Home

Search

Collections

Journals

About

Contact us

My IOPscience

Piezoelectric properties of zinc oxide nanowires: an ab initio study

This content has been downloaded from IOPscience. Please scroll down to see the full text. 2013 Nanotechnology 24 475401 (http://iopscience.iop.org/0957-4484/24/47/475401) View the table of contents for this issue, or go to the journal homepage for more

Download details: IP Address: 143.167.2.135 This content was downloaded on 07/07/2014 at 13:08

Please note that terms and conditions apply.

IOP PUBLISHING

NANOTECHNOLOGY

Nanotechnology 24 (2013) 475401 (5pp)

doi:10.1088/0957-4484/24/47/475401

Piezoelectric properties of zinc oxide nanowires: an ab initio study K K Korir1,2 , G Cicero1,2 and A Catellani2,3 1 2 3

Dipartimento Scienza Applicata e Tecnologia, Politecnico of Torino, Torino, Italy CNR–IMEM, Parco Area delle Scienze 37A, 1-431100 Parma, Italy CNR-NANO, Istituto Nanoscienze, Centro S3, I-41125 Modena, Italy

E-mail: [email protected]

Received 6 May 2013, in final form 13 September 2013 Published 31 October 2013 Online at stacks.iop.org/Nano/24/475401 Abstract Nanowires made of materials with non-centrosymmetric crystal structures are expected to be ideal building blocks for self-powered nanodevices due to their piezoelectric properties, yet a controversial explanation of the effective operational mechanisms and size effects still delays their real exploitation. To solve this controversy, we propose a methodology based on DFT calculations of the response of nanostructures to external deformations that allows us to distinguish between the different (bulk and surface) contributions: we apply this scheme to evaluate the piezoelectric properties of ZnO [0001] nanowires, with a diameter up to 2.3 nm. Our results reveal that, while surface and confinement effects are negligible, effective strain energies, and thus the nanowire mechanical response, are dependent on size. Our unified approach allows for a proper definition of piezoelectric coefficients for nanostructures, and explains in a rigorous way the reason why nanowires are found to be more sensitive to mechanical deformation than the corresponding bulk material. S Online supplementary data available from stacks.iop.org/Nano/24/475401/mmedia (Some figures may appear in colour only in the online journal)

failure and fractures of novel nanodevices. The working principle of nanogenerators relies on a beneficial coupling between piezoelectricity and semiconducting properties to produce a net current through Schottky contacts between the wide-gap semiconductor NWs and a metal coated AFM tip. Nonetheless, controversial results appear in the literature, depending on the material, on the nanostructure form, and on the growth direction [8–10]. From the experimental point of view, indeed, the characterization of mechanical properties of NWs can be severely affected by artifacts linked to the specific experimental setup, such as AFM on a single sample or a network, measurements performed on clamped or free standing samples, on conductive versus insulating substrates, etc (see e.g. Espinosa et al [9]). On the theoretical front, as well, while it has been assessed that spontaneous polarization of wires strongly depends on the nanostructure size [11], the question whether size affects piezoelectric properties still remains open. In particular, Cicero et al [11] established the existence of a

Zinc oxide (ZnO) has attracted extensive attention for half a century because of its excellent performance in optoelectronics, and spintronics [1, 2]. Most recently, we have been witnessing a significant increase of interest in ZnO nanostructures due to their unique physical, chemical, and electronic properties with potential applications in a wide range of technological fields, which include fundamental building blocks for fabrication of devices for energy harvesting [3], ultrasensitive gas sensors [4], nanoresonators [5], and nanocantilevers [6]. It has been recently proposed that ZnO nanowires (NWs) can be successfully employed for converting mechanical energy into electric power via atomic force microscopy [7], although this result has been questioned [8]: the underlying mechanisms indeed still await a microscopic description to enhance and take full advantage of the structural and electric properties of this material at the nanoscale size. Beyond this, a better knowledge of the elastic response of nanostructures to external deformations would be beneficial to prevent 0957-4484/13/475401+05$33.00

