8
First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO Rodica Plugaru a , Titus Sandu a , Neculai Plugaru b,a National Institute for R & D in Microtechnologies (IMT), Erou Iancu Nicolae Str. 126A, Bucharest 077190, P.O. BOX 38-160, Romania b National Institute of Materials Physics, Atomistilor Str. 105bis, Magurele-Bucharest 077125, P.O. Box MG-07, Ilfov, Romania article info Article history: Received 30 July 2012 Accepted 12 October 2012 Available online 23 October 2012 Keywords: Computer simulations Electronic properties Electrical transport Impurities in semiconductors Disordered systems abstract Based on first principles electronic structure calculations using the Coherent Potential Approximation (CPA) in the Blackman–Esterling–Berk (BEB) multiscattering formalism and the variable range hopping (VRH) model proposed by Mott, we evaluate the low temperature dc conductivity and its temperature dependence for n-doped wurtzite-type M:ZnO, with M = Al, Ti, Mn, at concentrations of 2, 5 and 10 at.% respectively. We theoretically determine the phenomenologic quantities in the expression of the hopping conductivity, as well as the temperature range in which the VRH model is applicable to the investigated compounds. We show that self-consistent CPA-BEB and LSDA+U calculations yield rea- sonable band gaps, dopant state localization and also spin magnetic moments for the Ti and Mn systems. These results are discussed in comparison with reported data obtained by supercell LSDA+U calculations for similar systems. The results in this study point to 2–5 at.% Ti and approximately 2 at.% Al codoping in wurtzite-type ZnO as an interesting option to obtain a material with an increased low temperature dc conductivity and ferromagnetic background. Ó 2012 Elsevier B.V. All rights reserved. 1. Introduction Over the past decade studies on transparent semiconductor oxi- des have brought considerable evidence leading to wurtzite-type zinc oxide (ZnO) as a promising candidate material for transparent electronics, sensors and solar cells [1–9]. The focus of experimental investigations has been on improving the control of both electrical [10] and spin-dependent conductivities [2], as well as the creation of an intrinsic stable ferromagnetic background above room tem- perature required in spintronic applications [11]. The real material shows a significant level of intrinsic n-doping to which the effects of dopant atoms sum up. Particularly, ZnO thin films and nano- structures show a great sensitivity of the targeted conduction and magnetic properties to dopant species used to tune their prop- erties, a fact which also determines a large scatter in the reported measured data. Thus, an early review of experimental results ob- tained on doped and undoped ZnO thin films deposited by a wealth of methods was presented in Ref. [12], with the aim to establish the physical limits of electrical resistivity in this material. Based on the findings that: (i) the lowest resistivities in ZnO thin films are in the range of 1.4–2 10 4 X cm, independently of the deposition method, (ii) an upper limit of carrier concentration of about 1.5 10 21 cm 3 , corresponding to a dopant solubility limit of about 4 at.%, and (iii) an upper limit of the carrier mobility in the region n >5 10 20 cm 3 of about 40 cm 2 V 1 s 1 , the author sug- gests that resistivities less than 1 10 4 X cm for doped polycrys- talline zinc oxide films are hard to achieve, due to a general limitation by ionized impurity scattering. Since then, a great effort has been undertaken to improve the electrical and optical properties of ZnO films by utilizing a diversity of material synthesis techniques, substrates, doping and processing routes. Controlling electrical conductivity by Al-doping has re- ceived a particular interest as a convenient way of carrier density and mobility changing [2,13], which also has useful effects on opti- cal properties in the blue and UV regions of the spectrum [14,15]. Recently, resistivity values of the order 10 3 –10 4 X cm were re- ported for Al-doped ZnO thin films with Al concentration between 0.25 and 3.5 at.% [16–19]. The investigations have proven that for a given Al concentration the film resistivity values strongly depend on the synthesis method and post-processing, which determine the nanostructure and the defective state. Also, some experiments underscore the fact that increasing Al concentration above approx- imately 4 at.% determines an increase in the resistivity [17,20]. Research work performed on Ti-doped (97% ZnO+3% Al 2 O 3 ), (ZATO), and undoped, (ZAO), films showed that the lowest achievable resistivity is 7.96 and 8.7 10 4 X cm for ZATO and ZAO as-processed films, respectively, whereas after annealing in air, the resistivity of both ZATO and ZAO films is higher than 2211-3797/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.rinp.2012.10.004 Corresponding author. Address: National Institute of Materials Physics, Labo- ratory of Magnetism and Superconductivity, Atomistilor Str. 105bis, Magurele- Bucharest 077125, P.O. Box MG-07, Ilfov, Romania. Tel.: +40 21 369 0170; fax: +40 21 369 0177. E-mail address: plug@infim.ro (N. Plugaru). Results in Physics 2 (2012) 190–197 Contents lists available at SciVerse ScienceDirect Results in Physics journal homepage: www.journals.elsevier.com/results-in-physics

First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

  • Upload
    neculai

  • View
    220

  • Download
    1

Embed Size (px)

Citation preview

Page 1: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

Results in Physics 2 (2012) 190–197

Contents lists available at SciVerse ScienceDirect

Results in Physics

journal homepage: www.journals .e lsevier .com/resul ts - in-physics

First principles study and variable range hopping conductivity in disorderedAl/Ti/Mn-doped ZnO

Rodica Plugaru a, Titus Sandu a, Neculai Plugaru b,⇑a National Institute for R & D in Microtechnologies (IMT), Erou Iancu Nicolae Str. 126A, Bucharest 077190, P.O. BOX 38-160, Romaniab National Institute of Materials Physics, Atomistilor Str. 105bis, Magurele-Bucharest 077125, P.O. Box MG-07, Ilfov, Romania

a r t i c l e i n f o

Article history:Received 30 July 2012Accepted 12 October 2012Available online 23 October 2012

Keywords:Computer simulationsElectronic propertiesElectrical transportImpurities in semiconductorsDisordered systems

2211-3797/$ - see front matter � 2012 Elsevier B.V. Ahttp://dx.doi.org/10.1016/j.rinp.2012.10.004

⇑ Corresponding author. Address: National Institutratory of Magnetism and Superconductivity, AtomiBucharest 077125, P.O. Box MG-07, Ilfov, Romania. Te21 369 0177.

