Alessandra Catellaniab,
Alice Ruiniac,
Marco Buongiorno Nardellide and
Arrigo Calzolari*ad
aIstituto Nanoscienze CNR-NANO-S3, I-41125 Modena, Italy. E-mail: arrigo.calzolari@nano.cnr.it; Fax: +39 059 367488; Tel: +39 059 2055627
bCNR-IMEM, Parco Area delle Scienze, 37A, I-43100 Parma, Italy
cDipartimento di Fisica, Informatica e Matematica, Universitá di Modena e Reggio Emilia, I-41125 Modena, Italy
dDepartment of Physics, University of North Texas, Denton, TX 76203, USA
eCenter for Materials Genomics, Duke University, Durham, NC 27708, USA
First published on 12th May 2015
Metal-doped ZnO nanowires exhibit the unique property of being simultaneously thermoelectric transparent conductors and low-loss plasmonic materials in the near-IR and visible range. Using calculations from first principles, we identify the mechanisms that regulate this behavior at the nanoscale and we describe how nanostructuring affects the optoelectronic, vibrational and transport properties of In:ZnO nanowires. Our results reveal that In doping imparts a good electrical conductivity and provides an injected free charge sufficient to sustain a surface-plasmon-polariton excitation. At the same time, surface scattering effects efficiently quench the thermal conductivity along the wire, improving the thermoelectric figure of merit of the system with respect to the bulk material. The coexistence of plasmonic and thermoelectric characteristics fosters the design of a novel class of coupled nanostructured devices for photothermal-electrical energy conversion.
The design of mixed architectures that could positively exploit both light and heat conversion would represent an important step forward in the realization of more efficient devices. Particularly promising are recent prototypical systems that couple a plasmonic heater to a thermoelectric device:4–6 the former acts as photothermal converter, through an excitation–dissipation process mediated by a plasmon,7,8 which activates a thermoelectric power generator.9,10 Along these lines, Xiong and coworkers4 demonstrated that the power conversion efficiency of a commercial thermoelectric unit is enhanced by ∼5% upon coating with plasmonic Ag nanoparticles.
Despite these promising results, this technology is, however, in its very infancy and many challenges must be addressed to improve the efficiency of these devices. In particular, the choice of the materials is one of the most crucial problems, because in order to minimize losses and to maximize the thermal transfer between the plasmonic heater and the thermoelectric converter, the two systems should be prepared with the same material.11 This constitutes a strict limitation: in order to promote plasmonic excitations in the visible and near-IR range (i.e. the part of the spectrum that corresponds to the maximum of solar radiation) the plasmon heater must include metallic systems with a huge amount of free electrons. On the other hand, noble metals like silver or gold are not good thermoelectric materials, as they are very efficient thermal conductors. On the contrary, standard bulk thermoelectric materials are doped semiconductors with high electron mobility, and low thermal conductivity, but no plasmon activity in the low-energy region.
The search for materials that are simultaneously plasmonic and thermoelectric is a tremendous challenge. Among the potential candidates, ZnO seems to uniquely fulfill all the requirements. ZnO is a direct bandgap wurtzite semiconductor with high electrical conductivity, efficient luminescence and strong excitonic effects even at room temperature. Due to its large band gap (3.4 eV), intrinsic ZnO is transparent and represents an attractive choice for applications in ultraviolet light emitters, field-effect transistors, sensors, piezoelectric devices and polariton laser.12 Upon metal doping (e.g. In, Al, Ga), ZnO acts as a transparent conductive oxide (TCO)13–15 with a large amount of free conduction charge (>1019 e cm−3), able to support bulk plasmon oscillations and surface plasmon polaritons (SPP) in the visible and near-IR range, including the standard telecommunication wavelength (1.5 μm).16–18 Very recently, it has been also demonstrated that plasmonic ZnO nanowires can be used to realize coherent sources of surface plasmon polaritons at the nanoscale and surface plasmon emitting diodes.19
In virtue of the wurtzite structure, ZnO may be easily grown in ordered arrays of nanostructures oriented and elongated along the polar axis,20,21 such as wires and tetrapods, that have been largely applied in optoelectronic devices like solar cells and LEDs.22,23 Even though the intrinsic thermoelectric properties of bulk ZnO are rather poor,24 ZnO nanowires have been recently used as active element in thermoelectric devices with enhanced ZT figure of merit.25–27 In particular, two-terminal transport measurements on individual suspended ZnO nanowires in vacuum28 demonstrated that moderate doping can increase the electrical conductivity without decreasing the Seebeck coefficient and that thermal conductivity of ZnO NWs is much smaller than the bulk one over a very large range of temperature (300–1000 K).
