of the planar defects.

97

It has been well documented that the catalyst composition and status are critical to

the growth of III-V nanowires,35,36 so that it is necessary to determine Ga concentration in

the catalysts during the nanowire growth. Accordingly, EDS analysis was performed

statistically on the catalysts of nanowires with different lateral dimensions in both samples.

Since EDS can only reveal the chemical composition in the post-growth catalysts and

since, during the cooling process of GaAs nanowire growth, the Ga can be expelled out

from the catalysts, we need to include the amount of Ga lost from the catalysts during the

cooling process. It has been well documented that, a “neck” region can be naturally formed

in grown nanowires during their cooling process,26,37 where the Ga in the catalysts can be

precipitated out in the catalyst-nanowire interface and combine with As to form a “neck”

section of nanowires underneath the catalysts, when the Ga source is switched off. In this

regard, the Ga presented in the “neck” region needs to be determined. Figure 3a shows a

TEM image of a top section of a typical nanowire, in which the nature of wurtzite structured

nanowire and a zinc-blende structured “neck” section can be seen. This structural feature

provides an opportunity to estimate the amount of Ga being expelled out from the catalyst

during the cooling process. In this study, we only focus on investigating the tendency of

the relationship between the Ga concentration in the catalysts and the catalyst size (in turn

the lateral dimension of the nanowires), so that we can simplify the estimation model by

assuming that the “neck” region has a cylinder shape and the catalyst has a hemisphere

shape, as illustrated in Figure 3b.

Since the total number of Ga atoms (NGa) in the catalyst during nanowire growth is

the sum of the number of Ga atoms (NN) in the “neck” section (zinc-blende structured

GaAs) and their number (NC) in the post-growth catalyst detected by EDS, that is, NGa =

NN + NC. Consequently, the Ga concentration during growth can be estimated as (NN +

NC)/ (NN + NC + NAu)38 with NAu being the number of Au atoms in the catalyst.

Since the “neck” region is zinc-blende structured GaAs, which contains 4 Ga atoms

in a unit cell, NN can be obtained by:

NN = 4VN/ , (1)

where is the volume of a unit cell of the zinc-blende structured GaAs, and VN is

the volume of the cylinder shaped “neck” region, which, according to Figure 3b, can be

determined by

2H (L being the lateral dimension of the nanowire and H the height

of the “neck” region).

98

To determine NC, we firstly determined the crystal structure of the catalysts. Figure 3c

is a selected area electron diffraction pattern taken from the interfacial region between a

catalyst and its underlying nanowire, in which two sets of electron diffraction patterns can

be seen. Using the diffraction spots of the GaAs nanowire as a reference, the diffraction

pattern of the catalyst can be determined as a [ 12] zone-axis of a face-centered cubic

structure with a lattice parameter of aC ≈ 0.41 nm. This is the Au crystal structure,

suggesting that the catalysts adopt the Au crystal structure. Accordingly, the volume of

each unit cell ( ) can be determined as . Based on the facts that (1) each unit

cell contains 4 atoms and (2) the total volume of catalyst is

(L/2 denotes the

radius of the hemispheric catalyst, as illustrated in Figure 3b), NC can be expressed as

, (2)

where CGa is the determined Ga atomic concentration in the post-growth catalysts by EDS.

Similarly, NAu can be determined using Equation 2, by replacing CGa with CAu. Figure 3d

shows an example of EDS profile taken from a typical catalyst, in which both Au and Ga

peaks are seen (noted that the observed Cu peaks are due to the Cu TEM grid). With our

careful compositional investigations over a dozen of nanowires, Ga concentration in the

post-growth catalysts is generally less than 10 at.%, which is smaller than the Ga solubility

in Au according to the Au-Ga phase diagram.39 Therefore, the catalysts adopt the Au

crystal structures.

Based on this approach, we can determine the Ga concentrations in the catalysts

during nanowire growth. Consequently, we examined a dozen of nanowires with different

diameters for each sample and the results are displayed in Figure 3e. As can be seen, a

clear tendency is shown as the thinner the nanowire, the higher the Ga concentration in

the catalyst during nanowire growth, which is consistent with previous studies.36,40

Interestingly, this tendency is seen to be more remarkably for the high V/III ratio, indicated

by the pink arrow in Figure 3e, possibly owing to the strong competition of Ga in catalysts

caused by its lower partial pressure in the growth chamber.

To understand our observations, two questions need to be answered. (1) Why, at a

relative high V/III ratio, nanowire lengths are varied and they contain planar defects with a

tendency of thinner nanowires being longer and having less planar defects? (2) Why are

nanowires grown at a low V/III ratio short, but have the same height and defect-free

structure?

99

To answer the first question, we note that, based on the classic theory of crystal

growth, the growth rate of nanowire ( ) can be expressed as:41,42

⁄ (3)

where is the Boltzmann constant, is the growth temperature andis the chemical

potential difference between the catalyst and nanowire nucleus, also known as

supersaturation in the catalyst, and can be presented as:27

=

, (4)

Figure 3 (a) A HRTEM image showing the top region of a typical GaAs nanowire grown at

the V/III ratio of 20. (b) A schematic model showing the nanowire top containing a

hemisphere catalyst, a cylinder disk-shaped “neck” region and the nanowire section. (c)

SAED pattern taken from catalyst/nanowire interface confirming the crystal structure of the

catalyst. (d) EDS spectra taken from the catalyst. (e) Estimated Ga concentration in the

catalysts as a function of catalysts size (dimension of nanowires) during nanowire growth.

100

where and are the Ga concentrations in the catalyst during nanowire growth

and at equilibrium with the GaAs crystal,27 respectively. Due to the low As solubility in Au,

As in the vapor would incorporate into the nanowire through the triple phase line,26 thus

and here indicate the As concentrations in the vapor. According to Equation 4, increases

with increasing , suggesting that the more Ga in the catalysts, the higher the chemical

potential difference,28,36,40,43 which in turn promotes fast nanowire growth according to

Equation 3.44 In this regard, with high Ga concentrations, small catalysts can induce

longer nanowires due to their high chemical potential difference and hence high growth

rate, resulting in the length variation of nanowires with different catalyst sizes.

It is has been well documented that high supersaturation in the catalyst can promote

the formation of wurtzite structure whereas low supersaturation is favored for the

generation of zinc-blende structure.26 Therefore, thinner nanowires with higher Ga

concentrations in the catalysts and hence with high catalyst supersaturation tend to

nucleate in wurtzite structure with better structural quality. In addition, from energy point of

view, the total free energy of nanowires, mainly including the surface energy and the

cohesive energy inside nanowires, tends to be minimized during nanowire growth. For

thinner nanowires, owing to the large surface-to-volume ratio, the surface energy plays a

crucial role in the total energy. In this situation, for thinner nanowires, wurtzite structure is

more energetically favorable due to their lower lateral surface energy compared to zinc-

blende structure, as suggested by the theoretical prediction.45,46 However, when the

diameter exceeds a critical value, the cohesive energy becomes dominating, which is

lower for the zinc-blende structure than for the wurtzite structure.33,47 As a result, for the

thicker nanowires, the difference of total free energy between wurtzite and zinc-blende

structured nanowires can be reduced, leading to the easy introduction of planar defects.

To answer the second question, we noted that, the group-III concentration in the

catalyst depends on the group-III partial pressure in the neighboring vapor,13,43 which is

determined by the direct absorption of group-III atoms by the catalyst and surface diffusion

of group-III atoms to the catalyst.48 It has been demonstrated that the diffusion length of

group III increases with decreasing the V/III ratio, resulting in higher group-III concentration

in the catalyst.49 During the MBE growth, only the growth species were introduced in the

growth chamber, hence the V/III ratio can be used to determine the partial pressure of

each growth species. Therefore, with the lower V/III ratio, the Ga partial pressure in the

growth chamber should be relatively increased in our study. Under this situation, almost all

Au catalysts can absorb more Ga regardless of their original sizes, resulting in a high

101

supersaturation in the catalysts. According to Equation 3, the increased catalyst

supersaturation should give rise to a high axial growth rate of nanowires. However, this is

not consistent with our observations, suggesting that the Ga concentration in the catalysts

is not the dominating factor for nanowire growth in our case. Therefore, we consider the

role of As during our nanowire growth under a low V/III ratio situation. Due to lowered As

partial pressure under a low V/III ratio, the chances of As atoms arriving at the growth front

and incorporating with Ga into the nanowire would be weakened,48 leading to a reduced

axial growth rate of nanowires.44 On this basis, we anticipate that the As flux can be

regarded as a limiting factor for our nanowire growth. Even though the Ga concentration in

catalysts varied from size to size, nanowires can still reach the same height due to the

limited As available in the vapor. Moreover, the dramatically slow growth of nanowires at

the low V/III ratio can be considered as a near equilibrium process,50 where the formation

of low energy {1100} facets (refer to Figure 1a)51 and the formation of perfect crystal

structure are favorable.

So far, we have achieved defect-free GaAs nanowires due to their slowed growth

rate under a low V/III ratio of 10, grown at a temperature of 500°C. Since, at a higher

temperature, As diffusion can be enhanced, which may enhance the nanowire growth, it is

necessary to clarify this point. For this reason, we repeat the GaAs nanowire growth with

the low V/III ratio of 10, but increasing the growth temperature to 550°C. Figure 4 shows

the electron microscopy characterization of this new designed sample. Figure 4a is the

side-view SEM image of the grown nanowires. Similar to its counterparts grown at 500°C,

many short nanowires with diverse nanowire sizes are grown on the substrate. Figure 4b

is a BF TEM image showing perfect crystal structure, and Figure 4c is a HRTEM image

showing the wurtzite crystal structure. Our extensive TEM investigations indicate that all

these nanowires are indeed defect-free wurtzite structured GaAs nanowires, suggesting

that under such an As limited environment, increasing their diffusibility is not sufficient to

promote the growth rate, which in turn benefit to adopt perfect crystal structure. This result

is consistent with the trend found in our previous study12 where the low V/III ratio promoted

defect-free crystal structure, although the V/III ratio was as low as 4. It is of interest to note

that, in our previous study,12 both higher Ga and As flux were used, in which As is no

longer as a limiting factor under the V/III ratio of 10. This conclusion suggests that the

absolute As flux should be as minimized in order for As becoming a limiting factor for

growing defect-free nanowires.

102

Figure 4 (a) Side–view SEM image of GaAs nanowires grown at 550°C with a V/III ratio of

10. (b) BF TEM image of a typical nanowire. (c) Corresponding HRTEM image showing

the structural characteristics.

Based on our comprehensive investigations and analysis outlined above, we believe

that our nanowires were grown under different mechanisms, controlled by the absolute As

flux. For a relatively low Ga flux of 1.2 10-7 Torr, Figure 5a illustrates the growth of GaAs

nanowires under a low V/III ratio and in turn a high Ga partial pressure in the vapor.

Therefore, all catalysts are able to absorb sufficient Ga to reach the supersaturation level

for nanowire nucleation and growth. With As being the limited factor for the nanowire

growth in this case, nanowires can only grow slowly and arrive at almost the same height

eventually. The slow growth of nanowires can be regarded as a thermodynamic process,

which results in the defect-free crystal structure. On the other hand, Figure 5b depicts

nanowire growth under a high V/III ratio, where Ga plays a dominated role for nanowire

growth in the As-enriched ambient. With competition of Ga, smaller catalysts are able to

absorb more Ga and would nucleate fast and promote longer nanowires with fewer planar

defects due to the higher supersaturation. As a consequence, nanowire growth under a

high V/III ratio can be regarded as a kinetic process.52

103

Figure 5 Schematic illustration of GaAs nanowires grown under varied V/III ratios. (a) As-

limited growth. (b) As-rich growth. The cuboid in red represents a three-dimensional

nucleus at the triple phase boundary and the stacking faults are marked as black lines in

the nanowire body.

Conclusions

In conclusion, we have demonstrated that the As flux can play an important role in the

growth behavior and structural quality of GaAs nanowires grown in MBE, and two different

growth mechanisms can be found under different V/III ratios when a low Ga flux is used.

The slow growth of GaAs nanowires under a low V/III ratio is As-limited, bringing in

uniform length and perfect crystal phase as a result of a thermodynamic process. In

contrast, under a high V/III ratio, Ga should be dominating factor for the GaAs nanowire

growth, which can be classified as a kinetic process. This study verifies the impact of the

As flux on the GaAs nanowires growth and realizes the controllable growth of defect-free

nanowires by limiting the overall As in growth chamber. Therefore, this study provides a

new approach for achieving phase perfection of III-V nanowires in MBE.

104

Author Information

Chen Zhou,† Kun Zheng,*,‡,§ Zhenyu Lu,∥ Zhi Zhang,† Zhiming Liao,† Pingping Chen,∥ Wei

Lu,∥ and Jin Zou*,†,‡

†Materials Engineering, ‡Centre for Microscopy and Microanalysis, and §Australian Institute

for Bioengineering and Nanotechnology, The University of Queensland, St. Lucia,

Queensland 4072, Australia

∥National Laboratory of Infrared Physics, Shanghai Institute of Technical Physics, Chinese

Academy of Science, 500 Yu-Tian oad, Shanghai 200083, People’s epublic of China

Corresponding Authors

*E-mail: [email protected].

*E-mail: [email protected].

References

(1) Xia, H.; Lu, Z.-Y.; Li, T.-X.; Parkinson, P.; Liao, Z.-M.; Liu, F.-H.; Lu, W,; Hu, W.-D.;

Chen, P.-P; Xu, H.-Y., et al., Distinct Photocurrent Response of Individual GaAs

Nanowires Induced by N-Type Doping. ACS Nano 2012, 6, 6005-6013.

(2) Guo, N.; Hu, W.; Liao, L.; Yip, S.; Ho, J. C.; Miao, J.; Zhang, Z.; Zou, J.; Jiang, T.; Wu,

S., et al., Anomalous and Highly Efficient InAs Nanowire Phototransistors Based on

Majority Carrier Transport at Room Temperature. Adv. Mater. 2014, 26, 8203-8209.

(3) Krogstrup, P.; Jorgensen, H. I.; Heiss, M.; Demichel, O.; Holm, J. V.; Aagesen, M.;

Nygard, J.; Fontcuberta i Morral, A., Single-Nanowire Solar Cells Beyond the Shockley-

Queisser Limit. Nat Photon 2013, 7, 306-310.

(4) Gudiksen, M. S.; Lauhon, L. J.; Wang, J.; Smith, D. C.; Lieber, C. M., Growth of

Nanowire Superlattice Structures for Nanoscale Photonics and Electronics. Nature 2002,

415, 617-620.

(5) Dayeh, S. A.; Aplin, D. P. R.; Zhou, X.; Yu, P. K. L.; Yu, E. T.; Wang, D., High Electron

Mobility InAs Nanowire Field-Effect Transistors. Small 2007, 3, 326-332.

(6) Wolfe, C. M.; Stillman, G. E.; Lindley, W. T., Electron Mobility in High-Purity GaAs. J.

Appl. Phys. 1970, 41, 3088-3091.

(7) Borg, B. M.; Wernersson, L.-E., Synthesis and Properties of Antimonide Nanowires.

Nanotechnology 2013, 24, 202001 .

105

(8) Mani, R. G.; Smet, J. H.; von Klitzing, K.; Narayanamurti, V.; Johnson, W. B.;

Umansky, V., Zero-Resistance States Induced by Electromagnetic-Wave Excitation in

GaAs/AlGaAs Heterostructures. Nature 2002, 420, 646-650.

(9) Dheeraj, D. L.; Patriarche, G.; Zhou, H. L.; Hoang, T. B.; Moses, A. F.; Gronsberg, S.;

van Helvoort, A. T. J.; Fimland, B. O.; Weman, H., Growth and Characterization of

Wurtzite GaAs Nanowires with Defect-Free Zinc Blende GaAsSb Inserts. Nano Lett. 2008,

8, 4459-4463.

(10) Wagner, R. S.; Ellis, W. C., Vapor-Liquid-Solid Mechanism of Single Crystal Growth.

Appl. Phys. Lett. 1964, 4, 89-90.

(11) Cho, A. Y.; Arthur, J. R., Molecular Beam Epitaxy. Prog. Solid State Chem. 1975, 10,

157-191.

(12) Lu, Z. Y.; Chen, P. P.; Liao, Z. M.; Shi, S. X.; Sun, Y.; Li, T. X.; Zhang, Y. H.; Zou, J.;

Lu, W., Impact of Growth Parameters on the Morphology and Microstructure of Epitaxial

GaAs Nanowires Grown by Molecular Beam Epitaxy. J. Alloys Compd. 2013, 580, 82-87.

(13) Zhang, Z.; Lu, Z.; Xu, H.; Chen, P.; Lu, W.; Zou, J., Structure and Quality Controlled

Growth of InAs Nanowires through Catalyst Engineering. Nano Res. 2014, 7, 1640-1649.

(14) Zhang, Z.; Lu, Z.-Y.; Chen, P.-P.; Lu, W.; Zou, J., Controlling the Crystal Phase and

Structural Quality of Epitaxial InAs Nanowires by Tuning V/III Ratio in Molecular Beam

Epitaxy. Acta Mater. 2015, 92, 25-32.