E-mail address: [email protected] (N. Plugaru).

a b s t r a c t

Based on first principles electronic structure calculations using the Coherent Potential Approximation(CPA) in the Blackman–Esterling–Berk (BEB) multiscattering formalism and the variable range hopping(VRH) model proposed by Mott, we evaluate the low temperature dc conductivity and its temperaturedependence for n-doped wurtzite-type M:ZnO, with M = Al, Ti, Mn, at concentrations of 2, 5 and10 at.% respectively. We theoretically determine the phenomenologic quantities in the expression ofthe hopping conductivity, as well as the temperature range in which the VRH model is applicable tothe investigated compounds. We show that self-consistent CPA-BEB and LSDA+U calculations yield rea-sonable band gaps, dopant state localization and also spin magnetic moments for the Ti and Mn systems.These results are discussed in comparison with reported data obtained by supercell LSDA+U calculationsfor similar systems. The results in this study point to 2–5 at.% Ti and approximately 2 at.% Al codoping inwurtzite-type ZnO as an interesting option to obtain a material with an increased low temperature dcconductivity and ferromagnetic background.

� 2012 Elsevier B.V. All rights reserved.

1. Introduction

Over the past decade studies on transparent semiconductor oxi-des have brought considerable evidence leading to wurtzite-typezinc oxide (ZnO) as a promising candidate material for transparentelectronics, sensors and solar cells [1–9]. The focus of experimentalinvestigations has been on improving the control of both electrical[10] and spin-dependent conductivities [2], as well as the creationof an intrinsic stable ferromagnetic background above room tem-perature required in spintronic applications [11]. The real materialshows a significant level of intrinsic n-doping to which the effectsof dopant atoms sum up. Particularly, ZnO thin films and nano-structures show a great sensitivity of the targeted conductionand magnetic properties to dopant species used to tune their prop-erties, a fact which also determines a large scatter in the reportedmeasured data. Thus, an early review of experimental results ob-tained on doped and undoped ZnO thin films deposited by a wealthof methods was presented in Ref. [12], with the aim to establish thephysical limits of electrical resistivity in this material. Based on thefindings that: (i) the lowest resistivities in ZnO thin films are in therange of 1.4–2 � 10�4 X cm, independently of the deposition

ll rights reserved.

e of Materials Physics, Labo-stilor Str. 105bis, Magurele-l.: +40 21 369 0170; fax: +40

method, (ii) an upper limit of carrier concentration of about1.5 � 1021 cm�3, corresponding to a dopant solubility limit ofabout 4 at.%, and (iii) an upper limit of the carrier mobility in theregion n > 5 � 1020 cm�3 of about 40 cm2 V�1 s�1, the author sug-gests that resistivities less than 1 � 10�4 X cm for doped polycrys-talline zinc oxide films are hard to achieve, due to a generallimitation by ionized impurity scattering.

Since then, a great effort has been undertaken to improve theelectrical and optical properties of ZnO films by utilizing a diversityof material synthesis techniques, substrates, doping and processingroutes. Controlling electrical conductivity by Al-doping has re-ceived a particular interest as a convenient way of carrier densityand mobility changing [2,13], which also has useful effects on opti-cal properties in the blue and UV regions of the spectrum [14,15].Recently, resistivity values of the order 10�3–10�4 X cm were re-ported for Al-doped ZnO thin films with Al concentration between0.25 and 3.5 at.% [16–19]. The investigations have proven that for agiven Al concentration the film resistivity values strongly dependon the synthesis method and post-processing, which determinethe nanostructure and the defective state. Also, some experimentsunderscore the fact that increasing Al concentration above approx-imately 4 at.% determines an increase in the resistivity [17,20].

Research work performed on Ti-doped (97% ZnO+3% Al2O3),(ZATO), and undoped, (ZAO), films showed that the lowestachievable resistivity is 7.96 and 8.7 � 10�4 X cm for ZATO andZAO as-processed films, respectively, whereas after annealingin air, the resistivity of both ZATO and ZAO films is higher than

Page 2: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

R. Plugaru et al. / Results in Physics 2 (2012) 190–197 191

105 X cm [21]. The authors attributed this behavior mainly tothe increase of oxygen defect levels and the disorder promotedby impurities.

The properties of Ti-doped ZnO films deposited on Al2O3 (0001)substrates with varying temperature, TS, were also investigated[22]. The Ti and carrier concentrations were found to increase withincreasing TS, whereas the band-gap increased for TS = 100 and200 �C from that of pure ZnO and decreased for TS = 300 and400 �C. These authors rationalize that the observed band-gap vari-ation along with a large increase in carrier concentration betweenTS = 200 and 300 �C are due to a merging of the donor band withthe conduction band minimum, which determines a semiconduc-tor-metal transition. The increase in the donor density was tenta-tively attributed to the increase of the defect density at higher Tidoping levels [22].

In another investigation of ZnO:Ti films, a resistivity value of3.82 � 10�3 X cm was determined and a semiconductor-metaltransition was also reported for Ti concentration of 1.3% [23]. Theenhancement of conductivity and the semiconductor-metal transi-tion are likely attributed to the increase in the free carrier concen-tration, along with band-gap shrinkage effects caused by Ti doping.

Several studies of Al-doped ZnO thin films [24,13,25,26] havesuggested that the conduction mechanism below 220 K may beinterpreted within the variable-range hopping (VRH) model pro-posed by Mott for the carrier transport in disordered systems[27,28]. For the proper carrier concentration range [29] the VRH re-gime was identified by measuring the film resistivity as a functionof temperature and fitting the experimental data to the relationln(rT1/2) / T�1/4 in the low temperature limit. A straight-line fitestablishes the VRH behavior for T < T0, with the fitting parameterT0 giving an indication of the hopping probability: a high value ofT0 corresponds to a low probability for hopping and to a low con-ductivity [27]. It has been suggested that the VRH mechanism isalso effective in pulsed laser deposited films obtained in a nitrogenatmosphere [30]. The observed enhancement of the conductivitywas assigned to both the nitrogen ions and nitrogen induced de-fects in the zinc oxide lattice, that behave as localized hopping cen-ters, as well as carrier suppliers.

A much lower temperature range for the hopping conductionmechanisms was determined in Mn-doped ZnO, in the dilute Mnconcentration range of 0.1, 0.3, and 0.6% Mn [31]. The electricalconduction was studied by analyzing the dc and ac conductivitydata. The abrupt change in the temperature dependence of conduc-tivity at approximately 18 K and a much lower activation energyfor conduction below 18 K were attributed to a transition fromband, above 18 K, to hopping conduction between 18 to 10 K[32]. Since Mn contributes with deep donor states at about2.0 eV below the conduction band bottom at room temperature,an increase in Mn concentration determines a decrease of the shal-low donor concentration and therefore of the conductivity. The cal-culated critical temperature for the transition from a nearestneighbor hopping (NNH) mechanism to the VRH regime was eval-uated below approx. 2 K [31].

