Next Article in Journal
Polyvinyl Alcohol-Few Layer Graphene Composite Films Prepared from Aqueous Colloids. Investigations of Mechanical, Conductive and Gas Barrier Properties
Next Article in Special Issue
Development of Monolithically Grown Coaxial GaInN/GaN Multiple Quantum Shell Nanowires by MOCVD
Previous Article in Journal
Hybrid Electro-Optical Pumping of Active Plasmonic Nanostructures
Previous Article in Special Issue
CPP-GMR Performance of Electrochemically Synthesized Co/Cu Multilayered Nanowire Arrays with Extremely Large Aspect Ratio
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photoluminescence of ZnO Nanowires: A Review

by
Andres Galdámez-Martinez
1,
Guillermo Santana
1,*,
Frank Güell
2,
Paulina R. Martínez-Alanis
2 and
Ateet Dutt
1,*
1
Instituto de Investigaciones en Materiales, Universidad Nacional Autónoma de México, Coyoacan, México City 04510, Mexico
2
ENFOCAT-IN2UB, Universitat de Barcelona, C/Martí i Franquès 1, 08028 Barcelona, Catalunya, Spain
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(5), 857; https://doi.org/10.3390/nano10050857
Submission received: 18 February 2020 / Revised: 6 March 2020 / Accepted: 11 March 2020 / Published: 29 April 2020
(This article belongs to the Special Issue Growth and Characterization in Nanowires)

Abstract

:
One-dimensional ZnO nanostructures (nanowires/nanorods) are attractive materials for applications such as gas sensors, biosensors, solar cells, and photocatalysts. This is due to the relatively easy production process of these kinds of nanostructures with excellent charge carrier transport properties and high crystalline quality. In this work, we review the photoluminescence (PL) properties of single and collective ZnO nanowires and nanorods. As different growth techniques were obtained for the presented samples, a brief review of two popular growth methods, vapor-liquid-solid (VLS) and hydrothermal, is shown. Then, a discussion of the emission process and characteristics of the near-band edge excitonic emission (NBE) and deep-level emission (DLE) bands is presented. Their respective contribution to the total emission of the nanostructure is discussed using the spatial information distribution obtained by scanning transmission electron microscopy−cathodoluminescence (STEM-CL) measurements. Also, the influence of surface effects on the photoluminescence of ZnO nanowires, as well as the temperature dependence, is briefly discussed for both ultraviolet and visible emissions. Finally, we present a discussion of the size reduction effects of the two main photoluminescent bands of ZnO. For a wide emission (near ultra-violet and visible), which has sometimes been attributed to different origins, we present a summary of the different native point defects or trap centers in ZnO as a cause for the different deep-level emission bands.

Graphical Abstract

1. Introduction

Similar to many binary compounds, zinc oxide (ZnO) has been widely used in the last century in a large number of daily applications such as cosmetics (UV screening), antibacterial applications, pigments, catalysts, piezoelectric transducers, and so on [1]. However, as the scientific community has focused so much attention on the synthesis and applications of semiconductor nanostructures, in the last few decades, ZnO nanostructures have gained much popularity due to their belonging to the rich family of possible nanostructures with tunable morphology and properties [2].
ZnO is a II-VI oxide semiconductor with a strong ionic character of chemical bonds [3]. Under ambient conditions, ZnO has a wurtzite crystalline structure with hexagonal unit cell and two lattice parameters, a and c with c/a ratio between 1.5393 and 1.6035 that is transformed into a metastable rock salt structure at high pressures [4]. With an exciton binding energy of 60 meV and wide direct bandgap of 3.4 eV at room temperature [5], ZnO can be considered an alternative to GaN for the construction of short-wavelength optoelectronic devices such as light-emitting diodes, UV lasers, and UV detectors, among other [6].
ZnO presents several advantages over GaN in these kinds of applications, not only because of its lower production cost but also its excellent radiation resistance [7], high electron mobility [8], greater exciton binding energy, and the ability to grow crystal structures. Nonetheless, the high performance and much more mature technology based on GaN puts severe limitations on the commercial use of ZnO in these applications.
Also, ZnO is a unique material that, besides its optoelectronic semiconductor behavior, exhibits exciting properties such as piezoelectricity [9] and pyroelectricity [10], which further make it a versatile, functional material.
As mentioned before, a wide diversity of different nanostructure morphologies can be achieved with ZnO using chemical or physical methods including thermal evaporation [11], chemical vapor deposition (CVD) [12,13], sol-gel method [14], hydrothermal [15], molecular beam epitaxy (MBE) [16], metalorganic vapor-phase epitaxy (MOVPE) [17], and solvothermal growth [18]. Among the most reported ZnO nanostructures are zero dimensional systems 0D (nanoparticles NPs [19]), one-dimensional (1D; nanowires NWs [20], nanorods NRs [21], nanotubes NTs [22]), two-dimensional (2D; nanobelts [23], thin films [24]), and more complex morphologies such as nanohelix [25], nanorings [26], nanocages [27], nanoflowers [28], or comb-like structures [29].
Among the above-mentioned nanostructures, nanowires and nanorods represent a significant and broad class of 1D nanostructures and, therefore, are at the forefront of nanoscience and nanotechnology. NWs and NRs can present a polycrystalline or monocrystalline structure and highly anisotropic properties, which could result in rapid growth in one particular direction [20,21]. NWs typically have a cylindrical, hexagonal, square, or triangular cross-section, which is much smaller than the overall length of these 1D nanostructures.
Such ZnO nanostructures have attracted much attention due to excellent charge carrier transport properties and high crystalline quality, which are ideal for optoelectronic and magnetic applications. Recent work on ZnO has demonstrated ferromagnetism properties by doping with transition metals for spintronic devices [30]. Also, detection and photocatalytic applications have been extensively studied due to the large surface area and volume ratio and the ability to adjust the bandgap through substitution with suitable dopants [31,32].
As applications of 1D ZnO nanostructures often rely on the optoelectronic properties of the system, a deep understanding of the mechanisms of light interaction (which are heavily influenced by the lattice dynamics of the system [33]) is desired.
As mentioned before, the literature on ZnO 1D nanostructures is vast; therefore, there are many reviews describing the different synthesis methods [34,35], optical [36] and electrical [37] properties, and practical applications of ZnO nanowires [38,39] and nanorods [40], including photocatalyst material [41], sensing platforms (chemical [42] and biological [43,44]), antibacterial element [45], light emissions sources(LEDs [46] and lasers [47]), dye-sensitized or thin film solar cells [48], field effect transistors [49], and piezoelectric nanogenerators [50].
Furthermore, their non-toxic nature makes ZnO nanowires a good candidate for biological applications such as urea, glucose, or DNA biosensors [44] and electric cardiac biomarker detection [43]. Also, the reported antibacterial activities of ZnO nanostructures have inspired a wide range of antimicrobial applications.
However, to our best knowledge, review reports are lacking where the photoluminescence properties and emission mechanism of ZnO nanowires have been discussed. For that reason, this paper presents an extensive review of the latest research of the luminescence properties of ZnO nanowires. A discussion of the sometimes controversial or contradictory light-emitting mechanisms in both near-band edge excitonic emission (NBE) and deep-level emission (DLE) bands is made as well as their respective contribution to nanoscale systems. Lastly, a discussion of the size reduction effects over the two main photoluminescent bands of ZnO is presented. For a broad emission (near UV and visible), which has sometimes been attributed to different origins, we present a summary of the different native point defects or trap centers in ZnO as a cause for the different deep-level emission bands.

2. ZnO Structure and Main Growth Techniques

2.1. Crystalline Structure of ZnO

Under ambient pressure and temperature, ZnO crystallizes in the hexagonal wurtzite structure (Figure 1), where each Zn2+ ion is surrounded by four O2− ions in tetrahedral coordination (an indicator of the presence of a covalent sp3 bond) and vice versa. This hexagonal structure belongs to the C6V4 class (Schoenflies notation) and the space group P63mc (Hermann–Mauguin notation). The spatial distribution of the interpenetrated Zn2+ and O2− tetrahedrons and the mixed ionic-covalent character of the Zn-O bond (with nearly equal contribution from each bonding type) are responsible for the intrinsic polarity in the ZnO crystal along the c-axis.
This polarity is responsible for the piezoelectric and spontaneous polarization properties of ZnO and is also a critical factor in crystalline growth and defect generation [41]. This feature gives rise to some interesting chemical reactions that are possible with this semiconductor oxide.
The ZnO hexagonal lattice has four characteristic crystal faces, and the c-axis orientated ones are the Zn-terminated [ 0001 ] and the O-terminated [ 000 1 ¯ ] , both exhibiting a polar character. The non-polar [ 11 2 ¯ 0 ] and [ 10 1 ¯ 0 ] are a-axis oriented and have the same number of Zn and O atoms [51]. Despite the inherent instability of the polar faces, the most common crystal orientations of the ZnO structure correspond to the [ 0001 ] and [ 000 1 ¯ ] faces; therefore, additional positive or negative charges are required to stabilize the surface [28].
The lattice parameters determined at room temperature by X-ray diffraction (XRD) are a = 3.25 Å and c = 5.20 Å with the ratio c/a = 1.6 in a perfect wurtzite structure. The values of the a-parameter usually range from 3.2475 to 3.2501 Å and from 5.2042 to 5.2075 Å for the c-parameter. The variations in the lattice parameters are the result of doping, defects, strain, or temperature effects [52].
On the other hand, at high pressures, a halite (rock salt cubic lattice) crystal structure can be achieved. In this phase, the ZnO exhibits an indirect semiconductor character with a bandgap of 2.7 eV [53].

2.2. ZnO Nanowires

One-dimensional semiconductor nanostructures such as nanowires or nanorods have received significant attention due to their versatile set of applications, among which nanoelectronic devices, photovoltaic systems, and biosensors are found. Furthermore, the valuable properties of these kinds of nanomaterials, including a direct bandgap and large exciton binding energy, make them a promising candidate for the fabrication of efficient optoelectronic nanodevices.
Several chemical and physical methods have been used for the synthesis of ZnO nanowires. In most of the reported methodologies, a bottom-up approach is preferred due to better control in the morphological parameters such as the aspect ratio, orientation, defects, surface density, and the compositional flexibility of obtaining heterostructures [4]. These properties have a significant influence over the efficiency and performance of the diverse applications of the nanowires. Among the most used growth methods are physical deposition techniques such as sputtering, atomic layer deposition, or chemical ones as spray pyrolysis, sol-gel, electrochemical deposition, hydrothermal technology, or all-inkjet-printing approach.
The linear growth of ZnO NWs on Zn metal substrates is also quite promising due to its ease and control over the growth process [54]. A more detailed exposition of these different growth techniques can be found in the work of Rong et al. [35]. However, among these techniques, the vapor-liquid-solid technique (VLS) and hydrothermal technique are the most commonly used bottom-up ZnO nanowires growth processes [49]. These two techniques are discussed below.