Many fundamental issues, however, remain unclarified. For instance, there is no consensus on the possibility to highly dope nanostructures without compromising their structural and electronic stability29,30 or to induce a TCO behavior in 1D structures. Furthermore, it is not confirmed that metal doping may provide enough free conduction charge to the host to support plasmon excitations in nanowires.31 Finally, the effect of the boundary surfaces on the thermal transport of the wire and its effect on the thermoelectric properties is still an open question.
In this paper, we provide a first principles characterization of the electronic, plasmonic and thermoelectric properties of In-doped ZnO (IZO) nanowires, focusing on the effects of doping and of the presence of boundary surfaces on the intrinsic optoelectronic and transport properties of the wires. Our results confirm that arrays of IZO wires are suitable candidates to realize both plasmon heaters and thermoelectric generators, which can be integrated in a unique device, opening the way for a new generation of photothermal-electrical energy converters.
The experimental results38 show that within the solubility limit (∼3.5%) group IIIA metals (Al, Ga, In) dope ZnO in Zn-substitutional sites, with very similar effects. Indium, with its large ionic radius, is frequently used in lab experiments because it is easily recognizable through microscopy techniques.39 With respect to bulk case, when doping a nanowire we have to distinguish between external and internal defect sites. Here, we considered IZO wires obtained substituting one Zn atom from the center (position 1) or from the surface (position 2) of the wire, as shown in Fig. 1b. In both cases, the inclusion of the dopant does not perturb the crystalline order of the ZnO host, in agreement with what observed for the 3D bulk phases. This is a first fundamental result as the dopability of the nanostructures is a well-known challenge.40,41 Due to the high surface-to-bulk ratio and the high reactivity of uncapped nanostructures, deriving from the abundance of frustrated bonds, the inclusion of dopants may indeed cause strong geometrical deformations that destabilize the structure and/or destroy the bond properties of the material under-/over-coordinating the atoms the host. This generally introduces trap states in the gap that are detrimental for the optical and transport properties of system. Thus, the structural stability of IZO nanowires is a fundamental prerequisite for the applications described above.
The DOS plots of IZO wires are displayed in panel (b) and (c) of Fig. 2, respectively; panel (d) shows the DOS for the 3D IZO bulk, in the wurtzite structure, included for comparison. From Fig. 2 it is evident that the DOS of the doped and undoped wires are qualitatively very similar to the 3D case, and the original bandgap of ZnO (ΔE1) is easily recognizable in all systems. However, while the presence of In in the 3D bulk does not change the shape and the curvature of the conduction band minimum, band modifications are more pronounced in 1D systems.
This is also reflected by the small but not-negligible differences induced by doping in the ΔE1 values (Table 1) for the wires, which accounts for a reduction of the internal ionicity, due to the presence of the In atoms. The inclusion of the dopant causes a decrease of the original ZnO gap (ΔE1), which partially compensates the gap opening induced by quantum confinement. IZO wires further exhibit a flattening of the lowest conduction band, which corresponds to an increase of the electron effective masses m*e (Table 1). This effect is larger in the case of surface dopant (2), whose effective mass is almost twice the one of bulk-like defects (1).
ΔE1 (eV) | ΔE2 (eV) | m*e/m0 | ωp (eV) | ne (cm−3) | Eoptg | |
---|---|---|---|---|---|---|
ZnO(wire) | 3.23 | — | 0.34 | — | — | 3.23 |
IZO(1) | 2.94 | 0.67 | 0.38 | 0.64 | 1.1 × 1020 | 3.85 |
IZO(2) | 2.61 | 0.39 | 0.66 | 0.34 | 5.3 × 1019 | 3.38 |
ZnO(bulk) | 3.10 | — | 0.29 | — | — | 3.10 |
IZO(bulk) | 3.13 | 0.80 | 0.29 | 1.24 | 3.1 × 1020 | 4.07 |
For all IZO systems (wires and bulk), no defect states appear in the gap: indium donates its 5p electron to the ZnO host, shifting the Fermi level into the original conduction band of ZnO. However, the amount of free electron charge injected in the ZnO matrix depends on the system. In order to quantify this effect, we defined the index ΔE2, which is the energy difference between the Fermi level of IZO and the conduction band minimum of ZnO, as shown in Fig. 2. The deeper the Fermi level is shifted into the conduction band (i.e. larger ΔE2), the larger is the charge injection. The best defect-to-host electron transfer is for IZO bulk (Table 1), followed by IZO(1) and IZO(2), whose ΔE2 is reduced by ∼50%, with respect to the bulk case. We can thus conclude that In-doping imparts a n-type conductive character to ZnO wires, while the details of the electronic properties depend on the specific doping site: the inner site (1), thanks to the optimal coordination with the next-neighbor oxygen atoms, displays a very good charge delocalization, very similar to the 3D case. On the contrary, the under coordinated In-atom on the surface (IZO(2)) favors a charge localization around the defect site and a reduction of the available free conduction charge. This is confirmed also by a residual magnetism for IZO(2) system, not detected in the other cases.