It is worth to note that under certain conditions the experimen-tal investigations may also be hindered by the formation of an elec-tron accumulation layer near the surface which can alter themeasured conductivity values in the underlying bulk or film (see[2] and references cited therein).

A systematic, clear cut study of the relationship between con-ductivity and magnetism in Mn, Co, Al -doped ZnO films with awide range of carrier densities was presented in [33]. Theseauthors identify three conductivity regimes which have a criticalrole on the magnetic behavior also, as follows: (1) an insulatingphase, at low carrier densities, in which the low temperature con-ductivity arises from a variable range hopping process. In this re-gime, the least conducting films are the most magnetic at room

temperature due to the interaction of the localized spins with sta-tic localized states; (2) an intermediate carrier concentration re-gime, where the samples satisfy neither the conditions for VRHat low temperatures nor for metallic behavior at room tempera-ture. In this case, magnetism is quenched, due to carriers in thelocalized states becoming mobile; (3) a metallic phase at high car-rier densities, in which free carrier-mediated exchange determinesferromagnetism, as predicted by theory [34].

In another investigation of the electrical transport and magne-toresistance in Co heavily doped TiO2 and ZnO [35] it has beenhighlighted that, due to significant Coulomb interaction betweenlocalized carriers at high doping levels, both the spin–spin ex-change between the carrier spins and the Coulomb interactionshould be taken into account in the same frame. A linear relationof lnq versus T�1/2 in the low temperature range was observedexperimentally, in agreement with the theoretical model in[36,37] and a quantitative explanation in a spin-dependent vari-able range hopping model was proposed in [35].

Several first principles studies have approached the physical as-pects involved in the description of the electronic conductionmechanisms in defective/doped ZnO. Extensive analyses of the de-tails of the electronic structure in defective and/or doped ZnO haveattempted to find justifications for the intrinsic n-type conductiv-ity, the role of defects and dopants, as well as the difficulty inachieving the p-type conductivity (see, e.g., [2,10,38,9]). Althoughthere is a conceptual agreement of these studies as regards the roleof defects on the electrical and optical properties, no quantitativeevaluation of the electrical conductivity is made based on calcula-tion results, as a function of doping or temperature. Thus, one maynot draw specific conclusions of practical relevance for the controlof conductivity in the real material unless establishing, by experi-ment and theory, an unambiguos relationship between the sources– dopant and/or defect – and the electronic effect they produce.

In this contribution we present electronic structure results ob-tained within the Coherent Potential Approximation (CPA) andBlackman–Esterling–Berk (BEB) multiscattering formalism imple-mented in the FPLO5 code, in wurtzite-type Zn1�xMxO crystal, withM=Al, Ti and Mn. We calculate the low temperature dc conductiv-ity in the regime where electron hopping on localized impurity lev-els is the dominating mechanism described in Mott’s VariableRange Hopping (VRH) model [32]. We theoretically determinethe phenomenologic quantities in the expression of the hoppingconductivity, as well as the temperature range in which the VRHmodel is applicable to the investigated compounds. A discussionof the magnetism promoted by Ti and Mn, and a prediction ofthe best material for a dilute magnetic semiconductor is also in-cluded. The paper is organized as follows. In Section 2 we presentthe details of our calculations. Then, an account on the electronicstructure results obtained in this work for the Al, Ti and Mn-dopedZnO systems is given in Section 3. We also contrast our present re-sults obtained using the CPA-BEB-LSDA+U formalism with re-ported calculated data on similar systems. In Section 4 we usethese results to discuss the dc conductivity in the underlying sys-tems analyzed in the variable range hopping model. We establishthe dopant concentration and temperature ranges that are ade-quate for the applicability of this model in Zn1�xMxO systems. Sec-tion 5 summarizes the conclusions of this work and a prediction ona suitable semiconductor combining a high dc conductivity andmagnetism is made.

2. Calculation method

We carried out self consistent calculations in the Local SpinDensity Approximation (LSDA) using the FPLO5.00-20 band struc-ture code [39,40]. The exchange and correlation potential was trea-

Page 3: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

Fig. 1. CPA-BEB-LDA results for the total DOS of Zn1�xAlxO, with x = 0, 2, 5 and 10%.Inset: Enlarged image of the conduction band for x = 0% and x = 5%.

192 R. Plugaru et al. / Results in Physics 2 (2012) 190–197

ted in the parameterization of Perdew and Wang [41]. The integra-tion in the Brillouin zone was performed at 480 irreducible k-points from 4096 (using a 16 � 16 � 16 k-mesh); this ensured bothan accuracy of 10�5 eV in the total energy and a proper treatmentof the Fermi surface effects if an insulator-to-metal transitionoccured as a result of electron doping.

The partial substitutional disorder brought about by the ran-dom dopant distribution at Zn sites in Zn1�xMxO, M=Al, Ti or Mn,was treated in the Coherent Potential Approximation (CPA) [42],in which the disordered crystal is treated as an effective medium[43]. In FPLO the CPA uses the multiscattering formalism of Black-man–Esterling–Berk (BEB) [44] implemented in a non-orthogonallocal orbitals scheme [42]. Some properties of the CPA-BEB imple-mentation in FPLO are, [42,45]: (i) The scheme is all-electron, themultiple scattering formalism is consistent with a single site CPAcondition and fully charge self-consistent. The scalar relativisticmodel of [46] is used; (ii) Both the diagonal and the off-diagonaldisorder are included; (iii) The atomic centers are distributed onperiodic lattice points where the atomic potential is site-depen-dent; still, a partial long-range order is assumed on sublattices,(iv) One may introduce several components and several sublatticeswith different compositions. (v) The symmetry as in the orderedcase is preserved, but instead of having one atom per Wyckoff po-sition one may introduce several atoms with certain site-concen-trations for every Wyckoff position. The advantage of using theCPA over the supercell method materializes in saving computa-tional effort, at low and also at high dopant concentrations CPA,by its nature, produces the electronic structure of a configuratio-naly averaged crystal, whereas in the supercell approach oneshould firstly consider all possible inequivalent dopant distribu-tions and then average the relevant physical properties.