2.3. Vapor-Liquid-Solid Technique

In a CVD process, the precursors are applied to the heated surface of a substrate where specific reactions take place for the formation of the desired material. For the synthesis of nanostructures, the process is performed on a substrate with previously deposited catalyst metal nanoparticles.
Accordingly, due to the phases presented in the system, the growth techniques can be divided into two groups: vapor-liquid-solid (VLS) and vapor-solid-solid (VSS) mechanisms [39].
In the early 1960s, Wagner and Ellis developed the VLS technique to obtain micrometric whiskers. A standard VLS process starts with the diffusion of the reagent gas species in a nanometric “liquid” metal particle (e.g., Au, Cu, Ni, and Sn are employed as catalyst metals [55]) forming a eutectic with the material for the formation of the nanowire. The continuous absorption of the gaseous reagents into the catalyst droplet contributes to a super-saturated solution resulting in the nucleation, crystallization, and growth of a single nanowire from each drop (Figure 2a,b).
By changing the gaseous reagents in the system, it is possible to obtain both radial or axial heterostructures, which makes this technique quite flexible [56]. The first ZnO nanowires with perfect vertical alignment to the substrate were reported by Huang [57], where a laser emission of the system was also found under UV at room temperature. However, to obtain ZnO nanowires with proper alignment, a low growth rate (5.5 nm/s) regime is required [58].
The alignment of the ZnO nanowires is strongly related to the epitaxial relationship between ZnO and the substrate. Small variations between the ZnO crystallographic network and some materials have allowed NWs to be obtained with excellent alignment on sapphire substrates [57] and several nitrides (GaN, AlGaN, AlN) [59].
Figure 2. (a) Basic principle of the vapor-liquid-solid (VLS) growth mechanism of ZnO NWs using gold as a metal catalyst. (b) 45° side-view SEM image of hexagonally ordered ZnO nanorod grown by VLS method on patterned Au-covered substrates. (c) Schematic representation of the growth of nanowires by the hydrothermal technique [60]. Copyright 2015 Elsevier.
Figure 2. (a) Basic principle of the vapor-liquid-solid (VLS) growth mechanism of ZnO NWs using gold as a metal catalyst. (b) 45° side-view SEM image of hexagonally ordered ZnO nanorod grown by VLS method on patterned Au-covered substrates. (c) Schematic representation of the growth of nanowires by the hydrothermal technique [60]. Copyright 2015 Elsevier.
Nanomaterials 10 00857 g002
Even though the growth process in the VLS technique is relatively simple, the high number of experimental parameters makes the method non-trivial. The source temperature, substrate temperature, growth time, carrier gas concentration and flow, the type of substrate, and the characteristics of the catalyst are among the leading parameters to control for precise growth of ZnO nanowires [61]. In particular, the total chamber pressure, partial oxygen pressure, catalyst thickness, and the growth temperature are the most critical parameters of the process [4].

2.4. Hydrothermal Technique

Although the VLS growth technique allows obtaining nanostructures with an excellent crystalline structure, the high temperatures at which the growth must be performed present a disadvantage when using organic substrates, which could lead to the manufacture of flexible devices. Due to this, different methods of fabrication of ZnO NWs from solutions have been investigated [62]. ZnO, as well as ordinary divalent metals, can crystallize through the hydrolysis of metal precursor salts (e.g., ZnAc dihydrate, zinc nitrate, among others) in a basic solution (e.g.,potassium hydroxide (KOH )and sodium hydroxide (NaOH)). The temperature and pH of the solution determine the formation of tetrahedrally coordinated Zn2+ ions, whose dehydration (by proton transfer) will eventually lead to the growth of ZnO nanostructure. The minimization of the energy of the system is considered the driving force for crystalline growth, while the reversible equilibrium determines the chemical reactions in the aqueous system. The (0001) ZnO polar surface has a strong affinity to attract and absorb ions, creating opposite Zn and O terminal surfaces, which lead to the anisotropic growth of the nanowires.
Since chemical reactions take place in an aqueous medium, the term hydrothermal is frequently used to describe the process.
The typical growth stages of the hydrothermal process are shown in Figure 2c. First, a seed layer of ZnO nanoparticles must be deposited on the substrate. To this end, a thin film of ZnO commonly deposited by spin coating is subdued to a thermal decomposition (>200 °C) to form the ZnO islands. This seed layer facilitates the nucleation of ZnO nanowires by decreasing the thermodynamic barrier for the formation. In addition to the already mentioned nutrient solution (alkaline solution containing a Zn2+), a guiding agent (e.g., polyethyleneimine (PEI)) is used in order to decrease the lateral growth of the NWs and maximize their length. The substrate with the ZnO seed layer is kept in the nutrient solution at a constant temperature to ensure stable growth regimes. Finally, heat treatment removes the organic residues from the ZnO surface.
In the same way as for the VLS method, in order to obtain NW arrangements with controlled morphology, orientation, surface density, and aspect ratio, the parameters involved in growth must be optimized. Some of the parameters affecting the morphology of the NWs are the reagents, seed layer characteristics, growth time, bath temperature, pH of the solution, guiding agents, and mechanical agitation [4].
A summary of different synthesis methods of 1D ZnO nanostructures and reported photoluminescent centers is compared and presented in Table 1. In this table, we present some of the most relevant works on the photoluminescence properties of ZnO nanowires in recent years. Some older works are also included due to their relevance in the area and for a better comparison.

3. Properties of ZnO

3.1. Optoelectronic Properties of ZnO

Broadband metal oxide semiconductors have characteristic properties, such as larger bandgap, better electron mobility, and proper light absorption. As a result, they are suitable for high temperature and power electronic devices and short UV optoelectronics. Within this family of broadband semiconductors, ZnO is found as an appropriate material with many favorable properties for these kinds of applications. Also, its excellent resistance to radiation makes it a promising candidate for space and nuclear applications.
The direct valence band of the wurtzite ZnO structure is split into three states Figure 3a, commonly called A (also called the light hole band), B (also called the heavy hole band), and C (also called the crystal-field split-band) sub-bands, due to crystal field and spin-orbit interactions [63].
Table 1. 1D ZnO nanostructures: morphology, methods of preparation, size, and photoluminescence emission.
Table 1. 1D ZnO nanostructures: morphology, methods of preparation, size, and photoluminescence emission.
ZnO NanostructureSynthesis MethodGrowth ReagentsSubstrateCharacteristic SizePL (eV)
NBEDLE
Randomly orientated NWs [64]Catalyst-free chemical vapor deposition (CVD)ZnOGraphene quantum dotsd = 10–30 nm3.26 Free excitonCentered 2.48
O vacancy
Zn interstitial
O interstitial
_
Nanowire [65]Chemical vapor depositionMetallic zincSilicond = 50–200 nm3.263
Near-band edge electron transition
1.8, 2.2 and 2.5
O vacancies
Zn vacancies
l = 6–8 μm
Nanowire [66]Preheating hydrothermalZinc nitrate hexahydrateGlass coated with ZnO thin filmd = 60–80 nm3.36–3.27 D0X, A0X, two-
electron satellite (TES), donor-acceptor pairs (DAP)
1.5–2.5 Absorbed hydroxyl group
l = 10 μm
Nanowire [15]Low temperature hydrothermalZinc chloride, Zinc nitrate hexahydrate, Zinc acetate dihydrate-d = 43–62 nm3.2982.066-1.937
l = 1.1–3.3 μm
Nanowire [67]Chemical vapor transport (CVT)Zinc gas precursor
Gold as catalyst
α sapphired = 100 nm3.2632.48
l = 15–20 μm
Nanowire [68]Vapor-liquid-solidZnO, graphiteQuartz, sapphire, and SiO2/Si substratesd = 40–200 nm3.38-3.332.81-1.55
O vacancy, Au impurities, Zn interstitial
l = 0.2–2.0 μm
Nanowire [69]ElectrochemicallyZinc nitrateZinc crystald = 170 nm3.172.43
2.17
2.00
l = 2.3 μm
Nanorods [70]Vapor phase techniqueZn powderglass substratesd = 50 nm3.25
band-to-band transitions bound excitons
2.5
Zn vacancies
l = 0.8–1.0 μm
Nanowire [71]Hydrothermal methodAmmonium persulphate, sodium hydroxideZn foild = 10–250 nm3.2632.851 Electronic defect
l > 10 μm
Nanowire [72]Reactive magnetron sputteringZnO targetPaper and PET substratesd = 20–40 nm3.36 D0X2.30
-
Nanowire [61,73]Vapor–liquid–solidZnO; graphiteAZO filmd = 30–70 nm
l = 7 μm
3.28-
3.25,
3.09
2.49-2.13
O vacancy
Zn vacancy
Nanowire [74]Catalyzed vapor–liquid–solid (VLS) using pulsed laserZnO ceramic targetAu-coated Al2O3 (0001)d = 70 nm
l = 8 μm
3.2632.38 Native point defect
Nanowire [75]Hydrothermal methodZinc acetate dihydrateITO coated glassd = 90 nm
l = 630 nm
3.2632.361-1.984 O vacancy
Nanowire [76]Plasma-assisted molecular beam epitaxy (PAMBE)Oxygen plasma and Zn vaporc-plane sapphire (0001) substrated = 50–100 nm3.42.1 O vacancy 1.8 Zn interstitial
l = 0.5–2.0 μm
Flower-like nano [77]Ultrasonic assisted hydrothermal methodZinc acetate dihydrate-nano-scale diameter
micro-scale length
-3.02 Zn vacancy 2.33 O vacancy 2.19 O interstitial
Nanowire [78]Vapor–solid Vapor–liquid–solidZnO, graphiteSilicon wafersd= 15–80 nm3.352.34, 2.57 O and Zn vacancies, ionized Zn interstitial
Nanowire [79]Metal–vapor deposition methodMetal zinc powderAlumina substratel several μm3.26 Free exciton2.44 singly ionized oxygen vacancy
d = 55 nm
However, it is possible to make measurements of the excitonic band of the material with a resolution of 40 μeV; the symmetry associated with the sub-band A (Г9 or Г7) remained an object of debate in the scientific community for more than 50 years [33]. However, recent studies of the magneto-optic exciton fine structure made by Ding et al. confirmed the Г7 the symmetry of the valence band according to the unique selection rules of Zeeman splitting lines [80]. There have been some comprehensive studies to establish the refractive index and dielectric constant of ZnO [81,82,83]. In the work of Yoshikawa et al. [81], the high frequency and static dielectric constants were measured using spectroscopic ellipsometry. The measurements were carried out on the surface parallel to the c axis, which allowed them to determine the perpendicular (Ec) and parallel (E||c) components of the optical constants. The values for the dielectric constants are ε = 3.68 and εs = 7.40 for Ec (polarization perpendicular to c-axis) and ε = 3.72 and εs = 8.49for E||c (parallel to c-axis). In the work of Gencyilmaz, the refractive index and extinction coefficient for the wurtzite structure of ZnO were calculated by using the envelope method, and the results are shown in Figure 3b [84]. The electron mobility in a single ZnO crystal measured by Hall effect at room temperature is 239 cm2 V−1s−1 and exhibits an N-type conductivity [85]. Characteristic photoluminescence of n-type bulk ZnO is shown (Figure 3c), measured at 4.2 K using a 325 nm He-Cd laser [33]. The emission of the material spans the UV to the green/orange spectral range. In the spectrum, the lines originated from bound exciton (BE), longitudinal optical (LO) phonon replicas, and also, donor-acceptor-pair (DAP) transitions are presented in the UV region. Also, the defect-induced emission band centered around 2.45 eV extending from the blue to the green range is presented.
Figure 3. (a) Band structure splitting of the hexagonal ZnO structure caused by the crystal field and spin-orbit splitting [63]. Copyright 2004 John Wiley and Sons (b) Refractive index (n) and extinction coefficient (k) of bulk ZnO obtained by the envelope method [84]. Copyright 2012 AIP Publishing. (c) Photoluminescence spectrum of bulk ZnO obtained with He-Cd excitation. Excitonic, donor-acceptor pair (DAxP), and longitudinal optical phonon replicas (LO) are indicated [33]. Copyright 2004 John Wiley and Sons.
Figure 3. (a) Band structure splitting of the hexagonal ZnO structure caused by the crystal field and spin-orbit splitting [63]. Copyright 2004 John Wiley and Sons (b) Refractive index (n) and extinction coefficient (k) of bulk ZnO obtained by the envelope method [84]. Copyright 2012 AIP Publishing. (c) Photoluminescence spectrum of bulk ZnO obtained with He-Cd excitation. Excitonic, donor-acceptor pair (DAxP), and longitudinal optical phonon replicas (LO) are indicated [33]. Copyright 2004 John Wiley and Sons.
Nanomaterials 10 00857 g003

3.2. Photoluminescence of ZnO Nanowires

Since the synthesis of crystals of ZnO can be carried out by different methodologies, obtained systems can be from tens of nanometers to a few millimeters; therefore, the optical properties can vary enormously.
The luminescence properties of single and collective ZnO nanostructures have been intensively studied using photoluminescence (PL) spectroscopy. The typical photoluminescence spectrum of the nanostructured zinc oxide (ZnO) shows its band edge and exciton luminescence in the UV region. It also presents a green-centered broadband (GL) commonly related to deep-level defects [46]. The first of these bands is usually reported in the literature as NBE (near-band edge excitonic), while the second is known as the DLE band (deep-level emission) [86]. The sizeable binding exciton energy of the ZnO (60 meV, nearly three times that of gallium nitride [87]) is responsible for the room temperature luminescence of the material and that it persists at temperatures as high as 700 K [88]. Usually, the ratio between the integrated spectral intensity of NBE and DLE bands is used to estimate the contribution of recombination due to defect levels [61].
Nonetheless, in ZnO nanostructures, the interaction between these bands becomes more complex, where the high surface-volume ratios result in an increase in surface defects, which could finally affect the processes of photoluminescence emission.