1

c 2013 IOP Publishing Ltd Printed in the UK & the USA

Nanotechnology 24 (2013) 475401

K K Korir et al

minimum diameter below which the NW polarization field is inverted with respect to bulk because of large surface effects in small nanostructures. As for the piezoelectric response of ZnO NWs, Xiang et al [12] proposed that NWs with diameter larger than 2.8 nm tend to have almost constant effective piezoelectric constants (close to the bulk value), while for small NWs, they found a non-trivial dependence of electromechanical coupling on the radius as a result of two competitive effects, i.e., increase of the lattice constant along with decrease of the NW radius. In contrast a giant piezoelectric response was predicted by Agrawal et al [13] of approximately two orders of magnitude larger than bulk for NWs with diameter lower than 1 nm; such an increase should be attenuated for NWs with diameters exceeding 1.5 nm. It is noteworthy that effective piezoelectric coefficients of ZnO NWs reported in the above mentioned studies [12–14] were determined using similar computational approaches, yet leading to dramatically different results. The origin of the above mentioned debate is due to the sources of uncertainty, that, although often neglected, severely alter the volume definition of a nanostructure: these are, namely, surface relaxation (dx) and ionic radius (see figure 1). While the first source of error can be easily accounted for via accurate calculations or dedicated experimental techniques, the second issue is more tricky. The volume definition used to calculate polarization of nanowires is usually not mentioned in previously published papers. However, the nanowire size can be defined through its diameter either by considering the atoms as point charges or as finite spheres (e.g. via the ionic radius): this is enough to obtain an uncertainty of few ˚ in the case of ZnO). This Angstroms on its value ('2 A uncertainty on the diameter has increasing relevance at the nanometer scale, and indeed leads to a large uncertainty in the volume occupied by the nanostructure (up to 70% in the case of the smallest NW considered in this work): the smaller the structures the larger is the uncertainty. As such, all the physical quantities normalized to volume, like spontaneous polarization and piezoelectric constants, present large variability depending on the particular volume definition employed in the calculations4 . In this paper, we propose a rigorous method to overcome the problem: in our study we adopt a scheme that passes through a normalization in terms of number of formula units per supercell (a ZnO pair in the case of zinc oxide nanosystems) without loss of generality. This approach was successfully employed in [11] to describe the size effect on the spontaneous polarization of nanowires, while correctly reproducing bulk values at increasing diameter. In particular, we take advantage of the Wannier functions (WFs) definition that allows one to calculate local contributions (one for each ZnO couple) to the average dipole of a system. In this way, we show that the piezoelectric properties of NWs are only mildly influenced by size reduction to the nanoscale. The procedure allows us to define in a unique rigorous way the different contributions (bulk versus surface) relevant to

Figure 1. Top: cross section view of ZnO NWs with different radii as indicated by dashed circles and labeled NW1, NW2, NW3, and NW4. Central panel: side view of the largest wire, NW4, (two repetitions of the unit cell are shown along the [0001] wire axis). Gray (red) spheres represent zinc (oxygen) atoms; d and D represent two possible choices of the NW diameter, calculated either when atoms are considered as point charges or as spheres, respectively; dx indicates the uncertainty induced by surface relaxation. Bottom: radial LD analysis in NWs: each point represents the average LD in concentric cylindrical shells at increasing distance from the NW center located at the zero of the abscissa axis.

evaluate the nanostructure response to external deformation, and thus finally opens the possibility of an optimized design of novel piezo-devices at the nanoscale. Density functional theory (DFT) in conjunction with maximally localized Wannier function (MLWF) was used to perform a detailed study of piezoelectric properties of ZnO NWs. All calculations were performed with reference to the theoretical DFT equilibrium lattice parameters of the ZnO wires [11], using the ultrasoft pseudopotential plane wave implementation of DFT in the Quantum Espresso code [15], with energy cutoff of 28 Ryd (280 Ryd) for the wavefunctions (charge density); Wannier functions for ZnO bulk and NWs were evaluated using the wanT [16] code. The exchange–correlation energies were calculated by

4 The approach used in determining the volume of nanostructures is usually

not discussed in publications. 2

Nanotechnology 24 (2013) 475401

K K Korir et al

Table 1. Structural, mechanical, and piezoelectric properties of ZnO in bulk and NW form. N is the number of hexagonal shells of the NWs, d (nm) is the diameter of the ZnO NWs with atoms taken as point charge, c represents the lattice periodicity along the NW axis, hLDi/strain is the rate change of average LD with respect to strain, and U 00 is the effective strain energy (see equation (6)).