The site-centered potentials and densities were expanded inspherical harmonic contributions up to lmax = 12. The basis setsused in the present calculations are: (i) 3s3p:4s4p3d;5s5p4d, forZn/Ti/Mn, (ii) 1s:2s2p;3s3p3d for O, and (iii) 2s2p:3s3p;4s4p3dfor Al, corresponding to semicore: valence; polarization states,respectively. The basis optimization was achieved for each dopantspecies and concentration, independently. We also performed aCPA-like ordered calculation for ZnO to allow a comparison ofthe results obtained for disordered systems with those in the refer-ence compound. Switching on the spin polarization in the CPA-LSDA calculations showed no relevance with respect to the LDAcase for Al doping. However, a magnetic ground state was indi-cated for the Ti and Mn containing systems. Therefore, for thesematerials, starting from the converged CPA calculations, we alsointroduced a Hubbard-type Coulomb repulsive term, U, and per-formed self-consistent CPA-LSDA+U calculations, using the AtomicLimit functional for the Ti and Mn correlated 3d orbitals [47]. Thevalues of the Slater integrals, Fk, and the Hund’s rule exchangeparameter, J, were established following the arguments elaboratedin [48–51]. Accordingly, we have taken for the screened Coulombparameter, U(=F0)=6 eV and J = 0.76 and 0.86 eV for Ti and Mn,respectively. As such, one obtains for the Slater parameters F2

and F4 the values 6.6 and 4.1 eV for Ti and 7.4 and 4.6 eV for Mn,respectively. Prior to performing all calculations with U=6 eV, wealso tested several values in the range 4–8 eV. This parameter ismore affected by screening effects and usually, in the semi-empir-ical LSDA+U method, one sets its value in an optimum range wherethe predicted band gap is close to the experimental one and alsothe least sensitivity of the results on U values is determined[51,52].

Although the present LDA results on Zn1�xAlxO, particularly theband gaps, are affected by the self-interaction error [53,48] we stillinclude them in our discussion of the dc conductivity, both for thesake of comparison with the Ti and Mn effects in this work, as wellas for their relevance when refering to reported data. Also, the CPA-

LSDA+U results obtained in this work will be analyzed in compar-ison with available data obtained in the LSDA+U and supercell ap-proach for similar systems.

All calculations were carried out on bulk Zn1�xMxO with spacegroup P 63 M C (no. 186) structure, with an impurity concentrationx = 2, 5 and 10 at.% and at the experimental lattice constants [54].Previous experimental investigations of Mn doped wurtzite-typeZnO in a wide concentration range showed variations of only0.32% and 0.44% for the a and c lattice constants, for Mn concentra-tion up to 10 at.% [55]. However, in the case of Al and Ti one can notderive such a straight conclusion, due to largely scattered reporteddata. In this work, due to our neglect of the lattice relaxation, thediscussion of the dopant effects on the dc conductivity refers tothe electronic contribution, only. In the density of states (DOS)plots throughout this work the DOS are given in states/eV/f.u.and the Fermi energy is set at zero energy.

3. Results

3.1. Zn1�xAlxO

The CPA-BEB-LDA results showing the overall effect of electrondoping on the total density of states (DOS) when Al gradually sub-stitutes for Zn are plotted in Fig. 1. The conduction band crosses theFermi level (EF) as x increases and a metallic state is already real-ized for x = 2% Al, in agreement with experimental results[56,57]. Increasing the Al concentration shifts the valence band(VB) to more negative energy values and has only a slight effecton the VB features, which are determined mainly by the Zn 3dand O 2p states. Nevertheless, one may observe that Al has a strongeffect on the conduction band (CB) details. In the pure compound,the CB is split by the crystal field in two regions separated by a gapbetween 6.5 and 7.0 eV. Increasing x determines an insulator-to-metal transition to take place. As the metallic state becomes dom-inant the increase in screening determines a decrease in crystalelectric field and therefore in the CB splitting, see inset in Fig. 1for a comparison between x = 0 and x = 5%. Al electronic configura-tion is 3p1 in ZnO and is equivalent to one electron doping persubstituted Zn atom. The Al 3s and 4s states form a narrow bandsituated at the valence band bottom; Al 3s also has the main con-tribution to DOS at the Fermi energy, whereas Al 3p to DOS in theconduction band, see Fig. 2 for x = 10%.

3.2. Zn1�xTixO

The CPA-LSDA density of states of Zn1�xTixO show some com-mon features for all Ti concentrations, see, e.g. the graphic forZn0.98Ti0.02O in Fig. 3. Thus, both Ti 3d subbands lie just at the con-duction band minimum, being merged with the CB. The partiallyfilled Ti eg majority spin band is located at the Fermi energy. At2% Ti doping the material is already metallic and the band gap be-

Page 4: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

Fig. 2. Al angular momentum-projected DOS for Zn0.95Al0.05O.

Fig. 3. CPA-BEB-LSDA total DOS and Ti 3d states-projected DOS for Zn0.98Ti0.02O.Note: Ti 3d states are scaled.

Fig. 4. CPA-BEB-LSDA+U total DOS and Ti 3d states-projected DOS of Zn1�xTixO,with x = 0.02 (a), x = 0.05 (b) and x = 0.10 (c). Note: Ti 3d states are scaled.

R. Plugaru et al. / Results in Physics 2 (2012) 190–197 193

comes a gap in the valence band, of about 2.47 eV. A spin magneticmoment appears localized on the partially occupied Ti 3d states, asresulted from the Mulliken population analysis. Table 1 summa-rizes the values (at 0 K) of the local spin magnetic moment andthe magnetic moment per formula unit for the Ti concentrationsstudied in this work.

Since the CPA-LSDA results are affected by the insufficienciesspecific to the LDA method [58,48] we also carried out extensiveself consistent calculations using the CPA-LSDA+U approach to ac-count for the correlations within the Ti 3d shell. The resulting plotsfor the density of states are presented in Fig. 4. In the case x = 2% Ti(see Fig. 4a), the Fermi level is situated in the band gap (of about2.98 eV) at 0.45 eV below the conduction band minimum. The Tieg majority spin states are localized in the band gap just belowthe Fermi level, with the tail crossing EF. The DOS plot for 5% Ti,(see Fig. 4b) shows a band gap of 2.63 eV since the Ti eg majorityspin states already merge with the conduction band minimum.The band gap becomes a gap in the valence band, actually, and thissituation is even more evident in the case of 10% Ti (see Fig. 4c).

The CPA-LSDA+U spin magnetic moment on Ti 3d states takeshigher values than those in CPA-LSDA calculations, see Table 1,which indicate a Ti ion with S = 1.