3.2.1. Near-Band Edge Excitonic Emission (NBE)

ZnO is a bright emitter in the UV region, so it is not hard to wonder why there are numerous photoluminescence investigations of the near-band gap recombinations. The ZnO UV emission (also called near-band emission, NBE) is located close to its absorption edge (3.35 eV [33]) and is produced by excitonic or band-band recombinations. Bound excitons are extrinsic transitions and are related to dopants, native defects, or complexes, which usually create discrete electronic states in the bandgap [52] and at low temperatures are the dominant radiative channels [51]. When the ZnO structure has high crystallinity, the UV emission is more intense than the emission in the visible region [89].
In the work of Qiu et al. [66], a detailed study of the near-band excitonic emission of ultralong ZnO nanowires is presented. The nanostructures were synthesized using a hydrothermal preheating technique, obtaining nanowires as long as 10 μm with a diameter varying from 50 to 80 nm. In Figure 4, the position of the different exciton transitions reported in [66] is schematized.
Although it is known that the NBE emission peak is composed of several peaks and shoulders of less intensity, in many works, its study is carried out indistinctly, analyzing only the sum of these emissions.
However, some studies focus on high-resolution photoluminescence in the area of output emission in detail due to the high consistency that exists between the measured values with respect to those previously reported [33,66,72,90,91,92,93,94].
In the 6K spectra reported by Qiu et al. [66], three peaks can be observed at 3.363, 3.303, and 3.228 eV, and another four shoulders at 3.367, 3.357, 3.353, and 3.342 eV. In high bandgap materials such as ZnO, excitons can be bound to neutral or charged donors and acceptors, due to the strong coupling between them [52]. The central peak at 3.363 and its two lateral shoulders (3.367 and 3.357) are tentatively ascribed (based in the previous work of Teke et al. [95]) to the exciton bound to neutral donors and are labeled as D0X/I4, D0X/I2, and D0X/I9, respectively. These notations come from the early characterization of bulk ZnO, where 12 transitions labeled from I0 to I11, were identified and form the basic set of bound exciton transitions characteristic of ZnO [90]. For the shoulder presented in 3.353 eV, Qiu associates an exciton bound to a neutral acceptor denoted as A0X [95]. For the shoulder in 3.342 eV, the authors associate a two-electron satellite transition (TES) [52]. Such recombinations are commonly found in the 3.32–3.34 eV spectral region and involve the radiative recombination of an exciton bound to a neutral donor, leaving the donor in the excited state (2s, 2p states). The TES transitions are red-shifted with respect to the DBE energy by the difference between the first excited (2s, 2p) and ground states (1s) of the donor [96]. An identification of the different species in the material can be made to determine the donor binding energies by using the energy difference between the ground state of the donor and relative excited TES states [97]. Finally, the last pair of peaks at 3.303 eV and 3.228 eV were identified as a donor-acceptor pair transitions (DAP) and first-order longitudinal optical (LO) replica (DAP-LO), respectively. The shift between these two peaks is near the theoretical LO phonon energy in ZnO (70–75 meV) [98]. The broadening of the NBE peak as the temperature is increased (Figure 4) is attributed to the thermal decomposition and ionization of bound excitons as well as to the strong coupling of phonons and excitons [66]. In the work of Cao [99], a low-temperature study of the NBE band of ZnO nanowires is presented. As in the work of Qiu, the photoluminescence intensity increases as the temperature is lowered to 4.2K as a consequence of temperature quenching effects [100]. The main peak they observed at low temperatures was blue-shifted from 3.281 to follow the type of tendency explained by Varshni relationship [101]. The phonon optical replicas (longitudinal LO and transversal TO) were also presented in the excitonic spectra and, as in the work of Qiu, also coincide with the theoretical phonon energies for bulk ZnO [98,102,103,104,105].
A time-resolved study of the NBE photoluminescence of ZnO nanowires with different sizes is presented in the work of Reparaz et al. [94], Figure 5.
They measured the lifetime of the neutral bound exciton at 3.365 eV as a function of the surface/volume ratio of the nanowires and demonstrated that the lifetime decreases for nanowires with smaller diameters. They assume that the systemic increase in the lifetime of the exciton as the nanowire diameter increases is due to surface-induced recombination, i.e., the surface proximity affects the lifetime of the D0X exciton. Also, the intensities and recombination dynamics of two broadband transitions in 3.341 (S1) and 3.314(S2) eV were studied. The increasing intensity of the I(S1)/I(D0X) and I(S2)/I(D0X) ratios when the nanowire diameter is decreased supports the hypothesis of surface states as the origin of these transitions. Also, free-electron to bound-hole transition (e, A0) has been assigned to transition S2 related to the intersection of basal plane stacking faults on the material’s surface [106].

3.2.2. Deep Level Emission (DLE)

As mentioned before, the photoluminescence (PL) of ZnO nanostructures exhibits a near-band-edge (NBE) and visible luminescence usually associated with exciton transitions and defect emissions, respectively, Figure 6a [107,108,109,110,111,112]. Several photoluminescence emission centers in the visible region are dependent on the synthesis technique and hence, on the vacancies, surface defects, and morphology of NWs [36]. However, the clear vision of the ZnO visible PL mechanism is still far from being fully understood [113]. The emissions in the region from 3.1 eV down to 1.653 eV are commonly referred to as deep-level emissions (DLE), Figure 6b. The deep levels denote the allowed levels inside the bandgap of the semiconductor that produces transitions with energy in the visible range of the spectrum. The band broadness presumably comes from a superposition of many different deep levels (yellow peak, green peak, and blue peak; Figure 7) that emit at the same time [114]. The green and yellow peaks have been observed in some ZnO systems [115,116,117,118], and the blue peak has been presented in thin films and nanostructured systems [119,120,121]. Some previous work attributed the green and orange luminescence to extrinsic impurities such as Cu, Li, or Fe [122]. Nonetheless, new research shows that undoped ZnO also presents photoluminescence peaks in the visible region and is generally ascribed to the electronic transition from single ionized VO+ centers to the valence band edge [123,124].
As many reports attribute the origin of these emissions to the presence of defects in the ZnO structure, a brief review of the most common defects in the ZnO lattice must be introduced. Defects in the ZnO structure can be divided into extrinsic and intrinsic levels [49]. The intrinsic, sometimes called native levels, are oxygen vacancy (VO), zinc vacancy (VZn), oxygen interstitial (Oi), zinc interstitial (Zni), oxygen anti-site (OZn), and zinc anti-site (ZnO) (Figure 6a) [125]. These native defects can exist in different charged states or a neutral state and may form clusters with another intrinsic defect as well as with extrinsic elements.
One example is the VOZni cluster, which is formed by an oxygen vacancy and zinc interstitial and is situated 2.16 eV below the conduction band minimum [136]. Also, these native defects are the most common explanation for the fact that ZnO always exhibits a high n-type conductivity [137]. Therefore, minority carrier lifetime, doping, compensation, and luminescence efficiency are frequently directly or indirectly determined by native point defects in semiconductors [138]. However, the reported position of these point defects inside the bandgap often varies [53]. A summary of the position of various energy levels inside the bandgap of ZnO due to intrinsic/extrinsic point defects in the lattice is presented in the work of Willander et al. [114].
Additionally, different widely reported levels are also summarized in this work in Figure 8. In order to understand the contribution of the different point defects in the photoluminescence, the formation process of these luminescence centers (traps) has been represented by different color codes in Figure 8 for the easy understanding of the readers.
According to Janotti et al. [107] at thermodynamic equilibrium and in no defect-defect interaction regime, the concentration of a point defect (c) depends on its formation energy as shown in the following equation:
c = Nsites(−Ef/kBT)
where Nsites represents the number of sites where the defect can be incorporated, kB is the Boltzmann constant, T is the temperature, and Ef is the formation energy of the specific point defect. It is essential to point out that the previous equation neglects the volume and formation of entropy contributions to the defect concentration that tends to become important at high pressures and temperatures. However, this equation shows that defects with high formation energies will be presented in lower concentrations as higher energy needs to be consumed in their formation.
The formation energy Ef of point defects in the ZnO lattice depends on the environment, which can be zinc-rich, oxygen-rich, or somewhere in between. Therefore, the abundance of oxygen and zinc atoms in the growth environment controls the concentration of point defects in ZnO. The formation of energy can also depend on the chemical potential (μ) if the defect state has an associate charge [139].
It is also important to note that the formation of energy is also a function of the morphology, as defect formation energies at the surface are lower than in bulk, resulting in high defect concentrations at the surface. As mentioned before, one of the characteristics of ZnO nanostructures is that they usually have a substantial surface-volume ratio. Then, the surface effects become dominant in the photoluminescence of ZnO nanowires [113].
Another surface effect is the absorption of gas or determined biomolecules on the surface of ZnO nanowires, which can strongly influence in photoluminescence properties of nanowires [2,140,141,142,143]. The earlier mentioned principle is employed for the use of ZnO nanowires as optical biosensor layers and gas detectors transducers [73,144]. The smaller the diameter of the nanowire, the larger surface-volume ratios it exhibits and the more substantial surface effects it can display [110]. While it is clear that the surface effects and surface band bending are important to ZnO-visible PL, the mechanism of how the surface effects influence the photoluminescence is still debatable [113].

3.2.3. Blue Photoluminescence Band (BPL)

The blue emission is commonly attributed in the literature to the transition between CB and zinc vacancies [145] or oxygen antisites OZn [146]. Oxygen antisites can be produced by irradiation or ion implantation to reach non-equilibrium conditions [138]. A brief summary of the band levels has been presented further in Figure 10.
Figure 8. A summary of different defect-related energy levels in the ZnO bandgap reported in the literature by different groups: (a) [129]. (b). [145]. (c). [133]. (d). [147]. (e). [148]. (f). [138] (g) [149]. (h) [146]. and (i). [150]. A color code has been used to point out the different defect states, e.g., the Zn defect is represented by orange lines, oxygen defects are presented by a dark violet color, the Li defect is presented as a yellow bar, the hydrogen level is represented by a light-blue bar, and the complex state attributed to the combination of oxygen and zinc defect states is represented by a green color.
Figure 8. A summary of different defect-related energy levels in the ZnO bandgap reported in the literature by different groups: (a) [129]. (b). [145]. (c). [133]. (d). [147]. (e). [148]. (f). [138] (g) [149]. (h) [146]. and (i). [150]. A color code has been used to point out the different defect states, e.g., the Zn defect is represented by orange lines, oxygen defects are presented by a dark violet color, the Li defect is presented as a yellow bar, the hydrogen level is represented by a light-blue bar, and the complex state attributed to the combination of oxygen and zinc defect states is represented by a green color.
Nanomaterials 10 00857 g008