N d (nm) ˚ c (A) hLDi/strain (D) U 00 (eV)

Bulk

NW4

NW3

NW2

NW1

∞ ∞ 5.315 9.53 30.75

4 2.30 5.361 9.95 22.37

3 1.65 5.385 10.19 21.94

2 0.99 5.423 10.81 19.44

1 0.38 5.397 13.11 15.94

in the case of a nanowire, this is a quantity that contains both bulk and surface effects. More specifically, local contributions (local dipoles—LDs) can be obtained by partitioning the WF set and ionic charges into neutral units (ZnO couples) in such a way to obtain zero polarization for the reference bulk ZB phase structure [11]. Thus, the polarization of the system can be featured simply by variations of local dipoles. Indeed, from a radial analysis of LDs, obtained by averaging LD contributions of ZnO pairs contained in concentric cylindrical shells at increasing distance from the NW center, one can see that for the largest wires considered in this study (NW3 and NW4) bulk behavior (hLDi = −0.24 D) is quickly recovered after the outer-most shell (see bottom panel of figure 1). In particular, for these two wires, one can clearly identify a ‘core’ part in which ZnO pairs have bulk-like behavior (points of the flat region of figure 1) and a surface shell that strongly deviates from bulk-like behavior. In this way, one eliminates inconsistencies and arbitrariness linked to the common relations that use diameter or volume of the system: the scheme provides a method to define in an unambiguous and rigorous way a normalization useful in particular for nanostructures, although recovering bulk limits at increasing size [11]. Within this approach, the piezoelectric response of the ZnO wires grown along the [0001] axis, which is also the direction of spontaneous polarization, can be calculated in terms of variation of hLDi with respect to strain. The results for longitudinal strain ranging from −3% to 3% are reported in figure 2(a). It is apparent that apart from the smallest wire, which structurally does not have any bulk-like ZnO pair, the behavior is well approximated by a linear fit whose slope is related to the piezoelectric response of the wire (see table 1). Also the analysis in terms of hLDi/strain reveals that NWs and bulk ZnO have similar piezoelectric behavior. Only NW1 is characterized by a fairly large increase of the slope, but this system is most probably unphysical. This close similarity between bulk and nanostructure piezoelectric responses reveals that the high enhancement of piezoelectric constants reported in [13] should not occur, and it is possibly related to different effects, other than the NW size. Taking advantage of the local analysis previously discussed, for NW3 and NW4 it is possible to separate the surface and core contribution to hLDi as a function of strain. The surface contribution is obtained by averaging LD on the ZnO pairs that deviate from bulk-like behavior and are unambiguously identified by the radial analysis reported in the bottom panel of figure 1, while the core contribution is obtained by averaging over bulk-like inner ZnO shells (results are reported in panel (c) and panel (b) of figure 2 respectively). Interestingly, also the surface contribution to the wire response is almost linear and the values of hLDi/strain have values of 1.79 D and 1.25 D for NW3 and NW4 surfaces respectively (to be compared with the values reported in table 1). This result indicates that the surface response to deformation is about one order of magnitude smaller than the bulk-like term and only slightly dependent on size; furthermore surface contributions become quickly negligible and the surface/volume ratio decreases. (See the

using the Perdew–Burke–Ernzerhof [17] approximation of the generalized gradient approach. The NWs were modeled in periodic supercells with the z direction corresponding to the wire axis (equilibrium c lattice constants are reported in table 1) and lateral lattice parameters (along the x and y axes) large enough to avoid interaction between periodic replicas. The convergence of k-space sampling was studied and a Monkhorst pack grid of 1 × 1 × 6 was used for the DFT calculations. All atoms were allowed to move, until ˚ −1 . This procedure leads to forces were less than 0.02 eV A fully relaxed nanostructures, including surface contribution5 . In this way, we were able to study NWs with diameter in the range 0.38–2.3 nm containing from 1 to 4 ZnO shells (NW1–NW4) as shown in figure 1. According to the modern theory [18, 19], the macroscopic polarization of a periodic structure, P, is defined only with respect to a reference system; for example, in the case of wurtzite (WZ) one can choose the zincblende (ZB) phase, since the two structures are equivalent up to third neighbors6 . In terms of WFs [20] P can be written as P = PWZ − PZB =

e X 2e X W ZI 1RI − 1ri  I  i

(1)

where we indicate with riW the positions of the Wannier centers, with RI (ZI ) the ionic positions (charges) and with  the volume of the system. For a periodic structure, P is defined modulo 2eRl / , Rl being a direct lattice vector. We stress again that in the presence of a surface, or for the case of nanostructures, the volume of the system is an ill-defined quantity, thus a different definition is required. In the previous formula, if one divides by the number of ZnO pairs that compose the system instead of dividing by , one gets a dipole averaged over the whole structure (hLDi) [11]: 5 Surface relaxation has been shown to affect dramatically the properties of

nanostructures [11, 13, 24]. We obtain a significant relaxation characterized by bond length contractions up to 6.7% for the outer-most ZnO dimers, accompanied by an intrinsic in-plane shrinking strain which decays significantly from the surface to the core region and finally approaches zero, in agreement with other results reported in literature [25, 26]. This surface relaxation is responsible for polarization inversion in small wires [11] and promotes the uncertainty in nanostructure volume definition, as evidenced in figure 1. 6 In experiment, only polarization changes are accessible. Similarly the polar behavior of a system is defined and calculated as the difference with respect to a reference structure (Pref ): one usually takes a non-polar phase as reference structure that has null polarization. 3