3.3. Zn1�xMnxO

The DOS plot of Zn0.98Mn0.02O shown in Fig. 5 stands as anexample for the CPA-LSDA density of states of the Zn1�xMnxOseries.

Table 1Spin magnetic moment, lM, and the magnetic moment per formula unit, M, inZn1�xMxO, M = Ti and Mn. The LSDA values are given in parentheses.

x(%) 2 5 10

lM(lB/at.)Ti 2.01 (0.90) 1.82 (1.13) 2.01 (1.28)Mn 4.97 (4.71) 4.97(4.71) 4.97(4.71)M(lB/f.u.)Ti 0.04 (0.02) 0.09 (0.06) 0.20 (0.13)Mn 0.10 (0.10) 0.25 (0.25) 0.50 (0.50)

The main features are related to the localization of the Mnmajority spin subband below the Fermi energy in the band gap,well separated from the minority subband which overlaps withthe conduction band minimum. The Mn 3d subbands broadenbut do not change position as Mn concentration increases. Forx = 5 and 10% the Mn majority spin eg shows a tail which extendsto the VB maximum and the Mn t2g merges with the CB minimum,so that a small density of states appears at the Fermi energy. TheMn majority spin subband also overlaps the oxygen 2p band andthat may provide paths for the Mn–Mn exchange coupling. Themanganese ion is in the high spin (S = 5/2) Mn2+ state. The valuesof the spin magnetic moment obtained in CPA-LSDA calculationsand the magnetic moment per formula unit are listed (betweenparentheses) in Table 1, also. The Mn magnetic moment takesthe value 4.71 lB, somewhat reduced relatively to the theoreticalone, 5 lB, for all Mn concentrations.

Switching on the Hubbard type repulsive term, U, produces theexpected effect on the Mn 3d subbands [51]. The Mn majority spinsubband is pushed to more negative energy, and as a result the Mneg majority spin strongly overlaps the valence band maximum, theMn t2g majority spin states merge with the VBM, whereas the Mnminority subband overlaps the conduction band at higher energy,see Fig. 6. The Mn magnetic moment increases to 4.97 lB/Mnion, in good agreement with the theoretical value at saturation.

To our knowledge, the use of the CPA-BEB module and DFT+UAtomic Limit functional has not been reported in any other publi-cation, hitherto. Therefore, in order to check the validity of thepresent approach and build confidence in its results, it is worthto compare the Mn/Ti:ZnO results in our work with other resultsproduced by different codes or experiment.

Thus, an inspection of the DOS plots in Figs. 5 and 6, this work,and the data in: (a) Fig. 3 in [51], obtained by supercell calculationsusing the FPLO code and LSDA/LSDA+U, (b) Figs. 5 and 6 in [59]using the PAW formalism as implemented in the VASP code, (c)Fig.1 in [60] obtained using CPA-KKR-LSDA/GGA, (d) Fig. 1 in

Fig. 5. CPA-BEB-LSDA total DOS and Mn 3d states-projected DOS of Zn0.98Mn0.02O.Note: Mn 3d states are scaled.

Page 5: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

Fig. 6. CPA-BEB-LSDA+U total DOS and Mn 3d states-projected DOS of Zn1�xMnxO,with x = 0.02 (a), x = 0.05 (b) and x = 0.10 (c). Note: Mn 3d states are scaled.

Table 2Spin magnetic moment, lM in Zn1�xMxO, M=Ti and Mn from selected references. Thevalues calculated within DFT schemes without Hubbard U are given in parenthesis.

M x(at.%) lM(lB/at.) References

Ti 2.78 0.81 (0.70) [62]2.75, 5.55 (0.0) [64]6.25 (0.5) [63]5.0 0.5⁄ [66]0.6 0.83 ± 0.01⁄ [65]

Mn 25 4.89 (4.62) [51]12.5 4.41 (4.31) [59]6.25 (4.9) [61]5.0 (4.0) [60]

⁄ Experimental value.

194 R. Plugaru et al. / Results in Physics 2 (2012) 190–197

[61], LSDA supercell calculations using SIESTA code and Ceperley-Alder exchange-correlation functional, shows a remarkable agree-ment between the details of the electronic structures in Mn dopedwurtzite type ZnO. We also note a fair agreement between the localMn magnetic moments listed in Table 1, and particularly in [51,61],see Table 2.

In the case of Ti doping, one finds rather spread results in theliterature. The DOS plot for x = 5% in Fig. 4b, present work, fairlycompares with the plot in Fig. 4 in [62], for x = 2.78%, obtained witha corrected-band-gap scheme (CBGS) and the DFT+U (CBGS+U).However, the Ti magnetic moments calculated in [62], see Table 2,are smaller than the present values. A previous study by supercell

Table 3Density of states at the Fermi level N(EF), the hopping distance R, hopping energy W, the n�

Ti/Mn:ZnO. The value of n is 12.1143 � 10�8 cm.

Dopant (at.%) T (K) N(EF) (eV�1cm�3) R

Al 2 1 2.769 � 1021 6100 2

Al 5 1 3.974 � 1021 5100 1

Al 10 1 5.158 � 1021 5100 1

Ti 2 1 3.805 � 1021 6100 1

Ti 5 1 9.512 � 1021 4100 1

Ti 10 1 6.553 � 1021 5100 1

Mn 2 1 2.113 � 1021 6100 2

Mn 5 1 2.325 � 1021 6100 2

Mn 10 1 2.748 � 1021 6100 2

calculations using the WIEN2k code, GGA and Perdew–Burke–Ern-zerhof for the exchange-correlation functional, gives 0.50 lB/Ti forTi0.0625Zn0.9375O [63]. At variance with these results, a zero mag-netic moment has been predicted in [64] using the VASP code,GGA and Perdew and Wang parameterization, see Table 2. Onthe experimental side, Ti-doped ZnO films deposited on Si(100)substrate were investigated in [65]. The largest saturation mag-netic moment, 0.83 ± 0.01 lB/Ti was observed for a ferromagneticinsulator Zn0.994Ti0.006O film, at room temperature, Also, a Ti mag-netic moment of 0.5 lB/Ti ion was determined in (110)-orientedZnO films at room temperature [66].

