Heterointerfaces in Semiconductor Nanowires

22
Nanowire heterointerfaces DOI: 10.1002/smll.200800556 Heterointerfaces in Semiconductor Nanowires Ritesh Agarwal* reviews R. Agarwal Examples of advanced nanowire heterostructures and their characterization. From the Contents 1. Introduction . . . 1873 2. Heterostructure NW Synthesis, Characterization, and Assembly . 1874 3. Semiconductor-NW– Metal-Contact Interfacial Properties .... 1875 4. NW–Dielectric Interfacial Properties .... 1878 5. Crossed-NW Devices ...... 1881 6. NW–Substrate Interface for Functional Devices ...... 1883 7. Axial Heterostructures 1884 8. Radial Heterostructures 1888 9. Conclusions . . . 1890 Keywords: electronics heterostructures interfaces nanowires photonics 1872 ß 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Transcript of Heterointerfaces in Semiconductor Nanowires

reviews R. Agarwal

1872

Nanowire heterointerfaces

DOI: 10.1002/smll.200800556

Heterointerfaces in Semiconductor NanowiresRitesh Agarwal*

Examples of advanced nanowire heterostructures and their characterization.

From the Contents

1. Introduction . . . 1873

2. Heterostructure

NW Synthesis,

Characterization,

and Assembly . 1874

3. Semiconductor-NW–

Metal-Contact

Interfacial

Properties . . . . 1875

4. NW–Dielectric

Interfacial

Properties . . . . 1878

5. Crossed-NW

Devices . . . . . . 1881

6. NW–Substrate

Interface for

Functional

Devices . . . . . . 1883

7. Axial

Heterostructures1884

8. Radial

Heterostructures1888

9. Conclusions . . . 1890

Keywords:� electronics

� heterostructures

� interfaces

� nanowires

� photonics

� 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Semiconductor nanowires have attracted considerable recent interest due to

their unique properties, including their highly anisotropic geometry, large

surface-to-volume ratio, and carrier and photon confinement. Currently,

tremendous efforts are devoted to the rational synthesis of advanced

nanowire heterostructures. Yet, if functional devices are to be made from

these materials, precise control over their composition, structure,

morphology, and dopant concentration must be achieved. Their funda-

mental properties must also be carefully investigated since the presence of a

large surface and interfacial area in nanowires can profoundly alter their

performance. In this article, the progress, promise, and challenges in the area

of nanowire heterostructured materials are reviewed, with particular

emphasis on the effect of different types of heterointerfaces on device

properties.

1. Introduction

The aggressive scaling down of devices to sub-100-nm

length scales remains the cornerstone for creating integrated

systems with enhanced functionality. The highly prolific

microelectronic industry has successfully utilized the top-

down approach for such device scaling,[1] which involves using

photolithographic and subsequent etching techniques to

reduce the device size, and has tremendously impacted almost

all aspects of our lives. However, it is becoming increasingly

difficult to continue this scaling trend due to the fundamental

limitations of photolithography, which uses tightly focused

optical radiation to transfer the electronic device design on a

planar substrate. This limitation comes from the classical

optical phenomena of diffraction, which limits the size of the

optical beam that can be focused on to small spot size, making

the goal of fabricating devices with arbitrarily small feature

sizes rather difficult. Possible approaches to reduce the critical

size of devices includes the use of smaller-wavelength

radiation (extreme-UV), higher-numerical-aperture lenses,

and aggressive etching; all of these are promising to continue

the scaling trends in the semiconductor industry but involve

prohibitively expensive procedures that may not be able to

maintain the lower costs of electronic devices that have

become essential commodities in the modern world.

One promising approach to overcome the fundamental

challenges inherent with the top-down fabrication method is

the bottom-up self-assembly technique, which mostly utilizes

chemical techniques for arranging and manipulating matter at

the atomic scale with exquisite control over the spatial

arrangement of atoms to form novel structures. The bottom-

up self-assembly technique is a unique approach for device

[�] Prof. R. Agarwal

Department of Materials Science and Engineering

University of Pennsylvania

Philadelphia, PA 19104 (USA)

E-mail: [email protected]

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

fabrication as the functionally of the device originates during

the growth of the structures and not from lithography, thereby

enabling functional structures that can be arbitrarily small

ranging from small molecules to mesoscopic structures such as

nanocrystals, nanotubes (NTs), and nanowires (NWs) in the

sub-100-nm length scale.[2] Impressive progress has beenmade

in the past decade in the area of nanostructured materials and

it is becoming increasingly clear that such non-traditional

approaches for assembling devices will be critical for achieving

the predicted metrics of next-generation computing, as well as

opening new opportunities across different scientific disci-

plines[2,3] by providing a methodology for scaling of devices to

the molecular scale with minimal surface roughness. Since the

bottom-up approach presents a new paradigm for assembling

functional devices, it also provides new challenges that need to

be overcome in order to make it a viable option for future

technologies. These challenges include generatingmethods for

rational synthesis of nanostructures with controlled properties

with minimum dispersion, understanding the novel physical

phenomena in quantum- or finite-sized structures, their

integration into devices, and finally the most challenging

requirement of developing new large-scale assembly tech-

niques for arranging these nanostructures into hierarchically

ordered and fully functional structures. Only when these

challenges are overcome would the bottom-up technology be

able to address interesting problems related to electronics,

photonics, information technology, medicine, sensors, and

diagnostics. This scheme for growing nanoscale components

and then assembling devices from the ‘‘bottom up’’ is also very

promising to overcome amajor obstacle faced by conventional

planar technology, where the ‘‘top down’’ nature of the

fabrication process imposes tremendous constraints on the

integration of vastly different materials. For example, even a

modest goal of integrating electronics and photonics on a

single chip is difficult in top-down processes owing to the

incompatibility of direct-bandgap semiconductors with Si.

However, since the nanoscale components can be synthesized

and then co-assembled on a common platform in separate

H & Co. KGaA, Weinheim www.small-journal.com 1873

reviews R. Agarwal

Ritesh Agarwal earned his under-

graduate degree from the Indian

Institute of Technology, Kanpur in

1996, and a master’s degree in

chemistry from the University of

Chicago in 1997. He received his PhD in

physical chemistry from the University

of California at Berkeley in 2001. After

completing his PhD, he was a post-

doctoral fellow at Harvard where he

studied the optical and photonic

properties of semiconductor nano-

wires. His work led to the development

of electrically driven single-nanowire

lasers and avalanche photodiodes. He is currently an assistant

professor in the Department of Materials Science and Engineering

at the University of Pennsylvania. His research interests include

size-dependent structural, optical, and electronic properties,

quantum-confined optics in nanowire heterostructures, and

studying phase transitions and electronic memory switching at the

nanoscale.

1874

processes, it opens up exciting opportunities to create complex

systems with multifunctional components with electronic,

optical, and magnetic properties.

Semiconductor NWs and NTs offer a versatile approach

for the bottom-up assembly of electronic and photonic devices

with the potential for integration of non-silicon-based

photonics with silicon electronics.[3–13] The unique geometries

of NWs and NTs enable them to function as both active device

elements and interconnects, which can be utilized to achieve

highly integrated device architectures. The inability to control

the chirality and hence the electronic properties of nanotubes

during synthesis and the difficulties associated with manip-

ulating individual NTs presents a significant challenge in

developing NT-based integrated devices. By contrast, the

ability to rationally synthesize NWs with tunable and

modulated chemical composition, size, structure, and mor-

phology and to accurately dope them with both p- and n-type

dopants has opened up opportunities for assembling almost

any kind of functional nanosystem ranging from photonics and

electronics to biological sensors.

One defining attribute of nanostructured devices is the

presence and significance of interfaces, which becomes

particularly important at sub-100-nm size scales. The role of

interfaces and interfacial states becomes all the more

important for NW heterostructures with electrical contacts.

An example of different types of interfaces in a typical core/

shell heterostructured NW transistor device is shown in

Figure 1. The different interfaces include the heterointerface

between the two semiconducting materials (e.g., Si and Ge),

semiconductor–gate-oxide interface, semiconductor–metal

contact interface, and semiconductor–vacuum interface, which

are in intimate contact with the NW, in addition to the gate-

oxide–gate-electrode interface not in direct contact with the

NW. The unique interplay of the properties of all these

interfaces has a profound effect in determining the properties

of the functioning NW device. For example, the size-

dependent band-edge alignments between the two semicon-

ductors, nature of the interface, crystallinity, doping levels,

defects, and other properties determine the unique properties

of just this one heterojunction interface! Similarly, the metal–

Figure 1. A schematic image of a core/shell nanowire transistor device

assembled on a dielectric substrate indicating the different hetero

interfaces. The interfaces include semiconductor-NW–bottom-substrate

interface, semiconductor-NW–metal contacts, core/shell interface

between different semiconductors, NW–gate-dielectric interface

(conformal or deposited on top), and a nondirect contact interface

between dielectric and gate metal.

www.small-journal.com � 2008 Wiley-VCH Verlag Gm

semiconductor contact will be influenced by the metal work

function, semiconductor-electron affinity or ionization poten-

tial, and also critically on the nature of surface states that can

lead to the pinning of the Fermi level at the contact region,

rendering contact properties independent of the type of metal

used. In this article, we will review the significant progress

made in the field of semiconductor NW devices, with emphasis

on the unique properties that emerge due to the presence of

heterojunctions and interfaces in different geometries and

their effect on the device properties.

2. Heterostructure NW Synthesis, Characterization,and Assembly

Semiconductor NWs are typically synthesized by the

vapor–liquid–solid (VLS) growthmechanism, in which ametal

nanoparticle functions as a catalyst for one-dimensional (1D),

anisotropic growth.[5,10,14–16] The gas-phase reactants mix with

the catalyst, which is kept at elevated temperatures to form a

supersaturated melt, at which point the reactants precipitate

via nucleation, and subsequent axial elongation of a crystalline

NW occurs. Single-crystalline NWs can be synthesized with

excellent control over diameter, chemical composition, and

dopant concentration from practically any semiconducting

material with diameters ranging from a few nanometers to

200 nm and with typical aspect ratios of 100–1000.

Manipulation of matter at the nanoscale should lead to

fascinating and novel behavior and should also impact device

properties. In analogy to planar semiconductor technology

where impressive advances have been made towards control-

ling heterostructures for various applications,[17] NW hetero-

structures hold tremendous promise for nanoelectronics and

photonics applications.[10] The ability to fabricate intrawire

heterostructures with well-defined crystalline interfaces

bH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

should greatly increase the versatility of NW-based photonic

devices and also reduce the load for subsequent assembly

processes with increased functionality within NWs.[18] Such a

synthetic control will be extremely useful for creating precisely

defined systems to investigate the effects of confinement on

properties of nanostructures resulting from themodification of

the electronic structure and density of states.

The unique geometry of NWs allows three types of

structurally coherent, epitaxially grown heterostructures:

1) axial, 2) radial (core/shell), and 3) branched heterostruc-

tures[19–22] (Figure 2). To synthesize axial heterostructures, the

addition of the first reactant is stopped during growth and a

second reactant is introduced, which grows along the NW axis.

To grow radial heterostructures, the VLS growth is minimized

and the second reactant grows epitaxially on the NW core.

Branched heterostructures are typically synthesized by

depositing the metal nanocatalysts after the growth of NW

backbones, which then can seed the growth of higher-

generation branches, preferably epitaxially from the first-

generation NWs. In addition to the epitaxially grown NW

heterostructures, it is also possible to fabricate functional

heterojunctions between two crossed NWs and also between

NWs assembled on a clean semiconductor substrate, both

formed due to van der Waals or other physical interactions.

Assembly of crossed-NW junctions can be reliably formed by

microfluidic-based flow,[25] electric field,[26] optical twee-

zers,[27] or Langmuir–Blodgett-based[28] alignment tech-

niques. The heterojunction formed between two crossed

NWs is unique because of the extremely small area of contact,

its ease of fabrication, and the ability to combine different

Figure 2. Nanowire heterostructures. An illustration of NW axial,

branched, and radial heterostructure growth schemes. Transmission

electron microscopy (TEM) images of InAs/InP axial [23] and Si/Ge

radial [24] heterostructures, and scanning electron microscopy (SEM)

images of CdS/ZnS NW branched heterostructures. [19] To grow single-

crystalline NW branched heterostructures, the main NW is seeded with

Au catalysts and the branches are grown from another material.

Reproduced with permission from References [19, 23]. Copyright 2007

and 2002, respectively, American Chemical Society.

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

materials. On the other hand, formation of junctions between

NWs assembled on planar semiconductor substrates leads to a

much larger contact area.

3. Semiconductor-NW–Metal-Contact InterfacialProperties

All semiconductor NW devices need to be electrically

connected by metallization processes usually achieved by

conventional lithography or direct contact writing with a

focused-ion beam (FIB). Broadly, two types of junction are

formed between semiconductor–metal interfaces: rectifying

Schottky barriers and non-rectifying Ohmic contacts. In

general, Schottky barriers are formed when the metal work

function is larger (smaller) than n (p)-type semiconductors,

while non-rectifying Ohmic contacts are formed for the

reverse case. Therefore, metals with high work functions

should form the best contacts to p-type semiconductors while

those with low work functions form the best contacts to n-type

semiconductors. However for real systems, the termination of

the semiconductor material at the surface leads to the

formation of surface states due to incomplete covalent bonds

and other effects. These surface states lie in the otherwise

forbidden electronic bandgap and are generally localized in

nature. In addition to the surface states originating from the

semiconductor, metal–semiconductor contact regions also

typically have thin interfacial layers, which may contain

oxides, clusters of semiconductor or metal phases, and other

alloys, all of which modulate the barrier to electron–hole

transport through the junction. For example, if the density of

surface states in the bandgap region is high, then it can pin the

Fermi level at some fixed position that is largely independent

of the metal used. The Schottky-barrier height is now

determined from the pinning of the Fermi level instead of

the metal work function.

Due to the increased surface-to-volume ratio in semicon-

ductor NWs, the properties of NW–metal contacts can be very

interesting on one hand but can also pose challenges to make

reliable contacts with desired properties. For most devices, one

would want to havemetal contacts to NWs that are Ohmic with

low resistance, high stability, reproducibility, and reliability.