Nanotechnology 24 (2013) 475401

K K Korir et al

Figure 2. (a) Plot of hLDi versus strain (3 ) for NWs with increasing diameter; (b) and (c) represent the core (surface) contribution alone, for NW3 and NW4 compared to bulk.

supplementary data section (available at stacks.iop.org/Nano/ 24/475401/mmedia) for a comparison with the more standard approach, in terms of bond deformations). These results are comparable to those reported in [12, 14], but significantly smaller than those reported in [13], probably due to the use of incorrect volume values in their calculated quantities: this supports the idea that different operating conditions motivate the spread in experimental results (see Espinosa et al [9] and references therein), and that surface defects may hinder a more trivial interpretation. Since in the core (bulk-like) region of the wires the volume that each ZnO couple occupies is a well defined quantity, one can recall the usual definition of polarization in terms of volume, and write the polarization change induced by strain, in terms of the bulk piezoelectric constants. For a WZ structure elongated along the spontaneous polarization axis [0001], and free to contract or expand along the perpendicular axes, as in the case of NWs, one has P − P◦ = 2e31 1 + e33 3 = (e33 − 2e31 ν)3 ,

V0 is the equilibrium volume/ZnO pair, i.e. at zero strain. We then use this expression to fit the data reported in figure 2(b) letting V0 , P◦ , e31 , e33 and ν act as free parameters. From the latter values it is possible to estimate eeff 33 for the wire cores: for NW3 and NW4 we obtain Poisson ratio, ν, values of 0.27 and 0.28 respectively that well compare to the value of 0.29 calculated for ZnO bulk. Correspondingly, the effective piezoelectric constants of NW3 and NW4 thus estimated are 1.19 and 1.21 C m−2 respectively, very close to the DFT bulk value of 1.28 C m−2 (experimental 1.22 C m−2 [21]). In conclusion, our analysis shows that piezoelectric constants expressed in terms of hLDi are negligibly influenced by the NW size: this is further demonstrated by a quick recovery of the bulk value. Surface effects and quantum confinement which are normally anticipated to dramatically modify properties of nanostructures seem to have minor influence especially on piezoelectric properties. Similar findings have been reported by Dai et al [22], who, using DFT, showed that for thin films the piezoelectric constant converges rapidly with increase in film thickness, in contradiction of classical molecular dynamics results where slow convergence was predicted [13]. In order to explain the energy harvesting mechanism observed in ZnO NWs [7], and to describe why it is expected that nanostructures perform better than bulk systems, we analyzed how nanowires respond to external mechanical deformation and evaluated their effective spring constants. In the elastic regime the strain energy, U, required to elongate the lattice parameter of a wire along its main axis of an amount 1c, can be written as  2 1 c0 1 (4) U(1c) = k1c2 = k1c2 2 2 c0

(2)

where P indicates the polarization for a strained structure, P◦ is the equilibrium structure polarization, e31 and e33 are the piezoelectric constants, i are the strains along the three lattice directions and ν is the Poisson ratio. In this case, the appropriate quantity to describe the response to strain is the effective piezoelectric constant, eeff 33 = (e33 − 2e31 ν). Multiplying equation (2) by the volume, V, occupied by a bulk ZnO pair, and expressing V as a function of the strained lattice parameters, it is possible to write an analytical formula for the dependence of hLDi on the applied strain, 3 , for the NW cores, hLDi = PV = (P◦ + 3 (e33 − 2e31 ν)) × V0 (1 − 3 ν)2 (1 + 3 ).

= 21 kc20 32 = 12 K32

(3) 4

(5)

Nanotechnology 24 (2013) 475401

K K Korir et al

where k is the elastic constant of the nanostructure, K = kc20 is the effective spring constant, c0 is the NW equilibrium lattice parameter and  is the strain applied along the wire axis. From the above equations we obtain the strain energy per ZnO pair, Us ,

availability of high performance computing resources and support.