By using a semiclassical Drude–Lorentz model, based on independent particle band-to-band transitions, we calculated the complex dielectric function = ε1 + iε2 and the corresponding electron energy loss (EEL) function L(ω) = Im{−1/
}. The spectral plots for the real and imaginary part of the dielectric function are shown in ESI (Fig. S5†). The results confirm that the undoped wire exhibits the typical features of a wide bandgap semiconductor. The inclusion of Indium imparts a metallic behavior in the infrared (i.e. negative ε1), while it induces a blue-shift of the absorption edge in the UV-range, which preserves the transparency of the doped systems. The transparency in the visible range along with the electrical n-type characteristics confirms the TCO behavior displayed by both wire and bulk IZO compounds. The demonstration of TCO-like properties for the wire systems is a particularly important result, as low-dimensional TCO materials may be exploited as transparent contacts in a large range of optoelectronic nanostructured devices.
Fig. 3 shows the EEL spectra for the IZO materials. When ε1 = 0 and ε2 ≪ 1, the loss function has a peak, corresponding to the plasma frequency ωp of the system (Table 1). As the plasma frequency lies in the near-IR for both wires and bulk, while the absorption energy is the UV, we expect a drastic reduction of the energy losses due to interband transitions in the NIR-vis operating range, the one with the maximum sunlight power distribution. This would optimize the conversion of the incoming radiation into thermal energy through the excitation/de-excitation of plasmon resonances, and minimize the energy dissipation via absorption/emission radiative processes.
![]() | ||
Fig. 3 EELS spectra of (a) undoped ZnO wire, (b) IZO wire, configuration 1, (c) IZO wire, configuration 2, (d) IZO bulk. |
In agreement with the analysis of the electronic structure, we identify some numerical differences in the energy position ωp, despite the fact that three systems have the same formal doping level. This can be directly related to the amount of injected free electron charge. Starting from the definition of the plasma frequency (where e is the electron charge, ε0 the dielectric permittivity of vacuum and m*e the electron effective mass), we extracted the free electron density ne. Consistently with the DOS results presented above, configuration (2) presents lower available charge than configuration (1) and IZO bulk. The resulting values for the IZO wires are however larger than the empirical lower concentration (1019 cm−3), required to sustain a plasma excitation.
The interface between IZO wires and vacuum can be considered as the simplest metal/dielectric interface, along which it is possible to excite a surface-plasmon polariton. As in our simulation we explicitly take into account the effect of the surface/vacuum interface, the calculated ωp can be directly assumed as the frequency of the surface plasmon polariton ωspp. Notably, the ideal SPP frequency obtained from the expression eV is only slightly larger than the value directly obtained from IZO wire (Table 1). This difference accounts for the confinement effect acting on the true 1D systems and absent in bulk calculations.
We can conclude that In-doped ZnO nanowires act as 1D TCO materials that can be exploited as low loss plasmonic elements in the near-IR and visible range. Thus, they are very promising candidates to realize a plasmon heater, i.e. converting the solar light into thermal radiation through the excitation/de-excitation of a SPP along the wire.
We simulated both electron and thermal coherent transport along ZnO nanowires by using the WanT code,44 which exploits the complete energy band structures and phonon dispersions to solve an extended Landauer problem, within a real-space Green's function framework. This approach allows us to directly link the transport properties (e.g. quantum conductance and I/V characteristics) to the dimensionality, and atomistic structure of the system. The method holds for a generic two-terminal open device (left-lead/conductor/right-lead). We here focus on the intrinsic conduction properties of the wires that are the theoretical prerequisite for any further two-terminal device simulation. This is easily done considering the two external leads made of the same material of the conductor (i.e. the ZnO wire). We refer to original papers45–48 for the complete description of the theory and the code implementation.