3.2.4. Green Photoluminescence Band (GPB)

For decades, the origin of the green photoluminescence band (2.4–2.5 eV) for undoped ZnO remained divisive [114]. As previously mentioned, most of the intrinsic point defects create different energy levels inside the bandgap of the semiconductor, and therefore, different photoluminescence emissions along the visible range are observed. While some authors attribute the origin of the green band due to the transitions from deep donor levels of oxygen vacancies (VO) [124,151] and interstitial Zn atoms (Zni) [118,152] to the valence band, it could also be due to the transition from a shallow donor state to zinc vacancies (VZn) [114]. In addition to these intrinsic point defects, some works also report the presence of impurities (mainly copper [153,154]) as responsible for the green emission of ZnO.
From the above-mentioned defects, the dominant native defects are VO or Zni since the growth of ZnO nanowires is usually carried out in a Zn-rich environment. The unintentional intrinsic n-type doping and conductivity are also ascribed to these native defects [155]. Since 1954 [156], VO has been the most widely reported defect responsible for the photoluminescent center in the green region. However, according to some of the theoretical works, such as that presented by Janotti et al. [138], the formation energy of VO is high (around 3.72 eV) in zinc-rich growth conditions (n-type ZnO), and according to the concentration of the point defect (c) equation presented earlier, their amount would not be significant.
Then, as the formation energy of VZn is the lowest of all the intrinsic defects in n- ZnO [138], some authors have proposed models taking this defect as the source of the green luminescence [70,157]. Zinc vacancies usually introduce bandgap states that are partially occupied states and are formed from broken oxygen bonds that introduce levels in the bandgap that are partially occupied and act as deep acceptors 0.9 eV above the valence band [119,138], Figure 8. The transition from a shallow donor or the conduction band edge to these zinc vacancies level would produce photoluminescence around 2.4 eV (green color) [114].
Corroborations of both oxygen vacancy and zinc vacancy models have been presented in the work of Fabbri. et al. [70] and Kennedy [76]. In these works, the spatial distribution of NBE of ZnO nanostructures by scanning electron microscopy cathodoluminescence techniques (SEM-CL) (Figure 9) is presented. Even though these authors also attribute the green emission to distinct point defects ([70] to VZn and [76] VO, there is a compromise that the defect responsible for the green emission is mainly located at the surface [136]. The layer responsible for the green emission was estimated to be 30 nm thick (Hsu et al. [158]) in polarized photoluminescent measurements of aligned ZnO nanorods. It has also been shown that coating ZnO nanostructures with a surfactant suppresses green emission [159].
These results are consistent with the spatial distribution of green emission centers seen in the SEM-CL images reported in [70,76], where it can be seen that the most intense green emission occurs near the surface of the nanostructures or at grain boundaries.
Finally, the third intrinsic defect proposed to be involved in the green emission is zinc interstitials (Zni). They generate a donor state level 0.22 eV [146,150] below the conduction band, then the energy level may inject electrons from the conduction band to VO [132]. However, their high formation energy (6 eV under Zn-rich conditions) indicates that Zni will be present in small concentrations and cannot contribute significantly to the green photoluminescence [138].
In addition to the above-mentioned intrinsic defect, other extrinsic deep levels, including Cu, have been studied [160]. The identification of the Cu defect as being responsible for the green band was made by measuring the PL spectrum as a function of the magnetic field [161]. In response to the applied magnetic field, one band (at around 2.85 eV) split into two with its characteristic fine structure due to the presence of CuZn defect. Also, the measured g-factors from the angle dependence of the splitting are nearly identical to those of a divalent copper ion Cu2+ in a zinc lattice site of ZnO measured by electron paramagnetic resonance (EPR) [161]. A representative scheme of these transition states is presented in Figure 10a–c.

3.2.5. Yellow Photoluminescence Band (YPB)

The yellow photoluminescence band has been ascribed to Oi defects [162,163] but most commonly to the lithium doping effect in the ZnO electronic structure [53,149,164]. Li generates states that act as deep acceptor levels, 0.8 eV over the valence band. The transition from the conduction band to these generated levels inside the bandgap leads to a yellowish emission. In the work of Zwingel [164], ZnO samples doped with lithium showed a yellow photoluminescence band centered at 2.2 eV measured from 4.2 K to 300 K.
He also reports a slower decay in the intensity of the yellow photoluminescence band (YPB) in comparison to the green band (GPB) after turning off the excitation laser. The temperature dependence studies of the GPB and YPB suggest that DAP-type transitions from the conduction band or shallow donor to the Li acceptor dominate at low temperatures [165]. It can be seen in ZnO: Li that YPB dominates over the GPB at low temperatures but is quenched at temperatures above 200 K.
On the other hand, the common proposed intrinsic defect responsible for the YPB is also the interstitial oxygen sites (Oi ) [129,166]. The position of the Oi level is approximately 2.28 eV below the conduction band, so the transition between the conduction band and this level will produce the yellow color. Orange emission involving the same Oi levels and zinc intersticial levels located 0.22 eV below the conduction band have been reported, so the transition between the levels of Zni to Oi is estimated to be approximately 2.06 eV [129,134,150], Figure 8 and Figure 10d–f.
The yellow photoluminescence band has been observed in chemically prepared hexagonal ZnO nanorods and is attributed to Oi defects [167] and Zn(OH)2 groups which could leave a residue of the growth technique on the surface of the nanostructures. It was later shown that the yellow emission could be substituted for the green and red bands by a heat-treatment desorption process of the hydroxyl groups [167].
Figure 10. Proposed deep level transition in the different colors indicating the defect involved (a) [132], (b) [157], (c) [154], (d) [165], (e) [166], (f) [134], (g) [130], (h) [168,169], (i) [145,146,149].
Figure 10. Proposed deep level transition in the different colors indicating the defect involved (a) [132], (b) [157], (c) [154], (d) [165], (e) [166], (f) [134], (g) [130], (h) [168,169], (i) [145,146,149].
Nanomaterials 10 00857 g010

3.2.6. Red Photoluminescence Band (RPB)

The photoluminescence emission of ZnO in the red region of the spectrum has been reported in the work of Özgür [63]. The band is centered at 1.75 eV and has an FWHM of 0.5 eV. Similar to the yellow emission, the red band also is quenched if the system is heated from 15 to 300 K, giving rise to the green band, which could subsequently dominate the photoluminescence spectra. The earlier discussion reflects the possible competition for recombination between the acceptors of the green and red bands. In the work of Alvi et al., oxygen vacancies situated 1.65 eV below the conduction band are considered responsible for the red emission of ZnO nanotubes [130], Figure 10g,h.
On the contrary, under oxygen-rich conditions, the ZnO structure can contain oxygen interstitial Oi that introduces acceptor states 0.72 and 1.59 eV above the valence band [138] that in theory can produce the transition in the orange-red region [169]. The emission in these regions has also been attributed to zinc vacancy complexes in the work of Gomi et al. [170] and with interstitial zinc atoms in the work of Djurisic et al. [171].

4. Summary

Despite the extensive literature published about zinc oxide nanostructures, as far as we know, there are no reviews centered on the discussion related to the emission mechanism in ZnO nanowires. These characteristic properties make ZnO an exciting candidate for many different practical applications such as biological, energy, and optoelectronic devices.
The effect of a size reduction over the main bands of emission has been discussed in this review. Furthermore, a brief summary of different native point defect states or trap centers has been presented in a simple tabular and schematic form for an easy and better understanding of the topic.

Author Contributions

Investigation and Writing—original draft, A.G.-M.; Supervision, G.S., F.G.; review and editing, F.G.; investigation, P.R.M.-A.; Supervision, review & editing, A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This project has received funding from the EU’s H2020 research and innovation program under GA no. 654148 Laserlab-Europe and from the second research collaboration fund 2019 of the Unión Iberoamericana de Universidades.