U(1c) 1 1 = K32 = U 00 32 (6) n 2n 2 where n is the number of ZnO pairs in the simulation supercell and U 00 = Kn is called the effective strain energy of the system. The latter quantity gives a quantitative estimate of the amount of energy required to deform a finite structure and contains both bulk and surface effects. As reported in table 1, the effective strain energy decreases with decreasing the size of the NWs, and for the smallest wire it is reduced by about 50%. The above results contribute to solving an existing debate on the mechanical properties of ZnO NWs still present in literature (see e.g. [9] and references therein), and support the conclusion that the Young’s modulus of ZnO wires decreases with decreasing size. This piece of information is relevant to open the way for the design and fabrication of efficient energy scavenging devices based on nanostructures. Furthermore, the results show that the energy requirements of these smart devices belong to the range of environmental noise, such as air, vibration, fluid flow and others [23]. In summary, in this work we solve the existing debate about the ‘size effects’ in the piezoelectric response of ZnO NWs with respect to bulk revealing that they are negligible. In contrast, effective strain energies of NWs are more sensitive to size reduction and are much lower compared to bulk. In particular, our theoretical predictions indicate that the advantage of using NWs for energy harvesting is due to their sensitivity to small mechanical agitation, thus making them ideal candidates for building efficient energy scavenging devices irrespective of the fact that NW nanogenerators (with typical size of 40 nm) are expected to have piezoelectric properties similar to bulk ones.

¨ ur U, ¨ Alivov Y, Liu C, Teke A, Reshchikov M, Do˘gan S, [1] Ozg¨ Avrutin V, Cho S and Morkoc H 2005 J. Appl. Phys. 98 041301 [2] Mofor A C et al 2005 Appl. Phys. Lett. 87 062501 [3] Wang X, Song J, Liu J and Wang Z 2007 Science 316 102 [4] Wan Q, Li Q H, Chen Y J, Wang T H, He X L, Li J P and Lin C L 2004 Appl. Phys. Lett. 84 3654–6 [5] Cheng-Yuan W and Adhikari S 2011 Phys. Lett. A 375 2171–5 [6] Zhou J, Lao C S, Gao P, Mai W, Hughes W L, Deng S Z, Xu N S and Wang Z L 2006 Solid State Commun. 139 222–6 [7] Wang Z and Song J 2006 Science 312 242 [8] Alexe M, Senz S, Scubert M A, Hesse D and Gosche U 2008 Adv. Mater. 20 4021 [9] Espinosa H, Bernal R A and Minary-Jolandan M 2012 Adv. Mater. 24 4656 [10] Bernal R A, Agraval R, Peng B, Bertness K A, Sanford N A, Davydov A V and Espinosa H D 2011 Nano Lett. 11 548 [11] Cicero G, Ferretti A and Catellani A 2009 Phys. Rev. B 80 201304(R) [12] Xiang H, Yang J, Hou J and Zhu Q 2006 Appl. Phys. Lett. 89 223111 [13] Agrawal R and Espinosa H 2011 Nano Lett. 11 786–90 [14] Mistrushchenkov A, Linquerri R and Chambaud G 2009 J. Phys. Chem. C 113 6883–6 [15] Giannozzi P et al 2009 J. Phys.: Condens. Matter 21 395502 [16] Ferretti A, Calzolari A, Bonferroni B and Di Felice R 2007 J. Phys.: Condens. Matter 19 036215 [17] Perdew J P, Burke K and Ernzerhof M 1996 Phys. Rev. Lett. 77 3865–8 [18] Resta R and Vanderbilt D 2007 Physics of Ferroelectrics (Berlin: Springer) (Theory of Polarization: A Modern Approach) pp 31–68 [19] Resta R 1994 Rev. Mod. Phys. 66 899–915 [20] Marzari N and Vanderbilt D 1997 Phys. Rev. B 56 12847 [21] Carlotti G, Socino G, Petri A and Verona E 1987 Appl. Phys. Lett. 51 1889 [22] Dai S, Gharbi M, Sharma P and Park H S 2011 J. Appl. Phys. 110 104305 [23] Dong-Hoon C, Chang-Hoon H, Hyun-Don K and Jun-Bo Y 2011 Smart Mater. Struct. 20 125012 [24] Agrawal R, Peng B, Gdoutos E E and Espinosa H D 2008 Nano Lett. 8 3668–74 [25] Fan W, Xu H, Rosa A L, Frauenheim T and Zhang R Q 2007 Phys. Rev. B 76 073302 [26] Kou L, Li C, Zhang Z, Chen C and Guo W 2010 Appl. Phys. Lett. 97 053104

References

Us =

Acknowledgments The research leading to these results has received funding from the European Union Seventh Framework Programme under grant agreement No. 265073 (ITN-Nanowiring). We acknowledge CINECA (ISCRA C project, 2011) for the

5

Piezoelectric properties of zinc oxide nanowires: an ab initio study.

Nanowires made of materials with non-centrosymmetric crystal structures are expected to be ideal building blocks for self-powered nanodevices due to t...
480KB Sizes 0 Downloads 0 Views