(15) Li, X.; Guo, H.; Yin, Z.; Shi, T.; Wen, L.; Zhao, Z.; Liu, M.; Ma, W.; Wang, Y.,

Morphology and Crystal Structure Control of GaAs Nanowires Grown by Au-Assisted MBE

with Solid As4 Source. J. Cryst. Growth 2011, 324, 82-87.

(16) Lu, Z.; Zhang, Z.; Chen, P.; Shi, S.; Yao, L.; Zhou, C.; Zhou, X.; Zou, J.; Lu, W.,

Bismuth-Induced Phase Control of GaAs Nanowires Grown by Molecular Beam Epitaxy.

Appl. Phys. Lett. 2014, 105.

(17) Zhang, Z.; Shi, S.-X.; Chen, P.-P.; Lu, W.; Zou, J., Influence of Substrate Orientation

on the Structural Quality of GaAs Nanowires in Molecular Beam Epitaxy. Nanotechnology

2015, 26, 255601.

(18) Liao, Z.-M.; Chen, Z.-G.; Lu, Z.-Y.; Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Zhang, Z.; Yang,

L.; Chen, P.-P.; Lu, W., et al., Au Impact on GaAs Epitaxial Growth on GaAs (111)B

Substrates in Molecular Beam Epitaxy. Appl. Phys. Lett. 2013, 102, 063106.

(19) Bauer, B.; Rudolph, A.; Soda, M.; Morral, A. F. i.; Zweck, J.; Schuh, D.; Reiger, E.,

Position Controlled Self-Catalyzed Growth of GaAs Nanowires by Molecular Beam

Epitaxy. Nanotechnology 2010, 21, 435601.

106

(20) Furthmeier, S.; Dirnberger, F.; Hubmann, J.; Bauer, B.; Korn, T.; Schüller, C.; Zweck,

J.; Reiger, E.; Bougeard, D., Long Exciton Lifetimes in Stacking-Fault-Free Wurtzite GaAs

Nanowires. Appl. Phys. Lett. 2014, 105, 222109.

(21) Plante, M. C.; LaPierre, R. R., Control of GaAs Nanowire Morphology and Crystal

Structure. Nanotechnology 2008, 19, 495603.

(22) Dheeraj, D. L.; Munshi, A. M.; Scheffler, M.; Helvoort, A. T. J. v.; Weman, H.; Fimland,

B. O., Controlling Crystal Phases in GaAs Nanowires Grown by Au-Assisted Molecular

Beam Epitaxy. Nanotechnology 2013, 24, 015601.

(23) Dubrovskii, V. G., Influence of the Group V Element on the Chemical Potential and

Crystal Structure of Au-Catalyzed III-V Nanowires. Appl. Phys. Lett. 2014, 104, 053110.

(24) Bao, X.-Y.; Soci, C.; Susac, D.; Bratvold, J.; Aplin, D. P. R.; Wei, W.; Chen, C.-Y.;

Dayeh, S. A.; Kavanagh, K. L.; Wang, D., Heteroepitaxial Growth of Vertical GaAs

Nanowires on Si (111) Substrates by Metal−Organic Chemical Vapor Deposition. Nano

Lett. 2008, 8, 3755-3760.

(25) Soci, C.; Bao, X.-Y.; Aplin, D. P. R.; Wang, D., A Systematic Study on the Growth of

GaAs Nanowires by Metal−Organic Chemical Vapor Deposition. Nano Lett. 2008, 8, 4275-

4282.

(26) Glas, F.; Harmand, J. C.; Patriarche, G., Why Does Wurtzite Form in Nanowires of III-

V Zinc Blende Semiconductors? Phys. Rev. Lett. 2007, 99, 146101.

(27) Joyce, H. J.; Wong-Leung, J.; Gao, Q.; Tan, H. H.; Jagadish, C., Phase Perfection in

Zinc Blende and Wurtzite III−V Nanowires sing Basic Growth Parameters. Nano Lett.

2010, 10, 908-915.

(28) Glas, F., Chemical Potentials for Au-Assisted Vapor-Liquid-Solid Growth of III-V

Nanowires. J. Appl. Phys. 2010, 108, 073506.

(29) Xu, H.; Wang, Y.; Guo, Y.; Liao, Z.; Gao, Q.; Jiang, N.; Tan, H. H.; Jagadish, C.; Zou,

J., High-Density, Defect-Free, and Taper-Restrained Epitaxial GaAs Nanowires Induced

from Annealed Au Thin Films. Cryst. Growth Des. 2012, 12, 2018-2022.

(30) Zhang, Z.; Zheng, K.; Lu, Z.-Y.; Chen, P.-P.; Lu, W.; Zou, J., Catalyst Orientation-

Induced Growth of Defect-Free Zinc-Blende Structured <001 > InAs Nanowires. Nano Lett.

2015, 15, 876-882.

(31) Lubinsky, A.; Duke, C.; Lee, B.; Mark, P., Semiconductor Surface Reconstruction: The

Rippled Geometry of GaAs(110). Phys. Rev. Lett. 1976, 36, 1058-1061.

107

(32) Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Liao, Z.-M.; Burgess, T.; Lu, H.-F.; Gao, Q.; Tan, H.;

Jagadish, C.; Zou, J., Quantitative Study of GaAs Nanowires Catalyzed by Au Film of

Different Thicknesses. Nanoscale Res. Lett. 2012, 7, 1-6.

(33) Shtrikman, H.; Popovitz-Biro, R.; Kretinin, A.; Houben, L.; Heiblum, M.; Bukała, M.;

Galicka, M.; Buczko, R.; Kacman, P., Method for Suppression of Stacking Faults in

Wurtzite III−V Nanowires. Nano Lett. 2009, 9, 1506-1510.

(34) Zou, J.; Paladugu, M.; Wang, H.; Auchterlonie, G. J.; Guo, Y.-N.; Kim, Y.; Gao, Q.;

Joyce, H. J.; Tan, H. H.; Jagadish, C., Growth Mechanism of Truncated Triangular III–V

Nanowires. Small 2007, 3, 389-393.

(35) Kuykendall, T. R.; Altoe, M. V. P.; Ogletree, D. F.; Aloni, S., Catalyst-Directed

Crystallographic Orientation Control of GaN Nanowire Growth. Nano Lett. 2014, 14, 6767-

6773.

(36) Zhang, Z.; Lu, Z.-Y.; Chen, P.-P.; Xu, H.-Y.; Guo, Y.-N.; Liao, Z.-M.; Shi, S.-X.; Lu,

W.; Zou, J., Quality of Epitaxial InAs Nanowires Controlled by Catalyst Size in Molecular

Beam Epitaxy. Appl. Phys. Lett. 2013, 103.

(37) Shtrikman, H.; Popovitz-Biro, R.; Kretinin, A.; Heiblum, M., Stacking-Faults-Free Zinc

Blende GaAs Nanowires. Nano Lett. 2008, 9, 215-219.

(38) Harmand, J. C.; Patriarche, G.; Ṕŕ-Laperne, N.; Ḿrat-Combes, M. N.; Travers, L.;

Glas, F., Analysis of Vapor-Liquid-Solid Mechanism in Au-Assisted GaAs Nanowire

Growth. Appl. Phys. Lett. 2005, 87, 1-3.

(39) Okamoto, H., Desk Handbook: Phase Diagram for Bianry Alloys, 2nd ed.; ASM

International: USA, 2000.

(40) Han, N.; Wang, F.; Hou, J. J.; Yip, S.; Lin, H.; Fang, M.; Xiu, F.; Shi, X.; Hung, T.; Ho,

J. C., Manipulated Growth of GaAs Nanowires: Controllable Crystal Quality and Growth

Orientations Via a Supersaturation-Controlled Engineering Process. Cryst. Growth Des.

2012, 12, 6243-6249.

(41) Givargizov, E. I., Fundamental Aspects of VLS Growth. J. Cryst. Growth 1975, 31, 20-

30.

(42) Biswas, S.; O’ egan, C.; Petkov, N.; Morris, M. A.; Holmes, J. D., Manipulating the

Growth Kinetics of Vapor–Liquid–Solid Propagated Ge Nanowires. Nano Lett. 2013, 13,

4044-4052.

(43) Guo, Y. N.; Xu, H. Y.; Auchterlonie, G. J.; Burgess, T.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C.; Shu, H. B.; Chen, X. S., et al., Phase Separation Induced by Au

Catalysts in Ternary InGaAs Nanowires. Nano Lett. 2013, 13, 643-650.

108

(44) Gil, E.; Dubrovskii, V. G.; Avit, G.; André, Y.; Leroux, C.; Lekhal, K.; Grecenkov, J.;

Trassoudaine, A.; Castelluci, D.; Monier, G., et al., Record Pure Zincblende Phase in

GaAs Nanowires Down to 5 nm in Radius. Nano Lett. 2014, 14, 3938-3944.

(45) Galicka, M.; Bukała, M.; Buczko, .; Kacman, P., Modelling the Structure of GaAs and

InAs Nanowires. J. Phys. Condens. Matter 2008, 20, 454226.

(46) Akiyama, T.; Sano, K.; Nakamura, K.; Ito, T., An Empirical Potential Approach to

Wurtzite–Zinc-Blende Polytypism in Group III–V Semiconductor Nanowires. Jpn. J. Appl.

Phys. 2006, 45, L275.

(47) Dubrovskii, V. G.; Sibirev, N. V.; Cirlin, G. E.; Bouravleuv, A. D.; Samsonenko, Y. B.;

Dheeraj, D. L.; Zhou, H. L.; Sartel, C.; Harmand, J. C.; Patriarche, G., et al., Role of

Nonlinear Effects in Nanowire Growth and Crystal Phase. Phys. Rev. B: Condens. Matter

2009, 80, 205305.

(48) Johansson, J.; Svensson, C. P. T.; Mårtensson, T.; Samuelson, L.; Seifert, W., Mass

Transport Model for Semiconductor Nanowire Growth. J. Phys. Chem. B 2005, 109,

13567-13571.

(49) Dayeh, S. A.; Yu, E. T.; Wang, D., III−V Nanowire Growth Mechanism:  V/III atio and

Temperature Effects. Nano Lett. 2007, 7, 2486-2490.

(50) Guo, Y. N.; Zou, J.; Paladugu, M.; Wang, H.; Gao, Q.; Tan, H. H.; Jagadish, C.,

Structural Characteristics of GaSb/GaAs Nanowire Heterostructures Grown by Metal-

Organic Chemical Vapor Deposition. Appl. Phys. Lett. 2006, 89, 231917.

(51) Ren, X.; Huang, H.; Dubrovskii, V. G.; Sibirev, N. V.; Nazarenko, M. V.; Bolshakov, A.

D.; Ye, X.; Wang, Q.; Huang, Y.; Zhang, X., et al., Experimental and Theoretical

Investigations on the Phase Purity of GaAs Zincblende Nanowires. Semicond. Sci.

Technol. 2011, 26, 014034.

(52) Dubrovskii, V. G.; Sibirev, N. V.; Harmand, J. C.; Glas, F., Growth Kinetics and

Crystal Structure of Semiconductor Nanowires. Phys. Rev. B 2008, 78, 235301.

109

Chapter 5

Effect of growth time on binary GaAs nanowires

5.1 Introduction

In this chapter, the impact of growth duration on the growth rate and structural quality of

binary GaAs nanowires were investigated. By growing nanowires for 1h, 1.5h and 2h

respectively under the same growth temperature of 500°C, we found that nanowires grew

faster and their structural qualities were enhanced with longer the growth duration. This

tendency was further confirmed by repeating the growth under 400°C, in which nanowires

were grown for 20min and 60min. This study demonstrates that the growth duration can

110

play a critical role in semiconductor nanowires growth and modulating their structural

quality. Correponsing results were published on Journal of Materials Chemistry C

(http://pubs.rsc.org/en/content/articlepdf/2017/tc/c6tc05209f) and intergrated in the Section

5.2 of this chapter.

5.2 Phase purification of GaAs nanowires by prolonging the growth

duration in MBE

In this study, we demonstrated that by merely prolonging the growth duration, the growth

behavior and phase purity of GaAs nanowires can be manipulated in molecular beam

epitaxy. Through careful morphological, structural and chemical characterizations of grown

nanowires, it was found that under the group-III dominated regime, by increasing the

growth duration, the Ga concentration in the catalysts increases due to formation of

nanowire shoulders, which provides sites for Ga diffusion towards the catalysts. This fact

leads to the enhancement of the nanowires growth and improved phase purity. In

particular, single-phase nanowire sections have been observed with prolonged nanowire

growth duration. This study provides a new insight into the role of growth duration that can

play in III-V nanowires growth and a possible approach for nanowire phases engineering,

which is critical for nanowires applications.

Introduction

Semiconductor nanowires, with the built-in one dimensional configuration enabling

quantum confinement,1 show many interesting properties in mechanical,2 thermal,3

electronic4 and optical areas,5 which are not available in their bulk counterparts. However,

these properties can be affected by the phase purity of nanowires.2-4 In particular, for the

case of III-V semiconductor nanowires, lattice defects, such as twins/stacking faults are

commonly observed, which can be considered as alternating layers of wurtzite and zinc-

blende phases,6 leading to the multi-phase nanowires that show different properties from

the single-phase nanowires.7 For example, Chen et al.8 showed that the Young’s modulus

of multi-phase GaAs nanowires was increased compared with that of the single-phase

nanowires, due to the varied bonding configuration. Additionally, Han et al.7 demonstrated

the relationship between the phase purity and electronic transport properties of GaAs

111

nanowires, in which multi-phase nanowires exhibit n-type conductivity while single-phase

nanowires show p-type conductivity. Therefore, understanding the impact factors and

realizing manipulation of nanowires phase purity and defects are critical for nanowires

applications with desired properties.

As one of the important III-V semiconductor nanowires, GaAs nanowires have been

investigated extensively,9 predominately due to their advantages in high electron mobilities

(compared with Si nanowires),10 direct band gap11 and heterojunction formation with other

III-V semiconductor nanowires, such as InAs12 and AlGaAs,13 making them great potential

for applications in future devices. To date, one of the most popular approaches for

fabricating III-V nanowires is the vapor-liquid-solid growth mechanism using Au

nanoparticles as catalysts to induce axial nanowire growth, which was first described by

Wagner and Ellis in 1960s.14 Over the years, many theoretical models and numerous

experiments have studied the growth behavior and phase engineering of III-V nanowires

by tuning the basic growth parameters, such as the growth temperature15 and the V/III

ratios.16 For example, it has been established that a relative low growth temperature can

promote high-quality nanowires because of the thermodynamic controlled growth.17,18 In

addition, for nanowires grown at III-dominated regime, their growth rate and phase purity

can be manipulated by tuning the group-III vapor pressure.19 Particularly, for GaAs

nanowires grown in MBE, the axial growth and phase purity are driven by the catalyst

supersaturation,20 which is determined by the Ga partial pressure in the vapor and thus the

Ga concentration in the catalysts during nanowire growth.21

On the other hand, it has been demonstrated that the average III-V nanowire length

increases linearly within a certain period of growth duration.22,23 However, in these studies,

nanowires were grown at relatively low V/III ratios, which is under V-dominated regime, in

which the nanowire growth is determined by the group-V vapor pressure in the

environment.21 Additionally, very little study so far was reported on the impact of the

growth duration on nanowire phase purity, which is a key factor for the grown nanowires.

In this study, we present the impact of the growth duration on GaAs nanowires grown

under a relative high V/III ratio of ~20 in MBE. Through extensive morphological, structural

and chemical investigations on the grown nanowires, we found an extraordinary nanowire

behavior: with increasing the growth duration, nanowires grow faster and have improved

structural qualities. More notably, the nanowire phase purity has been significantly

improved from multi-phase to single-phase structures during prolonged nanowire growth,

where the Ga concentration in the catalysts is found to be remarkably enhanced. The

112

fundamental reasons behind these extraordinary phenomena were carefully investigated

by electron microscopy, from which the observed extraordinary nanowire behavior was

clarified.

Experimental section

The epitaxial growth of Au catalyzed GaAs nanowires was carried out on GaAs {111}B

substrates in a Riber 32 MBE system. Degassing and deoxidization of GaAs {111}B

substrates were performed before nanowire growth to remove any contaminants in the

growth chamber. To ensure that nanowires can grow on the atomically flat surface, GaAs

buffer layers were firstly deposited on the pre-cleaned GaAs substrates at 580°C. The

substrates were then transferred to the preparation chamber and Au thin films with a

thickness of ~1 nm were deposited directly on the buffer layer surfaces. After that, the Au

coated substrates were transferred back to the growth chamber, where they were

annealed at 550°C for 5 min under the As source protection and solid Au nanoparticles

were formed to act as catalysts. The substrate temperature was then dropped to 500°C, at

which the growth of GaAs nanowires was initiated by introducing Ga source to the growth

chamber with a beam equivalent pressure of 1.9 10-7 Torr and a V/III ratio of ~20. In this

study, GaAs nanowires were grown with different growth durations, namely 1h, 1.5h and

2h. Then, the nanowire growth was terminated by switching off the Ga source while

keeping the As source until the system was cooled to 300°C.