Therefore, it is important to understand the properties of

electrical contacts between semiconductor NWs and different

metals, which can then be used to understand the complex

transport properties of these interesting systems. Many initial

studies of electrical transport of single semiconducting NWs

typically showed very high resistances for a large variety of

NWs with different metals, mostly attributed to poor contacts

to NWs. Even though such problems still persist for many

systems, progress has been made in the past few years to

control and systematically understand the nature of contacts

and the transport properties ofNWs.Of the large body of work,

the notable examples are the study of contacts to NWs

fabricated by FIB and the subsequent study of the micro-

structure and composition at the contact area and its influence

on the transport properties.[29] We will use this example to

illustrate the interplay of different interfacial properties in

determining the nature of electrical contacts to NWs.

H & Co. KGaA, Weinheim www.small-journal.com 1875

reviews R. Agarwal

1876

Nam et al. reported a detailed study of FIB-based Pt

contacts to GaNNWs of different diameters that were directly

written by the decomposition of an organometallic precursor

by using a Ga-ion beam.[29] Even though a Schottky barrier of

height 1.5 eV is expected with n-type GaN under ideal

conditions, a much lower barrier height of 0.18 eV was

observed for larger-diameter NWs, while Ohmic contacts were

observed for NWs of �60-nm diameter (Figure 3). Transport

through large-diameter (>150 nm) wires was explained with

back-to-back Schottky barriers, while the variable-range

hopping mechanism was responsible for transport in small-

diameter NWs at all temperatures (Figure 3). The resistance of

the back-to-back Schottky-diode device (large-diameter NWs)

originates from the reverse-bias contact, which upon break-

down leads to conduction due to thermionic emission.

However, carrier transport through the barrier can also be

facilitated by tunneling via the interface states or pure

tunneling through the barrier. In general, thermionic emission

over a barrier follows a linear ln(I) versus V1/4 relationship at

all temperatures, while applied-bias-dependent effective

barrier height can be determined from activation energy plots

(slope of ln(I/T2) versus 1/kT ). In addition, the presence of

tunneling due to the presence of interface states increases the

measured current and in effect lowers the Schottky barrier

Figure 3. Two-probe I–V characteristics measured from 40 to

300 K on FIB Pt-contacted GaN NW (a–c) and a model fit (d). a) For a

184-nm-diameter NW, I–V is nonlinear and the low-bias resistance

increases rapidly with decreasing T. b) For a 66-nm-diameter NW, I–V is

linear, the low-bias resistance also increases rapidly with decreasing T,

and the apparent zero-bias resistivity at 300 K is two decades smaller

(0.06 V cm) than that for the 184-nm NW (6.4V cm). c) Log(I) versus

V1/4 plot for the 184-nm-diameter NW; the linear behavior at and

above 160 K is consistent with formation of back-to-back Schottky

barrier contacts. d) Conversely, the T-dependent behavior of the

66-nm-diameter NW suggests 2D variable-range hopping. Inset: model

fit based on 2D Mott variable-range hopping. Reproduced with

permission from Reference [29]. Copyright 2005, American Chemical

Society.

www.small-journal.com � 2008 Wiley-VCH Verlag Gm

height measured from the procedure described above. Using

this analysis, Nam et al. showed that the transport mechanism

of large-diameter GaN NWs contacted with Pt can be

explained by thermionic emission over the barrier; however,

the barrier height was estimated to be 0.18 eV (low bias), much

lower than the ideal value of 1.5 eV. The reason for the large

lowering of the Schottky barrier was attributed to the

formation of interfacial states due to nitrogen vacancies

originating from the direct electrode-writing process with a

high-energy FIB, which pins the Fermi level and also provides

an effective tunneling pathway. Interestingly, the formation of

Ohmic contacts to small-diameter NWs (60 nm) was attributed

to the ion-beam-induced formation of amorphous (disor-

dered) GaN, which assumes significance for smaller-diameter

NWs. The conduction at the contact region was observed to

originate from variable-range hopping of carriers from highly

localized states following Ohmic conduction. These results

were also corroborated through careful TEM microscopy of

the contact region, which clearly showed predominance of

amorphized and compositionally inhomogeneous GaN at the

contact region observed to be similar for all NWs but

important for small-diameter NWs (Figure 4).[30] Similar

results were observed by Hernandez et al. for Pt contacts via a

FIB technique on SnO2 NWs, which also showed significant

lowering of the observed Schottky barrier due to the formation

of the interfacial layer.[31]

Kim et al. also studied the electrical properties of metal

contacts fabricated on GaN NWs and reported the formation

of Schottky barriers with Al and Ohmic contacts with Ti/Au

metal, both deposited by thermal evaporation.[32] They too

found evidence of formation of an insulating interfacial layer

at the contact region leading to non-negligible formation of

built-in potential, causing a significant shift in the forward-bias

threshold voltage. In a related study on GaN NWs contacted

with Ti/Al metal without annealing, Koley et al. observed both

Schottky- and Ohmic-contact formation by measurements

performed over many devices.[33] In their study, the reported

Schottky contacts were formed only on one of the two contacts

producing asymmetric current–voltage (I–V) behavior. The

variation in contact properties was attributed to the non-

uniform surface barrier, which was also confirmed by scanning

probe measurements. The effect of different lithographic

techniques on contact properties for GaN NW/Ni systems was

studied by Stern et al.[34] and it was observed that electron-

beam (e-beam) lithography tends to damage the surface of the

NW, leading to back-to-back Schottky diodes, while optical

lithography, which does not modify the surface, produces

Ohmic contacts. However, post annealing of the e-beam

lithographically fabricated devices converted the Schottky

contacts to Ohmic contacts. In another intriguing report by

Lao et al., asymmetric contacts were observedwhenZnONWs

were aligned on pre-fabricated bottom Au electrodes by

dielectrophoresis.[35] The asymmetry in contacts was described

by the asymmetric heating of the electrodes generated by the

AC signal during the alignment process; NW surface at the

higher-temperature electrode was believed to be degraded

slightly to produce Schottky barriers, while the other relatively

lower-temperature electrode produced Ohmic contacts due to

formation of oxygen vacancies, which is known to enhance the

bH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Figure 4. a) Bright-field scanning TEM cross-sectional images of two

points along a nanowire device. The pristine GaN nanowire is �70 nm

wide and the e-beam-deposited Pt layer is atop the nanowire; dark Pt

nanocrystals were embedded in a bright amorphous matrix. b) The

contact region. The top �30 nm of the nanowire was sputtered away

during FIB–Pt contact formation. The FIB–Pt contact contains a higher

density of Pt crystallites and appears darker than the e-beam-deposited

Pt layer. c) High resolution (HR) TEM image obtained from the region

within the dashed box in (a). The interface between the GaN nanowire

and the e-beam-deposited Pt capping layer is marked with a dashed

line. Two boxes (b,c) are marked. d) Fast Fourier transform (FFT)

diffractogram of the GaN region box labeled b. e) FFT diffractogram of the

GaN region in the box labeled c. Sharp spots in both diffractograms

show the high crystallinity of the nanowire all the way out to the

interface. f) HRTEM image obtained from the region within the dashed

box in Figure 1b. An amorphous GaN region (a-GaN) is marked between

dashed lines. Two boxes (e,f) are marked. f) FFT diffractogram of the

a-GaN region in box e, showing a diffuse-ring characteristic of

amorphous material. g) FFT diffractogram of the GaN region in box

(f), with diffuse disks, indicative of a disordered crystalline lattice.

Reproduced with permission from Reference [30].

conductivity. Bulk n-type InAs is an interesting case as it is

known to form Ohmic contacts with most metals due to the

pinning of its Fermi level above the conduction band to form

an accumulation layer of electrons. Similar Ohmic contact

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

properties were reported by Suyatin et al. with n-type InAs

NWs attributed to Fermi-level pinning.[36] Suyatin et al. also

studied the effect of sulfur passivation of InAs NW contact

properties and found that surface passivation further lowers

the contact resistance, believed to be due to the removal of

native oxides and the formation of covalent bonds with sulfur

with the surface atoms.[36]

The formation of Schottky diodes between Si NW and

annealed Ni contacts has also been studied carefully by Ahn

et al. with spatially resolved photocurrent experiments, in an

field-effect transistor (FET) geometry.[37] Large photocurrents

were observed at the source and drain contacts even at zero

appliedbias,whichwasattributedtobandbending,while littleor

nophotocurrentwas observed from the center of thewire due to

flat-bandpotential.The spatial dependenceofphotocurrentwas

further investigatedatdifferentgate-anddrain–sourcebiases, to

measure the bending steepness of the band of the Si NW at the

contacts (Figure 5).[37] From thesemeasurements, the Schottky-

barrier height (0.57 eV, close to bulk value) could also be

extracted. In another report by Gu et al., CdS NWs contacted

with Ti were probed by photocurrent microscopy techniques,

which showed enhanced photoconductivity at the contact

region, implying the formation of Schottky barriers.[38]

It is well known from bulk systems that annealing of

metallic contacts with Si forms silicides, which typically lower

the contact resistance and form contacts with well-defined

chemical composition. It is believed that similar silicidation

occurs for Si NWs. In an interesting experiment, Wu et al.

reported solid-state transformation of Si NWs to single-

crystalline NiSi, which was also used to form axial hetero-

structures between NiSi and Si NWs with near atomically

abrupt interfaces.[39] Formation of well-defined contacts

having the same length scale as the NW itself, with clean

interfaces and well-defined chemical composition, is essential

to understand the nature of nanoscale contacts and also to

develop a true bottom-up paradigm for device fabrication that

can lead to highly integrated device architectures. Electrical

measurements of NiSi/Si NW heterostructures showed linear

behavior at room temperature but with a small Schottky

behavior at lower temperatures.[39] Recently, progress has

been made to control the length of the Si NW segment by

longitudinal formation of NiSi from both ends of the NW,

thereby forming lithography-free FETs with extremely small

channel lengths (<100 nm).[40] Such precise control of the

active area of the FETNW channel with excellent control over

the contact properties allows very detailed understanding of

transport properties of sub-lithographic structures assembled

in unique geometries.[41] A systematic study of the effect of

annealing of Ni contacts on Si NWs has been recently reported

by Byon et al., which showed improvement of device

properties followed by conversion of p-type NW FETs to

ambipolar behavior.[42] In a two-step annealing process, the

first at 200 8C, essential to prevent excessive silicide formation,

produces Ni2Si, followed by a second annealing step at higher

temperatures to produce NiSi. The device properties

improved during the second annealing step in the 400–

500 8C range, while higher annealing temperatures degraded

the performance due to agglomeration of the contact metal.

The transition from unipolar p-channel to inversion was

H & Co. KGaA, Weinheim www.small-journal.com 1877

reviews R. Agarwal

Figure 5. a) A scanning photocurrent image overlaid on a confocal

reflection image simultaneously taken at V¼VG¼ 0 with a light

intensity of 100 kW cm�2. The image size is 20�50mm. The direction

of the positive current is indicated in the upper inset. Red (blue)

corresponds to a positive (negative) current. b) An enlarged photo-

current image from part (a) with energy-band diagrams describing the

mechanism of the Schottky photocurrent generation. Reproduced with

permission from Reference [37]. Copyright 2005, American Chemical

Society.

Figure 6. PL spectra of a single nanowire before (gray curve) and after

(blue curve) photoetching with defocused laser light (442 W cm�2) for

25 min in a 0.1% HF/5 g L�1 TOPO/butanol solution. For comparison,

the normalized PL spectrum of a macroscopic InP:Se crystal is plotted

(black curve). b,c) Dark-field optical and PL images of this nanowire,

respectively. Reproduced with permission from Reference [56]. Copy-

right 2005, American Chemical Society.

1878

attributed to the improvement of contacts by annealing, which

lowers the barrier heights for majority (holes) and minority

(electrons) carriers.[42] In another study of undoped Ge/Se

core/shell NWs by Xiang et al.,[43] Coulomb-blockade

behavior was observed from unannealed contacts due to the

Schottky barrier, while ballistic transport through the device

was achieved after annealing the contacts.

The ability to synthesize NWs with few defects coupled

with the formation of transparent contacts has also enabled

tunable supercurrents being observed through semiconductor

NWs.[44] Low-resistance superconducting Al electrodes on

InAs NWs were demonstrated to produce proximity-induced

supercurrents below 1K.[44] Supercurrent reversal in quantum

dots created by local gating of InAs NWs contacted by

superconducting Al electrodes was also achieved.[45] In a

recent paper, hole-carrier gas formed at the interface of a Si/

Ge core/shell NW structure was used to produce supercurrents

with Al electrodes.[46] All of these reports clearly demonstrate

the unique properties that can be achieved by using novel

nanostructures in combination with top-down fabrication

technology and open up exciting new opportunities to tune the

physical behavior of systems by chemical composition, size,

and geometry control along with controlling the contact

properties by choosing the right metal and processing

conditions.

4. NW–Dielectric Interfacial Properties

The presence of a large surface in NWs influences its

electronic and optical properties and the semiconductor–

dielectric interface becomes one of the most important

interfaces that includes interaction of the NW with air–

vacuum and all kinds of dielectrics. Therefore, the under-

www.small-journal.com � 2008 Wiley-VCH Verlag Gm

standing and control of the semiconductor–dielectric interface

becomes crucial in fabricating devices with the desired

electronic/optical response.

Photoluminescence (PL) data collected from direct-

bandgap NWs reveal a wealth of information about the

optical and electronic properties such as band-edge emission,

trap states, radiative efficiency, and carrier and photon

confinement, which are related to the surface properties. A

variety of NWs made from materials including ZnO,[47,48]

ZnS,[49] GaN,[50] CdS,[51] CdSe,[52]ZnSe,[53] InP,[54–57]

GaAs,[58] and InN[59] have been extensively studied by PL

spectroscopy. High-quality NWs typically emit brightly with a

spectral peak near the band edge. The width of the PL spectra

reveals critical information about the quality of the NWs with

narrow widths (�20–30 nm) corresponding to high optical

quality of NWs mostly due to better surface properties, and

fewer shallow impurities.

The systematic study of the influence of a NW’s surface on

its optical properties has until recently been a relatively

neglected research area.[60,61] The combination of a large

number of defect sites and high mobility of charge carriers and

surface charges[62,63] leads to the unfortunate situation that a

single surface-defect site can quench the PL of the wire over

distances in the micrometer range.[64] Efforts to (photo)-

chemically passivate the surface of NWs, analogous to

quantum-dot surface passivation,[65,66] have been mainly

focused on InP NWs. Chemisorption of oxygen on the surface

of InP NWs leads to the formation of sub-bandgap states,

creating non-radiative recombination pathways.[67] The PL

yield of InP NWs with diameters in the 50–100-nm range could

successfully be enhanced using a photochemical treatment

with HF and TOP/TOPO ligands[55,56] (Figure 6), although a

more recent study shows that this approach does not work for

InP NWs with diameters in the quantum-confinement

regime.[68] For ZnO NWs, it has been shown that addition

of a capping layer of SnO2 enhances the UV band-edge

emission.[69] Additionally, several studies have also shown the

influence of surface charges on PL peak positions.[56,63]

Whereas for light-emitting applications the influence of the

NW surface on the luminescent properties can be a

bH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

disadvantage, it can be advantageous for utilizing NWs as

optical sensors. In a study on the effect of adsorption of gasses

on the PL of ZnSe NWs, it was found that H2 promoted the

band-edge emission whereas Ar and N2 lowered its intensity.