In the Landauer approach, a key quantity to be calculated is the quantum transmittance, which represents the probability that a carrier (either electrons or phonons) may cross the conductor at a certain energy. The electron (el) and phonon (
ph) transmittance plots are shown in panels (a–c) and (d) of Fig. 4, respectively (see ESI† for further details). The electronic contributions have been explicitly evaluated for the ZnO and the two IZO wires under investigation. In the case of thermal transport, as the In-substitution only slightly perturbs the atomic structure, we calculated the
ph function only for the undoped ZnO wire, assuming as negligible the effect of Indium. In order to prove this statement, we checked a posteriori that the vibrational properties and, thus, the thermal transport were not changed, when the inclusion of In was considered in the mass defect approximation.49
![]() | ||
Fig. 4 Electron transmittance for (a) undoped ZnO wire, (b) IZO wire, configuration 1, (c) IZO wire, configuration 2. (d) Phonon thermal transmittance for un-doped ZnO wire. |
All the spectra in Fig. 4 have a step-like behavior typical of periodic systems. In the absence of external leads, at a given value of energy (wavenumber), the quantum transmittance is constant and proportional to the number of transmitting channels available for charge (phonon) mobility, which are equal to the number of conducting bands at the same energy (see also Fig. S6, ESI†). The electron quantum transmittance in particular well describes the electron donor effect due to doping, along with the quantitative differences in the position of the Fermi level between internal and external doping sites.
Although in periodic systems there is a direct correlation between transmittance and band structure, not all the states contribute to transport. This is particularly important in the case of thermal transport, that should be as low as possible in order to maximize the thermoelectric response of the system. In order to understand the features of the thermal conductance, we have carried out a microscopic analysis of the transmittance in terms of the most relevant contribution from the vibrational normal modes. The phonon DOS of the ZnO wire is shown in Fig. 5a, along with the corresponding bulk one, taken as reference. Following the irreducible representation of the wurtzite symmetry group C6v, the phonon modes of ZnO bulk can be classified as Γ = 2A1 + 2B1 + 2E1 + 2E2. One low energy A1 and one double E1 modes correspond to the transverse and longitudinal acoustic branches, while the others are optical modes.
In the nanowire, the presence of the surface breaks the atom equivalence of the bulk, introducing new phonon modes. This results in the spread of optical phonon branches at about ∼350–430 and at 600–700 cm−1 and in a general redistribution of the spectral weight in the entire wavenumber range, in agreement with Raman experimental data.50 The rotational invariance along the wire axis gives rise to a fourth zero-frequency mode at q = 0. This can be directly detected in the transmittance, which is ph = 4, in the zero-frequency limit. From the analysis of the single phonon displacements we can also distinguish between core-longitudinal and surf-transverse modes. The former contribute to coherent thermal transport, the latter do not transmit along the wire. Two representative phonon modes are displayed in panels (b and c) of Fig. 5. It is evident that the surface acts as an extended scattering defect that blocks part of the phonon modes. This is particularly true for ultrathin nanowires, where the ratio between surface and core modes is very high. The phonon component of the thermal conductance (Kph), resulting from the integration of the corresponding thermal transmittance
ph (see Fig. S6, ESI†) at T = 300 K is ∼2 nW K−1, i.e. one order of magnitude less than the corresponding bulk value. This goes in the direction of intrinsic reduction of the phonon thermal transport, as desired for thermoelectric application.
In the coherent transport regime, the figure of merit reduces to ZT = S2GelT/Kt, where Gel, Kt are the electron and thermal conductances, respectively (see ESI†). Once the thermal transmittance functions el and
ph are known, the thermopower coefficient and the thermoelectric ZT figure of merit can be straightforwardly obtained by using standard kinetic relations51 (see ESI†). The calculated ZT plots for the undoped and doped wires are displayed in Fig. 6 in the temperature range 200–400 K. The figure of merit generally increases with temperature, while the dependence on the chemical potential depends on the original position of the Fermi level of the system. For instance, the edge position of the ZnO wire at ∼1.75 eV reflects the presence of an energy gap. The absolute value of ZT is instead directly related to the intrinsic electron conductance (e.g. effective mass) of the system, so the ZT value of the undoped wire is almost double the value of IZO wires. The calculated ZT values for the wire are about one order of magnitude larger than the corresponding measured value for the ZnO bulk (ZT < 0.001) and are in very good agreement with the experimental results for Al-doped ZnO wires at the same temperatures.52 This confirms that: (i) the coherent transport approximation is accurate enough to describe the thermoelectric properties of ultrathin nanowires and (ii) nanostructuring is a promising way to increase the thermoelectric response of materials.