The morphological characteristics of grown GaAs nanowires were investigated using

field emission scanning electron microscopy (FE-SEM, JEOL 7800F, operated at an

accelerating voltage of 5kV) and their structural and chemical characteristics were

investigated by TEM (Philips Tecnai F20 equipped with EDS for compositional analysis).

SEM specimen for both tilted-view and side-view investigations were cleaved along the

{110} cleavage plane of GaAs {111}B substrates. Individual nanowires for TEM observation

were prepared through ultrasonicating the nanowire samples in ethanol and dispersion

individual nanowires onto the Cu grids coated with holey carbon films.

Results and discussion

Figure 1a-c is typical SEM images taken from the samples in which GaAs nanowires were

grown respectively for 1h, 1.5h and 2h (tilted 30° with regard to the surface normal), and

113

Figure 1 (a-c) 30°-tilted SEM images of GaAs nanowires grown for 1h, 1.5h and 2h,

respectively. (d-f) Corresponding side-view SEM images. The inset in (d) plots the

nanowires length as a function of the growth duration. All scale bars are 1µm.

shows that the grown nanowires in three samples have a similar density and the majority

nanowires are straightly aligned. It is of interest to note that the nanowire diameter is

relatively uniform for 1h growth, while nanowires grown for 1.5h and 2h exhibit shoulder

morphology, resulting from nanowire lateral growth.24 To better understand the nanowire

growth status, side-view SEM images were taken from three samples, respectively

(described in Figure 1d-f), where most of them grow perpendicularly to the substrate

surface, that is, along the <111>B direction. More notably, the average length of 1h grown

nanowires is ~1.5 µm, while 1.5h and 2h grown nanowires have ~3 µm and ~5.5 µm in

length, respectively. These results were confirmed by our extensive SEM investigations of

different areas in each sample. It is of interest to note that, through compared the average

lengths of nanowires grown under different durations (see Fig. 1d inset), we can clearly

witness the tendency that the nanowire average growth rate is accelerated at the later

stage of the nanowire growth.

To understand the structural characteristics of grown GaAs nanowires, extensive

TEM investigations were performed over a dozen nanowires for nanowires grown for

different durations. Figure 2a is a BF TEM image of a typical GaAs nanowire grown for 1h,

114

in which the nanowire with a uniform diameter can be observed. Figure 2b,c is the

corresponding <11 2 0> zone-axis HRTEM images taken from the frames in Figure 2a,

showing the phase purity of the grown nanowire. As can be seen, planar defects are

randomly distributed along the entire nanowire, with the defect density counted as

~240/μm. The inset in Figure 2b is the corresponding SAED pattern taken from the

nanowire bottom, confirming the wurtzite structured nanowire. Notably, pronounced

streaks along the 000l* diffraction spots can be seen in the SAED pattern, reflecting a

high-density of planar defects lying on the {0001} planes – in an excellent agreement with

the HRTEM image. On the other hand, Figure 2d is a BF TEM image of a typical nanowire

grown for 1.5h, which has a longer length when compared with nanowires grown for 1h.

Figure 2e,f is the corresponding <11 2 0> zone-axis HRTEM images taken respectively

from bottom and top regions of the nanowire. Similar to the case of nanowires grown for

1h, a large density of planar defects is witnessed in the nanowire bottom region, echoed

by the diffraction streaks along the 000l* diffraction spots shown in the corresponding

SAED pattern (see the inset in Figure 2e). However, it can be noted that the defects

density in the nanowire top region decreases notably to ~ 133/μm. Similarly, Figure 2g

shows a BF TEM image of a typical nanowire grown for 2h, and Figure 2h is the

corresponding HRTEM image taken from the nanowire bottom region, showing a high

density of planar defects (verified by the inset of corresponding SAED pattern), which is

consistent with other two samples, indicating that at the beginning of the nanowire growth,

planar defects exist indeed. Surprisingly, Figure 2i shows a section of defect-free nanowire

in the nanowire top region, suggesting that phase purity of the nanowire has been

significantly improved at the late stage of nanowire growth (verified by the inset of

corresponding SAED pattern). These observations are further confirmed by our extensive

TEM investigations from more than a dozen of nanowires for each sample. These

experimental results indicate that the phase purity of nanowires can be improved with

increasing the growth duration.

115

Figure 2 BF TEM images showing the typical GaAs nanowire grown for (a) 1h, (d) 1.5h

and (g) 2h. Corresponding HRTEM images taken from the bottom and top regions of the

nanowires grown for (b,c) 1h, (e,f) 1.5h and (h,i) 2h. The insets in (b,e,h,i) are

corresponding <11 2 0> zone-axis SAED pattern showing the phase purity of the nanowire

regions.

Since the growth behavior of III-V nanowires can be strongly affected by the catalyst

composition,25 particularly the group-III concentration,26 statistical EDS analysis was

performed on the post-growth catalysts of grown GaAs nanowires. Figure 3a-c is typical

BF TEM images taken from typical nanowire tips in nanowires grown for 1h, 1.5h and 2h,

respectively. Figure 3d-f is their corresponding EDS spectra taken from their catalysts, in

which a very small amount of Ga atoms were left in these catalysts after nanowire growth

116

(noted the Cu peaks come from the Cu grids used to support nanowires for TEM

investigations). This conclusion can be confirmed by extensive EDS analyses from over a

dozen post-growth catalysts of nanowires grown under different growth durations. To

determine the true Ga concentrations in the catalysts during the nanowire growth, we need

to take account the Ga atoms expelled out during the cooling stage (the necking region)

and those left in the post-growth catalysts.21 As can be seen from Figure 3a-c, the neck

region (the zinc-blende structured necking region below the catalyst) in Figure 3c is larger

than those in Figure 3a,b (confirmed by the their insets of corresponding HRTEM images),

suggesting that more Ga atoms were expelled out from the catalysts of nanowires grown

for 2h during the cooling process.21 Our statistic estimations suggest that, at the end of

nanowire growth, the Ga concentration in the catalysts are ~16 at.% (for 1h grown

nanowires), ~22 at.% (for 1.5h grown nanowires) and ~37 at.% (for 2h grown nanowires),

respectively (Figure S1). These results indicate that the Ga concentrations in the catalysts

vary during the nanowire growth - increasing with prolonged nanowires growth.

Figure 3 (a-c) High-magnified BF TEM images of the typical tip region in GaAs nanowires

grown for 1h, 1.5h and 2h, respectively. The yellow arrows mark the length of necking

regions. The insets are corresponding HRTEM images showing the structure of the

necking region. (d-f) Corresponding EDS spectrums taken from the catalysts.

117

Based on our detailed TEM investigations, the growth behavior of GaAs nanowires

can be summarized as follows. For the first hour of nanowire growth, nanowires are short

and have multi-phase structure; while, with increasing the growth duration, nanowires are

longer and their phase purities have been improved, which were significantly enhanced

during further prolonged nanowire growth. Particularly, for nanowires grown for 2h, single-

phase nanowire section can be found in the final growth stage. To understand this

tendency, we noted from our previous study21 that for GaAs nanowires grown in a Ga-

dominated growth regime (note that in this study, V/III ratio is ~20), catalysts containing

higher Ga concentration, and hence the resultant high supersaturation, enhance the axial

nanowires growth and wurtzite phase purity. It has also been well demonstrated that Ga in

the Au catalysts can affect the catalysts surface energy, leading to the phase purification

of GaAs nanowires.27 Therefore, to understand the differences of nanowire growth rate

and phase purity in our different growths, it is necessary to understand why Ga

concentrations in the catalysts increase with increasing the growth duration.

It has been generally acknowledged that in III-V nanowire growth, the size of catalysts

may vary the III concentration in catalysts due to their different surface tensions.21,28

Therefore, to understand the impact of the catalyst size on the catalyst composition and

the subsequent nanowire growth, we statistically measured the catalyst diameters in

nanowires grown under different durations. Figure 4 summaries our measurements and

shows that the distributions of catalyst sizes in three cases are similar (30-50 nm),

verifying the identical pre-growth (Au film deposition and annealing process). Based on

this part of the study, the impact of catalyst size on the nanowire growth can be eliminated.

On the other hand, it should be noted that the temperature difference between the hot

substrate and catalyst with longer the nanowire length (less than 10 µm) is minor;29

therefore, the temperature influence on the Ga concentration in the catalysts during

nanowire growth can be negligible.

118

Figure 4 Histograms statistically illustrate the catalysts diameters of GaAs nanowires

grown for 1h, 1.5h and 2h, respectively.

To further understand why Ga concentrations in the catalysts increase with increasing

the nanowire growth duration, , we note that Ga in the catalysts can be collected via three

ways: 23 directly absorption by catalysts, substrate diffusion and nanowires sidewall

diffusion. Since Ga absorption by catalysts is similar in the three samples which were

grown under the same V/III ratio, and Ga diffusion length is relatively short on GaAs {111}B

surface,30 the Ga concentration in catalysts is mainly determined by its sidewall diffusion.

On this point, we note that with increasing the growth duration, nanowires formed the

shoulder morphology. It has been theoretically demonstrated that these shoulders can be

considered as numerous atomic step sites (shown in Figure 5a) that increase with larger

shoulder diameter, where atoms can be collected in these sites and further diffuse towards

the catalysts.23 Since the incident angle (α) of Ga flux in the MBE system is less than 90°

(see Figure 5b), Ga atoms are not only collected by the nanowire sidewall (perpendicular

to the substrate surface) but also collected by the step sites (parallel to the substrate

surface, proportional to cosα),23 the fraction of Ga atoms reaching the catalysts can be

increased with increasing the step sites, which is in excellent agreement with our

experimental results. To comprehensively understand the growth mechanism of GaAs

nanowires with increasing the growth duration and their phase evolution, schematic growth

models were proposed as described in Figure 5c. As can be noted, nanowires grown for

1h have relatively uniform diameter, while, with increasing the growth duration to 1.5h, the

formed shoulders provide step sites for Ga atoms collection and assisting their further

diffusion towards the catalysts, and thus more Ga atoms can reach the catalysts (note:

119

under the relative high V/III ratio of 20 in this study, As diffusion is not a limiting factor for

nanowires growth). However, since the nanowires lateral growth occurs with increasing the

growth duration, we anticipate that the competition between Ga being consumed by

nanowires lateral growth and Ga being collected by the step sites is significant at this

stage (Figure S2). Consequently, Ga concentration in the catalysts remains relatively

stable during nanowire growth from 1h to 1.5h. Nevertheless, with further prolonging the

growth duration to 2h, the thicker shoulders provide more step sites, and leading more Ga

Figure 5 (a) HRTEM image of a typical shoulder area as marked in the inset. The red line

describes the shoulder with atomic steps. (b) Sketch of the typical shoulder and the

incident angle of Ga flux in the MBE system. (c) Schematic growth models illustrate GaAs

nanowires growth with increasing the growth time and corresponding Ga concentration in

the catalysts during the growth.

120

atoms to diffuse to the catalysts and in turn increasing the Ga concentration in the

catalysts. As a consequence, nanowires axial growth rate increase and their phase purity

are significantly improved. This impact of growth duration on nanowire growth was further

confirmed when we repeated the growth with varied growth parameters (see Figure S3-

S4). The multi-phase GaAs nanowires have been demonstrated to have a type II band

alignment and unique photoluminescence behavior,31,32 which shows advantages in

photovoltaics applications.33

Conclusion

In conclusion, we showed that growth duration can play a critical role in manipulating the

growth behavior and phase purity of GaAs nanowires grown in MBE. By increasing the

growth duration, the Ga concentration in the catalysts increases which in turn increases

the nanowire growth rate, caused by shoulder assistance sites for Ga diffusion towards the

catalysts. These lead to the growth of high phase-purity GaAs nanowires and the unique

configuration of multi-phase to single-phase structure opens up possibilities for band

structure alignment in nanowires. This study proposes a new mechanism for

understanding the growth behavior and structural characteristics of GaAs nanowires

through varying the growth duration, which may be universally applicable in other III-V

nanowires systems.

Author Information

Chen Zhou,† Kun Zheng,*, ♩, ¶ Zhi-Ming Liao,† Ping-Ping Chen,§ Wei Lu,§ and Jin Zou*, †, ♩

†Materials Engineering, ♩Centre for Microscopy and Microanalysis, ¶Australian Institute for

Bioengineering and Nanotechnology, The University of Queensland, St. Lucia,

Queensland 4072, Australia

§National Laboratory of Infrared Physics, Shanghai Institute of Technical Physics, Chinese

Academy of Science, 500 Yu-Tian oad, Shanghai 200083, People’s epublic of China

Corresponding Author

*Email: [email protected],

*Email: [email protected].

121

References

(1) Xia, Y.; Yang, P.; Sun, Y.; Wu, Y.; Mayers, B.; Gates, B.; Yin, Y.; Kim, F.; Yan, H.,

One-Dimensional Nanostructures: Synthesis, Characterization, and Applications. Adv.

Mater. 2003, 15, 353-389.

(2) Dai, S.; Zhao, J.; He, M.-R; Wang, X.; Wan, J.; Shan, Z.; Zhu, J., Elastic Properties of

GaN Nanowires: Revealing the Influence of Planar Defects on Young’s Modulus at

Nanoscale. Nano Lett. 2015, 15, 8-15.

(3) Murphy, K. F.; Piccione, B.; Zanjani, M. B.; Lukes, J. R.; Gianola, D. S., Strain- and

Defect-Mediated Thermal Conductivity in Silicon Nanowires. Nano Lett. 2014, 14, 3785-

3792.

(4) Thelander, C.; Caroff, P.; Plissard, S.; Dey, A. W.; Dick, K. A., Effects of Crystal Phase

Mixing on the Electrical Properties of InAs Nanowires. Nano Lett. 2011, 11, 2424-2429.

(5) Xia, H.; Lu, Z.-Y.; Li, T.-X.; Parkinson, P.; Liao, Z.-M.; Liu, F.-H.; Lu, W.; Hu, W.-D.;

Chen, P.-P.; Xu, H.-Y.; Zou, J.; Jagadish, C., Distinct Photocurrent Response of Individual

GaAs Nanowires Induced by n-Type Doping. ACS Nano 2012, 6, 6005-6013.

(6) Shtrikman, H.; Popovitz-Biro, .; Kretinin, A.; Houben, L.; Heiblum, M.; Bukała, M.;

Galicka, M.; Buczko, R.; Kacman, P.,Method for Suppression of Stacking Faults in

Wurtzite III−V Nanowires. Nano Lett. 2009, 9, 1506-1510.

(7) Han, N.; Hou, J. J.; Wang, F.; Yip, S.; Yen, Y.-T.; Yang, Z.-X.; Dong, G.; Hung, T.;

Chueh, Y.-L.; Ho, J. C., GaAs Nanowires: From Manipulation of Defect Formation to

Controllable Electronic Transport Properties. ACS Nano 2013, 7, 9138-9146.6-841.

(8) Chen, Y.; Burgess, T.; An, X.; Mai, Y.-W.; Tan, H. H.; Zou, J.; Ringer, S. P.; Jagadish,

C.; Liao, X., Effect of a High Density of Stacking Faults on the Young’s Modulus of GaAs

Nanowires. Nano Lett. 2016, 16, 1911-1916.

(9) Law, M.; Goldberger, J.; Yang, P. D., Semiconductor Nanowires and Nanotubes. Ann.

Rev. Mater. Res. 2004, 34, 83-122.

(10) Wolfe, C. M.; Stillman, G. E.; Lindley, W. T., Electron Mobility in High-Purity GaAs. J.

Appl. Phys. 1970, 41, 3088-3091.

(11) Bi, W. G.; Tu, C. W., Bowing parameter of the band-gap energy of GaNxAs1−x Appl.

Phys. Lett. 1997, 70, 1608-1610.

(12) Paladugu, M.; Zou, J.; Guo, Y.-N.; Zhang, X.; Kim, Y.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C., Nature of Heterointerfaces in GaAs/InAs and InAs/GaAs Axial Nanowire

Heterostructures Appl. Phys. Lett. 2008, 93, 101911.

122

(13) Mani, R. G.; Smet, J. H.; von Klitzing, K.; Narayanamurti, V.; Johnson, W. B.;

Umansky, V., Zero-resistance States Induced by Electromagnetic-wave Excitation in

GaAs/AlGaAs Heterostructures Nature 2002, 420, 646-650.

(14) Wagner, R. S.; Ellis, W. C., Vapor-Liquid-Solid Mechanism of Single Crystal Growth.

Appl. Phys. Lett. 1964, 4, 89-90.

(15) Plante, M. C.; LaPierre, R. R., Au-assisted Growth of GaAs Nanowires by Gas Source

Molecular Beam Epitaxy: Tapering, Sidewall Faceting and Crystal Structure.

Nanotechnology 2008, 19, 495603.

(16) Dheeraj, D. L.; Munshi, A. M.; Scheffler, M.; Helvoort, A. T. J. V; Weman, H.; Fimland,

B. O.,Controlling Crystal Phases in GaAs Nanowires Grown by Au-assisted Molecular

Beam Epitaxy. Nanotechnology 2013, 24, 015601.