These changes proved reversible, leading to the proposition of

using these NWs as gas sensors.[70] Likewise, the PL fromZnO

NWs could be reversibly quenched by adsorption of NO2 on its

surface[71] and Cu2þ ions in aqueous solution proved to

specifically quench the PL from CdTe NWs.[72] While the

aforementioned optical NW sensors rely on a global intensity

effect as a detection signal, a more advanced sensing scheme

allows for spectroscopy of molecules in the vicinity of the NW

surface. When light is guided through a NW, a large fraction of

the electromagnetic field travels outside of the NW as

evanescent field. Using this evanescent field, only molecules

present in the surface-interface region are sampled, leading to

high spatial selectivity. It was shown using SnO2 NWs with

evanescent field, absorption, and surface-enhanced Raman

spectroscopy (SERS) are feasible, promising to further reduce

analyte detection limits.[73–75]

Interface states have pronounced effects on the nature and

efficacy of charge-carrier transport in FETs, an effect more

significant for nanostructures.[76] The high density of surface

states can lead to Fermi-level pinning, thereby depleting the

carriers from near the surface, an effect with profound

consequences for small-diameter NWs.[60,77] Due to the

presence of surface states and charges, the efficacy of the

gate control becomes smaller, leads to surface recombination

of carriers, and also causes hysteresis due to charge–discharge

transients from the surface states. Interface traps are produced

due to a variety of reasons including interruption of the

periodic lattice structure at the surface, broken bonds,

impurities, vacancies, and non-stoichiometric compounds,

which can trap charges. Calculations of the electronic states

of donor and acceptor impurities in semiconductor NWs

revealed that their electronic structure strongly depends on

their dielectric surrounding and on the diameter of the

wires.[78] It was predicted that the ionization energy of the

impurities in NWs is enhanced with respect to the bulk

material when the wires are embedded in a low-dielectric-

constant material, while for very small diameter NWs

(<10 nm), the ionization of impurities at 300K is reduced

and heavy doping would be necessary to obtain large

conductivity.[78] For a Si/SiO2 interface, one of the most

important interfaces for metal–oxide semiconductor field-

effect transistor (MOSFET) devices, one can distinguish

between charges located deep inside the oxide (oxide-trapped

charges and mobile ionic charges) and charges located at the

interface (fixed oxide charges and interface trap charges).[79]

The fixed oxide charges at the interface do not interact with Si

through charge transfer and therefore do not depend on the

Fermi level. On the other hand, the interface trap charges

interact with Si via charge transfer and depend on the density

of the interface states. Schmidt et al. have recently discussed

the influence of the Si/SiO2 interface on the charge-carrier

density of n-Si NWs by using simple theoretical models.[80]

Their model predicted that a NW can be fully depleted when

its radius is smaller than their calculated critical radius, which

depends on the trap-level density and doping density. For

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

example, according to their model a 50-nm Si NW is expected

to get fully depleted assuming a moderate trap density of

1011 eV�1 cm�2 if the doping is below 3� 1016 cm�3.[80]

Depletion of carriers in Si NWs would occur due to the

trapping of the free carriers at the Si/SiO2 interface and should

be avoided in order to design efficient devices. Schmidt et al.

also derived a numerical expression for the carrier density in

the presence of trap states and fixed oxide charges and their

model provides important insights regarding the effect of the

Si/SiO2 interface on transport properties of NWs, with the

most important conclusion that the carrier density in Si NWs is

in general lower than the doping concentration.[80]

Wang et al. performed a 3D quantum-transport simulation

of small (sub-10 nm) Si NWs to study the effect of Si/SiO2

interface roughness on device characteristics.[81] Their simula-

tion results predicted that the surface scattering reduces the

electron density of states in the channel, which increases the

threshold voltage. In general, the surface-roughness scattering

was found to be less detrimental in NWs that are in the

quantum-transport regime due to the availability of fewer

propagating modes in comparison to planar MOSFETs with

many occupied modes.[81] Paul et al. did a comparative study

of the impact of process variability inherent in any

semiconductor processing on the performance of NW, NT,

and MOSFET devices.[82] They observed that NTs and NWs

were least affected by almost every process and geometry-

related parameter variation such as oxide thickness and gate

width, while the MOSFET devices (45- and 32-nm technol-

ogies) were dramatically affected. They attributed the

resilience of nanostructure devices on their unique cylindrical

electrostatic geometry and novel device structures.[82]

Typically, capacitance or conductance measurements are

used to measure the interface-trap density.[79] The conduc-

tance measurements are known to provide more accurate

results while the capacitance measurements provide a rapid

estimate of the flat-band potential and the interface-trap

charge. However, the capacitance of NWs is extremely small

due to their small size, which makes capacitance–voltage

(C–V) measurements very difficult, although recent advances

in measurement technology have made such experiments

possible.[83] Dayeh et al. studied the effect of surface states on

the transport properties of InAs NW transistors with ZrO2/

Y2O3 gate dielectric by steady and time-resolved I–V

measurements.[84] InAs, as mentioned previously, forms

Ohmic contacts due to Fermi-level pinning in the conduction

band caused by donor-type surface states and surface

reconstruction, and is therefore a good system to study the

effect of surface states on transport properties. Transport

characteristics of InAs NWs were found to be heavily

dependent on the gate-voltage sweep rate, while the time-

resolved data showed surface trapping and detrapping time

scales of 45 seconds. Slower gate-voltage sweep leads to a

charge-neutral interface with reduced capacitance and higher

transconductance with small hysteresis, all of which are

desired properties. The surface trap density was estimated to

be �1012 cm�2 and the importance of passivating the surface

was clearly identified.[85] The effect of surface traps on

transport properties of ZnO NW transistors were also studied

carefully by Xiong et al. by measuring the random telegraph

H & Co. KGaA, Weinheim www.small-journal.com 1879

reviews R. Agarwal

Figure 7. a) Raw drain current random telegraph signals for a time interval of 300 s observed

in a ZnO nanowire FET at 4.2 K as a function of gate bias. The drain bias is kept constant at 2 V.

b) Histograms of the time-domain RTS data. The large and small peaks represent the empty or

filled trap states, respectively. c) Band diagram for back gate voltage at 9 V with two near

interface oxide (border) traps. Reproduced with permission from Reference [86]. Copyright

2007, American Institute of Physics.

1880

signals (RTS) for defect characterization at low temperatures

(Figure 7).[86] It is known from work done on MOSFETs that

the presence of a large number of trapping and detrapping

processes with a broad distribution of timescales produces a

1/f-type noise spectra, while in submicrometer devices the

individual traps assume importance and produce discrete

switching events and RTS.[87] For the case of n-ZnO NW

transistors assembled on a 100-nm-thick SiO2 layer with the

bottom-gate electrode, Xiong et al. observed 1/f-type noise

characteristics at room temperature while at lower tempera-

tures the noise spectra changed to Lorentzian type with

current traces displaying RTS attributed to correlated carrier

number and mobility fluctuations due to trapping and

detrapping of carriers.[86] Three-level switching events were

also observed, therefore allowing the determination of two

distinct trap states close to the Fermi level in the ZnO NW

channel. This study demonstrated the importance of quantify-

ing the noise spectra to understand the energetics and spatial

location of interface states and their effects on the transport

characteristics of NW devices, an area that still requires

detailed research.

It is expected that high-k gate dielectric materials should

increase the coupling of the gate electrode with the NW,

thereby enhancing the transistor performance. Gnani et al.

simulated the mobility and on-currents for cylindrical NWs

with SiO2 and HfO2 gate dielectrics of equal thickness and

predicted that HfO2 leads to a slight degradation of the short-

channel effect compared with the SiO2 gate but can give an

improved on-current due to lateral capacitive-coupling

effects.[88] Dai and co-workers were the first to report Ge

NW transistors with HfO2 gate dielectric deposited by atomic-

layer deposition;[89] however, the performance was degraded

in comparison to SiO2, which they attributed to interfacial

traps. Recently, the Lieber group has demonstrated superior

transistor performance by using conformal ZrO2 gate

dielectric in comparison to SiO2.[43] In addition, some

attention has been paid to other gate materials in thin films,

www.small-journal.com � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinhe

such as Nb2O5, Dy2O3,[90–94] although

research on the effects of these novel

dielectrics on NWs is still lacking.

The unique cylindrical geometry of

NWs can be used to create a surrounding

gate structure that can give much better

transistor properties such as improved

subthreshold behavior and higher currents

through better electrostatic gate control

through higher capacitance. It has been

experimentally demonstrated that even

conformal top gate structure can elicit

improved response fromNW transistors.[43]

Dai and co-workers have recently demon-

strated that completely surround-gated

NW transistors exhibit higher capacitance

and better electrostatic gate control than

top- and bottom-gated devices.[83] Recently

there has been interest in configuring

vertical-surround-gate NW FETs as it

allows in addition to superior electrical

properties the potential to assemble high-

density devices by utilizing the sub-lithographic dimensions of

NWs (diameters).[95–98] Even though the research in this area

is still emerging, the results are promising and will be of

continuing interest to the scientific community.

The interface between inorganic nanowires with organic or

polymeric systems is also important for fundamental under-

standing and novel applications. Combining two different

classes of material with the idea of using superior electronic

properties of inorganics with the processing versatility of

organics is exciting but needs considerable scientific attention

for developing techniques to assemble these hybrids and study

the novel interfacial properties. For example, single ZnONW-

FETs fabricated using self-assembled organic gate dielectric

displayed outstanding device-performance metrics demon-

strating enhancement-mode ZnO NW-FETs operating at sub-

1V with exceptionally high on/off current ratios.[99] It was

observed that ozone-treated ZnO NW-FETs consistently

retained the enhanced device-performance metrics after SiO2

passivation. Two interesting polymer/nanowire systems cur-

rently under investigation for photovoltaic applications are

ZnO/P3HT/PCBM[100–105] and CdS(CdSe)/P3HT[106–113] with

a staggered band alignment. In ZnO systems[102,103] the proper

blending of P3HT and PCBM was speculated to decrease the

path that excitons travel before being split, thereby decreasing

their chance of recombination and improving the performance

of the device. Developments in CdS and CdSe/polymer

nanocomposites have centered around the role of ligands on

the semiconductor wire in the development of photovoltaic

devices. Specifically, the predominant message in the current

literature indicates that polymer/nanowire composites that

include pyridine in the solution processing will create devices

with improved fill factor and conversion efficiency.[106–112]

Wang et al.[106] studied MEH-PPV/CdS nanowire systems and

found that adding pyridine to their chlorobenzene solutions

increased short-circuit current in the resulting devices by a

factor of 6 over systems without pyridine. They speculate that

pyridine improves the dispersion of CdS in solution and thus

im small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

increases charge transfer and exciton dissociation, as seen by

improved PL quenching over devices prepared without

pyridine. Complimentary research done by Huynh and co-

workers[107] suggests thermal treatment may drive off pyridine

adsorbed on the surface of the nanowire, thus creating a more

efficient path for exciton splitting. The nature of the

conduction mechanism at the interface and in the nanocom-

posite in general has been investigated by several groups. Xi

et al.[109] grew CdS nanowires within PANI sheaths by

electrodeposition in order to study the charge-transfer

mechanism between polymer and CdS. Intensity of the PL

of the core/shell composite was seen to increase by 4.5 times.

They concluded that the increase in PL could not be explained

by the Forster or Dexter transfer mechanisms and that the

actual mechanism of PL enhancement is the proper alignment

of band edges in PANI and CdS, allowing for efficient

transport of photogenerated electrons to the CdS core and

thereby increasing its PL. Yu et al.[112] and Lai et al.[113] have

shown tunable light-emitting diode (LED) and FET nanowire/

polymer systems. Overall, research in nanowire/polymer

composites has shown promise in creating viable LED and

photovoltaic (PV) devices. However, the current research

underscores the importance of blend homogeneity and

processing conditions on the interface in these devices in

order to eliminate potential sites for undesirable recombina-

tion.

Control of NW surface properties is therefore extremely

important for all kinds of devices. For example, for NW

chemical/biological sensors based on FETs the surface has to

be appropriately functionalized with chemical or biological

receptors to achieve high selectivity and sensitivity for analyte

detection, which in turn requires knowledge about the surface

properties.[114,115] The polar surfaces of ZnO NWs are critical

for the generation of piezoelectric effects, which are getting

attention currently.[116,117] In a recently reported electric-field-

induced reversible crystalline-to-amorphous phase change in

chalcogenide NWs[118–120] the interaction of the NW with its

surroundings determines the efficiency of the phase transition.

The surrounding material has to have large heat capacity and

thermal conductivity in order to rapidly cool the NW

undergoing the phase change. As of now no detailed attention

has been devoted to the effect of NW surface and encapsula-

tion on the nature and efficiency of phase transition, an

important consideration for fundamental studies of nanoscale

heat generation and propagation and also for fabrication of

next-generation phase-change memory devices.

5. Crossed-NW Devices

Cross-bar technology from top-town lithographically

defined structures has been utilized for memory switches in

conventional electronics. A similar concept can also be

applied by creating architectures of crossed NWs, resulting

in truly nanoscale junctions at each crossing, which can

have unique encoded functionalities including transistors,

diodes, memory, address decoder, or tunneling junctions. The

device functionality, which originates from lithography-free

nanoscale junctions, can be as small as a few hundred atoms

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

and can be further enhanced by combining NWs of different

compositions and doping profiles. One of the first demonstra-

tions of crossed NW structure was reported by Heath and co-

workers,[121] where they assembled two Si NWs to fabricate a

four-terminal device. The resistivity through the crossed-wire

junction was measured to be similar to the resistivity of

individual wires forming the crossed device, indicating little or

no tunneling barrier at the junction of the two wires,

suggesting the absence of SiO2 at the junction. Thus, these

two crossed wires functioned as a single conducting unit.

The Lieber group has developed an impressive set of tools

to reliably manipulate NWs to assemble them into crossed

architectures[25,26] and has also demonstrated unique device

functionalities originating from combining NWs from a variety

of materials. One of the first devices reported by the Lieber

group was a p–n-junction diode from Si NWs, with controlled

dopant concentrations.[122,123] Subsequently, LEDs from p-

and n-doped crossed InP NWs (Figure 8) were reported[26]

and, more recently, from n-GaN/p-GaN,[125] which demon-

strated that it is possible to produce light from such nanoscale

junctions. Electrical measurements from the nanoscale diodes

showed clear rectification behavior, and electroluminescence

(EL) spectra showed emission from the band edge. In

addition, the demonstration of band-edge emission from cross

junction formed between a direct-bandgap NW with a Si NW

clearly demonstrated the versatility of this approach in

combining different materials to assemble functional archi-

tectures. For example, nanoscale LEDs assembled from n-

CdS, n-GaN, and n-CdSe NWs with p-Si NWs have been

successfully demonstrated.[124] In addition, integration of

nanoscale LEDs from these NWmaterials with a common p-Si

NW was also configured, which showed the potential of

assembling multicolor LEDs on a common platform

(Figure 8), a task which is otherwise difficult for top-down

fabricated devices.[124] Recently, Hayden et al. have also

demonstrated the use of localized light emission from crossed

NW LEDs from p-Si/n-CdS for fluorescence spectroscopy and

cellular imaging.[126] The extremely localized excitation source

can in principle excite femtoliter quantities of fluids, thereby

opening up possibilities for highly sensitive imaging and

spectroscopy using NW devices.

Crossed NW diodes can also be used as photodetectors

under reverse bias. In a recent demonstration, crossed p-Si/n-

CdS NW diodes were assembled with the dopant concentra-

tion in a p-Si NWkept very low to increase the depletion width

at the junction to engineer avalanche multiplication and

breakdown under reverse bias (Figure 9).[127] The device

under reverse bias showed exceptionally low dark currents

(<100 pA) with a sharp breakdown occurring at ��9.0V

(Figure 9a). Upon illuminating the device with light, a large

reverse-bias-dependent photocurrent was observed, which is

indicative of avalanche breakdown phenomenon with a

multiplication factor estimated to be �5� 104. Spatially

resolved photocurrent measurements (Figure 9b) performed

on the devices revealed a single, highly localized photocurrent

peak located at the position of the crossed NW p–n junction

with a spatial resolution of �250 nm, thereby demonstrating

that the detector has subwavelength spatial-resolution cap-

abilities with an estimated sensitivity of �75 photons. This

H & Co. KGaA, Weinheim www.small-journal.com 1881

reviews R. Agarwal

Figure 8. a) I–V behavior of n–n, p–p, and p–n junctions of crossed n- and p-InP nanowires.

The green and blue curves correspond to the I–V behavior of individual n- and p-type

nanowires in the junction, respectively. The red curves represent the I–V behavior across the

junctions. Reproduced with permission from Reference [26]. Copyright 2001, Nature Pub-

lishing Group. b) EL spectra from crossed p–n diodes of p-Si and n-CdS, CdSSe, CdSe, and InP

(top to bottom, respectively). Insets to the left are the corresponding EL images for CdS,

CdSSe, CdSe, and InP nanowire LEDs. The top-right inset shows representative I–V and SEM

data recorded for a p-Si/n-CdS crossed NW junction (scale bar 1mm). c) Schematic and

corresponding SEM image of a tricolor nano-LED array. d) Normalized EL spectra and color

images from the three elements. Reproduced with permission from Reference [124].

1882

very high sensitivity is due to the large avalanche multi-

plication effect, which compensates for the low photon-

absorption cross section of nanoscale detectors. Larger arrays

with independently addressable avalanche photodiodes

(APDs) (Figure 9c) were also demonstrated, an essential

feature to enable future-generation integrated photonics.[127]

Crossed-NW architecture has also been utilized by Huang

et al. to assemble FETs, with one NW configured as the active

channel with a dielectric gate-oxide shell and the other NWs

functioning as local gate electrodes.[128] For example,

transistor-like properties were obtained from a Si NW

functioning as the p-channel with a SiO2 shell crossed with

a GaN NW acting as the gate electrode. Extending this

approach, threeGaNNWswere subsequently crossed with a Si

NW to assemble logic gates such as a two-input and two-output

AND and NOR logic gates and other basic computation was

demonstrated with signal gain (Figure 10a).[128] This experi-

Figure 9. a) Characterization of nanowire APDs. a) I–V characteristic of the APD in dark (black li

inset shows an SEM image of the crossed n-CdS/p-Si device (scale bar 4mm). b) Plot of the spa

measured in proportional mode using a diffraction-limited laser; the bias voltage, laser power, and

(in x and y), respectively. The nanowire positions are indicated on the plot by solid lines. c) Spatiall

(a). Both devices were biased at�10 V and excited at 488 nm (200 nW, Arþ ion laser) with a scannin

device consisting of an n-CdS nanowire (horizontal) crossing two p-Si nanowires (vertical); the la

contacts (scale bar 10mm). Reproduced with permission from Reference [127]. Copyright 2006,

www.small-journal.com � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinhe

ment demonstrated the feasibility of using

crossed semiconductor NW architectures in

assembling functional nanoelectronic cir-

cuits.

Even though the device functionality

originating from crossed NW architectures

is extremely small in length scale, the

assembly of integrated architectures for

nanoelectronics and photonics will require

schemes for addressing individual elements

within the nanoscale arrays. Zhong

et al.[129] have reported an approach to

circumvent this problem based upon a

scalable crossed-NW FET architecture, in

which chemical modification of specific

cross points within the array was used to

define address codes that enables NW input

lines to turn on/off specific output lines

(Figure 10d).[129] This basic array structure

thus functions as an address decoder, with

electronic gain originating from the crossed

NWFETs that can restore the electrical signal. Beckman et al.

demonstrated an innovative demultiplexer architecture for

bridging from submicrometer to nanometer-scale dimensions

for the selective addressing of ultrahigh-density NW circuits

on an array of 150 Si NWs patterned at widths of 13 nm and a

pitch of 34 nm crossed with metal NWs.[130]

Although important demonstrations have been made

utilizing the crossed-NW architecture, there is not much

information about the nature of carrier transport across the

crossed junction. In order to quantitatively model the junction

behavior, information regarding the geometry, oxide thick-

ness, defects, and impurities present at the junction need to be

determined, which is experimentally challenging. Lack of such

detailed information has impeded efforts on the theoretical

front to understand the nature of charge transport at the

crossed junction.Wei et al.[131] have attempted to qualitatively

model the junction behavior by using a tight-binding model

ne) and illuminated (red line) conditions. The

tially resolved photocurrent from the nano-APD

scanning step size were�7 V, 200 nW, and 250 nm

y resolved photocurrent measured from the array in

g step size of 1mm. Inset: Optical micrograph of the

rger rectangular features correspond to metal

Nature Publishing Group.

im small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Figure 10. a) Schematic image of logic AND gate assembled from a

1� 3 crossed NW junction array. Inset: A typical SEM (scale bar 1mm) of

the assembled AND gate and symbolic electronic circuit. b) The output

voltage versus the four possible logic address level inputs. Inset: The

V0–Vi, where the solid and dashed red (blue) lines correspond to V0–Vi1and V0–Vi2 when the other input is 0 (1). c) The experimental truth table

for the AND gate. d) SEM image of a 4� 4 crossed Si NW array address

decoder, with four horizontal NWs (I1 to I4) as inputs and four vertical

NWs (O1 to O4) as signal outputs. The four diagonal cross points in the

array were chemically modified to differentiate their responses from the

input gate lines (scale bar 1mm). e) Real-time monitoring of the Vg

inputs (blue) and signal outputs (red) for the four-by-four decoder.

Reproduced with permission from References [128, 129]. Copyright

2001 and 2003, respectively, American Association for the Advance-

ment of Science.

Figure 11. CdS NW electrically pumped laser assembled on p-Si sub-

strate. a) Schematic image showing the NW electrical-injection laser

device structure. In this structure, electrons and holes can be injected

into the CdS nanowire along the entire length from the top metal contact

and the bottom p-Si substrate, respectively. b) Top panel shows an

optical image of a laser device and the arrow highlights the exposed

CdS nanowire end (scale bar 5mm). Bottom panel shows an EL image

recorded from the device. The dashed line highlights the nanowire

position. f) EL spectra obtained from the nanowire end with injection

currents of 120mA (red, below lasing threshold) and 210mA (green,

above lasing threshold). The black arrows highlight Fabry–Perot cavity

modes with an average spacing of 1.83 nm. The green spectrum is

shifted upwards by 0.15 intensity units for clarity. d) Integrated output

light intensity as a function of forward bias voltage. Inset: I–V

characteristic of the laser device. Reproduced with permission from

Reference [132]. Copyright 2003, Nature Publishing Group.

and Green’s function approach, and calculated the transmis-

sion coefficients through the crossed junction. They argued

from their model that the formation of bound and quasi-bound

states at the junction, which depend on the coupling strength

of the two NWs, determine the transport properties. One

interesting observation from their model was that at low

electron concentrations, an increase in the interwire coupling

strength may not increase the electron transmission coefficient

between the NWs.[131] Much still remains to be explored both

experimentally and theoretically in order to understand the

charge-transport properties of the crossed nanojunctions,

which would require dedicated efforts.

6. NW–Substrate Interface for Functional Devices

Crossed-NW devices as discussed above can be configured

as efficient p–n-junction diodes but with localized injection of

charges. For certain applications such as injection NW lasers,

uniform injection of carriers along the entire NW length may

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

be required. To enable uniform injection of carriers in a NW

optical cavity, Duan et al.[132] assembled n-CdS NWs on

heavily doped p-Si substrates with contacts to CdS fabricated

lithographically (the p-Si substrate functioned as the second

electrode) as shown in Figure 11a and b.[132] A thin layer of

Al2O3 was deposited in between the top and bottom contacts

to force the current through the p-Si/n-CdS diode formed at

the NW-substrate junction. The device under low forward bias

produced EL with a broad featureless spectrum from the

exposed NW end (Figure 11c and d). Upon increasing the bias,

the EL spectrum quickly collapsed into narrow lasing modes,

clearly demonstrating electrically pumped lasing in single NW

optical cavities. Electrically driven lasing in single NWs

represents a novel and powerful approach for assembling

highly integrated photonic devices with the potential to co-

assemble diverse materials such as GaN, ZnO, CdSe, and

InP[124] on a common platform to produce lasers and other

photonic devices.

A similar concept was used by Bao et al. to fabricate

broadband LEDs from single ZnO NWs assembled on p-Si

substrates.[133] The device characteristics and broadband

emission was explained based on the p–i–n junction that is

formed due to the depletion of the NW originating from

H & Co. KGaA, Weinheim www.small-journal.com 1883

reviews R. Agarwal

Figure 12. SEM image of a) n-GaN nanowire/p-Si light-emitting diode.

Inset: zoomed-in image of the nanowire end. b) Forward-bias EL spectra

of the device at 77 K (black) and at 295 K (gray). Both traces were

obtained with the same applied bias of þ5.4 V and currents of 150mA at

77 K and 165mA at room temperature. Inset: plot of integrated intensity

versus current at 77 K (black) and 295 K (gray). c) Reverse-bias EL

spectrum for n-GaN nanowire/p-Si device. Reproduced with permission

from Reference [134]. Copyright 2007, Institute of Physics and IOP

Publishing Limited.

1884

oxygen adsorption on the NW surface. The broadband

emission was attributed to defects and surface states although

some emission from exciton recombination was also observed.

Zimmler et al.[134] recently performed a systematic electrical

study of n-GaNNWassembled on an n-Si substrate where they

observed that by reversing the polarity of the applied voltage

the EL can be selectively obtained from either the NW or the

substrate (Figure 12). For one polarity of the applied voltage,

UV and visible light were generated in the GaNNW, while for

the opposite polarity infrared light was emitted from the Si

substrate. A model, which took into account the oxidation of

GaN NW surface and Si substrate and was able to explain the

key features of the data based on electron tunneling from the

valence band of one semiconductor into the conduction band

of the other semiconductor. For example, for one polarity of

the applied voltage, given a sufficient potential-energy

difference between the two semiconductors, electrons can

tunnel from the valence band of GaN into the Si conduction

band, resulting in the creation of holes in GaN, which can

recombine with conduction-band electrons generating GaN

band-to-band luminescence.[134] This device structure affords

an additional experimental handle to the study of EL in single

NWs and, furthermore, could be used as a novel approach to

two-color light-emitting devices.

7. Axial Heterostructures

One of the first demonstrations of modulation of dopant

concentrations along a NW was reported by Wagner et al. in

1970s.[135] In 1992 Haraguchi et al.[136] reported assembly of p–

n junctions in GaAs NWs of �100-nm diameter using gold

particles as a catalyst, which showed rectification and EL that

was polarized along the NW long axis. C–V measurements

were also performed, which indicated formation of abrupt

junctions, while analysis of I–V behavior suggested formation

of junctions with 60-nm diameter due the formation of a

surface-depletion layer. However, the measurements were not

performed on individual NWs but on large arrays, which

limited the detailed understanding of the behavior of

nanoscale p–n junctions in axial geometry.

There was renewed interest in NW axial heterostructures

when their controlled synthesis was reported in 2002

by the Lieber,[137] Yang,[138] and Samuelson groups.[23]

The Samuelson group reported axial heterostructures with

atomically sharp interfaces, using chemical beam epitaxy

(CBE). The laser ablation and chemical vapor deposition

(CVD) techniques used by the Lieber[137] andYang groups[138]

produced compositionally graded interfaces. The main reason

for compositionally broadened interfaces is the retention of

the first constituent in the catalyst causing mixing of the two

components.

The nature of the heterointerface determines the proper-

ties of the NW heterostructure device. The Si-SiGe super-

lattice structures reported by Yang and co-workers did not

have abrupt interfaces but a broad alloyed region, which

was determined to be responsible for phonon scattering

instead of the desirable scattering from the heterointerfaces

for thermoelectric applications.[139] The axial heterostructures

www.small-journal.com � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

synthesized by Lieber and co-workers (GaP-GaAs)[137] were

also not abrupt with the alloying region directly proportional

to the diameter of the nanocatalyst used for the NW growth.

PL spectra from the multisegmented GaP-GaAs NW showed

emission from GaAs components only with no emission

observed from the indirect bandgap of GaP segments.

The Lieber group also fabricated axial p–n junctions in Si

NWs and measured the electrical properties of single NW

devices.[137] The I–V curve displayed rectification (Figure 13a),

the spatial origin of which was characterized by scanning

probe measurements indicating abrupt transition (instrument-

resolution limited) in carrier type and the accompanying

built-in potential at the physical p–n junction in the NW.