4. Discussion

First, we make an assessment of the low temperature dc con-ductivity and its temperature dependence in Al/Ti/Mn-doped ZnOsystems with VRH model [27,28]. According to this model, at lowtemperature the electrical transport consists of phonon-assistedhopping of charge carriers from occupied to unoccupied localizedstates. The theoretically predicted nearest-neighbor hopping con-ductivity, r, is given by the formula:

r ¼ r0exp � T0

T

� �14

!; ð1Þ

where r0 is the pre-exponential factor and T0 is a parameter relatedto the disorder in the system. These quantities are expressed as:

r0 ¼3q2mph

ð8pkBÞ12

!NðEFÞaT

� �12

; ð2Þ

T0 ¼ka3

kBNðEFÞ; ð3Þ

where N(EF) is the localized state density at the Fermi level, a is aparameter such that n ¼ 1

a is the localization length of the wavefunction for localized states, kB is the Boltzmann constant,mph = 8.67 � 1012 s�1 is the phonon frequency at the Debye temper-ature (hD = 416 K for ZnO) [67] associated with the hopping process,and k = 16 is related to the hopping probability [68,25]. In the fol-lowing we approximate the localization length by the effective Bohrradius (aB) of a Wannier exciton [68], applicable to semiconductorswith high static dielectric constant and exciton energy of the orderof tens of meV, expressed as:

1R parameter calculated at T = 1 K and T = 100 K and the critical temperature, Tc for Al/

(cm � 10�7) W (meV) n�1R Tc (K)

.52982 0.309 5.39 166.7

.06491 9.792 1.70

.96588 0.282 4.92 116.2

.88658 8.946 1.55

.58935 0.265 4.61 89.5

.76751 8.382 1.45

.03105 0.286 4.97 121.3

.90719 9.044 1.57

.79638 0.227 3.95 48.5

.51675 7.193 1.25

.26468 0.249 4.34 70.5

.66484 7.895 1.37

.98646 0.331 5.76 218.5

.20931 10.477 1.82

.82144 0.323 5.63 198.6

.15713 10.230 1.78

.54226 0.310 5.40 168.0

.06884 9.811 1.70

Page 6: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

(a)

(b)

(c)

Fig. 7. Calculated ln(r � T1/2) versus T�1/4 dependence for Al (a), Ti (b) and Mn (c)doped ZnO, between 1 and 100 K.

(a)

(b)

(c)

Fig. 8. Calculated temperature dependence of hopping activation energy for Al (a),Ti (b) and Mn (c) in Zn1�xMxO. The kBT versus T is also plotted to allow adetermination of the temperature range for Mott’s VRH model validity, according tothe constraint W P kBT.

R. Plugaru et al. / Results in Physics 2 (2012) 190–197 195

n ¼ 1a¼ aB ¼

4per�h2

m�ee2 ; ð4Þ

where er = 8.15 is the relative dielectric constant of wurtzite-typeZnO and m�e ¼ 0:38 me is the effective electron mass [67]. Thus, avalue aB = 12.1143 � 10�8 cm is calculated using Eq. (4). Being ameasure of the electron wavelength localization at an impurity site,a may extend over a few near neighbors. The hopping distance, R, isgiven by

R ¼ 98pakBNðEFÞ

� �14

ð5Þ

and the hopping energy, W, is given by

W ¼ 43pR3NðEFÞ

: ð6Þ

We calculated the hopping distance R, the hopping energy W, thedimensionless parameter n�1R (or aR), and critical temperature,Tc, below which the Mott mechanism of conductivity is valid, theirvalues being listed in Table 3. Also, the calculated graphs ln(rT1/2)versus T�1/4 are displayed in Fig. 7a–c, which shows the variationof the conductivity over the temperature range 1–100 K. In orderto analyze these results, one should recall that the behavior indi-cated by the VRH model is assured if two conditions are fulfilled.According to Mott, the values of W should be larger than kBT andthose of aR should be greater than unity. The parameter aR, givenin Table 3, complies with the condition aR > 1, in the temperaturerange 1–100 K. Next, the calculated temperature dependence ofthe hopping activation energy in Al, Ti, Mn doped ZnO is plottedin Fig. 8a–c. The kBT versus T is also plotted to allow a determinationof the temperature range for the Mott’s VRH model validity. Tc isdetermined by the intersection of the W(T) and kB(T) curves.

Page 7: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

196 R. Plugaru et al. / Results in Physics 2 (2012) 190–197

Thus, one may observe that Al based alloys have a Tc from 166 Kat 2% concentration to 89 K at 10%. Ti based alloys have the small-est Tc, with the smallest value 48.5 K at 5% Ti and the largest one,121 K, at 2% Ti. Finally, the Mn based alloys have the largest Tc, be-tween 218.5 K at 2% Mn and 168 K at 10% Mn.

It is not a straightforward task to discuss the results in Table 3with reference to experiment, see, e.g. [13,26,16] for Al doped ZnO,due to the large scatter of the experimentally employed conditionsand reported quantities. Nevertheless, we tentatively compare ourresults for 2% Al doped ZnO, at 100 K, with the parameters deducedfrom experiment for Al doped ZnO films at 150 K, see Table 1 in[25]. First, one may observe a significant dependence of the hopingdistance, activation energy and aR parameter on the annealingtemperature, for the same nominal stoichiometry [25]. This maybe related to effects associated with modifications at the grainboundaries, namely defects which play a role either as hoppingcenters, or as periodic barriers [30]. Taking into account the dataspreading for x = 1 and x = 2 in Ref. [25], one may only concludethat there is a qualitative agreement between the values of thehopping range, activation energy and aR parameter derived in[25] and the calculated ones in this work.

Then, using Eq. (1) and the parameters listed in Table 3, we cal-culated the dc conductivity at several temperatures for each spe-cies, below Tc. The 77 K was chosen for its relevance to resistivitymeasurements 1 K and 48 K to allow a comparison between theimpurity effects on conductivity at the same temperature.

These results are presented in Fig. 9. The plots of r versus thedopant concentration x, at all selected temperatures, are qualita-tively in agreement with the experimental observations for Al[13,26,16,25], for Ti [69,23,70,22] as well as for Mn doping [71,72].

Thus, the VRH model shows that Al doping up to 10% deter-mines a gradual increase in the dc conductivity, the effect beingstronger as temperature increases, see Fig. 9. Since the LDA DOSplot in Fig. 1 suggests that Zn0.98Al0.02O is already metallic, oneshould only consider Al concentrations less than about 2% to obtaina semiconductor with higher conductivity than that of ZnO wurtz-ite phase.