(17) Lu, Z. Y.; Chen, P. P.; Liao, Z. M.; Shi, S. X.; Sun, Y.; Li, T. X.; Zhang, Y. H.; Zou, J.;

Lu, W., Impact of Growth Parameters on the Morphology and Microstructure of Epitaxial

GaAs Nanowires Grown by Molecular Beam Epitaxy. J. Alloys Compd. 2013, 580, 82-87.

(18) Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Kim, Y.; Zhang, X.; Guo, Y.; Zou, J.,

Twin-Free Uniform Epitaxial GaAs Nanowires Grown by a Two-Temperature Process.

Nano Lett. 2007, 7, 921-926.

(19) Joyce, H. J.; Qiang, G.; Wong-Leung, J.; Kim, Y.; Tan, H. H.; Jagadish, C., Tailoring

GaAs, InAs, and InGaAs Nanowires for Optoelectronic Device Applications. IEEE J. Sel.

Top. Quantum Electron. 2011, 17, 766-778.

(20) Johansson, J.; Karlsson, L. S.; Dick, K. A.; Bolinsson, J.; Wacaser, B. A.; Deppert, K.;

Samuelson, L., Effects of Supersaturation on the Crystal Structure of Gold Seeded III−V

Nanowires. Cryst. Growth Des. 2009, 9, 766-773.

(21) Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting As Flux in Molecular Beam Epitaxy. J.

Phys. Chem. C 2015, 119, 20721-20727.

(22) Soci, C.; Bao, X.-Y.; Aplin, D. P. R.; Wang, D., A Systematic Study on the Growth of

GaAs Nanowires by Metal−Organic Chemical Vapor Deposition. Nano Lett. 2008, 8, 4275-

4282.

(23) Plante, M. C.; LaPierre, R. R., Analytical Description of the Metal-assisted Growth of

III-V Nanowires: Axial and Radial Growths J. Appl. Phys. 2009, 105, 114304.

123

(24) Zou, J.; Paladugu, M.; Wang, H.; Auchterlonie, G. J.; Guo, Y.-N.; Kim, Y.; Gao, Q.;

Joyce, H. J.; Tan, H. H.; Jagadish, C., Growth Mechanism of Truncated Triangular III–V

Nanowires. Small 2007, 3, 389-393.

(25) Glas, F.; Harmand, J.-C.; Patriarche, G., Why Does Wurtzite Form in Nanowires of III-

V Zinc Blende Semiconductors? Phys. Rev. Lett. 2007, 99, 146101.

(26) Zhang, Z.; Lu, Z.; Xu, H.; Chen, P.; Lu, W.; Zou, J., Structure and Quality Controlled

Growth of InAs Nanowires Through Catalyst Engineering. Nano Res. 2014, 7, 1640-1649.

(27) Soda, M.; Rudolph, A.; Schuh, D.; Zweck, J.; Bougeard, D.; Reiger, E., Transition

from Au to Pseudo-Ga Catalyzed Growth Mode Observed in Gaas Nanowires Grown by

Molecular Beam Epitaxy. Phys. Rev. B 2012, 85, 245450.

(28) Zhang, Z.; Lu, Z.-Y.; Chen, P.-P.; Xu, H.-Y.; Guo, Y.-N.; Liao, Z.-M.; Shi, S.-X.; Lu,

W.; Zou, J., Quality of Epitaxial InAs Nanowires Controlled by Catalyst Size in Molecular

Beam Epitaxy. Appl. Phys. Lett. 2013, 103, 073109.

(29) Krogstrup, P.; Popovitz-Biro, R.; Johnson, E.; Madsen, M. H.; Nygård, J.; Shtrikman,

H., Structural Phase Control in Self-Catalyzed Growth of GaAs Nanowires on Silicon (111).

Nano Lett. 2010, 10, 4475-4482.

(30) López, M.; Nomura, Y., Surface Diffusion Length of Ga Adatoms in Molecular-Beam

Epitaxy on GaAs(100)–(110) Facet Structures. J. Cryst. Growth 1995, 150, 68-72.

(31) Vainorius, N.; Jacobsson, D.; Lehmann, S.; Gustafsson, A.; Dick, K. A.; Samuelson,

L.; Pistol, M.-E., Observation of Type-II Recombination in Single Wurtzite/Zinc-Blende

GaAs Heterojunction Nanowires. Phys. Rev. B 2014, 89, 165423

(32) Heiss, M.; Conesa-Boj, S.; Ren, J.; Tseng, H.-H.; Gali, A.; Rudolph, A.; Uccelli, E.;

Peiró, F.; Morante, J. R.; Schuh, D.; Reiger, E.; Kaxiras, E.; Arbiol, J.; Fontcuberta i

Morral, A., Direct Correlation of Crystal Structure and Optical Properties in Wurtzite/Zinc-

Blende GaAs Nanowire Heterostructures. Phys. Rev. B 2011, 83, 045303.

(33) Lo, S. S.; Mirkovic, T.; Chuang, C.-H.; Burda, C.; Scholes, G. D., Emergent Properties

Resulting from Type-II Band Alignment in Semiconductor Nanoheterostructures. Adv.

Mater. 2011, 23, 180-197.

124

Supporting information

It has been well demonstrated in our previous study1 that Ga concentration in the catalysts

during GaAs nanowires growth is the sum of Ga atoms (NN) in the “neck” and Ga atoms

(NC) in the post-growth catalyst detected by EDS: NGa = NN + NC. Since there is almost no

Ga left in post-catalysts of three samples (Figure 3d-f), NC ~ 0. Therefore, Ga

concentration during nanowires growth is estimated as NN / (NN + NAu) with NAu being the

number of Au atoms in the catalyst.

Since the “neck” region is zinc-blende structured GaAs, which contains 4 Ga atoms

in a unit cell, NN can be obtained by:

NN = 4VN/ , (1)

where is the volume of a unit cell of the zinc-blende structured GaAs, and VN is

the volume of the cylinder shaped “neck” region, which, according to Figure S1, can be

determined by

2H (L being the lateral dimension of the nanowire and H the height

of the “neck” region). Similarly, NAu can also be determined with Equation 1, by replacing

VN with catalyst volume

(L/2 denotes the radius of the hemispheric catalyst, as

illustrated in Figure 1Sb); and with unit cell volume of Au ( ) with face-

centered cubic structure: (aC ≈ 0.41 nm, the lattice parameter).

Figure S1 (a) A HRTEM image showing the top region of a typical GaAs nanowire grown

for 2 h. (b) A schematic model showing the nanowire top containing a hemisphere catalyst,

a cylinder disk-shaped “neck” region and the nanowire section.

125

Figure S2 shows the statistical measurements of GaAs nanowires lengths and

corresponding lateral dimension at nanowires bottom regions with the growth durations. As

can be noted, the average length of 1h grown nanowires is ~1.5 µm, while 1.5h and 2h

grown nanowires have ~3 µm and ~5.5 µm in length respectively, suggesting that

nanowire axial growth increases with the growth duration. The corresponding nanowires

diameters at the bottom regions are ~60 nm, 110 nm and 140 nm, suggesting the

nanowire lateral dimension increases with increasing the growth duration as well.

Figure S2 Plot of nanowires lengths and diameters in the bottom regions with the growth

durations.

To further confirm the impact of growth duration on nanowire growth behavior and phase

purity, we repeated the growth with different growth parameters (growth temperature of

400°C, a thinner Au film deposition (1100°C, 10s), growth duration of 20min and 60min,

respectively). (Figure S3-S5)

Figure S3a,b shows the typical side-view images of the grown sample, suggesting the

similar nanowire density and the major nanowire growth orientation is perpendicular to the

substrate surface. The average length of nanowires grown for 20min is ~500 nm, while

nanowires grown for 0min are as ~2 μm high. Additionally, it can be noted that the

diameter of 20min grown nanowires is uniform along the nanowire (see the inset in Figure

S3a), while nanowires grown for 60min formed the shoulders.

126

Figure S3 (a,b) Typical side-view SEM images of GaAs nanowires grown for 20min and

60min at 400°C. The inset in (a) is corresponding 30°-tilted SEM image.

Figure S4 shows the TEM analyses of the 20min-grown nanowires sample. Figure S4a

shows BF TEM image of a typical grown nanowire and Figure S4b,c is corresponding

HRTEM images of the nanowire top and bottom regions, indicating the defected structure

along the whole nanowire, which is confirmed by the our statistical analysis of this

nanowire sample, as shown by the BF TEM image (Figure S4d) of another typical grown

nanowire and the corresponding HRTEM image (Figure S4e) from the nanowire top.

Figure S4f shows a typical EDS point analysis profile taken from the catalyst and a table

summarizing the catalysts composition containing 5-8 at.% Ga of nanowires shown in

Figure S4a,b.

127

Figure S4 BF (a) and (b,c) HRTEM images of a typical 20min-grown GaAs nanowire and

its top and bottom regions. BF (d) and (e) HRTEM images of another typical nanowire and

its top region. (f) EDS point analysis profile taken from a typical catalyst and the table

show the catalysts composition of (a,d).

Figure S5 shows the TEM analyses of the 60min-grown nanowire sample. Figure S5a is

the BF TEM image of a typical nanowire, which has shoulder morphology. Figure S5b,c is

corresponding HRTEM images of the nanowire top and bottom regions, where the bottom

region is defected as the 20min-grown nanowire is, but the top region is defect-free. This

phenomenon is further confirmed by our statistical analyses on this sample, as typically

depicted in Figure S5d,e. Figure S5f shows a typical EDS point analysis profile taken

from the catalyst and a table summarizing the catalysts composition containing 14-17 at.%

Ga of nanowires shown in Figure S5a,b, which is higher than the Ga concentration in the

catalysts of the 20min grown nanowires.

128

Figure S5 BF (a) and (b,c) HRTEM images of a typical 60min-grown GaAs nanowire and

its top and bottom regions. BF (d) and (e) HRTEM images of another typical nanowire and

its top region. (f) EDS point analysis profile taken from a typical catalyst and the table

show the catalysts composition of (a,d).

References

(1) Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting As Flux in Molecular Beam Epitaxy. J.

Phys. Chem. C 2015, 119, 20721-20727.

129

Chapter 6

Effect of growth time on ternary InGaAs nanowires

6.1 Introduction

In this chapter, the growth of Au-catalyzed ternary InGaAs nanowires on GaAs {111}B

substrates was studied. The nominal In concentration in group-III (In/(In+Ga)) for InGaAs

nanowires growth is 50 at.%. By extensive cross-sectional study of nanowires tip, middle

and bottom regions which have typical varied chemical characteristics, we found that

grown nanowire formed the extraordinary hierarchical structure. By repeating nanowires

growth with shorter growth durations, the growth mechanism of such unique structured

130

nanowires was unveiled. A manuscript has been drafted based on our careful observations

and investigations, and was descried in Section 6.2 of this chapter.

6.2 Unexpected formation of hierarchical structure in ternary InGaAs

nanowires

The optoelectronic application of semiconductor nanowires largely depends on their

nanostructures and related chemical characteristics. Although the spontaneously formed

core-shell structure in ternary nanowires makes them often unpredictable, such versatile

growth with varied chemical characteristics may open up opportunities for widening their

applications. In this study, we present extraordinary phenomena observed during ternary

InGaAs nanowires growth by molecular beam epitaxy. It was unexpectedly found that

nanowires spontaneously formed the hierarchical structure during the growth: pure core

structure at nanowires tip, core-shell structure at middle and core-multishell structure at

bottom regions. By careful electron microscopy investigations on nanowires growth at the

early stages, the growth mechanism of such unique hierarchical structure was unveiled.

This study provides new insights for ternary III-V nanowires growth with the hope of

assistance in designing new nanomaterials and nanostructures for future optoelectronic

devices.

Introduction

One-dimensional semiconductor nanowires, due to their small dimensions and peculiar

properties,1 are the building blocks for nanoelectronic,2 catalysis3 and energy-conversion

devices.4 The rational design of nanowire structures for applications in nanotechnology

requires an improved understanding of their phase behavior in nanoscale systems.

However, the predictable growth of compound nanowires, particularly ternary and more

complex nanowires, is more difficult than their counterparts of elemental Si or Ge

nanowires, due to the complexity of ternary and quaternary phase diagrams.5 In this

regard, systematic and detailed investigations of semiconductor compound nanowires with

a wide range of multicomponent are required to gain a full understanding of one-

dimensional nanowire systems.

131

Ternary III-V nanowires, as an important class of one-dimensional nanomaterials,

have attracted a great attention, due to their band gaps tunability by tuning the

composition of nanowires.6-8 Among them, ternary InGaAs nanowires, with tunable band

gaps ranging from the near-infrared to the infrared region (0.34 1.42 eV),7 can find wide

applications in optoelectronics.9,10 The epitaxial growth of ternary III-V nanowires is

typically catalyzed by Au nanoparticles in MBE or metal-organic chemical vapor deposition

systems, in which the nanowire growth direction and quality can be secured for device

integrations.11,12 However, in these grown systems, complex nanostructures, such as core-

shell structure, can be formed spontaneously during the growth of a wide range of ternary

III-V nanowires, such as In/AlGaAs,6,13,14 GaAsN/P/Sb15-17 and InAsSb18 nanowires,

resulted from phase segregations. The structural and chemical variations in these ternary

nanowire systems have made them unpredictable, which in turn, however, may breed

diverse and exciting opportunities to create new nanomaterials and nanostructures in the

future. Therefore, understanding the versatile growth of ternary III-V nanowires has the

potential to answer fundamental questions about one-dimensional systems, which is

crucial to pursue towards understanding fundamental properties of nanostructures and

designing nanostructured materials with desired properties.

In this study, we present a new class of nanostructures with increasing complexity

formed in ternary InGaAs nanowires. It was found that during the “one-pot” growth,

nanowires spontaneously formed the hierarchical structure respective core, core-shell

and core-multishell structure at the nanowire tip, middle and bottom regions. This unique

configuration enables interfacial states and quantum effects in the nanowires, which would

exhibit peculiar physical properties and could also be obtained via the “step-by-step”

growth.19 With careful investigation through advanced electron microscopy, the growth and

evolution of this complex nanostructure were detailedly unveiled. This study, for the first

time, report the high complexity of nanostructure in ternary III-V nanowires via the “one-

pot” growth, which provides valuable insights in the field of ternary semiconductor

nanowires and possibilities for creating new nanomaterials and nanostructures via an

energy-saving and low-cost approach..

Experimental section

Au-catalyzed InGaAs nanowires were epitaxially grown on GaAs {111}B substrates in a

Riber 32 MBE system. After a GaAs {111}B substrate was degassed and deoxidized to

132

remove any contaminants in the growth chamber, a GaAs buffer layer was grown on the

substrate at 580°C to secure an atomically flat surface for nanowire growth. Subsequently,

the substrate was transferred to the preparation chamber, in which an Au thin film with a

thickness of ~1 nm was deposited on the top of the GaAs buffer layer. The Au-coated

substrate was then transferred back to the growth chamber, and annealed at 550°C for 5

min under As ambient (As source was switched on throughout the nanowire growth), in

which the Au thin film aggregates into nanoparticles.20,21 After annealing, the substrate

temperature was dropped to 400°C, where In and Ga vapor sources were introduced to

induce the nanowire growth, with the same beam equivalent pressure of ~2.2 10-7 Torr

to achieve the nominal In concentration (In/(In+Ga)) of ~50%, and As vapor pressure of

~4.4 10-6 Torr to achieve a V/III (As/(In+Ga)) ratio of ~10. The nanowire growth was

terminated after 60 min and cooled down in the As ambient naturally until 300°C.

The morphology, structure and chemical composition of grown nanowires were

investigated using FE-SEM (JEOL-7800F, operated at 5 kV) and TEM (Philips Tecnai-F20

operated at 200 kV equipped with EDS for compositional analysis), respectively. Individual

nanowires for TEM investigations were prepared by taking an ultrasonic bath in ethanol

and dispersing individual nanowires onto the Cu mesh grids coated with holey carbon

films. The cross-sections of grown nanowires for TEM investigations were prepared by first

embedding the nanowire sample in epoxy resins and then sectioning by an ultramicrotome

(Leica EM UC6). The sectioning was performed from nanowire bottoms to tops, and each

sectioned slice was parallel to the substrate surface. The individual thin slices from

different regions of nanowires were collected and transferred onto the Cu grids for TEM

investigations. On the other hand, to clarify the growth of the planar layer, the cross-

section of nanowires with their substrate was also prepared by using FIB (FEI-SCIOS),

after the grown nanowires were shook off. The prepared lamellar cross-section specimen

was then attached to the Cu lift-out grid for TEM investigations.