Moreover, intra-NW p-n junction in single InP NW devices

also showed EL under forward bias from the junction region.

Cheng et al.[141] also demonstrated the formation of axial p–n

junctions in single GaN NW devices, with temperature-

dependent measurements of I–V characteristics being con-

igure 13. Axial modulation-doped Si nanowires. a) Rectification behavior observed for

n axial p–n Si nanowire junction. Inset: electrical characterization by scanning probe

icroscopy and a scanning electron micrograph of the silicon nanowire device with source (S)

nd drain (D) electrodes. b) Scanning gate microscopy images of nþ-(n-nþ)N nanowires where

) N¼ 3, (B) N¼6, and (C) N¼ 8 are the periods. c) SEM image of a 2�2 decoder configured

sing two modulation-doped silicon nanowires as outputs (Out1 and Out2) and two Au metal

ates, which were deposited using e-beam lithography functioning as inputs (In1 and In2;

cale bar 1mm). Plots of input (blue) and output (red) voltages for the 2�2 decoder,

emonstrating selective addressing by Inputs 1 and 2. d) Coupled quantum-dot structures

ith a variable width barrier defined by synthesis; top) the device with the largest barrier

xhibits a single Coulomb oscillation period showing that the two quantum dots are weakly

oupled; middle) the intermediate-width barrier exhibits a splitting of each of the Coulomb

scillation peaks into doublets, which is the signature of enhanced tunneling conductance

etween the quantum dots; bottom) further reduction of the barrier width again produces

ingle Coulomb oscillation period with the effective quantum-dot size twice that of the

dividual quantum-dot regions, implying that the structures are fully delocalized.

eproduced with permission from Reference [140]. Copyright 2005, American Association

r the Advancement of Science.

F

a

m

a

(A

u

g

s

d

w

e

c

o

b

s

in

R

fo

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

sistent with electron tunneling through a voltage-dependent

barrier; however, the EL was not reported.

There have been multiple reports of formation of single-

crystalline axial heterojunctions between Si NW and NiSi by

solid-state transformation reactions (discussed in Section 3),

where the nature of nanoscale junctions has been shown to

produce superior device characteristics.[39,40] The concept of

modulation doping of Si NWs along their lengths with

different dopant types and concentrations was utilized by

Yang et al.[140] to assemble address decoders to bridge the

macro- to nanoscale size scale. Growth of nþ–n–nþ repeat

structures axially along Si NW length with control over size

and periodicity was achieved and studied by scanning gate

microscopy, which demonstrated well-defined dopant regions

of varying concentration (Figure 13b). A prototype address

decoder device was demonstrated with two Si-modulation-

doped NWs and microscale address wires (lithographically

defined) functioning as inputs and outputs, respectively

H & Co. KGaA, Weinheim

(Figure 13c). In the same publication, Yang

et al. also demonstrated control over the

size and separation of doped regions to

confine single and multiple quantum dots

by exploiting the different band offsets

formed by dopant variations.[140] By vary-

ing the separation between the confined

dots and the barrier heights, different

intradot coupling strengths were clearly

observed (Figure 13d), which demonstrated

the advantage of chemical synthesis over

lithography to control and encode different

device functionalities. In another report,

Yang et al.[142] also demonstrated the

formation of controlled axial p–i–n junc-

tions in Si NWs. Temperature-dependent

I–Vmeasurements recorded from individual

p–i–n devices showed an increase in the

breakdown voltage with temperature, char-

acteristic of avalanche breakdown. Spa-

tially resolved photocurrent measurements

showed that the largest photocurrent is

generated at the intrinsic region located

between the electrode contacts, with multi-

plication factors of 30.[142] Electron- and

hole-initiated avalanche gain measure-

ments performed by localized photoexcita-

tion of the p-type and n-type regions

yielded multiplication factors of 100 and

20, respectively, implying that the ioniza-

tion rate for electrons is significantly higher

than that for holes, in agreement with

conventional planar silicon APDs. One can

envision changing the dopant concentration

and length of different segments to pre-

cisely engineer the electric-field profiles in

these 1D NWs to produce the desired

electronic/optoelectronic response.

One of the first demonstrations of axial

NW heterostructures with near atomically

abrupt interfaceswas reportedbySamuelson

www.small-journal.com 1885

reviews R. Agarwal

1886

and co-workers in 2002 and in their subsequent publica-

tions.[23,97,143,144] Bjork et al. were able to produce InP–InAs

superlattice structures with abrupt interfaces using ultrahigh

vacuum CBE with NW growth catalyzed by Au particles

(Figure 14).[23] The reason for abrupt interfaces was attributed

to the high vapor pressures of the group-V precursors in

combination with the low growth rate (�1 monolayer s�1). In

later papers, detailed analysis was performed in which it was

found that the gold catalyst remains solid during the entire

growth process, leading to slow growth rates and hence

excellent control over heterointerfaces with no obvious

structural defects.[145] Utilizing this level of control over the

axial heterointerfaces, InP barrier segments of varying

lengths were embedded in 40-nm-diameter InAs NWs.[143]

Temperature-dependent electrical data from these hetero-

structured NWs were modeled by assuming thermionic

Figure 14. a) HRTEM image of an InAs NW with two InP barriers. Double-

barrier resonant tunneling data for devices fabricated inside a nano-

wire. b) TEM image of the top end of a NW with the double barrier clearly

visible with a barrier thickness of �5 nm (scale bar 30 nm). c) Energy-

band diagram for the device with the electronic states in the emitter

region indicated (left). In between the two InP barriers the fully

quantized levels of the central quantum dot are indicated. d) I–V data

for the same device as shown in (b) and (c) recorded at 4.2 K, revealing a

sharp peak in the characteristics, reflecting resonant tunneling into the

ground state with a voltage width of �5 mV. This width can be

translated into an energy width of the transition of �2 meV, corre-

sponding to the width of the shaded energy band in the emitter from

which electrons tunnel. The inset provides a magnified view of the

resonance peak for increasing voltage (gray) and for decreasing voltage

(black), hence, confirming negligible hysteresis effects. Reproduced

with permission from Reference [143]. Copyright 2002, American

Institute of Physics.

www.small-journal.com � 2008 Wiley-VCH Verlag Gm

emission over the InP barrier from which a barrier height

of �0.6 eV was determined, in close agreement with the

estimated barrier height of bulk systems.[143] By changing the

barrier thickness, the tunnel resistance was shown to vary over

a very large range. The Samuelson group also synthesized a

resonant tunnel diode (RTD) embedded in an InAs NW with

an InAs quantum dot sandwiched between two InP barriers in

axial geometry (Figure 14).[143] The observation of sharp

features in the tunneling spectra, corresponding to well-

defined quantum states of the InAs dot, was enabled only due

to the abrupt interfaces between InAs and InP segments.

Thelander et al.[144,146] then assembled single electron

transistors (SETs) by changing the RTD device by increasing

the InAs dot dimensions and fabricating a gate electrode

coupled to the InAs dot. By changing the InAs dot size it was

shown by Bjork et al. that it is possible to tune the device

performance from SET to a few electron quantum dots

(Figure 15), with the tunneling spectra obtained from

transport measurements displaying unique shell structure

owing to spin and orbital degeneracy.[144] Bjork et al. reported

tunneling spectroscopy measurements of the Zeeman spin

splitting in InAs; few electron quantum dots formed between

two InP barriers.[147] They observed that the values of the

electron g factors of the first few electrons entering the dot

strongly depend on the dot size ranging from the bulk value of

InAs in large dots (jg�j¼13) down to jg�j¼2.3 for the smallest

dots.[147] Fabrication of intra-NW SETs allows continuous

tunability of device characteristics, which is otherwise not

possible with SET behavior observed in homogeneous

NWs.[148] Fabrication of RTDs and SETs with the desired

device characteristics provides an impressive demonstration

that it is possible to form novel confinement potentials using

axial interfaces within NWs.

By extending the concept of growing InAs dots between

InP barriers, Samuelson and co-workers demonstrated storage

of single and few electrons in multiple tunnel junctions

incorporated at the end of InAs wires, where the Au seed

particle functioned as the storage mode.[149] By placing

another NW in close proximity, the mechanism of charge

transfer in the NW was studied and trapping of electrons was

confirmed. In another demonstration of NWmemory based on

stacks of axial quantum dots, Nilsson et al.[150] showed

memory operation by tunneling of electrons through multiple

dots separated by barriers. The transport through the

heterostructured device was shown to be tunable by applied

bias due to misalignment of the energy levels leading to

hysteresis of charging/discharging of the storage island.[150]

One of the challenges in developing devices based on quantum

states in NWs for quantum-information applications is the

ability to detect and measure the charge states on the quantum

dots. Wallin et al.[151] demonstrated the operation of a charge-

readout scheme for InAs dots defined between InP barriers in

InAs NW using a quantum-point contact defined in a GaAs/

AlGaAs 2D electron gas (2DEG) beneath the NW. Applying

negative voltages to two split-gate electrodes aligned to the

NW was shown to induce a quantum point contact in the

2DEG such that charging of quantum dots in the NW

modulated the quantum point contact transmission, resulting

in the desired detector response.[151]

bH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Figure 15. a) left: Dark-field scanning TEM image of a nanowire with a 100-nm-long InAs

quantum dot between two very thin InP barriers (scale bar 20 nm). Right: corresponding

image of a 10-nm-long InAs dot. The InP barrier thickness is 3 and 3.7 nm, respectively. Effect

of reduced quantum-dot length on transport properties; b) gate characteristics of a

single-electron transistor (SET) with a 100-nm-long dot. The oscillations are periodic. c) When

the dot length is 30 nm, the level spacing at the Fermi energy is comparable to the charging

energy and the Coulomb oscillations are no longer completely periodic. d) A 10-nm dot results

in a device depleted of electrons at zero gate voltage. By increasing the electrostatic potential

electrons are added one by one. For some electron configurations the addition energy is

larger, corresponding to filled electron shells. All data were recorded at 4.2 K. Reproduced

with permission from Reference [144]. Copyright 2004, American Chemical Society.

Optical spectroscopy has also been utilized to study the

effect of quantum confinement of charge carriers in axial NW

heterostructures. PL spectra have been studied in direct-gap

systems, including ZnSe–CdSe,[152] InAs–GaAs,[153] InAs–

InP,[154] and ZnO–ZnMgO.[155] Poole et al. reported

InP–InAs–InP NW heterostructures on InP substrates without

the use of a catalyst with growth based on the preferential

diffusion of In along facets.[154] The InAs quantum-well

thickness was estimated to be 0.6 nm while quantitative

estimates of interface quality was not reported. At low

optical power, a single well-defined peak was observed at

921meV in PL spectrum. At higher powers, excited-state

peaks were observed at energy spacings consistent with the

quantum confinement of excitons corresponding to the

estimated well widths.[154] Park et al.[155] demonstrated

excellent control over the axial heterointerface between

repeated segments of ZnO–ZnMgO nanorods grown using

catalyst-free metal–organic vapor-phase epitaxy (Figure 16).

ZnO–ZnMgO interface widths of �1 nm were measured

using high-resolution transmission electron microscopy

(HRTEM) and no significant Mg diffusion. They observed a

systematic blueshift of emission with decreasing quantum-well

width in ZnO–ZnMgO multiple quantum-well heterostruc-

tures (ensemble measurement), which was modeled assuming

periodic square-well potentials for 10 period structures

(Figure 16c and d).[155] Radial confinement was not observed

as the wire diameters averaged �40 nm, which is much

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim

greater than the exciton Bohr radius of

ZnO. In a later publication, Park et al.[156]

also measured time-resolved excited-state

decay characteristics from ZnO–ZnMgO

NW single-quantum-well samples and

observed slower decay with two time

constants (170 ps and 580 ps) from the

quantum-well emission peaks in compar-

ison to the remaining wire (56 ps and

330 ps). They attributed slower decay from

the quantum-well to size-dependent light–

matter interaction in low-dimensional sys-

tems, leading to formation of exciton–

polaritons with longer recombination life-

time than in bulk systems.[156] Even though

it is an interesting observation, more

detailed experiments and theoretical mod-

eling are required to make such claims in

complex heterostructured systems at length

scales that are smaller than the exciton

Bohr radius. Yatsui et al.[157] reported low-

temperature near-field spectroscopy of

isolated ZnO–ZnMgO single-quantum-

well structures grown on the end of ZnO

nanorods. The absorption spectra of iso-

lated quantum-well structures displayed a

small Stokes shift of 3 meV and sharp PL

peaks. Furthermore, polarization spectro-

scopy of isolated nanorod quantum wells

revealed valence-band anisotropy in PL

spectra.[157] Since the exciton in a quantum

structure is an ideal two-level system with

long coherence times, their results provided criteria for

designing nanophotonic switches, which was indeed reported

in a more recent publication.[158] By using time-resolved near-

field spectroscopy of ZnO–ZnMgO nanorod double-quantum-

well structures, they observed nutation of the population

between the resonantly coupled exciton states of the coupled

quantum dots. Optical switching dynamics were also demon-

strated by controlling the exciton excitation in the dipole-

inactive state via an optical near field.[158] Their results are a

promising step toward designing a nanometer-scale photonic

switch and related devices that utilize the quantum confine-

ment in stand-alone nanorod structures.

Kim et al.[159] have reported high-efficiency LEDs using

dislocation-free InGaN–GaN multiquantum-well NW arrays

grown on sapphire substrate. The NWs were tapered with a

diameter in the center of �70 nm and a length of �1mm and

with typical quantum-well widths of �4.8 nm with near

atomically abrupt interfaces. Due to the lack of dislocations

and the large surface areas provided by the sidewalls of NWs,

both internal and extraction efficiencies were significantly

enhanced with about 4.3 times more light than the conven-

tional broad-area LEDs. The present method of utilizing

dislocation-free NW LEDs is promising for the fabrication of

superbright-white LEDs and for improving their total external

efficiency.[159]

Panev et al.[153] observed sharp exciton-like emission from

InAs quantum dots embedded inside GaAs NW hosts grown

www.small-journal.com 1887

reviews R. Agarwal

Figure 16. a) Schematic of axial multiple quantum-well NWs consisting of 10 periods of

Zn0.8Mg0.2O/ZnO. Inset: corresponding electronic-band diagram. b) Z-contrast TEM image of a

2.5-nm-width sample clearly showing the composition variation (bright layers are ZnO wells).

c) Pl spectra at 10 K for a single-period heterostructure (top) and 10-period multiple quantum

well with varying well widths. The corresponding band diagrams are given. d) Quantum-well-

dependent peak positions (squares) along with theoretical prediction. Reproduced with

permission from Reference [155].