Acknowledgments

A.G.M acknowledges CONACyT Scholarship CVU 860916. G.S. gratefully acknowledges the CONACyT project PN 4797 and IG100320. A.D. thanks the DGAPA project PAPIIT IA100219. The authors are grateful to project MAT2017-87500-P and to the Programa Ajut a la Recerca Transversal de l’IN2UB 2018 for financial support.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Espitia, P.J.P.; Otoni, C.G.; Soares, N.F.F. Zinc Oxide Nanoparticles for Food Packaging Applications; Elsevier Inc.: Cambridge, MA, USA, 2016; ISBN 9780128007235. [Google Scholar]
  2. Wang, D.; Reynolds, N. Photoluminescence of Zinc Oxide nanowires: The effect of surface band bending. ISRN Condens. Matter Phys. 2012, 2012. [Google Scholar] [CrossRef] [Green Version]
  3. Butcher, P.N.; March, N.H.; Tosi, M.P. Crystalline Semiconducting Materials and Devices, 1st ed.; Springer: New York, NY, USA, 1986; ISBN 9781475799026. [Google Scholar]
  4. Petkov, P.; Tsiulyanu, D.; Popov, C.; Kulisch, W. Nanoscience Advances in CBRN Agent Detection, Information and Energy Security: An Introduction; Springer: Dordrecht, The Netherlands, 2015; ISBN 9789401796972. [Google Scholar]
  5. Marotti, R.E.; Guerra, D.N.; Bello, C.; Machado, G.; Dalchiele, E.A. Bandgap energy tuning of electrochemically grown ZnO thin films by thickness and electrodeposition potential. Sol. Energy Mater. Sol. Cells 2004, 82, 85–103. [Google Scholar] [CrossRef]
  6. Ozgur, Ü.; Hofstetter, D.; Morkoç, H. ZnO devices and applications: A review of current status and future prospects. Proc. IEEE 2010, 98, 1255–1268. [Google Scholar] [CrossRef]
  7. Lv, J.; Li, X. Defect evolution in ZnO and its effect on radiation tolerance. Phys. Chem. Chem. Phys. 2018, 20, 11882–11887. [Google Scholar] [CrossRef] [PubMed]
  8. Kaidashev, E.M.; Lorenz, M.; Von Wenckstern, H.; Rahm, A.; Semmelhack, H.C.; Han, K.H.; Benndorf, G.; Bundesmann, C.; Hochmuth, H.; Grundmann, M. High electron mobility of epitaxial ZnO thin films on c-plane sapphire grown by multistep pulsed-laser deposition. Appl. Phys. Lett. 2003, 82, 3901–3903. [Google Scholar] [CrossRef]
  9. Chen, G.; Wang, P.S. Surface-induced enhancement of piezoelectricity in ZnO nanowires. Chin. Phys. Lett. 2018, 35, 127701. [Google Scholar] [CrossRef]
  10. Wang, X.; Dai, Y.; Liu, R.; He, X.; Li, S.; Wang, Z.L. Light-triggered pyroelectric nanogenerator based on a pn-junction for self-powered near-infrared photosensing. ACS Nano 2017, 11, 8339–8345. [Google Scholar] [CrossRef]
  11. Lee, J.S.; Kang, M., II; Kim, S.; Lee, M.S.; Lee, Y.K. Growth of zinc oxide nanowires by thermal evaporation on vicinal Si(100) substrate. J. Cryst. Growth 2003, 249, 201–207. [Google Scholar] [CrossRef]
  12. Güell, F.; Martínez-Alanis, P.R.; Roso, S.; Salas-Pérez, C.I.; García-Sánchez, M.F.; Santana, G.; Monroy, B.M. Plasma versus thermal annealing for the Au-catalyst growth of ZnO nanocones and nanowires on Al-doped ZnO buffer layers. Mater. Res. Express 2016, 3, 065013. [Google Scholar] [CrossRef]
  13. Ma, S.; Kitai, A.H. ZnO nanowire growth by chemical vapor deposition with spatially controlled density on Zn2GeO4: Mn polycrystalline substrates. Mater. Res. Express 2017, 4, 065012. [Google Scholar] [CrossRef]
  14. Ameer, A.A.; Suriani, A.B.; Jabur, A.R.; Hashim, N.; Physics, F.; Zaid, K. The fabrication of zinc oxide nanorods and nanowires by solgel immersion methods. J. Phys. Conf. Ser. 2019, 1170, 012005. [Google Scholar] [CrossRef]
  15. Alshehri, N.A.; Lewis, A.R.; Pleydell-Pearce, C.; Maffeis, T.G.G. Investigation of the growth parameters of hydrothermal ZnO nanowires for scale up applications. J. Saudi Chem. Soc. 2018, 22, 538–545. [Google Scholar] [CrossRef] [Green Version]
  16. Isakov, I.; Panfilova, M.; Sourribes, M.J.L.; Warburton, P.A. Growth of ZnO and ZnMgO nanowires by Au-catalysed molecular-beam epitaxy. Phys. Status Solidi Curr. Top. Solid State Phys. 2013, 10, 1308–1313. [Google Scholar] [CrossRef]
  17. Park, W.I.; Kim, D.H.; Jung, S.W.; Yi, G.C. Metalorganic vapor-phase epitaxial growth of vertically well-aligned ZnO nanorods. Appl. Phys. Lett. 2002, 80, 4232–4234. [Google Scholar] [CrossRef]
  18. Rai, P.; Kwak, W.K.; Yu, Y.T. Solvothermal synthesis of ZnO nanostructures and their morphology-dependent gas-sensing properties. ACS Appl. Mater. Interfaces 2013, 5, 3026–3032. [Google Scholar] [CrossRef] [PubMed]
  19. Guo, L.; Yang, S.; Yang, C.; Yu, P.; Wang, J.; Ge, W.; Wong, G.K.L. Highly monodisperse polymer-capped ZnO nanoparticles: Preparation and optical properties. Appl. Phys. Lett. 2000, 76, 2901–2903. [Google Scholar] [CrossRef]
  20. Bagga, S.; Akhtar, J.; Mishra, S. Synthesis and applications of ZnO nanowire: A review. In AIP Conference Proceedings; AIP Publishing LLC: Baltimore, MD, USA, 2018; Volume 1989. [Google Scholar]
  21. Vayssieres, L. Growth of arrayed nanorods and nanowires of ZnO from aqueous solutions. Adv. Mater. 2003, 15, 464–466. [Google Scholar] [CrossRef]
  22. Wu, J.J.; Liu, S.C.; Wu, C.T.; Chen, K.H.; Chen, L.C. Heterostructures of ZnO-Zn coaxial nanocables and ZnO nanotubes. Appl. Phys. Lett. 2002, 81, 1312–1314. [Google Scholar] [CrossRef]
  23. Li, Y.B.; Bando, Y.; Sato, T.; Kurashima, K. ZnO nanobelts grown on Si substrate. Appl. Phys. Lett. 2002, 81, 144–146. [Google Scholar] [CrossRef]
  24. Crupi, I.; Boscarino, S.; Strano, V.; Mirabella, S.; Simone, F.; Terrasi, A. Optimization of ZnO:Al/Ag/ZnO:Al structures for ultra-thin high-performance transparent conductive electrodes. Thin Solid Films 2012, 520, 4432–4435. [Google Scholar] [CrossRef]
  25. Gao, P.X.; Ding, Y.; Wang, Z.L. Electronic transport in superlattice-structured ZnO nanohelix. Nano Lett. 2009, 9, 137–143. [Google Scholar] [CrossRef] [PubMed]
  26. Hughes, W.L.; Wang, Z.L. Controlled synthesis and manipulation of ZnO nanorings and nanobows. Appl. Phys. Lett. 2005, 86, 2003–2006. [Google Scholar] [CrossRef]
  27. Haibo, Z.; Peisheng, L.; Weiping, C.; Shikuan, Y.; Xiaoxia, X. Controllable Pt/ZnO porous nanocages with improved photocatalytic activity. J. Phys. Chem. C 2008, 112, 19620–19624. [Google Scholar]
  28. Wang, Y.; Li, X.; Wang, N.; Quan, X.; Chen, Y. Controllable synthesis of ZnO nanoflowers and their morphology-dependent photocatalytic activities. Sep. Purif. Technol. 2008, 62, 727–732. [Google Scholar] [CrossRef]
  29. Gao, W.; Li, Z. Nanostructures of zinc oxide. Int. J. Nanotechnol. 2009, 6, 245–257. [Google Scholar] [CrossRef]
  30. Pearton, S.J.; Norton, D.P.; Heo, Y.W.; Tien, L.C.; Ivill, M.P.; Li, Y.; Kang, B.S.; Ren, F.; Kelly, J.; Hebard, A.F. ZnO spintronics and nanowire devices. J. Electron. Mater. 2006, 35, 862–868. [Google Scholar] [CrossRef]
  31. Sun, Y.F.; Liu, S.B.; Meng, F.L.; Liu, J.Y.; Jin, Z.; Kong, L.T.; Liu, J.H. Metal oxide nanostructures and their gas sensing properties: A review. Sensors 2012, 12, 2610–2631. [Google Scholar] [CrossRef] [Green Version]
  32. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent developments of zinc oxide based photocatalyst in water treatment technology: A review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef]
  33. Meyer, B.K.; Alves, H.; Hofmann, D.M.; Kriegseis, W.; Forster, D.; Bertram, F.; Christen, J.; Hoffmann, A.; Straßburg, M.; Dworzak, M.; et al. Bound exciton and donor-acceptor pair recombinations in ZnO. Phys. Status Solidid Basic Res. 2004, 241, 231–260. [Google Scholar] [CrossRef]
  34. Zhang, Y.; Ram, M.K.; Stefanakos, E.K.; Goswami, D.Y. Synthesis, characterization, and applications of ZnO nanowires. J. Nanomater. 2012, 2012. [Google Scholar] [CrossRef]
  35. Rong, P.; Ren, S.; Yu, Q. Fabrications and applications of ZnO nanomaterials in flexible functional devices—A review. Crit. Rev. Anal. Chem. 2019, 49, 336–349. [Google Scholar] [CrossRef] [PubMed]
  36. Rauwel, P.; Salumaa, M.; Aasna, A.; Galeckas, A.; Rauwel, E. A review of the synthesis and photoluminescence properties of hybrid ZnO and carbon nanomaterials. J. Nanomater. 2016, 2016. [Google Scholar] [CrossRef] [Green Version]
  37. Parihar, V.; Raja, M.; Paulose, R. A brief review of structural, electrical and electrochemical properties of zinc oxide nanoparticles. Rev. Adv. Mater. Sci. 2018, 53, 119–130. [Google Scholar] [CrossRef]
  38. Li, Y.; Meng, G.W.; Zhang, L.D.; Phillipp, F. Ordered semiconductor ZnO nanowire arrays and their photoluminescence properties. Appl. Phys. Lett. 2000, 76, 2011–2013. [Google Scholar] [CrossRef]
  39. Wahab, R.; Hwang, I.H.; Shin, H.S.; Kim, Y.S.; Musarrat, J.; Al-Khedhairy, A.A.; Siddiqui, M.A. Zinc oxide nanostructures and their applications. Intell. Nanomater. Process. Prop. Appl. 2012, 28, 183–212. [Google Scholar]
  40. Witkowski, B.S. Applications of ZnO nanorods and nanowires—A review. Acta Phys. Polon. A 2018, 134, 1226–1246. [Google Scholar] [CrossRef]
  41. Samadi, M.; Zirak, M.; Naseri, A.; Khorashadizade, E.; Moshfegh, A.Z. Recent progress on doped ZnO nanostructures for visible-light photocatalysis. Thin Solid Films 2016, 605, 2–19. [Google Scholar] [CrossRef] [Green Version]
  42. Kumar, R.; Al-Dossary, O.; Kumar, G.; Umar, A. Zinc oxide nanostructures for NO 2 gas–sensor applications: A review. Nano-Micro Lett. 2015, 7, 97–120. [Google Scholar] [CrossRef] [Green Version]
  43. Shanmugam, N.R.; Muthukumar, S.; Prasad, S. A review on ZnO-based electrical biosensors for cardiac biomarker detection. Future Sci. OA 2017, 3, FSO196. [Google Scholar] [CrossRef] [Green Version]
  44. Rackauskas, S.; Barbero, N.; Barolo, C.; Viscardi, G. ZnO nanowire application in chemoresistive sensing: A review. Nanomaterials 2017, 7, 381. [Google Scholar] [CrossRef] [Green Version]
  45. Sirelkhatim, A.; Mahmud, S.; Seeni, A.; Kaus, N.H.M.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on zinc oxide nanoparticles: Antibacterial activity and toxicity mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [Green Version]
  46. Rahman, F. Zinc oxide light-emitting diodes: A review. Opt. Eng. 2019, 58, 010901. [Google Scholar] [CrossRef]
  47. Gargas, D.J.; Toimil-Molares, M.E.; Yang, P. Imaging single ZnO vertical nanowire laser cavities using UV-laser scanning confocal microscopy. J. Am. Chem. Soc. 2009, 131, 2125–2127. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Consonni, V.; Briscoe, J.; Kärber, E.; Li, X.; Cossuet, T. ZnO nanowires for solar cells: A comprehensive review. Nanotechnology 2019, 30, 362001. [Google Scholar] [CrossRef] [PubMed]
  49. Ditshego, N.M.J. ZnO nanowire field effect transistor for biosensing: A review. J. Nano Res. 2019, 60, 94–112. [Google Scholar] [CrossRef] [Green Version]
  50. Wang, Z.L. Ten years’ venturing in ZnO nanostructures: From discovery to scientific understanding and to technology applications. Chin. Sci. Bull. 2009, 54, 4021–4034. [Google Scholar] [CrossRef]