Results and discussion

Figure 1a is a top-view SEM image of grown nanowires, showing that most of the

nanowires were grown vertically with respect to their substrate, and its inset showing the

nanowire has the hexagonal cross-section. Figure 1b is the corresponding side-view SEM

image taken along the <110> direction, from which we further verify that the majority of

nanowires are grown perpendicular to the substrate surface, and their average length is ~2

133

µm. Additionally, grown nanowires have the pencil shape, similar to those binary GaAs

nanowires with the wurtzite structure,22-24 caused by their lateral growth,25 and pyramid

bases (marked by arrows) resulted from atoms surface diffusion.9,26 Figure 1c is a BF)TEM

image of a typical nanowire, and Figure 1d is the corresponding TEM image taken from

the nanowire tip, showing the catalyst size of ~25 nm. Figure 1e is a high-resolution HR-

TEM image taken from the middle region of the nanowire and the inset is the

corresponding SAED pattern, indicating that the nanowire has the wurtzite structure with

no observable lattice defects and its growth direction is along the [000 ] direction. Our

extensive TEM investigations from over a dozen nanowires confirmed that our nanowires

are defect-free. The combination of SEM and TEM investigations indicates that the side-

facets of our nanowires are {11 0}. Figure 1f is a high-magnified TEM image taken from

the typical nanowire bottom region, in which the strain contrast can be witnessed,

suggesting the formation of core-shell structure in the nanowire. To further understand the

compositional characteristics of grown nanowires, corresponding EDS analyses were

performed. Figure 1g shows respectively EDS spectra taken from the typical catalyst and

nanowire tip region (refer to Figure 1d), in which no Ga is found in the catalyst while the In

concentration is measured as ~26 at.% (point 1); while the Ga concentration at the

nanowire tip is measured as ~72 at.% of group III elements (point 2), higher than the

nominal Ga concentration (50 at.% of group III elements), which is consistent with previous

study that Au has a stronger affinity with In,13,27-31 leading to Ga being expelled out from

the catalysts and form the Ga-enriched nanowire tip.

To understand the chemical characteristics of our nanowires along their lateral

direction, extensive TEM investigations were further performed on the nanowire cross-

sections taken from different regions. Figure 2a shows a BF TEM image of a typical

nanowire and Figure 2b-d is BF TEM images of typical cross-sections taken from nanowire

tip, middle and bottom regions, respectively (marked in Figure 2a, note the middle region

is located in the pencil body but close to the nanowires tip, refer Figure S1, Supporting

Information). In Figure 2b, the nanowire cross-section has a lateral dimension of ~27 nm,

close to the catalyst size. Furthermore, the relative uniform contrast indicates that no

nanowire shell is found in this region; while both cross-sections at the nanowire middle and

bottom regions have a similar lateral dimension of ~120 nm, indicating existence of

134

Figure 1 Typical (a) top-view and (b) side-view SEM images taken from grown nanowires.

The inset in (a) is the zoom-in top-view SEM image. The arrow in (b) shows the nanowire

base. (c) BF TEM image of a typical nanowire. (d-f) Corresponding TEM/HRTEM images

taken from nanowire tip, middle and bottom regions respectively, as marked in (c). The

inset in (e) is corresponding <11 0> zone-axis SAED pattern. (g) Typical EDS spectra

taken from the catalyst and nanowire, as marked in (d).

nanowire lateral growth. Figure 2e-g shows corresponding EDS line-scan profiles of the

cross-sections taken from the respective regions (note that no As peaks were shown due

to its constant composition in the nanowires). The uniform element distributions along the

line-scan across the cross-section at the nanowire tip confirm the nanowire core structure;

while the In intensity profile over the cross-section from the nanowire middle region

135

suggests a core-shell structure with In-poor core and In-enriched shell (Figure 2f); and

surprisingly, three In peaks are witnessed in the EDS line-scan profile of the cross-section

at the nanowire bottom region, suggesting the In-enriched outer-shell, In-poor inner-shell

and In-enriched core. This extraordinary finding is further verified by EDS elements maps

of the cross-section from this region (Figure 2g). These experimental findings form

different regions are further confirmed by our statistical EDS analyses of over a dozen

Figure 2 (a) BF TEM image of a typical InGaAs nanowire. (b-d) BF TEM images of typical

cross-sections at nanowires tip, middle and bottom regions, respectively, as marked in (a).

(e-g) Corresponding EDS line-scan profile of the cross-sections. (h) Measured In

concentrations at different sections and different regions from the nanowire cross-sections

(based on statistic EDS analyses at varied positions - typically marked in (b-d)). (i)

Schematic illustration of the grown hierarchical structure.

136

cross-sections from each respective region. Figure 2h summaries the In concentrations

measured in different regions and different sections. Accordingly, Figure 2i illustrates the

composition profile found in our InGaAs nanowires. Based on the obtained In

concentrations found in the nanowire cores at different regions, we found that the In

concentration in the nanowire cores decreases (Ga concentration increases) along the

nanowires from bottoms to tops.

To understand how the hierarchical structure forms, particularly the core-multishell

structure, we examine nanowire status in the early stage of the nanowire growth.

Accordingly, we repeated the nanowire growth by shortening the growth duration to 15 min

and keep same for the other growth parameters. Figure 3a shows a side-view SEM image

of nanowires grown for 15 min, in which the nanowires have their average length of ~500

nm. Figure 3b,c is BF TEM images taken from a typical nanowire and a typical cross-

section at the nanowire bottom (mark in Figure 3b), both show nanowire diameter of ~60

nm. Figure 3d is the corresponding EDS line scan and Figure 3e is element maps, both

taken from the nanowire cross-section, indicating a nanowire core-shell structure with In-

enriched core and In-poor shell. This is further verified by our statistical compositional

measurements of over a dozen cross-sections at nanowire bottoms (Figure 3f), in which

the In concentration is measured as 41 ± 3 at.% in nanowire cores while 18 ± 3 at.% in the

nanowire shells. These values are similar to the compositions at the nanowire cores and

their inner-shells at nanowire bottoms grown for the 60 min. Further TEM study was also

performed at cross-sections from nanowire middle regions, showing similar chemical

characteristics (Figure S2, Supporting Information). At the nanowire tips, Figure 3g shows

a corresponding EDS spectrum and indicates the In concentration of ~40 at.% in the

nanowire core. These results suggest that nanowires spontaneously form the core-shell

structure and the cores maintain a constant composition during their first 15 min growth.

Moreover, by comparing the chemical characteristics of the nanowire bottoms in two

growth durations (15 min and 60 min), the nanowire outer-shells found in the nanowires

grown for 60 min must be formed during the late stage of the nanowire growth.

137

Figure 3 Typical (a) side-view SEM and (b) BF TEM images of the InGaAs nanowire

grown for 15 min. (c) BF TEM image of the cross section from the nanowire bottom region

marked in (b). (d) Typical EDS line scan profile, (e) EDS elements maps and (f)

compositional measurements (marked in (c)) of the cross section. (g) Typical EDS

spectrum taken from the nanowire top (mark in (b)).

By correlating the experimental results shown above, we note the following facts.

With short growth duration, nanowires form a core-shell structure and the composition of

the nanowire cores remains as a constant. With increasing the growth duration, (1) In-

enriched outer-shells form in the nanowire bottom regions, resulting in nanowire core-

multishell structure, and (2) the Ga concentration in the nanowire cores increases from

nanowire bottoms to their tops. As a consequence, a nanowire hierarchical structure is

formed.

138

To understand the formation of this extraordinary hierarchical nanowire structure, we

note that, during ternary III-V nanowire growth, their composition can be affected by the

simultaneous two-dimensional layer growth on the substrate.15 In this regard, it is

necessary to clarify the nature of the planar layer directly grown on the substrate.

Accordingly, TEM investigations were performed on the cross-sections of two nanowire

samples to characterize the grown planar layers. Figure 4a,c is cross-section TEM images

taken respectively from the nanowire sample grown for 60 min and 15 min, in which the

planar layer grown for 60 min contains larger convex areas than those in the planar layer

grown for 15 min (see the inset in Figure 4c). By correlating with the SEM images (e.g.

Figure 1b), these convex areas can be assigned as the nanowire bases.26 Figure 4b,d is

the corresponding EDS spectra taken from locations of planar layers (P1, P3) and typical

nanowire bases (P2, P4), showing different In concentrations of 36 ± 2 at.% at P1 and 57

± 2 at.% at P2, and the In concentrations of 33 ± 2 at.% at P3 and 36 ±2 at.% at P4 (refer

the respective insets). This suggests that the In concentrations in the nanowire bases are

higher than those in the planar layers, which is further confirmed by the typical EDS line

scans marked in Figure 4a,c. The comparison of planar layers in two cases indicates that

(1) with the short growth duration, the grown InGaAs planar layer and nanowire bases are

Ga-enriched (compared with the nominal composition of Ga:In = 50:50) which should be

energetically favorable by reducing the lattice mismatch with the GaAs substrates;32 and

(2) with the prolonged growth duration, the nanowire bases grow significantly, and become

comparatively In-enriched InGaAs bases. It has been demonstrated that nanowires base

growth occurs at the early stage of the nanowire growth,26 which is determined mainly by

two factors: mass transport (depending upon atom diffusion on the substrate surface)23

and growth energetics (depending upon energy minimization of the growth).33 It has been

well documented that, for a In(Ga)As layer epitaxially grown on a GaAs substrate, the

misfit strain gradually builds up, which can be eventually relaxed through the generation of

misfit dislocations34,35 or island growth through mass transport.36,37 Therefore, in the case

of growing InGaAs nanowires on GaAs substrates, the nanowire bases are formed through

mass transport and result in the strain relaxation between the planar layer and the

substrate. Due to the lower surface energy of InAs than that of GaAs,38,39 the nanowire

bases can accommodate more In, leading to the In-enriched nanowire bases, particularly

at the later stage of the nanowire growth.

139

Figure 4 (a,c) TEM image of the cross-sections taken from the nanowires samples grown

for 60 min and 15 min, respectively. The insets are typical EDS line scan profiles in (a,c)

and enlarged TEM image of a typical nanowire base in (c). (b,d) Corresponding EDS

spectra and the insets of In concentration measurements taken from the cross-sections as

marked in (a,c).

Accordingly, our extraordinary observations can be explained by answering the

following two questions: (1) Why the core-multishell structure is formed and (2) why the

group III composition of nanowire cores varies, both when the nanowire growth duration is

prolonged.

To answer the first question, we note that the spontaneously formed core-shell

structure in ternary III-V nanowires can be ascribed to the phase segregation, in which

catalysts induce the nanowire core growth via the VLS growth mode, while the nanowire

shell growth results from nanowire lateral growth under the VS growth mode.14,15 It has

been documented that, during III-V nanowire shell growth, their composition depends on

the chance of group-III atoms bonding with group-V atoms on the nanowire sidewalls.40 In

this study, the fact is that the In-poor inner shell at the nanowire bottoms was formed

140

during the early stage of the nanowire growth (15 min), in which the grown nanowire bases

and planar layer are comparatively Ga-rich when compared with the nominal group-III

concentration (Ga:In=50:50). Therefore, excessive Ga atoms on the substrate surface can

diffuse to the nanowire sidewalls, and in turn incorporating with As atoms to form Ga-

enriched nanowire shells. Additionally, since the Ga-As bond energy is stronger than In-

As, the Ga-As bonding should be relatively stable from decomposition.41 On the other

hand, during the later stage of the nanowire growth, the fact of In-enriched nanowire bases

indicates that relatively more In atoms can diffuse from the substrate surface to the

nanowire sidewalls. As a result, the In concentration on the nanowire sidewalls should be

increased, which increases their chances to bond with As atoms. Consequently, more In

atoms can incorporate to the nanowire sidewalls and form the In-enriched outer shells.

To answer the second question, we also note that the In concentration in nanowire

cores are lower than the nominal In concentration (Ga:In = 50:50), which is consistent with

previous observation that Au catalysts prefer to affiliate In while expel Ga after the

supersaturation of group-III concentration in the catalysts is reached, leading to the high

Ga concentration in the nanowire cores.13,30,42,43 Furthermore, the fact that, during the later

stage of the nanowire growth, the growth of In-enriched InGaAs nanowire bases may

result in less In available for the nanowire core growth.15 Due to less In impingement on

the catalysts surface, less In can then incorporate to the catalysts, leading to the

decreased In concentration but increased Ga concentration in catalysts, so as in the

nanowire cores at later stage of the nanowire growth. Therefore, compositional varied

nanowire cores are achieved..

To understand the growth mechanism of our hierarchical InGaAs nanowires, we

further investigate nanowires grown for 5 min to clarify when and why the pencil-shape of

our nanowires are formed. Figure 5a shows side-view SEM images of nanowires grown for

different durations, in which the pencil-shape morphology can be witnessed at early stage

of the nanowire growth (5 min) and such morphology was maintained with prolonging the

growth duration. This suggests that nanowires prefer to maintain the vertical {11 0}

sidewalls (Figure S2, Supporting Information) rather than the tapered morphology caused

by the lateral growth.25 Since the sidewalls of tapered nanowires are no longer to be the

{11 0} sidewalls for the wurtzite structure or the {110} sidewalls for the zinc-blende

structure, the fact that our nanowires remain {11 0} sidewalls suggests that the growth of

the nanowires was controlled by thermodynamics as the {11 0} planes have low surface

141

energy.44 According to our observations of nanowires growth at early stages, we propose a

growth model45 to illustrate the formation of our hierarchical nanowire structure in Figure

5b. It should be noted that, during the early stage of the nanowire growth, nanowires form

an initial core-shell structure along the nanowires (except the tip region below the

catalysts) with an In-enriched core and an In-poor shell. Simultaneously, nanowire bases

and the planar layer were grown due to atom deposition on the substrate surface and

adatom diffusion. With prolonging the growth duration, due to the strain relaxation between

the InGaAs planar layer and the GaAs substrate, adatom diffusion on the substrate

surface increases, particularly for In atoms with higher mobility,9,13 leading to the an

increased In concentration in the nanowire bases. Consequently, more In adatoms can

diffuse to the nanowire sidewalls, resulting in the growth of (1) In-enriched outer shell on

the initially formed core-shell nanowire structure (forming the core-multishell structure at

nanowire bottoms), and (2) the In-enriched shell on the newly grown nanowire core

(forming the core-shell structure at nanowire middle regions). As a result, the observed

hierarchical core-multishell nanowires are formed.

Figure 5 (a) Side-view SEM images of InGaAs nanowires grown with varied durations. (b)

Schematic growth model describes the hierarchical structure formation in grown InGaAs

nanowires with prolonged growth duration.

142

We anticipate that our extraordinary hierarchical structured nanowires have great

potential in versatile optoelectronic applications. Such unique nanostructure can affect

carriers’ transports in the nanowire and thus selectively realize quantum confinement in

individual nanowires. In particular, the core-multishell structure at nanowire bottoms may

confine carriers in the nanowire cores and outer-shells, acting as a quantum nanotube. In

fact, the core-shell nanostructure integrated photo-detector and emitters have shown

prominently amplified spontaneous emissions.46 For this reason, the findings observed in

this study suggest the potential usefulness of our unique hierarchical structured nanowires

as optoelectronic components and pave a pathway for developing new quantum structures

for boarding their applications in advanced devices.

Conclusion

In conclusion, we demonstrated an extraordinary phenomenon of a hierarchical structure

formed in the ternary InGaAs nanowire. By tracing the nanowires growth at early stages,

we found that this complex structure was actually formed during the later stage of the

nanowire growth, in which the composition of InGaAs planar layer that simultaneously

grown on the GaAs substrate was changed, due to the requirements of stain relaxation

between the planar layer and substrate, and the energy minimized growth of the planar

layer. This fact in turn, affects the nanowires composition and induces the hierarchical

structured nanowire. The formation of such extraordinary hierarchical structured nanowire

is expected to show unique optoelectronic properties and thus could lead to wider

applications in nanoscale optoelectronic and quantum communications systems. This

study provides a new insight for ternary III-V nanowires growth and sheds light on the

fundamental understanding extended to the general one-dimensional nanomaterials for

optoelectronic and quantum applications.

Author Information

Chen Zhou,† Kun Zheng,*, ♩, ¶ Ping-Ping Chen,§ Wei Lu,§ and Jin Zou*, †, ♩

†Materials Engineering, ♩Centre for Microscopy and Microanalysis, ¶Australian Institute for

Bioengineering and Nanotechnology, The University of Queensland, St. Lucia,

Queensland 4072, Australia

143

§National Laboratory of Infrared Physics, Shanghai Institute of Technical Physics, Chinese

Academy of Science, 500 Yu-Tian oad, Shanghai 200083, People’s epublic of China

Corresponding Authors

*E-mail: [email protected].

*E-mail: [email protected].

References

1. Bierman, M. J.; Lau, Y. K. A.; Kvit, A. V.; Schmitt, A. L.; Jin, S., Dislocation-Driven

Nanowire Growth and Eshelby Twist. Science 2008, 320, 1060-1063.

2. Huang, Y.; Duan, X.; Wei, Q.; Lieber, C. M., Directed Assembly of One-Dimensional

Nanostructures into Functional Networks. Science 2001, 291, 630-633.

3. Tian, Z. R.; Voigt, J. A.; Liu, J.; McKenzie, B.; McDermott, M. J.; Rodriguez, M. A.;

Konishi, H.; Xu, H., Complex and Oriented ZnO Nanostructures. Nat. Mater. 2003, 2, 821-

826.

4. Luo, J.; Karuturi, S. K.; Liu, L.; Su, L. T.; Tok, A. I. Y.; Fan, H. J., Homogeneous

Photosensitization of Complex TiO2 Nanostructures for Efficient Solar Energy Conversion.

Sci. Rep. 2012, 2, 451.