1888

using CBE. The authors observed growth in both axial and

radial directions, thereby producing tapered NWs that

effectively encapsulated the InAs dot within a GaAs shell,

leading to effective surface passivation resulting in increased

radiative quantum yield. The emission from the InAs dots was

significantly blueshifted from the bulk InAs, indicating that

substantial alloying with the GaAs wire had occurred in

addition to any quantum-confinement effects, which made

estimation of confinement effects rather difficult.[153] In

another interesting study, Kim et al.[160] observed formation

of embedded quantum dots in ZnO–ZnMgO multiple

quantum-well samples due to partial strain relaxation if the

number of periods exceeded a certain number (four) or the

quantum-well width exceeded 2.5 nm in thickness. Emission

from embedded quantum dots was clearly observed and their

experiments[160] showed that efficient strain relaxation in

nanoscale heterostructures is not a universal feature, as

previously believed. Borgstrom et al.[161] reported assembly of

As-rich quantum dots of �15-nm length embedded in a P-rich

GaAs/GaP NW system. Significantly, photon antibunching

was observed from quantum-dot emission proving the zero-

dimensional nature of these heterostructures, which also

opens up interesting possibilities of fabricating electrically

driven single photon sources based on the bottom-up

www.small-journal.com � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinhe

approach.[161] An important breakthrough

towards realizing this goal has been

reported by Minot et al.[162] where they

have been able to achieve reproducible

fabrication of InP–InAsP LEDs in which

electron–hole recombination is restricted

to a quantum-dot InAsP section confined

between n- and p-doped InP segments

(Figure 17). Rectification was observed in

I–V measurements and EL corresponding

to InAsP quantum dots was also

observed.[162] Even though several other

challenges such as emission from n- and p-

doped InP, low quantum of emission, and

better device structure need to be over-

come to assemble an electrically driven

single photon source fom NWs, this

experiment is a very critical advance

towards the final goal.

8. Radial Heterostructures

The ability to create new systems by

manipulating the growth procedures can be

extended to create NW radial heterostruc-

tures with compositional modulation in the

radial direction. To grow radial hetero-

structures, the VLS growth is minimized

and the second component grows epitaxi-

ally on the NW core. Analogous to

colloidal core/shell quantum dots,[163]

NW radial heterostructures can passivate

the surface to remove the deleterious

effects of surface states, enabling enhanced

performance of devices. For example, a shell of high-k

dielectric can enhance the performance of FETs, a ferro-

electric oxide shell can be used to fabricate non-volatile

memory, and materials with different band offsets can be

combined to create a high-mobility carrier gas, or confine

excitons in the radial direction. Core/shell structures have also

been utilized to dope the NW core via deposition of a heavily

doped shell,[164] formation of axial p–n junctions,[165] and for

the synthesis of novel nanotubes.[166,167]

The first synthesis of core/shell NW heterostructures was

reported by Lauhon et al.[24] in 2003 by using a CVD approach

to grow homo- and heterostructures from Si and Ge with

different dopant concentration and types including i-Si–p-Si,

Si–Ge and Ge–Se core/shell NWs. Electrical transport

measurements on p-Si–i-Ge and i-Ge–p-Si core–shell wires

showed evidence of charge transfer due to the valence-band

offset between Si and Ge.[24] Lauhon et al. also coated a

conducting p-Si/i-Ge core/shell wire with silicon oxide, and

then with a heavily doped layer of p-Ge to fabricate a coaxially

gated FET, all from a single NW.[24] This device demonstrated

the possibility of making a modulation-doped radial core/–

shell FET with a 1D channel or a 1D carrier gas, which was

later demonstrated by Lu et al. (Figure 18),[168] who reported

the synthesis and transport studies of a 1D hole gas system

im small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Figure 17. a) TEM image of n-InP/i-InAs0.6P0.4/p-InP NW. Rotational

stacking defects are seen in the InP segments, with a higher density of

stacking defects in n-InP. Inset shows a HRTEM image of the InAsP crystal

structure. EL properties of an InP NW LED without InAsP section. b) Optical

microscopy image collected by CCD camera of a device under forward bias

(scale bar 2mm). c) EL spectrum at 10 K from an InP NW LED with a 12-nm

InAsP active region. The spectra correspond to 1.4, 1.6, 1.8, 2.0, and 2.2 V

bias and 60, 60, 30, 10, and 10 s integration times, respectively. The

vertical dashed line indicates the InP bandgap energy at 10 K. Inset shows

the I–Vbias characteristics of the device. Reproduced with permission from

Reference [162]. Copyright 2007, American Chemical Society.

Figure 18. Ge/Si core/shell nanowire FET device. a) Schematic image of

section through the Ge/Si core/shell structure and b) band diagram for

heterostructure. The dashed line indicates the position of the Fermi level,

inside the Si bandgap and below the Ge valance-band edge. c) HRTEM ima

Si core/shell nanowire with 15-nm Ge (dark gray) core diameter and 5-n

gray) shell thickness (scale bar 5 nm). d) Coulomb blockade in unannea

devices. I–Vg for a 10-nm core diameter Ge/Si NW (T¼1.5 K, VSD ¼0.5 m

L¼ 112 nm). Inset: I–VSD data taken at Vg¼�9.38 V showing the Coulom

gap. e) Ballistic transport in Ge/Si 1D hole gas device. G(conductance)–Vg

4.7 K for a 10-nm core Ge/Si nanowire with L¼350 nm. Left inset: I–VSD

recorded at Vg¼ 10 V. Right inset: I–VSD data recorded at Vg from 10 to �steps. Reproduced with permission from Reference [168]. Copyright 200

Academy of Sciences USA.

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

based on a free-standing Ge/Si core/shell NW heterostructure

with a 5-nm Si shell. The valence-band offset of 500meV

between Ge and Si at the interface was shown to serve as a

confinement potential for the holes that would accumulate in

the Ge NW. Room-temperature electrical measurements

showed hole accumulation in undoped Ge/Si NW heterostruc-

tures, in contrast to control experiments on homogeneousNWs.

Low-temperature studies showed well-controlled Coulomb-

blockade oscillations when the Si shell serves as a tunnel

barrier to the hole gas in the Ge channel (Figure 18b).[168]

They also observed conductance quantization at low tem-

peratures, corresponding to ballistic transport through 1D sub-

bands, where the measured sub-band energy spacings agreed

with calculations for a cylindrical confinement potential

(Figure 18c). The conductance was observed to be relatively

insensitive to temperature, consistent with their calculation of

reduced backscattering in 1D systems. Recently, Xiang

et al.[43] used the same device architecture to configure FETs

using high-k dielectrics in a top-gated geometry. The 1D hole

gas in the Ge/Si NW heterostructure and enhanced gate

coupling with high-k dielectrics displayed excellent perfor-

mance from their FETs with on-currents (2.1mA mm�1)

that are three to four times greater than state-of-the-art

MOSFETs. Xiang et al.[46] also utilized Ge/Si radial NW

heterostructures contacted by superconducting leads to study

proximity-induced superconductivity and presented studies of

mesoscopic Josephson junctions that demonstrated tunable

dissipationless supercurrents and signatures of multiple

Andreev reflections, suggesting coherent charge transport.

a cross

a Si–Ge–Si

which lies

ge of a Ge/

m Si (light

led Ge/Si

V,

b-blockade

recorded at

curve

10 V in 1 V

5, National

H & Co. KGaA,

Their study presents new opportunities to utilize

the scattering-free 1D potential of radial NW

heterostructures as superconducting FETs and

unique quantum interference for future devices.

The optical properties of core/shell NWs have

also been studied in some detail. Skold et al.[169]

synthesized GaAs–GaxIn1–xP (0.34< x< 0.69)

core/shell NWs by metal–organic vapor-phase

epitaxy. PL measurements on individual NWs at

5K showed that the emission efficiency increased

by 2 to 3 orders of magnitude compared to

uncapped samples due to the passivation of the

surface states by the epitaxially grown shell. Skold

et al.[169] also strained the GaAs core by using

lattice mismatched shells and were able to tune

the bandgap by �240meV and the measured

strain induced shifts were confirmed by calcula-

tions based on deformation-potential theory. The

excited state decayed rapidly (�100 ps), implying

additional non-radiative trap states in addition to

the surface states passivated by the epitaxial shell.

Park et al.[170] demonstrated radial heterostruc-

tures between ZnO/ZnMgO with a ZnO core of

�9 nm. PL measurements revealed a dominant

peak at 3.316 eV indicating a blueshift of 30meV

resulting from the quantum confinement of the

core. The peak intensity for the heterostructure

was much higher than bare ZnO NWs, demon-

strating the passivation of the surface from the

shell.[170] Jang et al.[171] and Bae et al.[172] reported

Weinheim www.small-journal.com 1889

reviews R. Agarwal

Figure 19. a) Schematic illustration of a ZnO/Mg0.2Zn0.8O/ZnO/Mg0.2Zn0.8O multishell

NW heterostructure and its band diagrams for different ZnO (quantum well) shell widths (LW).

b) PL spectra (10 K) of ZnO/Mg0.2Zn0.8O core/shell NW heterostructures and ZnO/

Mg0.2Zn0.8O/ZnO/Mg0.2Zn0.8O multishell NW quantum structures with different ZnO

(quantum-well) shell widths of 45, 30, 15, 8 A. c) Width-dependent PL peak positions in

ZnO/Mg0.2Zn0.8O coaxial quantum structures (solid circles) and theoretically calculated

values (open circles) in one period of 1D square potential wells. Reproduced with permission

from Reference [171]. Copyright 2006, American Institute of Physics.

1890

ZnO/ZnMgO multishell quantum-well NW heterostructures

and obtained systematic blueshift of the PL peaks with

decreasing quantum-well thickness and evidence of transitions

from sub-band levels (Figure 19). Zanolli et al.[173] reported

quantum confinement in InAs/InP core/shell NWs with

�25-nm core diameter. They observed two peaks at higher

excitation intensities, which were attributed to emission from

the ground and the first excited quantized levels due to

quantum confinement. Their results could only be explained

by taking into account the strain in the system, which again

shows that strain does build up in nanoscale heterostructures

without getting fully relaxed.[173] Choi et al.[174] demonstrated

the formation of GaN–AlGaN radial NW heterostructures,

which formed due to spontaneous Al–Ga–N phase separation

at the nanometer scale. The simultaneous excitonic and

photonic confinement within these heterostructures lead to

quantum-wire UV lasers with relatively low threshold in

comparison to single-component GaN NW lasers.[174]

Core/shell NWhave also been configured as p–n diodes and

LEDs. Qian et al.[175] reported efficient injection of carriers in

active nanophotonic devices involving the synthesis of well-

defined, defect-free n-GaN/InGaN/p-GaN core/shell/shell NW

heterotructures grown bymetal–organic CVD. PL data showed

strong emission from the InGaN shell centered at 448 nm.

Significantly, electrical injection devicesmade by contacting the

n-type core/p-type shell demonstrated that in forward bias these

individualNWs behave as LEDswith bright blue emission from

www.small-journal.com � 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinhe

the InGaN shell.[175] In a later paper, Qian

et al.[176] extended this approach to grow

InxGa1–xN/GaN/p-AlGaN/p-GaN struc-

tures where the EL wavelength was tunable

by simply changing the In content. The

ability to rationally synthesize radial NW

heterostructures therefore opens up new

opportunities for assembling novel nano-

photonicdevices, includingmulticolorLEDs

and lasers.

In order to understand the interfacial

properties of NW heterostructures and

correlate them with their function, further

development of characterization techniques

are required. For example, TEM-based

strain mapping,[177,178] TEM tomogra-

phy,[179] electron-beam induced current,[180]

electron holography,[181] grazing-incidence

X-ray,[182] scanning tunneling microscopy

(STM) techniques,[183] and atom probe

tomography[184] have been reported

recently, which are promising in detailing

the interfacial properties of NW hetero-

structures. Development of such techniques

in conjunction with electrical and optical

studies will be essential to correlate and

understand how the nature of the interfaces

determines the measured properties, some-

thing that is still largely missing.

9. Conclusions

Tremendous progress has beenmade in the past fewyears in

the area of semiconductor NW heterostructure growth and

characterization of structural, electronic, and optical properties

of theheterointerfaces andassemblyofdevices.However,many

challenges still remain to be overcome in order to enable unique

technologies in the future. More work is required to synthesize

NW heterostructures with control over their interfaces,

dimensions, and chemical composition from other materials

in addition to III–V semiconductors. Furthermore, a detailed

and systematic understanding of the effect of surface/interfaces

at the nanoscale needs to bedeveloped,which is still lacking and

would also require development of new characterization

techniques. Finally, and most significantly, techniques need

to be developed for the hierarchical assembly of nanostructures

with exquisite spatial control to enable continued progress

in this field to fabricate new and interesting nanosystems

with enhanced properties with applications ranging from

electronics, photonics, and biological and chemical sensing to

imaging.

Acknowledgements

I would like to thank many of my colleagues at the University of

Pennsylvania, particularly Lambert van Vugt, Bin Zhang, and

im small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

Christopher Rodd for helpful discussions. Funding from the

National Science Foundation (NSF-CAREER ECCS-0644737 and

NSF-DMR-0706381) is gratefully acknowledged.

[1] International Technology Roadmap for Semiconductors, 2005report. http://www.itrs.net/Common/2005ITRS/Exec-

Sum2005.pdf.

[2] R. Compano, L. Molenkamp, D. J. Paul, Technology Roadmap for

Nanoelectronics, European Commission IST Program: Future and

Emerging Technologies, Microelectronics Advanced Research

Initiative; http://nanoworld.org/NanoLibrary/nanoroad.pdf.

[3] J. R. Heath, P. J. Kuekes, G. S. Snider, R. S. Williams, Science 1998,

280, 1716.

[4] C. M. Lieber, Sci. Am. 2001, 285, 58.

[5] a) K. Hiruma, M. Yazawa, T. Katsuyama, K. Ogawa, K. Haraguchi,

M. Koguchi, H. Kakibayashi, J. Appl. Phys. 1995, 77, 447; b) M.

Law, J. Goldberger, P. D. Yang, Annu, Rev. Mater, Res. 2004, 34,

83.

[6] C. M. Lieber, MRS Bull. 2003, 28, 486.

[7] Y. N. Xia, P. D. Yang, Y. G. Sun, Y. Y. Wu, B. Mayers, B. Gates, Y. D.

Yin, F. Kim, Y. Q. Yan, Adv. Mater. 2003, 15, 353.

[8] C. Dekker, Phys. Today 1999, 52, 22.

[9] P. G. Collins, P. Avouris, Sci. Am. 2000, 283, 62.

[10] A. J. Mieszawska, R. Jalilian, G. U. Sumanasekera, F. P. Zamborini,

Small 2007, 3, 722.

[11] L. Samuelson, Mater. Today 2003, 6, 22.

[12] W. Lu, C. M. Lieber, Nat. Mater. 2007, 6, 841.

[13] C. Thelander, P. Agarwal, S. Brongersma, J. Eymery, L. F. Feiner, A.

Forchel, M. Scheffler, W. Riess, B. J. Ohlsson, U. Goesele, L.

Samuelson, Mater. Today 2006, 9, 28.

[14] R. S. Wagner, W. C. Ellis, Appl. Phys. Lett. 1964, 16, 761.

[15] A. M. Morales, C. M. Lieber, Science 1998, 279, 208.

[16] X. F. Duan, C. M. Lieber, Adv. Mater. 2000, 12, 298.

[17] C. Weisbuch, B. Vinter, Quantum Semiconductor Structures:

Fundamentals and Applications, Academic, San Diego 1991.[18] L. J. Lauhon, M. S. Gudiksen, C. M. Lieber, Philos. Trans. R. Soc. A

2004, 362, 1247.

[19] Y. Jung, D. K. Ko, R. Agarwal, Nano Lett. 2007, 7, 264.

[20] M. Paladugu, J. Zou, G. J. Auchterlonie, Y. N. Guo, Y. Kim, H. J.

Joyce, Q. Gao, H. H. Tan, C. Jagadish, Appl. Phys. Lett. 2007, 91.

[21] M. Paladugu, J. Zou, Y. N. Guo, G. J. Auchterlonie, H. J. Joyce, Q.

Gao, H. H. Tan, C. Jagadish, Y. Kim, Small 2007, 3, 1873.

[22] K. A. Dick, S. Kodambaka, M. C. Reuter, K. Deppert, L. Samuelson,

W. Seifert, L. R. Wallenberg, F. M. Ross, Nano Lett. 2007, 7, 1817.

[23] M. T. Bjork, B. J. Ohlsson, T. Sass, A. I. Persson, C. Thelander, M.

H. Magnusson, K. Deppert, L. R. Wallenberg, L. Samuelson, Nano

Lett. 2002, 2, 87.

[24] L. J. Lauhon, M. S. Gudiksen, C. L. Wang, C. M. Lieber, Nature

2002, 420, 57.

[25] Y. Huang, X. F. Duan, Q. Q. Wei, C. M. Lieber, Science 2001, 291,

630.

[26] X. F. Duan, Y. Huang, Y. Cui, J. F. Wang, C. M. Lieber, Nature 2001,

409, 66.

[27] R. Agarwal, K. Ladavac, Y. Roichman, G. H. Yu, C. M. Lieber, D. G.

Grier, Opt. Express 2005, 13, 8906.

[28] D. Whang, S. Jin, Y. Wu, C. M. Lieber, Nano Lett. 2003, 3,

1255.

[29] C. Y. Nam, D. Tham, J. E. Fischer, Nano Lett. 2005, 5, 2029.

[30] D. Tham, C. Y. Nam, J. E. Fischer, Adv. Mater. 2006, 18, 290.

[31] F. Hernandez-Ramirez, A. Tarancon, O. Casals, E. Pellicer, J.

Rodriguez, A. Romano-Rodriguez, J. R. Morante, S. Barth, S.

Mathur, Phys. Rev. B 2007, 76, 085429.

[32] J. R. Kim, H. Oh, H. M. So, J. J. Kim, J. Kim, C. J. Lee, S. C. Lyu,

Nanotechnology 2002, 13, 701.

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

[33] G. Koley, L. Lakshmanan, H. Wu, H. Y. Cha, Phys. Status Solidi A

2007, 204, 1123.

[34] E. Stern, G. Cheng, J. F. Klemic, E. Broomfield, D. Turner-

Evans, C. Li, C. Zhou, M. A. Reed, J. Vac. Sci. Technol. B 2006,

24, 231.

[35] C. S. Lao, J. Liu, P. X. Gao, L. Y. Zhang, D. Davidovic, R. Tummala,

Z. L. Wang, Nano Lett. 2006, 6, 263.

[36] D. B. Suyatin, C. Thelander, M. T. Bjork, I. Maximov, L. Samuelson,

Nanotechnology 2007, 18, 105307.

[37] Y. Ahn, J. Dunning, J. Park, Nano Lett. 2005, 5, 1367.

[38] Y. Gu, E. S. Kwak, J. L. Lensch, J. E. Allen, T. W. Odom, L. J. Lauhon,

Appl. Phys. Lett. 2005, 87, 043111.

[39] Y. Wu, J. Xiang, C. Yang, W. Lu, C. M. Lieber, Nature 2004, 430,

61.

[40] W. M. Weber, L. Geelhaar, A. P. Graham, E. Unger, G. S. Duesberg,

M. Liebau, W. Pamler, C. Cheze, H. Riechert, P. Lugli, F. Kreupl,

Nano Lett. 2006, 6, 2660.

[41] J. Appenzeller, J. Knoch, E. Tutuc, M. Reuter, S. Guha, in Electron

Devices Meeting, IEDM IEEE, 2006, p. 1.

[42] K. Byon, D. Tham, J. E. Fischer, A. T. Johnson, Appl. Phys. Lett.

2007, 90, 143513.

[43] J. Xiang, W. Lu, Y. J. Hu, Y. Wu, H. Yan, C. M. Lieber, Nature 2006,

441, 489.

[44] Y. J. Doh, J. A. van Dam, A. L. Roest, E. P. A. M. Bakkers, L. P.

Kouwenhoven, S. De Franceschi, Science 2005, 309, 272.

[45] J. A. Van Dam, Y. V. Nazarov, E. P. A. M. Bakkers, S. De Franceschi,

L. P. Kouwenhoven, Nature 2006, 442, 667.

[46] J. Xiang, A. Vidan, M. Tinkham, R. M. Westervelt, C. M. Lieber, Nat.

Nanotech. 2006, 1, 208.

[47] P. D. Yang, H. Q. Yan, S. Mao, R. Russo, J. Johnson, R. Saykally, N.

Morris, J. Pham, R. R. He, H. J. Choi, Adv. Funct. Mater. 2002, 12,

323.

[48] M. H. Huang, S. Mao, H. Feick, H. Q. Yan, Y. Y. Wu, H. Kind, E.

Weber, R. Russo, P. D. Yang, Science 2001, 292, 1897.

[49] Q. H. Xiong, G. Chen, J. D. Acord, X. Liu, J. J. Zengel, H. R. Gutierrez,

J. M. Redwing, L. C. L. Y. Voon, B. Lassen, P. C. Eklund, Nano Lett.

2004, 4, 1663.

[50] H. W. Seo, S. Y. Bae, J. Park, H. N. Yang, K. S. Park, S. Kim, J. Chem.

Phys. 2002, 116, 219492.

[51] R. Agarwal, C. J. Barrelet, C. M. Lieber, Nano Lett. 2005, 5, 917.

[52] C. Ma, Y. Ding, D. Moore, X. D. Wang, Z. L. Wang, J. Am. Chem. Soc.

2004, 126, 708.

[53] B. Xiang, H. Z. Zhang, G. H. Li, F. H. Yang, F. H. Su, R. M. Wang, J.

Xu, G. W. Lu, X. C. Sun, Q. Zhao, D. P. Yu, Appl. Phys. Lett. 2003,

82, 193330.

[54] M. S. Gudiksen, J. F. Wang, C. M. Lieber, J. Phys. Chem. B 2002,

106, 4036.

[55] M. Mattila, T. Hakkarainen, H. Lipsanen, H. Jiang, E. I. Kauppinen,

Appl. Phys. Lett. 2007, 90, 033101.

[56] L. K. van Vugt, S. J. Veen, E. P. A. M. Bakkers, A. L. Roest, D.

Vanmaekelbergh, J. Am. Chem. Soc. 2005, 127, 12357.

[57] H. Yu, J. B. Li, R. A. Loomis, L. W. Wang, W. E. Buhro, Nat. Mater.

2003, 2, 517.

[58] X. F. Duan, J. F. Wang, C. M. Lieber, Appl. Phys. Lett. 2000, 76,

091116.

[59] T. Stoica, R. J. Meijers, R. Calarco, T. Richter, E. Sutter, H. Luth,

Nano Lett. 2006, 6, 1541.

[60] L. Polenta, M. Rossi, A. Cavallini, R. Calarco, M. Marso, R. Meijers,

T. Richter, T. Stoica, H. Luth, ACS Nano 2008, 2, 287.

[61] A. Cavallini, L. Polenta, M. Rossi, T. Richter, M. Marso, R. Meijers,

R. Calarco, H. Luth, Nano Lett. 2006, 6, 1548.

[62] V. Protasenko, S. Gordeyev, M. Kuno, J. Am. Chem. Soc. 2007,

129, 13160.

[63] J. Muller, J. M. Lupton, A. L. Rogach, J. Feldmann, D. V. Talapin, H.

Weller, Phys. Rev. Lett. 2004, 93, 167402.

[64] Y. Nagamune, H. Watabe, F. Sogawa, Y. Arakawa, Appl. Phys. Lett.

1995, 67, 1535.

H & Co. KGaA, Weinheim www.small-journal.com 1891

reviews R. Agarwal

1892

[65] O. I. Micic, J. Sprague, Z. Lu, A. J. Nozik, Appl. Phys. Lett. 1996, 68,

223150.

[66] D. V. Talapin, N. Gaponik, H. Borchert, A. L. Rogach, M. Haase, H.

Weller, J. Phys. Chem. B 2002, 106, 12659.

[67] T. M. Schmidt, Appl. Phys. Lett. 2006, 89, 123117.

[68] F. Wang, W. E. Buhro, J. Am. Chem. Soc. 2007, 129, 14381.

[69] L. Shi, Y. Xu, S. Hark, Y. Liu, S. Wang, L-m. Peng, K. Wong, Q. Li,

Nano Lett. 2007, 7, 3559.

[70] K. M. Ip, Z. Liu, C. M. Ng, S. K. Hark, Nanotechnology 2005, 16,

1144.

[71] E. Comini, C. Baratto, G. Faglia, M. Ferroni, G. Sberveglieri, J. Phys.

D 2007, 40, 7255.

[72] B. Tang, J. Niu, C. Yu, L. Zhuo, J. Ge, Chem. Commun. 2005,4184.

[73] D. J. Sirbuly, M. Law, P. Pauzauskie, H. Yan, A. V. Maslov, K.

Knutsen, C.-Z. Ning, R. J. Saykally, P. Yang, Proc. Natl. Acad. Sci.

USA 2005, 102, 7800.

[74] D. J. Sirbuly, M. Law, H. Q. Yan, P. D. Yang, J. Phys. Chem. B 2005,

109, 15190.

[75] D. J. Sirbuly, A. Tao, M. Law, R. Fan, P. Yang, Adv. Mater. 2007, 19,

61.

[76] H. Hasegawa, J. Korean Phys. Soc. 2007, 50, 543.

[77] R. Calarco, M. Marso, T. Richter, A. I. Aykanat, R. Meijers, A. V.

Hart, T. Stoica, H. Luth, Nano Lett. 2005, 5, 981.

[78] M. Diarra, Y. M. Niquet, C. Delerue, G. Allan, Phys. Rev. B 2007,

75, 045301.

[79] S. M. Sze, Physics of Semiconductor Devices, Wiley Interscience,

1981.[80] V. Schmidt, S. Senz, U. Gosele, Appl. Phys. A 2007, 86, 187.

[81] J. Wang, E. Polizzi, A. Ghosh, S. Datta, M. Lundstrom, Appl. Phys.

Lett. 2005, 87, 043101.

[82] B. C. Paul, S. Fujita, M. Okajima, T. H. Lee, H. S. P. Wong, Y. Nishi,

IEEE Trans. Electron Devices 2007, 54, 2369.

[83] R. Tu, L. Zhang, Y. Nishi, H. J. Dai, Nano Lett. 2007, 7, 1561.

[84] S. A. Dayeh, C. Soci, P. K. L. Yu, E. T. Yu, D. L. Wang, Appl. Phys.

Lett. 2007, 90, 162112.

[85] X. Zhou, S. A. Dayeh, D. Wang, E. T. Yu, J. Vac. Sci. Technol. B

2007, 25, 1427.

[86] H. D. Xiong, W. Y. Wang, Q. L. Li, C. A. Richter, J. S. Suehle, W. K.

Hong, T. Lee, D. M. Fleetwood, Appl. Phys. Lett. 2007, 91, 053107.

[87] P. Dutta, P. M. Horn, Rev. Mod. Phys. 1981, 53, 497.

[88] E. Gnani, S. Reggiani, M. Rudan, G. Baccarani, IEEE Trans.

Nanotechnol. 2007, 6, 90.

[89] D. W. Wang, Q. Wang, A. Javey, R. Tu, H. J. Dai, H. Kim, P. C.

McIntyre, T. Krishnamohan, K. C. Saraswat, Appl. Phys. Lett. 2003,

83, 122432.

[90] K. Kukli, M. Ritala, M. Leskela, T. Sajavaara, J. Keinonen, J. Mater.

Sci. 2003, 14, 361.

[91] F. C. Chiu, J. Appl. Phys. 2007, 102.

[92] H. Y. Yu, S. Z. Chang, A. Veloso, A. Lauwers, C. Adelmann, B.

Onsia, P. Lehnen, T. Kauerauf, S. Brus, K. M. Yin, P. Absil, S.

Biesemans, IEEE Electron Device Lett. 2007, 28, 957.

[93] K. Kukli, M. Ritala, M. Leskela, J. Electrochem. Soc. 2001, 148,

F35.

[94] W. Gao, J. F. Conley, Y. Ono, Appl. Phys. Lett. 2004, 84, 4666.

[95] V. Schmidt, H. Riel, S. Senz, S. Karg, W. Riess, U. Gosele, Small

2006, 2, 85.

[96] T. Bryllert, L. E. Wernersson, T. Lowgren, L. Samuelson, Nano-

technology 2006, 17, S227.

[97] M. T. Bjork, O. Hayden, H. Schmid, H. Riel, W. Riess, Appl. Phys.

Lett. 2007, 90, 142110.