  51. Chennupati Jagadish, S.J.P. Zinc Oxide Bulk, Thin Films and Nanostructures: Processing, Properties and Applications; Elsevier Science: Amsterdam, The Netherlands, 2006. [Google Scholar]
  52. Morkoç, H.; Özgür, Ü. Zinc Oxide: Fundamentals, Materials and Device Technology; John Wiley & Sons: Hoboken, NJ, USA, 2009; ISBN 9783527408139. [Google Scholar]
  53. Rodnyi, P.A.; Khodyuk, I.V. Optical and luminescence properties of zinc oxide (Review). Opt. Spectrosc. Engl. Transl. Opt. Spektrosk. 2011, 111, 776–785. [Google Scholar] [CrossRef] [Green Version]
  54. Rackauskas, S.; Nasibulin, A.G.; Jiang, H.; Tian, Y.; Kleshch, V.I.; Sainio, J.; Obraztsova, E.D.; Bokova, S.N.; Obraztsov, A.N.; Kauppinen, E.I. A novel method for metal oxide nanowire synthesis. Nanotechnology 2009, 20, 165603. [Google Scholar] [CrossRef]
  55. Goyal, R.K. Nanomaterials and Nanocomposites; CRC Press: Boca Ratón, FL, USA, 2016; ISBN 9783527683772. [Google Scholar]
  56. Singh, Y.R.G.; Kataria, S.; Barshilia, H.C. Different morphologies of zinc oxide nanostructures grown under similar deposition conditions during vapor-liquid-solid growth. Nanosci. Nanotechnol. Lett. 2013, 5, 1224–1230. [Google Scholar] [CrossRef]
  57. Huang, M.H.; Mao, S.; Feick, H.; Yan, H.; Wu, Y.; Kind, H.; Weber, E.; Russo, R.; Yang, P. Room-temperature ultraviolet nanowire nanolasers. Science 2001, 292, 1897–1899. [Google Scholar] [CrossRef] [Green Version]
  58. Simon, H.; Krekeler, T.; Schaan, G.; Mader, W. Metal-seeded growth mechanism of ZnO nanowires. Cryst. Growth Des. 2013, 13, 572–580. [Google Scholar] [CrossRef]
  59. Wang, Z.L. Zinc oxide nanostructures: Growth, properties and applications. J. Phys. Condenced Matter 2004, 16, R829–R858. [Google Scholar] [CrossRef]
  60. Liu, X.; Huang, W.; Cheng, H.; Huang, B.; Bai, D.; Fu, F.; Wu, H.; Li, L. In-situ observation of hydrothermal growth of ZnO nanowires on patterned Zn substrate and their photocatalytic performance. Appl. Surf. Sci. 2015, 356, 240–248. [Google Scholar] [CrossRef]
  61. Serrano, A.; Arana, A.; Galdámez, A.; Dutt, A.; Monroy, B.M.; Güell, F.; Santana, G. Effect of the seed layer on the growth and orientation of the ZnO nanowires: Consequence on structural and optical properties. Vacuum 2017, 146, 509–516. [Google Scholar] [CrossRef]
  62. Xu, S.; Wang, Z.L. One-dimensional ZnO nanostructures: Solution growth and functional properties. Nano Res. 2011, 4, 1013–1098. [Google Scholar] [CrossRef] [Green Version]
  63. Özgür, Ü.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.A.; Doǧan, S.; Avrutin, V.; Cho, S.J.; Morkoç, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 11. [Google Scholar] [CrossRef] [Green Version]
  64. Biroju, R.K.; Tilak, N.; Rajender, G.; Dhara, S.; Giri, P.K. Catalyst free growth of ZnO nanowires on graphene and graphene oxide and its enhanced photoluminescence and photoresponse. Nanotechnology 2015, 26, 145601. [Google Scholar] [CrossRef]
  65. Lupan, O.; Emelchenko, G.A.; Ursaki, V.V.; Chai, G.; Redkin, A.N.; Gruzintsev, A.N.; Tiginyanu, I.M.; Chow, L.; Ono, L.K.; Roldan Cuenya, B.; et al. Synthesis and characterization of ZnO nanowires for nanosensor applications. Mater. Res. Bull. 2010, 45, 1026–1032. [Google Scholar] [CrossRef]
  66. Qiu, J.; Li, X.; He, W.; Park, S.J.; Kim, H.K.; Hwang, Y.H.; Lee, J.H.; Kim, Y.D. The growth mechanism and optical properties of ultralong ZnO nanorod arrays with a high aspect ratio by a preheating hydrothermal method. Nanotechnology 2009, 20, 155603. [Google Scholar] [CrossRef]
  67. Gargas, D.J.; Gao, H.; Wang, H.; Yang, P. High quantum efficiency of band-edge emission from ZnO nanowires. Nano Lett. 2011, 11, 3792–3796. [Google Scholar] [CrossRef] [Green Version]
  68. Güell, F.; Martínez-Alanis, P.R. Tailoring the Green, Yellow and Red defect emission bands in ZnO nanowires via the growth parameters. J. Lumin. 2019, 210, 128–134. [Google Scholar] [CrossRef]
  69. Fan, J.; Güell, F.; Fábrega, C.; Fairbrother, A.; Andreu, T.; López, A.M.; Morante, J.R.; Cabot, A. Visible photoluminescence components of solution-grown ZnO nanowires: Influence of the surface depletion layer. J. Phys. Chem. C 2012, 116, 19496–19502. [Google Scholar] [CrossRef] [Green Version]
  70. Fabbri, F.; Villani, M.; Catellani, A.; Calzolari, A.; Cicero, G.; Calestani, D.; Calestani, G.; Zappettini, A.; Dierre, B.; Sekiguchi, T.; et al. Zn vacancy induced green luminescence on non-polar surfaces in ZnO nanostructures. Sci. Rep. 2014, 4, 5158. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  71. Tai, G.A.; Liu, J.S. ZnO nanowire arrays: Controllable preparation and effect of annealing temperature on optical properties. Mater. Res. Innov. 2014, 18, 20–26. [Google Scholar] [CrossRef]
  72. Borysiewicz, M.A.; Gryglas-Borysiewicz, M.; Masłyk, M.; Wojciechowski, T.; Wzorek, M.; Kaczmarski, J.; Wojtowicz, T.; Kamińska, E. Room temperature sputter deposited catalyst-free nanowires with wurtzite/zinc blende ZnO superstructure and their application in electromechanical nanogenerators on polymer and paper substrates. Nanotechnology 2017, 28, 085204. [Google Scholar] [CrossRef] [PubMed]
  73. Galdamez, A.; Serrano, A.; Santana, G.; Arjona, N.; Arriaga, L.G.; Tapia Ramirez, J.; Oza, G.; Dutt, A. DNA probe functionalization on different morphologies of ZnO/Au nanowire for bio-sensing applications. Mater. Lett. 2019, 235, 250–253. [Google Scholar] [CrossRef]
  74. Choi, A.; Kim, K.; Jung, H.I.; Lee, S.Y. ZnO nanowire biosensors for detection of biomolecular interactions in enhancement mode. Sens. Actuators B Chem. 2010, 148, 577–582. [Google Scholar] [CrossRef]
  75. Belhaj, M.; Dridi, C.; Habba, Y.G.; Capo-Chichi, M.; Leprince-Wang, Y. Surface morphology evolution with fabrication parameters of ZnO nanowires toward emission properties enhancement. Phys. B Condens. Matter 2017, 526, 64–70. [Google Scholar] [CrossRef]
  76. Kennedy, O.W.; White, E.R.; Howkins, A.; Williams, C.K.; Boyd, I.W.; Warburton, P.A.; Shaffer, M.S.P. Mapping the origins of luminescence in ZnO nanowires by STEM-CL. J. Phys. Chem. Lett. 2019, 10, 386–392. [Google Scholar] [CrossRef] [PubMed]
  77. Lai, Y.; Meng, M.; Yu, Y.; Wang, X.; Ding, T. Photoluminescence and photocatalysis of the flower-like nano-ZnO photocatalysts prepared by a facile hydrothermal method with or without ultrasonic assistance. Appl. Catal. B Environ. 2011, 105, 335–345. [Google Scholar] [CrossRef]
  78. Mousavi, S.H.; Haratizadeh, H.; Minaee, H. Comparison of structural and photoluminescence properties of zinc oxide nanowires grown by vapor-solid and vapor-liquid-solid methods. Thin Solid Films 2012, 520, 4642–4645. [Google Scholar] [CrossRef]
  79. Lyu, S.C.; Zhang, Y.; Ruh, H.; Lee, H.J.; Shim, H.W.; Suh, E.K.; Lee, C.J. Low temperature growth and photoluminescence of well-aligned zinc oxide nanowires. Chem. Phys. Lett. 2002, 363, 134–138. [Google Scholar] [CrossRef]
  80. Ding, L.; Yang, C.; He, H.; Wang, J.; Tang, Z.; Foreman, B.A.; Jiang, F.; Ge, W. Verification of Γ7 symmetry assignment for the top valence band of ZnO by magneto-optical studies of the free A exciton state. New J. Phys. 2013, 15, 033015. [Google Scholar] [CrossRef] [Green Version]
  81. Adachi, S.; Yoshikawa, H. Optical #constants of ZnO. J. Appl. Phys. 1997, 36, 6237–6243. [Google Scholar]
  82. Zuo, J.; Erbe, A. Optical and electronic properties of native zinc oxide films on polycrystalline Zn. Phys. Chem. Chem. Phys. 2010, 12, 11467–11476. [Google Scholar] [CrossRef]
  83. Gadallah, A.S.; El-Nahass, M.M. Structural, optical constants and photoluminescence of ZnO thin films grown by sol-gel spin coating. Adv. Condens. Matter Phys. 2013, 2013. [Google Scholar] [CrossRef] [Green Version]
  84. Gencyilmaz, O.; Karakaya, S.; Atay, F.; Ozbas, O.; Akyuz, I. Structural and optical properties of transparent polycrystalline ZnO films. AIP Conf. Proc. 2012, 1476, 216–220. [Google Scholar]
  85. Lin, W.; Chen, D.; Zhang, J.; Lin, Z.; Huang, J.; Li, W.; Wang, Y.; Huang, F. Hydrothermal growth of ZnO single crystals with high carrier mobility. Cryst. Growth Des. 2009, 9, 4378–4383. [Google Scholar] [CrossRef]
  86. Khranovskyy, V.; Lazorenko, V.; Lashkarev, G.; Yakimova, R. Luminescence anisotropy of ZnO microrods. J. Lumin. 2012, 132, 2643–2647. [Google Scholar] [CrossRef] [Green Version]
  87. Muth, J.F.; Lee, J.H.; Shmagin, I.K.; Kolbas, R.M.; Casey, H.C.; Keller, B.P.; Mishra, U.K.; DenBaars, S.P. Absorption coefficient, energy gap, exciton binding energy, and recombination lifetime of GaN obtained from transmission measurements. Appl. Phys. Lett. 1997, 71, 2572–2574. [Google Scholar] [CrossRef]
  88. Chen, X.B.; Huso, J.; Morrison, J.L.; Bergman, L. The properties of ZnO photoluminescence at and above room temperature. J. Appl. Phys. 2007, 102, 2005–2008. [Google Scholar] [CrossRef]
  89. Hullavarad, S.; Hullavarad, N.; Look, D.; Claflin, B. Persistent photoconductivity studies in nanostructured ZnO UV sensors. Nanoscale Res. Lett. 2009, 4, 1421–1427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Reynolds, D.C.; Litton, C.W.; Collins, T.C. Zeeman effects in the edge emission and absorption of ZnO. Phys. Rev. 1965, 140, A1726. [Google Scholar] [CrossRef]
  91. Mohanta, S.K.; Kim, D.C.; Cho, H.K.; Soh, C.B.; Chua, S.J.; Tripathy, S. Influence of buffer layer on structural and optical properties of ZnO nanorods on glass substrates. Electrochem. Solid-State Lett. 2008, 11, 143–146. [Google Scholar] [CrossRef]
  92. Khranovskyy, V.; Yakimova, R.; Karlsson, F.; Syed, A.S.; Holtz, P.O.; Nigussa Urgessa, Z.; Samuel Oluwafemi, O.; Reinhardt Botha, J. Comparative PL study of individual ZnO nanorods, grown by APMOCVD and CBD techniques. Phys. B Condens. Matter 2012, 407, 1538–1542. [Google Scholar] [CrossRef] [Green Version]
  93. Khranovskyy, V.; Eriksson, M.O.; Radnoczi, G.Z.; Khalid, A.; Zhang, H.; Holtz, P.O.; Hultman, L.; Yakimova, R. Photoluminescence study of basal plane stacking faults in ZnO nanowires. Phys. B Condens. Matter 2014, 439, 50–53. [Google Scholar] [CrossRef]
  94. Reparaz, J.S.; Güell, F.; Wagner, M.R.; Hoffmann, A.; Cornet, A.; Morante, J.R. Size-dependent recombination dynamics in ZnO nanowires. Appl. Phys. Lett. 2010, 96, 053105. [Google Scholar] [CrossRef] [Green Version]
  95. Teke, A.; Özgür, Ü.; Doǧan, S.; Gu, X.; Morkoç, H.; Nemeth, B.; Nause, J.; Everitt, H.O. Excitonic fine structure and recombination dynamics in single-crystalline ZnO. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 70, 195207. [Google Scholar] [CrossRef] [Green Version]