Our calculations suggest a strong effect of Ti on conductivity inZn1�xTixO with an anomalously high value at about 5%, see Fig. 9.Tentatively, we attribute this behavior to an overlap of the impu-rity band tail with the conduction band minimum, see Fig. 4b,which determines an increased density of states at the Fermi en-ergy as well as an increase in the carrier concentration. Further,we attribute the decrease in conductivity for Ti concentration fromabout 5 to 10%, to the decrease of Ti localized states density at theFermi energy, as the impurity band shifts to more negative energy,

Fig. 9. Calculated dc conductivity at 1 K versus dopant concentration x for bulkZn1�xMxO, M=Al, Ti and Mn.

see Fig. 4c. Increasing x from 5 to 10% also leads to a decrease in theband gap, see Fig. 4b and c, and this is in qualitative agreementwith the observation of a red-shift in the band gap at 1.3% Ti in[23].

According to our results in Fig. 9, Mn doping determines only aslight increase in the dc conductivity in Zn1�xMnxxO, for x 6 10%.Very recently, a small increase in conductivity of Mn:ZnO filmshas been reported for Mn concentration up to 5% in Mn dopedZnO films deposited by atmospheric barrier torch discharge tech-nique on SiO2 substrates [72]. In another recent investigation, asuppression of the conductivity in oriented and doped up to 4%Mn:ZnO films was reported [71]. Variable-temperature Hall effectmeasurements indicated that the conductivity actually decreaseswith increasing Mn content. The authors suggest that the incorpo-ration of Mn atoms suppresses the formation of native defects suchas oxygen vacancies. Also, it was observed that Mn ions cause ablue-shift of photoluminescence PL energy [71], determined as ashift in the PL peak position to higher energy when Mn doping in-creases. This shift was presumably attributed to an increase in theband gap, which is not confirmed by our results in Fig. 6a–c. In fact,a decrease in the band gap may be observed when Mn concentra-tion increases, similarly to Ti doping.

It is interesting to note that the present results indicate Ti dop-ing as having a stronger effect than Al on conductivity at very lowtemperatures, e.g. 1 K, see Inset in Fig. 9, whereas at higher tem-peratures, e.g. 48 K, Al has the dominant effect, see Fig. 9.

Also, it is worth to comment on the differences between the cal-culated r values in the present work and the experimental ones re-ported in references. We performed the calculations on disordered,stoichiometric systems. However, the conductivity of the realmaterial is strongly determined by processing, see for examplethe scatter in the resistivity values in Table 2 in [73] for Al/Co:ZnOhigh quality epitaxial thin films. Not in the least, a significant con-tribution to conductivity may originate from defects, particularlyoxygen vacancies, as intrinsic defects, process-induced defects orsubsequent charge compensation mechanisms, [10,2,74,75].

5. Conclusions

CPA-BEB-LSDA(+U) calculations of the electronic structure inwurtzite-type Zn1�xMxO, M=Al, Ti, Mn and x = 0, 2, 5 and 10%, wereperformed using the FPLO5 code. To our knowledge, this is the firstreport on doped ZnO based on this approximation. The density ofstates plots presented herein are in fair agreement with those re-ported on similar systems obtained by using a variety of othertechniques, which sets confidence in our approach. On the founda-tion laid by the Mott’s VRH model, we calculated the low temper-ature dc conductivity and its temperature dependence, targeting tohighlight the electronic effects produced by the doping ions. Thephenomenologic parameters of the model are also calculated andthe temperature range of applicability of the theory is establishedfor each system. The hopping conductivity shows a significant in-crease with Al concentration at high temperature, in accordancewith experiment, and only a moderate increase at very low tem-perature (e.g. 1 K). In the case of Ti doping, r versus concentrationshows a maximum at about 5% which we relate to the localizationof the impurity band at the Fermi energy and its proximity to theconduction band. An almost constant dc conductivity is predictedfor Mn doping up to 10%. Our calculations also predict magneticground states for Ti and Mn doping for x 6 10%. From a practicalpoint of view, our results suggest that about 2% Al doping and 5%Ti co-doping in hexagonal ZnO may promote a material combininghigh conductivity and eventually stable magnetism, in a ratherwide temperature range.

Page 8: First principles study and variable range hopping conductivity in disordered Al/Ti/Mn-doped ZnO

R. Plugaru et al. / Results in Physics 2 (2012) 190–197 197

Acknowledgments

The work of one of us (T.S.) has been supported by the SectorialOperational Programme Human Resources Development, financedfrom the European Social Fund and by the Romanian Governmentunder the contract number POSDRU/89/1.5/S/63700. N. Plugaruacknowledges the financial support received in the frame of theCore Program Contract PN09-45, INCDFM, and the technical sup-port from IFIN-HH and the RomanianGRID CA.

References

[1] Djurišic A, Ng AC, Chen X. Quant Electron 2010;34:191–259.[2] Janotti A, Van de Walle CG. Rep Progr Phys 2009;72:126501.[3] Jagadish C, Pearton S, editors. Zinc oxide bulk, thin films and

nanostructures. Oxford: Elsevier; 2006.[4] Volintiru I, Creatore M, Sanden MCMV. J Appl Phys 2008;103:033704.[5] Wang X, Lei Q, Xu W, Zhou W, Yu J. Mater Lett 2009;63:1371–3.[6] Kirby SD, van Dover R. Thin Solid Films 2009;517:1958–60.[7] Kwak D-J, Park MW, Sung Y-M. Vacuum 2009;83:113–8.[8] Epurescu G, Birjega R, Galca A. Appl Phys A 2011;104:889–93.[9] Litton CW, Reynolds DC, Collins TC, editors. Zinc oxide materials for electronic

and optoelectronic device applications. Chichester West Sussex, UK: JohnWiley & Sons; 2011.

[10] Van de Walle CG. J Phys Condens Matter 2008;20:064230.[11] Zunger A, Lany S, Raebiger H. Physics 2010;3:53.[12] Ellmer K. J Phys D Appl Phys 2001;34:3097–108.[13] Sagar P, Kumar M, Mehra R. Solid State Commun 2008;147:465–9.[14] Ding J, Ma S, Chen H, Shi X, Zhou T, Mao L. Physica B 2009;404:2439–43.[15] Elilarassi R, Chandrasekaran G. Mater Chem Phys 2010;121:378–84.[16] Lin K, Chen Y, Chou K. J Sol-Gel Sci Technol 2009;49:238–42.[17] Lu J, Tsai S, Lu Y, Lin T, Gan K. Solid State Commun 2009;149:2177–80.[18] Qiang X, Hong D, Yan L, Hang C. Mat Sci Semicon Process 2006;9:132–5.[19] Zhu M, Gong J, Sun C, Xia J, Jiang X. J Appl Phys 2008;104:073113.[20] Lupan O, Chow L, Shishiyanu S, Monaico E, Shishiyanu T, Sontea V, Cuenya BR,