[98] J. Goldberger, A. I. Hochbaum, R. Fan, P. D. Yang, Nano Lett. 2006,

6, 973.

[99] S. Ju, K. Lee, M. H. Yoon, A. Facchetti, T. J. Marks, D. B. Janes,

Nanotechnology 2007, 18, 155201.

[100] R. Konenkamp, R. C. Word, C. Schlegel, Appl. Phys. Lett. 2004, 85,

246004.

www.small-journal.com � 2008 Wiley-VCH Verlag Gm

[101] C. Y. Chang, F. C. Tsao, C. J. Pan, G. C. Chi, H. T. Wang, J. J. Chen, F.

Ren, D. P. Norton, S. J. Pearton, K. H. Chen, L. C. Chen, Appl. Phys.

Lett. 2006, 88, 173503.

[102] D. C. Olson, J. Piris, R. T. Collins, S. E. Shaheen, D. S. Ginley, Thin

Solid Films 2006, 496, 26.

[103] K. Takanezawa, K. Hirota, Q. S. Wei, K. Tajima, K. Hashimoto,

J. Phys. Chem. C 2007, 111, 7218.

[104] A. M. Peiro, P. Ravirajan, K. Govender, D. S. Boyle, P. O’Brien,

D. D. C. Bradley, J. Nelson, J. R. Durrant, J. Mater. Chem. 2006, 16,

2088.

[105] S. R. C. Vivekchand, K. C. Kam, G. Gundiah, A. Govindaraj, A. K.

Cheetham, C. N. R. Rao, J. Mater. Chem. 2005, 15, 4922.

[106] L. Wang, Y. S. Liu, X. Jiang, D. H. Qin, Y. Cao, J. Phys. Chem. C

2007, 111, 9538.

[107] W. U. Huynh, J. J. Dittmer, A. P. Alivisatos, Science 2002, 295,

2425.

[108] W. U. Huynh, J. J. Dittmer, W. C. Libby, G. L. Whiting, A. P.

Alivisatos, Adv. Funct. Mater. 2003, 13, 73.

[109] Y. Y. Xi, J. Z. Zhou, H. H. Guo, C. D. Cai, Z. H. Lin, Chem. Phys. Lett.

2005, 412, 60.

[110] K. Dutta, S. De, S. K. De, J. Appl. Phys. 2007, 101, 093711.

[111] S. Bhattacharya, S. Malik, A. K. Nandi, A. Ghosh, J. Chem. Phys.

2006, 125, 174717.

[112] S. H. Yu, M. Yoshimura, J. M. C. Moreno, T. Fujiwara, T. Fujino, R.

Teranishi, Langmuir 2001, 17, 1700.

[113] Q. Lai, Z. Li, L. Zhang, X. Li, W. F. Stickle, Z. Zhu, Z. Gu, T. I. Kamins,

R. S. Williams, Y. Chen, Nano Lett. 2008, 8, 876.

[114] F. Patolsky, C. M. Lieber, Mater. Today 2005, 8, 20.

[115] F. Patolsky, G. F. Zheng, O. Hayden, M. Lakadamyali, X. W.

Zhuang, C. M. Lieber, Proc. Natl. Acad. Sci. USA 2004, 101,

14017.

[116] Z. L. Wang, X. Y. Kong, Y. Ding, P. X. Gao, W. L. Hughes, R. S. Yang,

Y. Zhang, Adv. Funct. Mater. 2004, 14, 943.

[117] Z. L. Wang, J. H. Song, Science 2006, 312, 242.

[118] S. H. Lee, Y. Jung, R. Agarwal, Nat. Nanotechnol. 2007, 2,

626.

[119] S. H. Lee, D. K. Ko, Y. Jung, R. Agarwal, Appl. Phys. Lett. 2006, 89,

223116.

[120] Y. Jung, S. H. Lee, D. K. Ko, R. Agarwal, J. Am. Chem. Soc. 2006,

128, 14026.

[121] J. Y. Yu, S. W. Chung, J. R. Heath, J. Phys. Chem. B 2000, 104,

11864.

[122] Y. Cui, X. F. Duan, J. T. Hu, C. M. Lieber, J. Phys. Chem. B 2000,

104, 5213.

[123] Y. Cui, C. M. Lieber, Science 2001, 291, 851.

[124] Y. Huang, X. F. Duan, C. M. Lieber, Small 2005, 1, 142.

[125] Z. H. Zhong, F. Qian, D. L. Wang, C. M. Lieber, Nano Lett. 2003, 3,

343.

[126] O. Hayden, C. K. Payne, Angew. Chem. Int. Ed. 2005, 44, 1395.

[127] O. Hayden, R. Agarwal, C. M. Lieber, Nat. Mater. 2006, 5, 352.

[128] Y. Huang, X. F. Duan, Y. Cui, L. J. Lauhon, K. H. Kim, C. M. Lieber,

Science 2001, 294, 1313.

[129] Z. H. Zhong, D. L. Wang, Y. Cui, M. W. Bockrath, C. M. Lieber,

Science 2003, 302, 1377.

[130] R. Beckman, E. Johnston-Halperin, Y. Luo, J. E. Green, J. R. Heath,

Science 2005, 310, 465.

[131] J. H. Wei, K. S. Chan, Phys. Lett. A 2005, 341, 244.

[132] X. F. Duan, Y. Huang, R. Agarwal, C. M. Lieber, Nature 2003, 421,

241.

[133] J. M. Bao, M. A. Zimmler, F. Capasso, X. W. Wang, Z. F. Ren, Nano

Lett. 2006, 6, 1719.

[134] M. A. Zimmler, J. Bao, I. Shalish, W. Yi, V. Narayanamurti, F.

Capasso, Nanotechnology 2007, 18, 235205.

[135] R. S. Wagner, Growth of Whiskers by Vapor-Phase Reactions,

Wiley, New York 1970.[136] K. Haraguchi, T. Katsuyama, K. Hiruma, K. Ogawa, Appl. Phys.

Lett. 1992, 60, 060745.

bH & Co. KGaA, Weinheim small 2008, 4, No. 11, 1872–1893

Heterointerfaces in Semiconductor Nanowires

[137] M. S. Gudiksen, L. J. Lauhon, J. Wang, D. C. Smith, C. M. Lieber,

Nature 2002, 415, 617.

[138] Y. Y. Wu, R. Fan, P. D. Yang, Nano Lett. 2002, 2, 83.

[139] D. Y. Li, Y. Wu, R. Fan, P. D. Yang, A. Majumdar, Appl. Phys. Lett.

2003, 83, 153186.

[140] C. Yang, Z. H. Zhong, C. M. Lieber, Science 2005, 310, 1304.

[141] G. S. Cheng, A. Kolmakov, Y. X. Zhang, M. Moskovits, R. Munden,

M. A. Reed, G. M. Wang, D. Moses, J. P. Zhang, Appl. Phys. Lett.

2003, 83, 1578.

[142] C. Yang, C. J. Barrelet, F. Capasso, C. M. Lieber, Nano Lett. 2006, 6,

2929.

[143] M. T. Bjork, B. J. Ohlsson, C. Thelander, A. I. Persson, K. Deppert,

L. R. Wallenberg, L. Samuelson, Appl. Phys. Lett. 2002, 81,

234458.

[144] M. T. Bjork, C. Thelander, A. E. Hansen, L. E. Jensen, M. W.

Larsson, L. R. Wallenberg, L. Samuelson, Nano Lett. 2004, 4,

1621.

[145] A. I. Persson, M. W. Larsson, S. Stenstrom, B. J. Ohlsson, L.

Samuelson, L. R. Wallenberg, Nat. Mater. 2004, 3, 677.

[146] C. Thelander, T. Martensson, M. T. Bjork, B. J. Ohlsson, M. W.

Larsson, L. R. Wallenberg, L. Samuelson, Appl. Phys. Lett. 2003,

83, 2052.

[147] M. T. Bjork, A. Fuhrer, A. E. Hansen, M. W. Larsson, L. E. Froberg, L.

Samuelson, Phys. Rev. B 2005, 72.

[148] S. De Franceschi, J. A. van Dam, E. P. A. M. Bakkers, L. F. Feiner, L.

Gurevich, L. P. Kouwenhoven, Appl. Phys. Lett. 2003, 83, 02344.

[149] C. Thelander, H. A. Nilsson, L. E. Jensen, L. Samuelson, Nano Lett.

2005, 5, 635.

[150] H. A. Nilsson, C. Thelander, L. E. Froberg, J. B. Wagner, L.

Samuelson, Appl. Phys. Lett. 2006, 89, 163101.

[151] D. Wallin, A. Fuhrer, L. E. Froberg, L. Samuelson, H. Q. Xu, S.

Hofling, A. Forchel, Appl. Phys. Lett. 2007, 90, 172112.

[152] R. Solanki, J. Huo, J. L. Freeouf, B. Miner, Appl. Phys. Lett. 2002,

81, 203864.

[153] N. Panev, A. I. Persson, N. Skold, L. Samuelson, Appl. Phys. Lett.

2003, 83, 2238.

[154] P. J. Poole, J. Lefebvre, J. Fraser, Appl. Phys. Lett. 2003, 83, 2055.

[155] W. I. Park, G. C. Yi, M. Kim, S. J. Pennycook, Adv. Mater. 2003, 15,

526.

[156] W. I. Park, S. J. An, J. L. Yang, G. C. Yi, S. Hong, T. Joo, M. Kim, J.

Phys. Chem. B 2004, 108, 15457.

[157] T. Yatsui, M. Ohtsu, J. Yoo, S. J. An, G. C. Yi, Appl. Phys. Lett. 2005,

87, 033101.

[158] T. Yatsui, S. Sangu, T. Kawazoe, M. Ohtsu, S. J. An, J. Yoo, G. C. Yi,

Appl. Phys. Lett. 2007, 90.

[159] H. M. Kim, Y. H. Cho, H. Lee, S. I. Kim, S. R. Ryu, D. Y. Kim, T. W.

Kang, K. S. Chung, Nano Lett. 2004, 4, 1059.

[160] C. Kim, W. I. Park, G. C. Yi, M. Kim, Appl. Phys. Lett. 2006, 89,

113106.

[161] M. T. Borgstrom, V. Zwiller, E. Muller, A. Imamoglu, Nano Lett.

2005, 5, 1439.

small 2008, 4, No. 11, 1872–1893 � 2008 Wiley-VCH Verlag Gmb

[162] E. D. Minot, F. Kelkensberg, M. van Kouwen, J. A. van Dam, L. P.

Kouwenhoven, V. Zwiller, M. T. Borgstrom, O. Wunnicke, M. A.

Verheijen, E. P. A. M. Bakkers, Nano Lett. 2007, 7, 367.

[163] M. A. Hines, P. Guyot-Sionnest, J. Phys. Chem. 1996, 100, 468.

[164] A. B. Greytak, L. J. Lauhon, M. S. Gudiksen, C. M. Lieber, Appl.

Phys. Lett. 2004, 84, 4176.

[165] E. Tutuc, J. Appenzeller, M. C. Reuter, S. Guha, Nano Lett. 2006, 6,

2070.

[166] J. Goldberger, R. R. He, Y. F. Zhang, S. W. Lee, H. Q. Yan, H. J. Choi,

P. D. Yang, Nature 2003, 422, 599.

[167] H. J. Fan, M. Knez, R. Scholz, K. Nielsch, E. Pippel, D. Hesse, M.

Zacharias, U. Gosele, Nat. Mater. 2006, 5, 627.

[168] W. Lu, J. Xiang, B. P. Timko, Y. Wu, C. M. Lieber, Proc. Natl. Acad.

Sci. USA 2005, 102, 10046.

[169] N. Skold, L. S. Karlsson, M. W. Larsson, M. E. Pistol, W. Seifert, J.

Tragardh, L. Samuelson, Nano Lett. 2005, 5, 1943.

[170] W. I. Park, J. Yoo, D. W. Kim, G. C. Yi, M. Kim, J. Phys. Chem. B

2006, 110, 1516.

[171] E. S. Jang, J. Y. Bae, J. Yoo, W. I. Park, D. W. Kim, G. C. Yi, T. Yatsui,

M. Ohtsu, Appl. Phys. Lett. 2006, 88, 023102.

[172] J. Y. Bae, J. Yoo, G. C. Yi, Appl. Phys. Lett. 2006, 89, 173114.

[173] Z. Zanolli, M. E. Pistol, L. E. Froberg, L. Samuelson, J. Phys.

Condens. Matter 2007, 19, 295218.

[174] H. J. Choi, J. C. Johnson, R. R. He, S. K. Lee, F. Kim, P. Pauzauskie, J.

Goldberger, R. J. Saykally, P. D. Yang, J. Phys. Chem. B 2003, 107,

8721.

[175] F. Qian, Y. Li, S. Gradecak, D. L. Wang, C. J. Barrelet, C. M. Lieber,

Nano Lett. 2004, 4, 1975.

[176] F. Qian, S. Gradecak, Y. Li, C. Y. Wen, C. M. Lieber, Nano Lett.

2005, 5, 2287.

[177] M. W. Larsson, J. B. Wagner, M. Wallin, P. Hakansson, L. E.

Froberg, L. Samuelson, L. R. Wallenberg, Nanotechnology

2007, 18, 015504.

[178] M. T. Borgstrom, M. A. Verheijen, G. Immink, T. de Smet, E. P. A. M.

Bakkers, Nanotechnology 2006, 17, 4010.

[179] M. A. Verheijen, R. E. Algra, M. T. Borgstrom, G. Immink, E. Sourty,

W. J. P. van Enckevort, E. Vlieg, E. P. A. M. Bakkers, Nano Lett.

2007, 7, 3051.

[180] A. Gustafsson, M. T. Bjork, L. Samuelson, Nanotechnology 2007,

18, 205306.

[181] J. H. Chung, L. Rabenberg, Appl. Phys. Lett. 2006, 88, 013106.

[182] J. Eymery, F. Rieutord, V. Favre-Nicolin, O. Robach, Y. M. Niquet, L.

Froberg, T. Martensson, L. Samuelson, Nano Lett. 2007, 7, 2596.

[183] L. Ouattara, A. Mikkelsen, N. Skold, J. Eriksson, T. Knaapen, E.

Cavar, W. Selfert, L. Samuelson, E. Lundgren, Nano Lett. 2007, 7,

2859.

[184] D. E. Perea, J. E. Allen, S. J. May, B. W. Wessels, D. N. Seidman, L. J.

Lauhon, Nano Lett. 2006, 6, 181.

H & Co. KGaA, Weinheim

Received: April 18, 2008Published online: October 17, 2008

www.small-journal.com 1893