![]() | ||
Fig. 6 (a) ZT figure of merit evaluated at T = 300 K for (a) undoped ZnO wire, (b) IZO wire, configuration 1, (c) IZO wire, configuration 2. |
This unusual coexistence of plasmonic and thermoelectric properties may be exploited to realize coupled surface-plasmon/thermoelectric power generators where both active elements are realized with arrays of IZO wires. An illustrative scheme of this kind of device is displayed in Fig. 7. The nano-heater converts the absorbed photons into heat (i.e. photothermal effect) through the excitation of a surface plasmon polariton (SPP) along the wire. The photothermal effect involves complex excitation–dissipation processes that include:7,8 the excitation of plasma oscillations in a quasi-free electron gas; the ultrafast thermalization of the hot-electron gas through inelastic electron–electron scattering; the energy transfer to crystal lattice via phonon emission (electron–phonon coupling)53 and thermal radiation (heat diffusion) between nanowires and towards the external environment. The thermal energy resulting from the plasmon decay can be exploited to generate a temperature gradient across the thermoelectric unit, whose temperature is controlled by the presence of thermal bath. An intermediate gap spacer between the two units of the device prevents the overheating of the thermoelectric part, which is detrimental for the energy conversion efficiency. The manufacturing of such a device would constitute a major step forward in the quest for efficient systems for green energy conversion.
The pathological underestimation of the ZnO bandgap (characteristic of the standard DFT semi-local approaches) is corrected by using the newly developed ACBN0 functional.33 ACBN0 is a pseudo-hybrid Hubbard density functional that is a fast, accurate and parameter-free extension of traditional DFT + U that has been proved to correct both the band gap and the relative position of the different bands in transition metal compounds, in particular the ones deriving from the d orbitals of transition metal atoms. Within ACBN0, the values of U and J are functionals of the electron density and depend directly on the chemical environment and crystalline field, thus providing a direct way of computing the Hubbard corrections for any individual atom in any local environment. ACBN0 self-consistent U values are 12.8 eV 5.29 eV for the 3d orbitals of zinc and 2p orbitals of oxygen, respectively, in perfect agreement with previous parameters reported by Calzolari et al.34 (U3d = 12.0, U2p = 6.5 eV) and Ma et al.56 (U3d = 10, U2p = 7 eV), both of which were found by a fitting procedure to reproduce the experimental bandgap and position of the 3d bands.
ZnO nanowires are simulated by using periodically repeated orthorhombic supercells (108 atoms). Parallel replica are separated by ∼12 Å of vacuum in the directions perpendicular to the wire axis. IZO wires are obtained substituting one Zn atom with an In one in different positions (see Fig. 1), this corresponds to formal doping of ∼1.0%. 10 k-points along the wire axis are used to sample the 1D Brillouin Zone (BZ). 3D IZO bulk is simulated by using a hexagonal (3 × 3 × 3) ZnO supercell, also including 108 atoms; i.e. corresponding to the same doping value. In this case, a (6 × 6 × 6) k-point mesh is used for the BZ sampling. All structures were relaxed until forces on all atoms were lower than 0.03 eV Å−1.
The complex dielectric function is calculated in the independent particle approximation, using the code epsilon.x, also included in the QUANTUM ESPRESSO suite. The code implements a band-to-band formulation of the Drude–Lorentz model for solids.57,58 The complete theoretical treatment and the accuracy tests for the case of Al-doped ZnO system can be found in ref. 18. In this case DFT calculations are performed employing norm conserving pseudopotentials with an energy cutoff of 100 Ry. Once the complete dielectric function is known, the electron energy loss function can be easily obtained as L(ω) = Im{−1/
}.
Electronic and thermal coherent transport characteristics are simulated by using the WanT package,44 which provides a unified real-space implementation of the Landauer theory based on Green's function technique,45,59 for both electrons46 and phonons.48 The real space electronic hamiltonian is obtained from the DFT calculation through a pseudo-atomic projection procedure, as described in ref. 47. The phonon spectrum and Interatomic Force Constant (IFC) matrix are simulated with a joint finite-differences/finite-fields approach, also implemented in the QUANTUM ESPRESSO/package.60 For the calculation of phonon modes we considered a (3 × 1 × 1) orthorhombic supercell (324 atoms) and we calculated forces (i.e. IFC) displacing only 108 atoms in primitive cell along the three spatial directions. This corresponds to a set of 324 DFT calculations.
Footnote |
† Electronic supplementary information (ESI) available: ESI includes complementary details on the electronic structure, the real and imaginary part of the dielectric function and transport properties of ZnO and IZO wires. See DOI: 10.1039/c5ra06199g |
This journal is © The Royal Society of Chemistry 2015 |