  96. Thonke, K.; Gruber, T.; Teofilov, N.; Schönfelder, R.; Waag, A.; Sauer, R. Donor-acceptor pair transitions in ZnO substrate material. Phys. B Condens. Matter 2001, 308–310, 985–988. [Google Scholar] [CrossRef]
  97. Rodina, A.V.; Strassburg, M.; Dworzak, M.; Haboeck, U.; Hoffmann, A.; Zeuner, A.; Alves, H.R.; Hofmann, D.M.; Meyer, B.K. Magneto-optical properties of bound excitons in ZnO. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 69, 125206. [Google Scholar] [CrossRef] [Green Version]
  98. Wagner, M.R. Fundamental Properties of Excitons and Phonons in ZnO: A Spectroscopic Study of the Dynamics, Polarity, and Effects of External Fields. Ph.D. Thesis, Technische Universität, Berlin, Germany, 2010. [Google Scholar]
  99. Cao, B.; Teng, X.; Heo, S.H.; Li, Y.; Cho, S.O.; Li, G.; Cai, W. Different ZnO nanostructures fabricated by a seed-layer assisted electrochemical route and their photoluminescence and field emission properties. J. Phys. Chem. C 2007, 111, 2470–2476. [Google Scholar] [CrossRef]
  100. Tong, Y.; Liu, Y.; Shao, C.; Liu, Y.; Xu, C.; Zhang, J.; Lu, Y.; Shen, D.; Fan, X. Growth and optical properties of faceted hexagonal ZnO nanotubes. J. Phys. Chem. B 2006, 110, 14714–14718. [Google Scholar] [CrossRef] [PubMed]
  101. Varshni, Y.P. Temperature dependence of the energy gap in semiconductors. Physica 1967, 34, 149–154. [Google Scholar] [CrossRef]
  102. Sakai, K.; Noguchi, K.; Fukuyama, A.; Ikari, T.; Okada, T. Low-temperature photoluminescence of nanostructured ZnO crystal synthesized by pulsed-laser ablation. Jpn. J. Appl. Phys. 2009, 48, 0850011–0850015. [Google Scholar] [CrossRef]
  103. Guo, X.; Tripp, C.P.; Chen, C.; Wang, Y.; Yin, S.; Qin, W. Temperature dependence of the photoluminescence from ZnO microrods prepared by a float zone method. CrystEngComm 2016, 18, 3130–3135. [Google Scholar] [CrossRef]
  104. Fang, F.; Zhao, D.; Li, B.; Zhang, Z.; Shen, D.; Wang, X. Bending-induced enhancement of longitudinal optical phonon scattering in ZnO nanowires. J. Phys. Chem. C 2010, 114, 12477–12480. [Google Scholar] [CrossRef]
  105. Hamby, D.W.; Lucca, D.A.; Klopfstein, M.J.; Cantwell, G. Temperature dependent exciton photoluminescence of bulk ZnO. J. Appl. Phys. 2003, 93, 3214–3217. [Google Scholar] [CrossRef]
  106. Rauch, C.; Gehlhoff, W.; Wagner, M.R.; Malguth, E.; Callsen, G.; Kirste, R.; Salameh, B.; Hoffmann, A.; Polarz, S.; Aksu, Y.; et al. Lithium related deep and shallow acceptors in Li-doped ZnO nanocrystals. J. Appl. Phys. 2010, 107, 024311. [Google Scholar] [CrossRef] [Green Version]
  107. Janotti, A.; Van de Walle, C.G. Fundamentals of zinc oxide as a semiconductor. Rep. Prog. Phys. 2009, 72, 126501. [Google Scholar] [CrossRef] [Green Version]
  108. Kong, Y.C.; Yu, D.P.; Zhang, B.; Fang, W.; Feng, S.Q. Ultraviolet-emitting ZnO nanowires synthesized by a physical vapor deposition approach. Appl. Phys. Lett. 2001, 78, 407–409. [Google Scholar] [CrossRef]
  109. Grabowska, J.; Meaney, A.; Nanda, K.K.; Mosnier, J.P.; Henry, M.O.; Duclère, J.R.; McGlynn, E. Surface excitonic emission and quenching effects in ZnO nanowire/nanowall systems: Limiting effects on device potential. Phys. Rev. B Condens. Matter Mater. Phys. 2005, 71, 115439. [Google Scholar] [CrossRef] [Green Version]
  110. Shalish, I.; Temkin, H.; Narayanamurti, V. Size-dependent surface luminescence in ZnO nanowires. Phys. Rev. B Condens. Matter Mater. Phys. 2004, 69, 245401. [Google Scholar] [CrossRef]
  111. Zhang, B.P.; Binh, N.T.; Segawa, Y.; Wakatsuki, K.; Usami, N. Optical properties of ZnO rods formed by metalorganic chemical vapor deposition. Appl. Phys. Lett. 2003, 83, 1635–1637. [Google Scholar] [CrossRef]
  112. Huang, M.H.; Wu, Y.; Feick, H.; Tran, N.; Weber, E.; Yang, P. Catalytic growth of Zinc oxide nanowires by vapor transport. Adv. Mater. 2001, 24, 113–116. [Google Scholar] [CrossRef]
  113. Liao, Z.M.; Zhang, H.Z.; Zhou, Y.B.; Xu, J.; Zhang, J.M.; Yu, D.P. Surface effects on photoluminescence of single ZnO nanowires. Phys. Lett. Sect. A Gen. At. Solid State Phys. 2008, 372, 4505–4509. [Google Scholar] [CrossRef]
  114. Willander, M.; Nur, O.; Sadaf, J.R.; Qadir, M.I.; Zaman, S.; Zainelabdin, A.; Bano, N.; Hussain, I. Luminescence from zinc oxide nanostructures and polymers and their hybrid devices. Materials 2010, 3, 2643–2667. [Google Scholar] [CrossRef]
  115. Zhang, B.; Zhou, S.; Liu, B.; Gong, H.; Zhang, X. Fabrication and green emission of ZnO nanowire arrays. Sci. China Ser. E Technol. Sci. 2009, 52, 883–887. [Google Scholar] [CrossRef]
  116. Kang, H.S.; Kang, J.S.; Kim, J.W.; Lee, S.Y. Annealing effect on the property of ultraviolet and green emissions of ZnO thin films. J. Appl. Phys. 2004, 95, 1246–1250. [Google Scholar] [CrossRef]
  117. Studenikin, S.A.; Golego, N.; Cocivera, M. Fabrication of green and orange photoluminescent, undoped ZnO films using spray pyrolysis. J. Appl. Phys. 1998, 84, 2287–2294. [Google Scholar] [CrossRef]
  118. Vanheusden, K.; Warren, W.L.; Seager, C.H.; Tallant, D.R.; Voigt, J.A.; Gnade, B.E. Mechanisms behind green photoluminescence in ZnO phosphor powders. J. Appl. Phys. 1996, 79, 7983–7990. [Google Scholar] [CrossRef]
  119. Xiong, G.; Pal, U.; Serrano, J.G.; Ucer, K.B.; Williams, R.T. Photoluminescence and FTIR study of ZnO nanoparticles: The impurity and defect perspective. Phys. Status Solidi Curr. Top. Solid State Phys. 2006, 3, 3577–3581. [Google Scholar]
  120. Xu, S.; Xu, C.; Liu, Y.; Hu, Y.; Yang, R.; Yang, Q.; Ryou, J.H.; Kim, H.J.; Lochner, Z.; Choi, S.; et al. Ordered nanowire array blue/near-UV light emitting diodes. Adv. Mater. 2010, 22, 4749–4753. [Google Scholar] [CrossRef]
  121. Zhu, H.; Iqbal, J.; Xu, H.; Yu, D. Raman and photoluminescence properties of highly Cu doped ZnO nanowires fabricated by vapor-liquid-solid process. J. Chem. Phys. 2008, 129, 124713. [Google Scholar] [CrossRef] [PubMed]
  122. Lauer, R.B. The I.R. photoluminescence emission band in ZnO. J. Phys. Chem. Solids 1973, 34, 249–253. [Google Scholar] [CrossRef]
  123. Leiter, F.H.; Alves, H.R.; Hofstaetter, A.; Hofmann, D.M.; Meyer, B.K. The oxygen vacancy as the origin of a green emission in undoped ZnO. Phys. Status Solidi Basic Res. 2001, 226, 5–6. [Google Scholar] [CrossRef]
  124. Quemener, V.; Vines, L.; Monakhov, E.V.; Svensson, B.G. Evolution of deep electronic states in ZnO during heat treatment in oxygen- and zinc-rich ambients. Appl. Phys. Lett. 2012, 100, 1–5. [Google Scholar] [CrossRef]
  125. Sokol, A.A.; French, S.A.; Bromley, S.T.; Catlow, R.A.; Dam, J.J.; Van Sherwood, P. Point defects in ZnO. Semicond. Semimet. 2007, 134, 267–282. [Google Scholar] [CrossRef] [PubMed]
  126. Tam, K.H.; Cheung, C.K.; Leung, Y.H.; Djurišić, A.B.; Ling, C.C.; Beling, C.D.; Fung, S.; Kwok, W.M.; Chan, W.K.; Phillips, D.L.; et al. Defects in ZnO nanorods prepared by a hydrothermal method. J. Phys. Chem. B 2006, 110, 20865–20871. [Google Scholar] [CrossRef]
  127. Jeong, S.H.; Kim, B.S.; Lee, B.T. Photoluminescence dependence of ZnO films grown on Si(100) by radio-frequency magnetron sputtering on the growth ambient. Appl. Phys. Lett. 2003, 82, 2625–2627. [Google Scholar] [CrossRef]
  128. Shan, F.K.; Liu, G.X.; Lee, W.J.; Shin, B.C. The role of oxygen vacancies in epitaxial-deposited ZnO thin films. J. Appl. Phys. 2007, 101, 053106. [Google Scholar] [CrossRef]
  129. Djurišić, A.B.; Leung, Y.H.; Tam, K.H.; Ding, L.; Ge, W.K.; Chen, H.Y.; Gwo, S. Green, yellow, and orange defect emission from ZnO nanostructures: Influence of excitation wavelength. Appl. Phys. Lett. 2006, 88, 28–31. [Google Scholar] [CrossRef] [Green Version]
  130. Alvi, N.H.; ul Hasan, K.; Nur, O.; Willander, M. The origin of the red emission in n-zno nanotubes/p-gan white light emitting diodes. Nanoscale Res. Lett. 2011, 6, 130. [Google Scholar] [CrossRef] [Green Version]
  131. Yamauchi, S.; Goto, Y.; Hariu, T. Photoluminescence studies of undoped and nitrogen-doped ZnO layers grown by plasma-assisted epitaxy. J. Cryst. Growth 2004, 260, 1–6. [Google Scholar] [CrossRef]
  132. Cao, B.; Cai, W.; Zeng, H. Temperature-dependent shifts of three emission bands for ZnO nanoneedle arrays. Appl. Phys. Lett. 2006, 88, 161101. [Google Scholar] [CrossRef]
  133. Lin, B.; Fu, Z.; Jia, Y. Green luminescent center in undoped zinc oxide films deposited on silicon substrates. Appl. Phys. Lett. 2001, 79, 943–945. [Google Scholar] [CrossRef]
  134. Ahn, C.H.; Kim, Y.Y.; Kim, D.C.; Mohanta, S.K.; Cho, H.K. A comparative analysis of deep level emission in ZnO layers deposited by various methods. J. Appl. Phys. 2009, 105, 013502. [Google Scholar] [CrossRef]
  135. Anantachaisilp, S.; Smith, S.M.; Ton-That, C.; Osotchan, T.; Moon, A.R.; Phillips, M.R. Tailoring deep level surface defects in ZnO nanorods for high sensitivity ammonia gas sensing. J. Phys. Chem. C. 2014, 118, 27150–27156. [Google Scholar] [CrossRef]
  136. Djurišić, A.B.; Leung, Y.H. Optical properties of ZnO nanostructu130res. Small 2006, 2, 944–961. [Google Scholar] [CrossRef] [PubMed]
  137. Liu, L.; Mei, Z.; Tang, A.; Azarov, A.; Kuznetsov, A.; Xue, Q.K.; Du, X. Oxygen vacancies: The origin of n-type conductivity in ZnO. Phys. Rev. B 2016, 93, 235305. [Google Scholar] [CrossRef] [Green Version]
  138. Janotti, A.; Van De Walle, C.G. Native point defects in ZnO. Phys. Rev. B Condens. Matter Mater. Phys. 2007, 76, 165202. [Google Scholar] [CrossRef]
  139. Van De Walle, C.G.; Neugebauer, J. First-principles calculations for defects and impurities: Applications to III-nitrides. J. Appl. Phys. 2004, 95, 3851–3879. [Google Scholar] [CrossRef]
  140. Yang, L.; Wang, Z.; Zhang, Z.; Sun, Y.; Gao, M.; Yang, J.; Yan, Y. Surface effects on the optical and photocatalytic properties of graphene-like ZnO:Eu3+ nanosheets. J. Appl. Phys. 2013, 113, 033514. [Google Scholar] [CrossRef] [Green Version]
  141. Wu, X.L.; Siu, G.G.; Fu, C.L.; Ong, H.C. Photouminescence and cathodoluminescence studies of stoichiometric and oxygen-deficient ZnO films. Appl. Phys. Lett. 2001, 78, 2285–2287. [Google Scholar] [CrossRef] [Green Version]
  142. Zhao, Q.X.; Klason, P.; Willander, M.; Zhong, H.M.; Lu, W.; Yang, J.H. Deep-level emissions influenced by O and Zn implantations in ZnO. Appl. Phys. Lett. 2005, 87, 211912. [Google Scholar] [CrossRef]
  143. Chen, C.; He, H.; Lu, Y.; Wu, K.; Ye, Z. Surface passivation effect on the photoluminescence of ZnO nanorods. ACS Appl. Mater. Interfaces 2013, 5, 6354–6359. [Google Scholar] [CrossRef]
  144. Agarwal, S.; Rai, P.; Gatell, E.N.; Llobet, E.; Güell, F.; Kumar, M.; Awasthi, K. Gas sensing properties of ZnO nanostructures (flowers/rods)synthesized by hydrothermal method. Sens. Actuators B Chem. 2019, 292, 24–31. [Google Scholar] [CrossRef]