5. Duan, X.; Lieber, C. M., General Synthesis of Compound Semiconductor Nanowires.

Adv. Mater. 2000, 12, 298-302.

6. Chen, C.; Shehata, S.; Fradin, C.; LaPierre, R.; Couteau, C.; Weihs, G., Self-Directed

Growth of AlGaAs Core-Shell Nanowires for Visible Light Applications. Nano Lett. 2007, 7,

2584-2589.

7. Hou, J. J.; Wang, F.; Han, N.; Xiu, F.; Yip, S.; Fang, M.; Lin, H.; Hung, T. F.; Ho, J. C.,

Stoichiometric Effect on Electrical, Optical, and Structural Properties of Composition-

Tunable InxGa1–XAs Nanowires. ACS Nano 2012, 6, 9320-9325.

8. Jacobsson, D.; Persson, J. M.; Kriegner, D.; Etzelstorfer, T.; Wallentin, J.; Wagner, J.

B.; Stangl, J.; Samuelson, L.; Deppert, K.; Borgström, M. T., Particle-Assisted GaXIn1− XP

Nanowire Growth for Designed Bandgap Structures. Nanotechnology 2012, 23, 245601.

9. Kim, Y.; Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Paladugu, M.; Zou, J.;

Suvorova, A. A., Influence of Nanowire Density on the Shape and Optical Properties of

Ternary InGaAs Nanowires. Nano Lett. 2006, 6, 599-604.

144

10. Treu, J.; Stettner, T.; Watzinger, M.; Morkötter, S.; Döblinger, M.; Matich, S.; Saller, K.;

Bichler, M.; Abstreiter, G.; Finley, J. J.; Stangl, J.; Koblmüller, G., Lattice-Matched

InGaAs–InAlAs Core–Shell Nanowires with Improved Luminescence and Photoresponse

Properties. Nano Lett. 2015, 15, 3533-3540.

11. Svensson, J.; Dey, A. W.; Jacobsson, D.; Wernersson, L.-E., III–V Nanowire

Complementary Metal–Oxide Semiconductor Transistors Monolithically Integrated on Si.

Nano Lett. 2015, 15, 7898-7904.

12. Mårtensson, T.; Svensson, C. P. T.; Wacaser, B. A.; Larsson, M. W.; Seifert, W.;

Deppert, K.; Gustafsson, A.; Wallenberg, L. R.; Samuelson, L., Epitaxial III−V Nanowires

on Silicon. Nano Lett. 2004, 4, 1987-1990.

13. Guo, Y.-N.; Xu, H.-Y.; Auchterlonie, G. J.; Burgess, T.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C.; Shu, H.-B.; Chen, X.-S.; Lu, W.; Kim, Y.; Zou, J., Phase Separation

Induced by Au Catalysts in Ternary InGaAs Nanowires. Nano Lett. 2013, 13, 643-650.

14. Guo, Y.-N.; Burgess, T.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou, J., Polarity-Driven

Nonuniform Composition in InGaAs Nanowires. Nano Lett. 2013, 13, 5085-5089.

15. Sun, W.; Guo, Y.-N.; Xu, H.-Y.; Liao, Z.-M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou, J.,

Unequal P Distribution in Nanowires and the Planar Layer During GaAsP Growth on GaAs

{111}B by Metal–Organic Chemical Vapor Deposition. J. Phys. Chem. C 2013, 117, 19234-

19238.

16. Yuan, X.; Caroff, P.; Wang, F.; Guo, Y.; Wang, Y.; Jackson, H. E.; Smith, L. M.; Tan,

H. H.; Jagadish, C., Antimony Induced {112}A Faceted Triangular GaAs1−XSbx/InP

Core/Shell Nanowires and Their Enhanced Optical Quality. Adv. Funct. Mater. 2015, 25,

5300-5308.

17. Yoshiaki, A.; Masahito, Y.; Fumitaro, I., Growth of Dilute Nitride GaAsN/GaAs

Heterostructure Nanowires on Si Substrates. Nanotechnology 2013, 24, 065601.

18. Potts, H.; Friedl, M.; Amaduzzi, F.; Tang, K.; Tütüncüoglu, G.; Matteini, F.; Alarcon

Lladó, E.; McIntyre, P. C.; Fontcuberta i Morral, A., From Twinning to Pure Zincblende

Catalyst-Free InAs(Sb) Nanowires. Nano Lett. 2016, 16, 637-643.

19. Ng, K. W.; Ko, W. S.; Chen, R.; Lu, F.; Tran, T.-T. D.; Li, K.; Chang-Hasnain, C. J.,

Composition Homogeneity in InGaAs/GaAs Core–Shell Nanopillars Monolithically Grown

on Silicon. ACS Appl. Mater. Interfaces 2014, 6, 16706-16711.

20. Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting As Flux in Molecular Beam Epitaxy. J.

Phys. Chem. C 2015, 119, 20721-20727.

145

21. Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Liao, Z.-M.; Burgess, T.; Lu, H.-F.; Gao, Q.; Tan, H.;

Jagadish, C.; Zou, J., Quantitative Study of GaAs Nanowires Catalyzed by Au Film of

Different Thicknesses. Nanoscale Res. Lett. 2012, 7, 1-6.

22. Zhang, Z.; Shi, S.-X.; Chen, P.-P.; Lu, W.; Zou, J., Influence of Substrate Orientation

on the Structural Quality of GaAs Nanowires in Molecular Beam Epitaxy. Nanotechnology

2015, 26, 255601.

23. Lu, Z. Y.; Chen, P. P.; Liao, Z. M.; Shi, S. X.; Sun, Y.; Li, T. X.; Zhang, Y. H.; Zou, J.;

Lu, W., Impact of Growth Parameters on the Morphology and Microstructure of Epitaxial

GaAs Nanowires Grown by Molecular Beam Epitaxy. J. Alloys Compd. 2013, 580, 82-87.

24. Plante, M. C.; LaPierre, R. R., Au-Assisted Growth of GaAs Nanowires by Gas Source

Molecular Beam Epitaxy: Tapering, Sidewall Faceting and Crystal Structure. J. Cryst.

Growth 2008, 310, 356-363.

25. Zou, J.; Paladugu, M.; Wang, H.; Auchterlonie, G. J.; Guo, Y.-N.; Kim, Y.; Gao, Q.;

Joyce, H. J.; Tan, H. H.; Jagadish, C., Growth Mechanism of Truncated Triangular III–V

Nanowires. Small 2007, 3, 389-393.

26. Liao, Z.-M.; Chen, Z.-G.; Lu, Z.-Y.; Xu, H.-Y.; Guo, Y.-N.; Sun, W.; Zhang, Z.; Yang, L.;

Chen, P.-P.; Lu, W.; Zou, J., Au Impact on GaAs Epitaxial Growth on GaAs (111)B

Substrates in Molecular Beam Epitaxy. Appl. Phys. Lett. 2013, 102, 063106.

27. Paladugu, M.; Zou, J.; Guo, Y.-N.; Zhang, X.; Kim, Y.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C., Nature of Heterointerfaces in GaAs/InAs and InAs/GaAs Axial Nanowire

Heterostructures. Appl. Phys. Lett. 2008, 93, 101911.

28. Dick, K. A.; Bolinsson, J.; Borg, B. M.; Johansson, J., Controlling the Abruptness of

Axial Heterojunctions in III–V Nanowires: Beyond the Reservoir Effect. Nano Lett. 2012, 12,

3200-3206.

29. Borg, B. M.; Dick, K. A.; Ganjipour, B.; Pistol, M.-E.; Wernersson, L.-E.; Thelander, C.,

InAs/GaSb Heterostructure Nanowires for Tunnel Field-Effect Transistors. Nano Lett. 2010,

10, 4080-4085.

30. Krogstrup, P.; Yamasaki, J.; Sørensen, C. B.; Johnson, E.; Wagner, J. B.; Pennington,

R.; Aagesen, M.; Tanaka, N.; Nygård, J., Junctions in Axial III−V Heterostructure

Nanowires Obtained Via an Interchange of Group III Elements. Nano Lett. 2009, 9, 3689-

3693.

31. Messing, M. E.; Wong-Leung, J.; Zanolli, Z.; Joyce, H. J.; Tan, H. H.; Gao, Q.;

Wallenberg, L. R.; Johansson, J.; Jagadish, C., Growth of Straight InAs-on-GaAs

Nanowire Heterostructures. Nano Lett. 2011, 11, 3899-3905.

146

32. Woicik, J. C.; Miyano, K. E.; Pellegrino, J. G.; Shaw, P. S.; Southworth, S. H.; Karlin,

B. A., Strain and Relaxation in Inas and InGaAs Films Grown on GaAs(001). Appl. Phys.

Lett. 1996, 68, 3010-3012.

33. Paladugu, M.; Zou, J.; Guo, Y.-N.; Zhang, X.; Joyce, H. J.; Gao, Q.; Tan, H. H.;

Jagadish, C.; Kim, Y., Formation of Hierarchical InAs Nanoring / GaAs Nanowire

Heterostructures. Angew. Chem. Int. Ed. 2009, 48, 780-783.

34. Zou, J.; Cockayne, D. J. H.; Usher, B. F., Misfit Dislocations and Critical Thickness in

InGaAs/GaAs Heterostructure Systems. J. Appl. Phys. 1993, 73, 619-626.

35. Zou, J.; Chou, C. T.; Cockayne, D. J. H.; Sikorski, A.; Vaughan, M. R., Misfit

Dislocations Lying Along <100> in [001] GaAs/In0.25Ga0.75As/GaAs Quantum Well

Heterostructures. Appl. Phys. Lett. 1994, 65, 1647-1649.

36. Liao, X. Z.; Zou, J.; Cockayne, D. J. H.; Wan, J.; Jiang, Z. M.; Jin, G.; Wang, K. L.,

Alloying, Elemental Enrichment, and Interdiffusion During the Growth of Ge(Si)/Si(001)

Quantum Dots. Phys. Rev. B 2002, 65, 153306.

37. Walther, T.; Cullis, A. G.; Norris, D. J.; Hopkinson, M., Nature of the Stranski-

Krastanow Transition During Epitaxy of InGaAs on GaAs. Phys. Rev. Lett. 2001, 86, 2381-

2384.

38. Messmer, C.; Bilello, J. C., The Surface Energy of Si, GaAs, and GaP. J. Appl. Phys.

1981, 52, 4623-4629.

39. Liu, W.; Zheng, W. T.; Jiang, Q., First-Principles Study of the Surface Energy and

Work Function of III-V Semiconductor Compounds. Phys. Rev. B 2007, 75, 235322.

40. Zhang, Y.; Aagesen, M.; Holm, J. V.; Jørgensen, H. I.; Wu, J.; Liu, H., Self-Catalyzed

GaAsP Nanowires Grown on Silicon Substrates by Solid-Source Molecular Beam Epitaxy.

Nano Lett. 2013, 13, 3897-3902.

41. Shiramine, K.-i.; Itoh, T.; Muto, S., Critical Cluster Size of InAs Quantum Dots Formed

by Stranski–Krastanow Mode. J. Vac. Sci. Technol. B 2004, 22, 642-646.

42. Zhang, X.; Zou, J.; Paladugu, M.; Guo, Y.; Wang, Y.; Kim, Y.; Joyce, H. J.; Gao, Q.;

Tan, H. H.; Jagadish, C., Evolution of Epitaxial InAs Nanowires on GaAs (111)B. Small

2009, 5, 366-369.

43. Paladugu, M.; Zou, J.; Guo, Y.-N.; Auchterlonie, G. J.; Joyce, H. J.; Gao, Q.; Hoe Tan,

H.; Jagadish, C.; Kim, Y., Novel Growth Phenomena Observed in Axial InAs/GaAs

Nanowire Heterostructures. Small 2007, 3, 1873-1877.

44. Leitsmann, R.; Bechstedt, F., Surface Influence on Stability and Structure of Hexagon-

Shaped III-V Semiconductor Nanorods. J. Appl. Phys. 2007, 102, 063528.

147

45. Chen, X. J.; Gayral, B.; Sam-Giao, D.; Bougerol, C.; Durand, C.; Eymery, J., Catalyst-

Free Growth of High-Optical Quality GaN Nanowires by Metal-Organic Vapor Phase

Epitaxy. Appl. Phys. Lett. 2011, 99, 251910.

46. Chen, R.; Ng, K. W.; Ko, W. S.; Parekh, D.; Lu, F.; Tran, T.-T. D.; Li, K.; Chang-

Hasnain, C., Nanophotonic Integrated Circuits from Nanoresonators Grown on Silicon. Nat.

Commun. 2014, 5, 4325.

148

Supporting Information

Figure S1a shows the side-view SEM image of the 60 min grown nanowires and Figure

S1b,c is the TEM images of the cross sections taken from the nanowires middle and

bottom regions, respectively. Since the cross sections were cut from nanowires bottom to

top with a step of ~ 50 nm, the small-size cross sections in Figure S1 indicate that they are

from the nanowires top (with lower height), while the larger-size cross sections indicate

that they are from the nanowire middle regions (as shown in Figure S1a).

Figure S1 (a) Side-view SEM image of the 60 min grown InGaAs nanowires. Typical TEM

images of the cross sections taken from the nanowires (b) middle and (c) bottom regions

marked in (a).

Figure S2a is a typical TEM image of the pencil-shape nanowire top and Figure S2b-d is

typical TEM images of the cross sections taken from the varied regions of the pencil top

with lateral dimensions of ~27 nm, 70 nm and 100 nm respectively (marked in Figure S2a).

Figure S2e-g is corresponding EDS line scan profiles of the cross sections, in which

nanowires keeps the core-shell structure of In-poor core and In-enriched shell until there is

no shell growth. This result is consistent with the core-shell structure observed at

nanowires middle regions (see Figure 2f). Figure S2h-i is corresponding [0001] zone-axis

SAED pattern, indicating that nanowires keep the {11 0} side-facets during the growth.

Figure S2j shows the corresponding composition measurements of the cores composition

taken from the cross sections, indicating that the In concentration ranging from 32 ± 2 at.%,

30 ± 2 at.% to 28 ± 2 at.% with the decreased diameter of the cross section. On the other

149

hand, Figure S2k is a typical TEM image of the 15 min grown nanowire and Figure S2m is

TEM images of the cross sections taken from the nanowires middle and top regions,

respectively. The corresponding EDS elements maps (Figure S2n) confirm that the grown

nanowires form core-shell structure with the In-enriched core and In-poor shell along the

nanowire (expect the region close to the catalyst). Figure S2l is corresponding [0001]

zone-axis SAED patterns, indicating the {11 0} side-facets of the nanowires.

Figure S2 (a) Typical TEM image of the 60 min grown nanowire top. (b-d) Typical TEM

images of the cross sections taken from the pencil-shape top as marked in (a). (e-g)

Corresponding EDS line scan profiles and (h-i) SAED patterns. (j) Compositional

measurements of the cores compositions marked in (b-d). (k) Typical TEM image of the 15

min grown nanowire. (m) Typical TEM images of the cross sections taken from the middle

and top regions marked in (k). Corresponding (n) EDS elements maps and (l) SAED

patterns.

150

Chapter 7

Composition characteristics of ternary InGaAs nanowires

7.1 Introduction

In this chapter, Au-catalzyed ternary InGaAs nanowires were grown with a very high

norminal In concentration (In/(In+Ga)) of 85 at.%. By extensive cross-sectional study of

grown nanowires, it was found that nanowires formed the core-shell structure with In-

enriched cores, and the composition in nanowire cores and shells were varied at varied

nanowires regions. This new phenomenon of composition gradient of nanowire cores and

shells is demonstrated and clarified in Section 7.2 of this chapter, as a drafted manuscript.

151

7.2 Composition gradient ternary InGaAs nanowires caused by strain

relaxation

Understanding the compositional distribution of heteroepitaxially grown ternary nanowires

has been pursued for their controllable growth due to their potential applications with

desired band gaps. In this study, through growing ternary InGaAs nanowires with high In

concentration on GaAs {111}B substrates in molecular beam epitaxy, we demonstrate the

extraordinary core-shell structured ternary InGaAs nanowires with In-rich cores and Ga-

enriched shells. Furthermore, both nanowire cores and shells showed composition

gradient. Our detailed structural and chemical characterizations using electron microscopy

suggest that In-rich nanowire cores are formed due to the Ga-limited growth environment,

caused by the competition with the spontaneous InGaAs planar layer growth on the

substrate that consumes more Ga. Moreover, the composition gradient in the nanowires

cores is a result of strain relaxation with increased shell thickness and the limited Ga

diffusion length with long nanowires. This new finding of compositional gradient ternary

nanowires provides possibilities for continuous band-gap tuning and carrier confinement in

individual nanowires, which may pave a way for novel optoelectronic applications.

Introduction

III-V nanowires have become versatile building blocks for future electronic and

optoelectronic devices due to their unique physical and optical properties.1,2 In particular,

ternary III-V nanowires, although less studied than their binary counterparts, have been

attracting an increasing interest as they allow for the tunability of desired band gap for

different applications by modulating composition fraction of their alloys. Among them,

InGaAs nanowires have demonstrated to be technologically important for a wide range of

applications, attributed to their tunable band gap ranging from the near-infrared region to

the infrared region, which makes them constitute the ideal material system for numerous

optoelectronic devices, such as light-emitting diodes3 and nanolasers.4 In addition, arising

from the tunable band gap, the electric properties of InGaAs nanowires can be modulated

intentionally as well, particularly to maintain superior electron mobility and thus enhance

performance in electronic devices.5 Such band-gap tunability can be realized by tuning the

InGaAs alloy composition. For example, by tuning the In composition x from 0 to 1, the

152

band gap of InxGa1-xAs can be lowered from ~1.42 to 0.34 eV,5 and the electron mobility

can be enhanced from ~1000 to 6000 cm2 V-1 s-1.6 The intrinsically lower band gap of In-

rich InGaAs than Ga-rich InGaAs or GaAs would realize carrier confinement when it forms

heterostructures with other III-V materials of larger band gap7 and its high electron mobility

makes it possible for applications in high electron mobility transistors8 and high-efficiency

photodetectors.9 Therefore, InGaAs nanowires with high In concentration are highly

desirable for performance and application of future nanoscale devices.

On the other hand, the capability of growing uniform ternary III-V nanowires on the

substrate with large lattice mismatch is a prerequisite for practical applications by the

combination of nanowire growth with semiconductor integration techniques. Within this

process, one of the critical factors to evaluate nanowire optoelectronic performance is the

composition control. However, due to the competition of the ternary nanowire and the

planar layer heteroepitaxial growth on the substrate, the composition of the grown

nanowire can be deviated from the nominal composition designs.10 On the other hand, for

Au-catalyzed ternary III-V nanowires grown by the VLS mechanism,11 the unexpected

lateral growth on nanowire sidewalls leads to nanowire tapering and phase

segregation,12,13 which makes the composition of ternary nanowires more complex. For

example, it has been discovered that the core-shell structure forms spontaneously in Au-

catalyzed ternary nanowires because of phase separation, such as AlGaAs,14 GaAsP15

and InGaAs16 nanowires. As for InGaAs ternary nanowires involving two group-III

elements that can be alloyed with Au, the stronger affinity of Au-In than that of Au-Ga

leads to In atoms staying in the catalysts rather than incorporating into the nanowires.17,18

Consequently, Ga-enriched cores and In-enriched shells were normally observed in

InGaAs nanowires grown under low or intermediate In/(In+Ga) ratios.19-21 So far, no study

has been reported for the nanowire formation of the core-shell composition configuration

under very high In/Ga ratios, so that the complete understanding of the InGaAs nanowires

growth mechanisms remains unclear.

In this study, we demonstrate the growth and the compositional characteristics of Au-

catalyzed InGaAs nanowires on GaAs {111}B substrates under a high In/(In+Ga) ratio of

~0.85 in molecular beam epitaxy (MBE). Through detailed electron microscopy

investigations on the grown InGaAs nanowires and their cross sections from different

regions, we found that InGaAs nanowires have the average In concentration of more than

80 at.% and they spontaneously formed In-rich cores and Ga-enriched shells. More

interestingly, we found that, during nanowire growth, the compositional difference between

153

the core and the shell decreased towards nanowires bottom regions and the Ga

concentration in the nanowire core decreases along the nanowire growth direction, which

can be derived from the enhancing strain relaxation with the increasing shell thickness

towards the nanowire bottom and the limited Ga diffusion length along the nanowires. The

fundamental reasons behind these new phenomena were investigated and the growth

mechanism of our ternary InGaAs nanowires was clarified.

Experimental section

Epitaxial InGaAs nanowires were grown on GaAs {111}B substrates using Au as catalysts

in a Riber 32 MBE system. After a GaAs {111}B substrate was degassed and deoxidized to

remove any contaminants in the growth chamber, GaAs buffer layer was grown on the

substrate at 580°C to ensure atomically flat surface for the nanowire growth.

Subsequently, the substrate was transferred to the preparation chamber, where a Au thin

film was deposited directly on the top of the GaAs buffer layer. The Au-coated substrate

was then transferred back to the growth chamber and annealed at 550°C for 5 min under

As ambient (As source was switched on throughout the nanowire growth), in which the Au

thin film aggregates into nanoparticles. After annealing, the substrate temperature was

dropped to 400°C, where In and Ga vapor sources were introduced to induce the nanowire

growth with the respective beam equivalent pressures of ~2.2 10-7 and ~4 10-8 Torr

(so the nominal In concentration (In/(In+Ga)) of ~85%), and a V/III (As/(In+Ga)) ratio of

~15. The nanowire growth was terminated after 60 min and cooled down in the As ambient

naturally until 300°C.

The morphological, structural and chemical characteristics of grown InGaAs

nanowires were investigated using FE-SEM (JEOL 7800F) and TEM (Philips Tecnai F20

equipped EDS for compositional analysis), respectively. Individual nanowires for TEM

investigations were prepared through ultrasonicating the nanowire sample in ethanol and

dispersing individual nanowires onto the Cu mesh grids coated with holey carbon films.

The cross sections from different regions of grown nanowires for TEM investigations were

prepared by first embedding the nanowire sample in epoxy resins and then sectioning by

an ultramicrotome (Leica EM UC6). The nanowire sample was well located to ensure that

the sectioning was initiated from the substrate towards the nanowire top and each

sectioned slice was parallel to the substrate surface. The individual thin slices from

different sections of nanowires were collected and transferred onto the Cu grids for TEM

154

investigations. On the other hand, to clarify the growth of planar layer, the cross section of

nanowire sample with substrate was also prepared for TEM investigation, using FIB (FEI-

SCIOS) with Ga ion source after the sample was ultrasonicated to remove the grown

nanowires and coated with a Pt layer. This lamellar cross-section sample was then

attached to the Cu lift-out grid for TEM investigations.

Results and discussion

Figure 1a is a side-view SEM image of grown InGaAs nanowires, showing that they have a

high density with an average length of ~8 µm. Figure 1b is a BF TEM image taken from a

typical InGaAs nanowire, in which slight tapering caused by nanowire lateral growth can

be witnessed. The insets are the respective SAED pattern and HRTEM image taken from

the nanowire body (along the <11 2 0> zone-axis), indicating that the nanowire has the

high-quality wurtzite crystal structure. Figure 1c-e is corresponding high-magnified TEM

images taken from the nanowire top, middle and bottom regions (as marked in Figure 1b),

respectively. It is of interest to note that Moiré fringes can be observed in both nanowires

middle and bottom regions, suggesting that the grown nanowires might have a core-shell

structure, where the core and shell have different chemical composition and thus different

lattice parameters, leading to the strain contrast.22 To understand the composition

distribution along the nanowires, EDS point analyses were performed along the nanowire

and the results are shown in Figure 1f. As can be noted, no Ga was found in the post-

growth catalyst while the In concentration can be estimated as ~30 at.% (point 1). Since

Au has a stronger affinity with In over Ga,16 and Ga, if there is any in the Au catalysts, can

be expelled from the catalysts during the cooling process, and if so, a necking region

should be formed.23 However, no necking region is seen below the catalyst (refer to Figure

1c), suggesting that no Ga was absorbed in the catalyst, at least in the later stage of the

nanowire growth. Nevertheless, a small amount of Ga can be observed in the nanowire

top region (see point 2), suggesting that Ga atoms indeed participate in the nanowire core

growth, possibly via the triple phase line as As does. On the other hand, it is of interest to

note that the Ga concentration in the nanowire increases towards the bottom region

(points 3 and 4). Our extensive EDS investigations over a dozen of nanowires supports

this trend of compositional variation along nanowires, but exact Ga concentrations

155

Figure 1 Typical (a) side-view SEM image and (b) BF TEM image of the grown InGaAs

nanowire. (c-e) High-magnified TEM images of the nanowire top, middle and bottom

regions as marked by the colored rectangle in (b), respectively. (f) Typical EDS point

analysis profiles taken from the catalyst towards the nanowire bottom region as marked in

(c-e).

measured for different nanowires are slightly different, within the range between 25–30

at.% Ga at the bottom of the nanowires (see statistic analysis below).

Since our nanowires may have a core-shell structure, it is necessary to clarify the

structural and chemical characteristics of nanowire cores and shells at the different

regions. Figure 2a shows a BF TEM image of a typical cross section taken from a

nanowire bottom region, which has a hexagon shape and a lateral dimension of ~80 nm

(equivalent to mark A in Figure 1b). As can be noted, the cross section indeed has the

core-shell structure, in which the core diameter can be found as ~30 nm, similar to the

156

catalyst size, verifying that the growth of the nanowire core was directly induced by the

catalyst under the VLS mechanism.11 On the other hand, the nanowire shell has ~25 nm in

thickness, due to the lateral nanowire growth via the VS mechanism.13 Figure 2b is the

corresponding SAED pattern of the cross section viewed from the [0001] zone-axis,

indicating the {1120} side-facets of the grown nanowires, both for the nanowire core and

the shell. To understand the compositional variations between the core and the shell, EDS

mappings were performed, and Figure 2c,d shows the measured Ga and In EDS maps,

from which the In-enriched core and Ga-enriched shell can be clearly witnessed. Figure 2e

shows EDS point analysis profiles taken from the nanowire core and shell, indicating that

Ga concentration is ~15 at.% in the core while ~25 at.% in the shell. These observations

are verified by our extensive TEM investigations over a dozen of nanowire cross sections

(see detailed statistic analysis later).

Figure 2 (a) BF TEM image of a typical cross section from the InGaAs nanowire

bottom region. The inset is the plot of Ga concentrations at different positions of the

cross section as denoted by the orange circles, which is based on the EDS point

analyses. (b) Corresponding SAED pattern viewed along the [0001] zone-axis. (c,d)

EDS Ga and In maps of the cross section, respectively. (e) Typical EDS point

analysis profiles taken from the nanowire core and the shell, respectively.

Similar TEM characterizations were employed on the cross sections of nanowire

middle regions. Figure 3a is a BF TEM image taken from the cross section from a typical

nanowire middle region (with an overall diameter of ~53 nm – equivalent to mark B in

157

Figure 1b), showing that the nanowire has also maintained the core-shell structure in this

region with the core diameter of ~30 nm but a much thinner shell. Figure 3b is the

corresponding SAED pattern taken from Figure 3a, indicating that the core and the shell

have kept the same {1120} side-facets in the nanowire middle region. Figure 3c,d

illustrates the typical results of Ga and In EDS maps over the cross section respectively,

indicating the In-enriched core and the Ga-enriched shell, where the elements segregation

is similar to that in the nanowire bottom region. Figure 3e shows the typical result of

composition profile taken from the nanowire core and shell, showing that Ga concentration

is ~8 at.% in the core while ~35 at.% in the shell, different from those in the nanowire

bottom region.

Figure 3 (a) BF TEM image of a typical cross section from the InGaAs nanowire

middle region. The inset is the plot of Ga concentrations at different positions of the

cross section as denoted by the orange circles, which is based on the EDS point

analyses. (b) Corresponding [0001] zone-axis SAED pattern. (c,d) The

corresponding EDS Ga and In maps, respectively. (e) Typical EDS point analysis

profiles taken from the nanowire core and shell, respectively.

To further examine the compositional characteristics of the nanowire top region,

cross-sections of nanowire tops were investigated by TEM. Figure 4a shows an HRTEM

image of a hexagonal cross section in the typical nanowire top region with the lateral

dimension of ~30 nm (equivalent to mark C in Figure 1b), which is close to the catalyst

158

Figure 4 (a) BF TEM image of the typical cross section from the InGaAs nanowire top

region. The inset is corresponding EDS line scan profile of the cross section. (b)

Corresponding [0001] zone-axis SAED pattern. (c) Typical EDS point analysis profile of

the cross section marked in (a).

size, implying that there is no shell at the nanowire top region. This can be confirmed by

the corresponding EDS line scan across the cross section, showing the uniform element

distribution (see the inset in Figure 4a, note that only Ga-K and In-L peaks were extracted

in this study, as the As concentration is considered as constant in nanowires). Figure 4b is

the corresponding SAED pattern. By correlating Figure 4a,b, the side-facets of the

nanowire core can be determined as six {1120} facets. Figure 4c shows the

Gaconcentration of ~5 at.%, which is lower than those in the nanowire middle and bottom

regions, and lower than the nominal Ga concentration.

Since the compositions of the nanowire cores and shells are both varied along the

nanowires, to quantify such variations, EDS measurements were carried out on the cross

sections from different regions of over a dozen nanowires. Figure 5a-c shows the

respective schematic models of the cross sections from the nanowire top to bottom

regions, demonstrating the elements segregation in each region. Figure 5d-f is the

histograms showing measured Ga concentrations in nanowire cores and shells for

different nanowires and different regions. As can be seen, in the nanowire top region, the

159

Ga concentration is 4-6 at.% in nanowire cores (much lower than that in the nominal 15

at.%), which is in excellent agreement with the EDS analysis shown above; in the

nanowire middle region, the Ga concentration is 6-8 at.% in nanowire cores and 30-35

at.% in the shells; and in the nanowire bottom region, the Ga concentration is 12-16 at.%

in the cores and 25-30 at.% in the shells.

The statistic results outlined above confirm the In-rich nanowire cores and Ga-

enriched shells and suggest that, along the nanowires from bottom to top, the Ga

concentration decreases in the core while increases in the shell. Since there is more Ga in

the nanowire shells than that in the nanowire cores, the difference of the Ga concentration

Figure 5 (a-c) Schematic illustration of the cross-section models from nanowire top to

bottom regions. (d-e) Corresponding histograms illustrate the Ga concentrations in

nanowire cores and shells of a number of nanowires studied, with the dashed line

indicating the averaged Ga concentrations.

160

between the nanowire cores and shells is decreased towards the nanowire bottom. To

understand these extraordinary observations, two issues need to be addressed: (1) Why

the grown InGaAs nanowires have the In-rich core and the Ga-enriched shell, and (2) why

the axial composition gradient formed in the core and the shell during nanowire growth

with the tendency of the difference of Ga concentration between the nanowire cores and

shells decreasing towards the nanowire bottom.

To address the first issue, we note that, during nanowire growth, the planar layer

growth on the substrate simultaneously takes place, competing with nanowires for the

growing materials. Particularly, for III-V ternary nanowires grown on the binary substrate,

such a competition may impact the consequent compositional distribution, and, in turn,

lead to compositional variations in the grown nanowires.10 Therefore, to understand the

role of planar layer growth that may affect the InGaAs nanowire compositions during their

growth, we investigated the composition of the planar layer grown on the substrate. Figure

6a is a typical BF cross-section TEM image, in which the planar layer grown on the

substrate is shown, in which several “hills” can be recognized and they are corresponding

to the nanowire bottoms. Figure 6b shows details of a “hill” region; and the inset is

corresponding EDS line-scan across the planar layer, indicating the uniformity of the Ga

composition across the grown planar layer. Figure 6c shows the EDS point analyses of

different regions, from which no Ga is found in the Pt coated layer (point 1), the Ga

concentration in the planar layer is determined as ~30 at.% (point 2), and the GaAs

substrate has their atomic ratio of ~1:1 (point 3). The comparison of these three EDS point

analyses suggests no Ga contamination during the preparation of cross-section TEM

specimen, so that our measured Ga concentration in the planar layer reflects the true

value, which is significantly higher than the nominal Ga concentration (~15 at.%) for the

nanowire growth. In fact, it is energetically favorable for the planar layer containing more

Ga (when competing with nanowire growth) as less misfit strain is introduced into the

system.24 Nevertheless, there is still a significant lattice mismatch existed between the

planar layer and the substrate, as verified by the split diffraction spots in the SAED pattern

taken at the interface between the substrate and the planar layer (the inset in Figure 6b).

Consequently, under the high In/Ga ratio growth environment, the enhanced Ga

incorporation in the planar layer growth will cause less Ga for the nanowire growth, and

consequently leads to less Ga to incorporate in the nanowire cores through the triple

phase line. As a result, the Ga concentration found in the nanowire cores is much less

than the nominal concentration. On the other hand, because the nanowire shells are

161

Figure 6 (a) BF TEM image showing the typical cross-section of the nanowire sample. (b)

High-magnified TEM image demonstrating the enlarged area of the cross section. The

insets are respective EDS line profile of the planar layer and the [011] zone-axis SAED

pattern taken from the area marked by the dashed square. (c) EDS point analysis profiles

taken from the cross-section sample as marked in (a).

grown under the VS mechanism, the incorporation of In and Ga into the nanowire sidewall

depends on their mobility and bonding ability with As. Since the bonding energy of In-As is

weaker than that of Ga-As,25 it is energy favorable for Ga atoms to be bonded with As

atoms during the nanowire shell growth. Additionally, since Ga atoms are less mobile than

In atoms and have shorter diffusion length,17 more Ga atoms diffusing from the substrate

and nanowire sidewall would accumulate increasingly towards the nanowire bottom region

and incorporate to the sidewall, leading to the Ga-enriched shell, which is in good

agreement with our observations that Ga concentration decreases along the nanowire

from bottom to top and so does the shell thickness.

Regarding the second issue, it has been found that the In concentration in nanowire

cores increases linearly during InGaP nanowire growth, due to the longer In diffusion

length compared with Ga.26 However, in this study, the Ga concentrations in nanowire

cores decrease slowly from nanowire middle to top regions, suggesting that the shorter Ga

diffusion length may be one of the reasons for the compositional variations in nanowire

162

cores. On the other hand, as can be noted that similar to the planar heteroepitaxy, the

lattice mismatch does exist at the interface between the nanowire cores and shells with

different compositions. It has been well demonstrated that the misfit strain can facilitate the

mass transport through the hetero-interface, and lead to the strain relaxation.27 This fact

can be further enhanced with increasing the temperature and increasing the shell

thickness. For example, it has been demonstrated that, by annealing the grown

Ge/Si multilayer sample at higher temperature, the varied compositions become

uniform due to the strain relaxation between the layers realized through the

increased atoms diffusivity at high temperature.28 On the other hand, as reported for

the InAs/GaAs core-shell nanowire growth, when the GaAs shell thickness exceeds

a critical value, the radial strain relaxation takes place and the residual strain in

nanowires decreases with increasing the shell thickness.29 In this study, with

increasing the shell thickness towards the nanowire bottom that has the longest

annealing time, atoms interdiffusion is increasingly strengthened towards this

region, resulted in the decreased compositional difference between nanowire cores

and shells. This is supported by the fact that the grown nanowire has more

significant phase segregation in the middle region, where less strain has been

released and the composition difference is more significant near the core/shell

interface, confirming that less atoms interdiffusion takes place at the core/shell

interface in the nanowire middle region than that in the nanowire bottom (see the

insets in Figures 2a and 3a).

Based on our experimental results and discussion above, a growth model of InGaAs

nanowires with a high In concentration is proposed. Figure 7a illustrates the nanowire core

growth, where the Ga-enriched planar layer growth causes less Ga to participate nanowire

growth and leads to the nanowire core with the In concentration higher than its nominal.

With further growth, nanowire lateral growth takes place and forms the Ga-enriched shell,

causing the lattice mismatch and thus the misfit strain between nanowire cores and shells.

Such a misfit strain is gradually released to minimize the system energy via the atoms

interdiffusion between nanowire cores and shells (as shown in Figure 7b). With the axial

nanowire growth, nanowire shells grow laterally, so that the atoms interdiffusion is

enhanced towards the nanowire bottom region (see Figure 7c), resulting in the decreased

composition difference between the nanowire cores and the shells. Additionally, the Ga

concentration in the core decreases while the In concentration increases during nanowire

163

Figure 7 Schematic diagram illustrating the growth of core-shell structured InGaAs

nanowires. (a) Growth of nanowire cores and the simultaneous planar layer on the

substrate. (b) Nanowire shell growth and the atoms interdiffusion between the core and

shell (marked by the colored arrows). (c) Nanowire shell thickness increases with further

growth, leading to the enhanced atoms interdiffusion between the core and shell. The long

dashed arrows indicate the atoms diffusion lengths.

growth, in which the longer In diffusion length than Ga should also be taken into

consideration.

It is expected that our grown core-shell structured InGaAs nanowires have potential

in optoelectronic applications in the near-infrared region. The varied composition of core

and shell enable them with varied band gaps, in which the low-bandgap In-enriched

InGaAs core would confine carriers and the large-bandgap Ga-enriched shell would act as

a passivation layer.30 Accordingly, the nanowire high intrinsic surface state density would

be suppressed, which would possibly enhanced the performance of nanowire-based

devices.

164

Conclusions

In conclusion, we demonstrated the composition variations in ternary InGaAs nanowires

grown under the high In/Ga ratio by MBE. Our detailed electron microscopy analysis

confirms the spontaneous formation of core-shell structure in grown nanowires. With the

competition of the epitaxial planar layer growth, nanowire growth is Ga-limited, leading to

the In-rich nanowire cores and Ga-enriched shells. Additionally, with increasing the shell

thickness towards nanowire bottoms, the enhanced strain relaxation leads to the

decreased compositional difference between nanowire cores and shells, and further

resulted in the Ga concentration in the core decreasing during nanowire growth. Such

unique compositional variations in nanowire cores and shells can realize carrier

confinement towards the nanowire core and in the top region with narrow band-gaps. This

study shows that the compositional characteristics of InGaAs nanowires with high In

concentration, and highlight the important role of strain relaxation played in the

compositional distribution of ternary nanowires, which is of great significance for

composition and band-structure engineering in ternary III-V nanowire system.

Author Information

Chen Zhou,† Kun Zheng,*, ♩, ¶ Ping-Ping Chen,§ Wei Lu,§ and Jin Zou*, †, ♩

†Materials Engineering, ♩Centre for Microscopy and Microanalysis, ¶Australian Institute for

Bioengineering and Nanotechnology, The University of Queensland, St. Lucia,

Queensland 4072, Australia

§National Laboratory of Infrared Physics, Shanghai Institute of Technical Physics, Chinese

Academy of Science, 500 Yu-Tian oad, Shanghai 200083, People’s epublic of China

Corresponding Authors

*E-mail: [email protected].

*E-mail: [email protected].

165

References

(1) Gudiksen, M. S.; Lauhon, L. J.; Wang, J.; Smith, D. C.; Lieber, C. M., Growth of

Nanowire Superlattice Structures for Nanoscale Photonics and Electronics. Nature 2002,

415, 617-620.

(2) Xia, H.; Lu, Z.-Y.; Li, T.-X.; Parkinson, P.; Liao, Z.-M.; Liu, F.-H.; Lu, W.; Hu, W.-D.;

Chen, P.-P.; Xu, H.-Y.; Zou, J.; Jagadish, C., Distinct Photocurrent Response of Individual

GaAs Nanowires Induced by N-Type Doping. ACS Nano 2012, 6, 6005-6013.

(3) Dimakis, E.; Jahn, U.; Ramsteiner, M.; Tahraoui, A.; Grandal, J.; Kong, X.; Marquardt,

O.; Trampert, A.; Riechert, H.; Geelhaar, L., Coaxial Multishell (In,Ga)As/GaAs Nanowires

for near-Infrared Emission on Si Substrates. Nano Lett. 2014, 14, 2604-2609.

(4) Chen, R.; Tran, T.-T. D.; Ng, K. W.; Ko, W. S.; Chuang, L. C.; Sedgwick, F. G.; Chang-

Hasnain, C., Nanolasers Grown on Silicon. Nat Photon 2011, 5, 170-175.

(5) Hou, J. J.; Wang, F.; Han, N.; Xiu, F.; Yip, S.; Fang, M.; Lin, H.; Hung, T. F.; Ho, J. C.,

Stoichiometric Effect on Electrical, Optical, and Structural Properties of Composition-

Tunable InxGa1–xAs Nanowires. ACS Nano 2012, 6, 9320-9325.

(6) Hannah, J. J.; Callum, J. D.; Qiang, G.; Tan, H. H.; Chennupati, J.; James, L.-H.;

Laura, M. H.; Michael, B. J., Electronic Properties of GaAs, InAs and InP Nanowires

Studied by Terahertz Spectroscopy. Nanotechnology 2013, 24, 214006.

(7) Treu, J.; Stettner, T.; Watzinger, M.; Morkötter, S.; Döblinger, M.; Matich, S.; Saller, K.;

Bichler, M.; Abstreiter, G.; Finley, J. J.; Stangl, J.; Koblmüller, G., Lattice-Matched InGaAs-

InAlAs Core-Shell Nanowires with Improved Luminescence and Photoresponse

Properties. Nano Lett. 2015, 15, 3533-3540.

(8) Dayeh, S. A.; Aplin, D. P. R.; Zhou, X.; Yu, P. K. L.; Yu, E. T.; Wang, D., High Electron

Mobility InAs Nanowire Field-Effect Transistors. Small 2007, 3, 326-332.

(9) Kim, Y.; Joyce, H. J.; Gao, Q.; Tan, H. H.; Jagadish, C.; Paladugu, M.; Zou, J.;

Suvorova, A. A., Influence of Nanowire Density on the Shape and Optical Properties of

Ternary InGaAs Nanowires. Nano Lett. 2006, 6, 599-604.

(10) Sun, W.; Guo, Y.-N.; Xu, H.-Y.; Liao, Z.-M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou, J.,

Unequal P Distribution in Nanowires and the Planar Layer During GaAsP Growth on GaAs

{111}B by Metal-Organic Chemical Vapor Deposition. J. Phys. Chem. C 2013, 117, 19234-

19238.

(11) Wagner, R. S.; Ellis, W. C., Vapor-Liquid-Solid Mechanism of Single Crystal Growth.

Appl. Phys. Lett. 1964, 4, 89-90.

166

(12) Givargizov, E. I., Fundamental Aspects of VLS Growth. J. Cryst. Growth 1975, 31, 20-

30.

(13) Zou, J.; Paladugu, M.; Wang, H.; Auchterlonie, G. J.; Guo, Y.-N.; Kim, Y.; Gao, Q.;

Joyce, H. J.; Tan, H. H.; Jagadish, C., Growth Mechanism of Truncated Triangular III-V

Nanowires. Small 2007, 3, 389-393.

(14) Lim, S. K.; Tambe, M. J.; Brewster, M. M.; Gradečak, S., Controlled Growth of

Ternary Alloy Nanowires Using Metalorganic Chemical Vapor Deposition. Nano Lett. 2008,

8, 1386-1392.

(15) Sun, W.; Huang, Y.; Guo, Y.; Liao, Z. M.; Gao, Q.; Tan, H. H.; Jagadish, C.; Liao, X.

Z.; Zou, J., Spontaneous Formation of Core-Shell GaAsP Nanowires and Their Enhanced

Electrical Conductivity. J. Mater. Chem. C 2015, 3, 1745-1750.

(16) Guo, Y.-N.; Xu, H.-Y.; Auchterlonie, G. J.; Burgess, T.; Joyce, H. J.; Gao, Q.; Tan, H.

H.; Jagadish, C.; Shu, H.-B.; Chen, X.-S.; Lu, W.; Kim, Y.; Zou, J., Phase Separation

Induced by Au Catalysts in Ternary InGaAs Nanowires. Nano Lett. 2013, 13, 643-650.

(17) Krogstrup, P.; Yamasaki, J.; Sørensen, C. B.; Johnson, E.; Wagner, J. B.;

Pennington, R.; Aagesen, M.; Tanaka, N.; Nygård, J., Junctions in Axial III-V

Heterostructure Nanowires Obtained Via an Interchange of Group III Elements. Nano Lett.

2009, 9, 3689-3693.

(18) Dick, K. A.; Bolinsson, J.; Borg, B. M.; Johansson, J., Controlling the Abruptness of

Axial Heterojunctions in III-V Nanowires: Beyond the Reservoir Effect. Nano Lett. 2012,

12, 3200-3206.

(19) Guo, Y.-N.; Burgess, T.; Gao, Q.; Tan, H. H.; Jagadish, C.; Zou, J., Polarity-Driven

Nonuniform Composition in InGaAs Nanowires. Nano Lett. 2013, 13, 5085-5089.

(20) Jung, C. S.; Kim, H. S.; Jung, G. B.; Gong, K. J.; Cho, Y. J.; Jang, S. Y.; Kim, C. H.;

Lee, C.-W.; Park, J., Composition and Phase Tuned InGaAs Alloy Nanowires. J. Phys.

Chem. C 2011, 115, 7843-7850.

(21) Wu, J.; Borg, B. M.; Jacobsson, D.; Dick, K. A.; Wernersson, L.-E., Control of

Composition and Morphology in InGaAs Nanowires Grown by Metalorganic Vapor Phase

Epitaxy. J. Cryst. Growth 2013, 383, 158-165.

(22) Paladugu, M.; Zou, J.; Guo, Y. N.; Zhang, X.; Joyce, H. J.; Gao, Q.; Tan, H. H.;

Jagadish, C.; Kim, Y., Evolution of Wurtzite Structured GaAs Shells around InAs Nanowire

Cores. Nanoscale Res. Lett. 2009, 4, 846-849.

167

(23) Zhou, C.; Zheng, K.; Lu, Z.; Zhang, Z.; Liao, Z.; Chen, P.; Lu, W.; Zou, J., Quality

Control of GaAs Nanowire Structures by Limiting as Flux in Molecular Beam Epitaxy. J.

Phys. Chem. C 2015, 119, 20721-20727.

(24) Zhang, X.; Zou, J.; Paladugu, M.; Guo, Y.; Wang, Y.; Kim, Y.; Joyce, H. J.; Gao, Q.;

Tan, H. H.; Jagadish, C., Evolution of Epitaxial InAs Nanowires on GaAs (111)B. Small

2009, 5, 366-369.

(25) Sato, M.; Horikoshi, Y., Effect of Indium Replacement by Gallium on the Energy Gaps

of InAs/GaAs Thin-Layer Structures. J. Appl. Phys. 1991, 69, 7697-7702.

(26) Fakhr, A.; Haddara, Y. M.; LaPierre, R. R., Dependence of InGaP Nanowire

Morphology and Structure on Molecular Beam Epitaxy Growth Conditions.

Nanotechnology 2010, 21, 165601.

(27) Liao, X. Z.; Zou, J.; Cockayne, D. J. H.; Wan, J.; Jiang, Z. M.; Jin, G.; Wang, K. L.,

Alloying, Elemental Enrichment, and Interdiffusion During the Growth of Ge(Si)/Si(001)

Quantum Dots. Phys. Rev. B 2002, 65, 153306.

(28) Liao, X. Z.; Zou, J.; Cockayne, D. J. H.; Wan, J.; Jiang, Z. M.; Jin, G.; Wang, K. L.,

Annealing Effects on the Microstructure of Ge/Si(001) Quantum Dots. Appl. Phys. Lett.

2001, 79, 1258-1260.

(29) Kavanagh, K. L.; Saveliev, I.; Blumin, M.; Swadener, G.; Ruda, H. E., Faster Radial

Strain Relaxation in InAs-GaAs Core-Shell Heterowires. J. Appl. Phys. 2012, 111, 044301.

(30) Treu, J.; Stettner, T.; Watzinger, M.; Morkötter, S.; Döblinger, M.; Matich, S.; Saller, K.;

Bichler, M.; Abstreiter, G.; Finley, J. J.; Stangl, J.; Koblmüller, G., Lattice-Matched

InGaAs–InAlAs Core–Shell Nanowires with Improved Luminescence and Photoresponse

Properties. Nano Lett. 2015, 15, 3533-3540.

168

Chapter 8

Conclusions and recommendations

8.1 Conclusions

In this PhD thesis, we demonstrate our new findings in the study of binary GaAs nanowires

and ternary InGaAs nanowires growth. By extensive investigations on the morphological,

strutrual and chemical characterizations of grown nanowires, the growth mechanism

behind these new finding is unveiled.

169

In the binary GaAs nanowires growth, we found that:

By tuning the As flux in the growth environment, the growth rate and structural

quality of GaAs nanowires can be modulated. Under the high As flux, nanowires

gowth is Ga-limited, in which thin nanowires grow longer and have defect-free

wurtzite structure, while thick nanowires are short and have defected structure.

In this case, nanowires growth is kinetics controlled. Under the low As flux,

nanowires growth is As-limited, in which nanowires are short and their structure

are defect-free. In this case, nanowires growth is thermodynamics controlled.

Growth duration can play a critical role in the nanowires growth rate and

structural quality. With short growth duration, nanowires diameters are uniform

and planar defects are randomly distributed along the nanowires. With

prolonged growth duration, nanowires grow much longer and form the shoulder

morphology, and their structure becomes defect-free at top regions. This

phemonon can ascribed to the increased Ga supersaturation in the catalysts

resulted from increased Ga diffusion from the shoulder surface.

In the ternary InGaAs nanowires growth, we found that:

Growth duration can also play a role in the core-shell strucutrue formation of

ternary nanowires. With the nominal In concentration of 50 at.% in the group-III,

nanowires formed the core-multishell structure at their bottoms, which was

formed during the later stage of the nanowires growth. This phenomenon was

attributed to the composition variation of planar layer on the substrate, in which

the strain relaxation between the InGaAs layer and GaAs substrate occurs.

With the nominal In concentration of 85 at.% in the group-III, nanowires formed

the core-shell structure, with In-rich cores and Ga-enriched shells. The

composition in nanowire cores and shells were varied at varied from nanowire

bottoms to tops. This composition gradient can be attributed to the atoms

interdiffusion during nanowires growth.

170

8.2 Recommendations

In this PhD thesis, the growth behaviour and strcutral quality of GaAs nanowires have

been studied by varying the group-V flux and the growth duration, which provides new

insights for binary III-V nanowires growth. Particularly, nanowires formed half defected and

half defect-free sections with the prolonged growth duration, in which the related

optoelectronic properties can be varied at varied nanowire sections.

On the other hand, the structural and compositional characterisics of ternary InGaAs

nanowires grown with varied In concentrations have been demonstrated. Based on this

study, a more systematic study of tenary InGaAs nanowires should be achieved. The

careful investigations of phase segregation induced core-shell or core-multishell structure

formed in ternary InGaAs nanowires provides a new understanding of complex one-

dimensional nanomaterials system. Besides, it is interesting to invesitugate the

correspongding optoelectronic properties of these unique nanowire structures.