Naitabdi A, Park S, Schulte A. Mat Res Bull 2009;44:63–9.[21] Jiang M, Liu X. Appl Surf Sci 2008;255:3175–8.[22] Park YR, Kim KJ. Solid State Commun 2002;123:147–50.[23] Lu J, Lu Y, Tsai S, Hsiung T, Wang H, Jang L. Opt Mater 2007;29:1548–52.[24] Brilis N, Tsamakis D, Ali H, Krishnamoorthy S, Iliadis A. Thin Solid Films

2008;516:4226–31.[25] Bandyopadhyay S, Paul G, Roy R, Sen S, Sen S. Mater Chem Phys

2002;74:83–91.[26] Liu H, Qiu H, Chen X, Yu M, Wang M. Current Appl Phys 2009;9:1217–22.[27] Mott N. J Non-Crystalline Solids 1968;1:1–17.[28] Mott N. Phil Mag 1969;19:835–52.[29] Natsume Y, Sakata H, Hirayama T. Phys Stat Sol 1995;148:485–95.[30] Heluani S, Braunstein G, Villafuerte M, Simonelli G, Duhalde S. Thin Solid Films

2006;515:2379–86.[31] Han MSJP, Cao W, Senos A, Mantas P. Appl Phys Lett 2003;82:67–9.[32] Mott N, Davis E. Electronic processes in non-crystalline

materials. London: Oxford University Press; 1979.[33] Behan AJ, Mokhtari A, Blythe HJ, Score D, Xu X-H, Neal JR, Fox AM, Gehring GA.

Phys Rev Lett 2008;100:047206.

[34] Chattopadhyay A, Sarma D, Millis A. Phys Rev Lett 2001;87:227202.[35] Yan SS, Liu JP, Mei LM, Tian YF, Song HQ, Chen YX, Liu GL. J Phys Condens

Matter 2006;18:10469–80.[36] Efros A, Shklovskii B. J Phys C Solid State Phys 1975;8:L49–51.[37] Shklovskii B, Efros A. Electronic properties of doped

semiconductors. Berlin: Springer; 1984.[38] Iusan D, Knut R, Sanyal B, Karis O, Eriksson O, Coleman V, Westin G, Magnus J,

Svedlindh P. Phys Rev B 2008;78:085319.[39] Koepernik K, Eschrig H. Phys Rev B 1999;59:1743–57.[40] Opahle I, Koepernik K, Eschrig H. Phys Rev B 1999;60:14035–41.[41] Perdew JP, Wang Y. Phys Rev B 1992;45:13244–9.[42] Koepernik K, Velicky B, Hayn R, Eschrig H. Phys Rev B 1997;55:5717–29.[43] Soven P. Phys Rev 1967;156:809.[44] Blackman JA, Esterling DM, Berk NF. Phys Rev B 1971;4:2412.[45] Eschrig H. Optimized LCAO method and the electronic structure of extended

systems. Berlin: Springer; 1989.[46] Richter M, Eschrig H. Solid State Commun 1989;72:263.[47] Eschrig H, Koepernik K, Chaplygin I. J Solid State Chem 2003;176:482–95.[48] Anisimov V, Zaanen J, Andersen O. Phys Rev B 1991;44:943–54.[49] Solovyev I, Dederichs P, Anisimov V. Phys Rev B 1994;50:16861–71.[50] Nemoshkalenko V, Antonov V. Computational methods in solid state

physics. C.R.C. Press; 1998.[51] Chanier T, Sargolzaei M, Opahle I, Hayn R, Koepernik K. Phys Rev B

2006;73:134418.[52] Zhang W, Koepernik K, Richter M, Eschrig H. Phys Rev B 2009;79:155123.[53] Gunnarsson O, Schonhammer K. Phys Rev Lett 1986;56:1968–71.[54] Sabine TM, Hogg S. Acta Cryst B 1969;25:2254–6.[55] Fukumura T, Jin Zhengwu, Ohtomo A, Koinuma H, Kawasaki M. Appl Phys Lett

1999;75:3366–8.[56] Han J, Mantas P, Senos A. J Eur Ceramic Soc 2001;21:1883–6.[57] Suchea M, Christoulakis S, Katsarakis N, Kitsopoulos T, Kiriakidis G. Thin Solid

Films 2007;515:6562–6.[58] Perdew JP, Parr RG, Levy M, Balduz Jr JL. Phys Rev Lett 1982;49:1691–4.[59] Gopal P, Spaldin N. Phys Rev B 2006;74:094418.[60] Iusan D, Sanyal B, Eriksson O. Phys Rev B 2006;74:235208.[61] Spaldin N. Phys Rev B 2004;69:125201.[62] Shao B, Liu H, Wu J, Zuo X. J Appl Phys 2012;111:07C301.[63] Osuch K, Lombardi E, Gebicki W. Phys Rev B 2006;73:075202.[64] Weng ZC, Huang ZG, Lin WX. Physica B 2012;407:743–7.[65] Yon Z, Liu T, Uruga T, Tanida H, Qi D, Rusydi A, Wee ATS. Materials

2010;3:3642–53.[66] Venkatesan M, Fitzgerald CB, Lunney J. Phys Rev Lett 2004;93:177206.[67] Lide DR, editor. Handbook of Chemistry and Physics. Boca Raton, Florida: CRC

Press Taylor & Francis; 2010.[68] Paasch G, Lindner T, Scheinert S. Synthetic Metals 2002;132:97–104.[69] Lu Y-M, Chang C-M, Tsai S-I, Wey T-S. Thin Solid Films 2004;447–448:56–60.[70] Lin S-S, Huang J-L, Sajgalik P. Surf Coat Technol 2005;191:286–92.[71] Oo WH, Saraf L, Engelhard M, Shutthanandan V, Bergman L, Huso J, McCluskey

M. J Appl Phys 2009;105:013715.[72] Chvostova D, Dejneka A, Hubicka Z, Churpita A, Bykov P, Jastrabik L, Trepakov

V. Phys Status Solidi A 2011;208:2140–3.[73] Kaspar TC, Droubay T, Heald SM, Nachimuthu P, Wang CM, Shutthanandan V,

Johnson CA, Gamelin DR, Chambers SA. New J Phys 2008;10:055010.[74] McCluskey MD, Jokela SJ. J Appl Phys 2009;106:071101.[75] de Walle CGV, Janotti A. Phys Status Solidi B 2011;248:19–27.