  145. Lima, S.A.M.; Sigoli, F.A.; Jafelicci, M.J.; Davolos, M.R. Luminescent properties and lattice defects correlation on zinc oxide. Int. J. Inorg. Mater. 2001, 3, 749–754. [Google Scholar] [CrossRef]
  146. Wang, D.F.; Zhang, T.J. Study on the defects of ZnO nanowire. Solid State Commun. 2009, 149, 1947–1949. [Google Scholar] [CrossRef]
  147. Nikitenko, V. Optics and spectroscopy of point defects in ZnO. In Zinc Oxide—A Material for Micro-and Optoelectronic Applications; Springer: Dordrecht, The Netherlands, 2005; pp. 69–81. [Google Scholar]
  148. Van De Walle, C.G. Hydrogen as a cause of doping in zinc oxide. Phys. Rev. Lett. 2000, 85, 1012–1015. [Google Scholar] [CrossRef] [Green Version]
  149. Schirmer, O.F.; Zwingel, D. The yellow luminescence of zinc oxide. Solid State Commun. 1970, 8, 1559–1563. [Google Scholar] [CrossRef]
  150. Kayaci, F.; Vempati, S.; Donmez, I.; Biyikli, N.; Uyar, T. Role of zinc interstitials and oxygen vacancies of ZnO in photocatalysis: A bottom-up approach to control defect density. Nanoscale 2014, 6, 10224–10234. [Google Scholar] [CrossRef] [Green Version]
  151. Vanheusden, K.; Seager, C.H.; Warren, W.L.; Tallant, D.R.; Voigt, J.A. Correlation between photoluminescence and oxygen vacancies in ZnO phosphors. Appl. Phys. Lett. 1996, 68, 403–405. [Google Scholar] [CrossRef]
  152. Bhaskar, R.; Lakshmanan, A.R.; Sundarrajan, M.; Ravishankar, T.; Jose, M.T.; Lakshminarayan, N. Mechanism of green luminescence in ZnO. Indian J. Pure Appl. Phys. 2009, 47, 772–774. [Google Scholar]
  153. Agarwal, D.C.; Singh, U.B.; Gupta, S.; Singhal, R.; Kulriya, P.K.; Singh, F.; Tripathi, A.; Singh, J.; Joshi, U.S.; Avasthi, D.K. Enhanced room temperature ferromagnetism and green photoluminescence in Cu doped ZnO thin film synthesised by neutral beam sputtering. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Dahan, P.; Fleurov, V.; Thurian, P.; Heitz, R.; Hoffmann, A.; Broser, I. Properties of the intermediately bound α-, β- and γ-excitons in ZnO:Cu. J. Phys. Condens. Matter 1998, 10, 2007–2019. [Google Scholar] [CrossRef]
  155. Look, D.C.; Hemsky, J.W.; Sizelove, J.R. Residual native shallow donor in ZnO. Phys. Rev. Lett. 1999, 82, 2552–2555. [Google Scholar] [CrossRef] [Green Version]
  156. Kröger, F.A.; Vink, H.J. The origin of the fluorescence in self-activated ZnS, CdS, and ZnO. J. Chem. Phys. 1954, 22, 250–252. [Google Scholar] [CrossRef]
  157. Chen, H.; Gu, S.; Tang, K.; Zhu, S.; Zhu, Z.; Ye, J.; Zhang, R.; Zheng, Y. Origins of green band emission in high-temperature annealed N-doped ZnO. J. Lumin. 2011, 131, 1189–1192. [Google Scholar] [CrossRef]
  158. Hsu, N.E.; Hung, W.K.; Chen, Y.F. Origin of defect emission identified by polarized luminescence from aligned ZnO nanorods. J. Appl. Phys. 2004, 96, 4671–4673. [Google Scholar] [CrossRef]
  159. Djurišić, A.B.; Choy, W.C.H.; Roy, V.A.L.; Leung, Y.H.; Kwong, C.Y.; Cheah, K.W.; Rao, T.K.G.; Chan, W.K.; Lui, H.F.; Surya, C. Photoluminescence and electron paramagnetic resonance of ZnO tetrapod structures. Adv. Funct. Mater. 2004, 14, 856–864. [Google Scholar] [CrossRef]
  160. Dietz, R.E.; Kamimura, H.; Sturge, M.D.; Yariv, A. Electronic structure of copper impurities in ZnO. Phys. Rev. 1963, 132, 1559–1569. [Google Scholar] [CrossRef] [Green Version]
  161. Dingle, R. Luminescent transitions associated with divalent copper impurities and the green emission from semiconducting zinc oxide. Phys. Rev. Lett. 1969, 23, 579–581. [Google Scholar] [CrossRef]
  162. Reshchikov, M.A.; Xie, J.Q.; Hertog, B.; Osinsky, A. Yellow luminescence in ZnO layers grown on sapphire. J. Appl. Phys. 2008, 103, 103514. [Google Scholar] [CrossRef] [Green Version]
  163. Lv, J.; Li, C.; Chai, Z. Defect luminescence and its mediated physical properties in ZnO. J. Lumin. 2019, 208, 225–237. [Google Scholar] [CrossRef]
  164. Zwingel, D. Trapping and recombination processes in the thermoluminescence of Li-doped ZnO single crystals. J. Lumin. 1972, 5, 385–405. [Google Scholar] [CrossRef]
  165. Cox, R.T.; Block, D.; Hervé, A.; Picard, R.; Santier, C.; Helbig, R. Exchange broadened, optically detected ESR spectra for luminescent donor-acceptor pairs in Li doped ZnO. Solid State Commun. 1978, 25, 77–80. [Google Scholar] [CrossRef]
  166. Alvi, N.H.; Willander, M.; Nur, O. The effect of the post-growth annealing on the electroluminescence properties of n-ZnO nanorods/p-GaN light emitting diodes. Superlattices Microstruct. 2010, 47, 754–761. [Google Scholar] [CrossRef]
  167. Li, D.; Leung, Y.H.; Djurišić, A.B.; Liu, Z.T.; Xie, M.H.; Shi, S.L.; Xu, S.J.; Chan, W.K. Different origins of visible luminescence in ZnO nanostructures fabricated by the chemical and evaporation methods. Appl. Phys. Lett. 2004, 85, 1601–1603. [Google Scholar] [CrossRef] [Green Version]
  168. Reshchikov, M.A.; Behrends, A.; Bakin, A.; Waag, A. Photoluminescence from ZnO nanowires. J. Vac. Sci. Technol. B Microelectron. Nanom. Struct. 2009, 27, 1688. [Google Scholar] [CrossRef]
  169. Greene, L.E.; Law, M.; Goldberger, J.; Kim, F.; Johnson, J.C.; Zhang, Y.; Saykally, R.J.; Yang, P. Low-temperature wafer-scale production of ZnO nanowire arrays. Angew. Chem. Int. Ed. 2003, 42, 3031–3034. [Google Scholar] [CrossRef]
  170. Gomi, M.; Oohira, N.; Ozaki, K.; Koyano, M. Photoluminescent and structural properties of precipitated ZnO fine particles. Jpn. J. Appl. Phys. 2003, 42, 481–485. [Google Scholar] [CrossRef]
  171. Djurišić, A.B.; Leung, Y.H.; Tam, K.H.; Hsu, Y.F.; Ding, L.; Ge, W.K.; Zhong, Y.C.; Wong, K.S.; Chan, W.K.; Tam, H.L.; et al. Defect emissions in ZnO nanostructures. Nanotechnology 2007, 18, 095702. [Google Scholar] [CrossRef]
Figure 1. (a) ZnO unit cell with wurtzite structure. (b) Various crystal planes of the ZnO wurtzite structure [42].
Figure 1. (a) ZnO unit cell with wurtzite structure. (b) Various crystal planes of the ZnO wurtzite structure [42].
Nanomaterials 10 00857 g001
Figure 4. The exciton-bound transition energy reported in [66] at 6 K. D0X (denotes exciton bound to neutral donors), A0X (exciton bounded to a neutral acceptor), TES (two-electron satellite transition), DAP (donor-acceptor pair transitions), and DAP-LO (first-order longitudinal optical). Quenching effect of the D0X/I2 emission as the temperature is increased from 6 to 180 K.
Figure 4. The exciton-bound transition energy reported in [66] at 6 K. D0X (denotes exciton bound to neutral donors), A0X (exciton bounded to a neutral acceptor), TES (two-electron satellite transition), DAP (donor-acceptor pair transitions), and DAP-LO (first-order longitudinal optical). Quenching effect of the D0X/I2 emission as the temperature is increased from 6 to 180 K.
Nanomaterials 10 00857 g004
Figure 5. Low-temperature time-resolved photoluminescence (TRPL) spectra at 3.365 eV for three ZnO NWs arrays with different diameters. A diameter dependence over the exciton lifetime was observed [94]. Copyright 2010 AIP Publishing.
Figure 5. Low-temperature time-resolved photoluminescence (TRPL) spectra at 3.365 eV for three ZnO NWs arrays with different diameters. A diameter dependence over the exciton lifetime was observed [94]. Copyright 2010 AIP Publishing.
Nanomaterials 10 00857 g005
Figure 6. (a) Illustration of different intrinsic point defects in the ZnO lattice. Oxygen vacancy (VO), zinc vacancy (VZn), oxygen interstitial (Oi), and zinc interstitial (Zni) in different charge configurations are presented [125]. Copyright 2006 Royal Society of Chemistry (b) Schematic band diagram of the DLE emissions in ZnO based on the full-potential linear muffin-tin orbital method and the reported data as described in references [126,127,128,129,130,131,132,133,134,135]. Also, oxygen vacancies situated 1.65 eV below the conduction band are denoted to contribute to the red emission [130].
Figure 6. (a) Illustration of different intrinsic point defects in the ZnO lattice. Oxygen vacancy (VO), zinc vacancy (VZn), oxygen interstitial (Oi), and zinc interstitial (Zni) in different charge configurations are presented [125]. Copyright 2006 Royal Society of Chemistry (b) Schematic band diagram of the DLE emissions in ZnO based on the full-potential linear muffin-tin orbital method and the reported data as described in references [126,127,128,129,130,131,132,133,134,135]. Also, oxygen vacancies situated 1.65 eV below the conduction band are denoted to contribute to the red emission [130].
Nanomaterials 10 00857 g006
Figure 7. Visible PL spectra of a single ZnO nanowire measured at different temperatures. Three characteristic emission bands (yellow, green, and blue) were observed [113]. Copyright 2008 Elsevier.
Figure 7. Visible PL spectra of a single ZnO nanowire measured at different temperatures. Three characteristic emission bands (yellow, green, and blue) were observed [113]. Copyright 2008 Elsevier.
Nanomaterials 10 00857 g007
Figure 9. (a) Secondary electron micrograph of the ZnO thin film, (b) CL monochromatic map taken under ZnO NBE emission, (c) Green emission CL monochromatic map at ZnO; all the scale bars are 500 nm long [70]. (d) STEM image of nanowire amplitude maps of (e) near-band-edge emission, (f) surface emission, (g) defect emission 1, and (h) defect emission 2 deconvolved components. (i) Counts integrated across all energies; for further permissions related to the material (d–i), please refer to ACS https://pubs.acs.org/doi/10.1021/acs.jpclett.8b03286 [76].
Figure 9. (a) Secondary electron micrograph of the ZnO thin film, (b) CL monochromatic map taken under ZnO NBE emission, (c) Green emission CL monochromatic map at ZnO; all the scale bars are 500 nm long [70]. (d) STEM image of nanowire amplitude maps of (e) near-band-edge emission, (f) surface emission, (g) defect emission 1, and (h) defect emission 2 deconvolved components. (i) Counts integrated across all energies; for further permissions related to the material (d–i), please refer to ACS https://pubs.acs.org/doi/10.1021/acs.jpclett.8b03286 [76].
Nanomaterials 10 00857 g009

Share and Cite

MDPI and ACS Style

Galdámez-Martinez, A.; Santana, G.; Güell, F.; Martínez-Alanis, P.R.; Dutt, A. Photoluminescence of ZnO Nanowires: A Review. Nanomaterials 2020, 10, 857. https://doi.org/10.3390/nano10050857

AMA Style

Galdámez-Martinez A, Santana G, Güell F, Martínez-Alanis PR, Dutt A. Photoluminescence of ZnO Nanowires: A Review. Nanomaterials. 2020; 10(5):857. https://doi.org/10.3390/nano10050857

Chicago/Turabian Style

Galdámez-Martinez, Andres, Guillermo Santana, Frank Güell, Paulina R. Martínez-Alanis, and Ateet Dutt. 2020. "Photoluminescence of ZnO Nanowires: A Review" Nanomaterials 10, no. 5: 857. https://doi.org/10.3390/nano10050857

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop