X ray diffraction and extended x ray absorption fine structure study of epitaxial mixed ternary bixbyite prxy2 xo3 x 0 2 films on si 111

7 0 0
X ray diffraction and extended x ray absorption fine structure study of epitaxial mixed ternary bixbyite prxy2 xo3 x 0 2 films on si 111

Đang tải... (xem toàn văn)

Thông tin tài liệu

X ray diffraction and extended X ray absorption fine structure study of epitaxial mixed ternary bixbyite PrxY2−xO3 (x = 0–2) films on Si (111) X ray diffraction and extended X ray absorption fine stru[.]

X-ray diffraction and extended X-ray absorption fine structure study of epitaxial mixed ternary bixbyite PrxY2-xO3 (x = 0–2) films on Si (111) G Niu, M H Zoellner, P Zaumseil, A Pouliopoulos, F d'Acapito, T Schroeder, and F Boscherini Citation: J Appl Phys 113, 043504 (2013); doi: 10.1063/1.4788982 View online: http://dx.doi.org/10.1063/1.4788982 View Table of Contents: http://aip.scitation.org/toc/jap/113/4 Published by the American Institute of Physics JOURNAL OF APPLIED PHYSICS 113, 043504 (2013) X-ray diffraction and extended X-ray absorption fine structure study of epitaxial mixed ternary bixbyite PrxY22xO3 (x 0–2) films on Si (111) G Niu,1 M H Zoellner,1 P Zaumseil,1 A Pouliopoulos,2 F d’Acapito,3 T Schroeder,1,4 and F Boscherini2,3 IHP, Im Technologiepark 25, 15236 Frankfurt (Oder), Germany Department of Physics and Astronomy, University of Bologna, viale C BertiPichat 6/2, 40127 Bologna, Italy Consiglio Nazionale delle Ricerche, Istituto Officina dei Materiali, Operative Group in Grenoble, c/o European Synchrotron Radiation Facility, B.P 220, 38043 Grenoble, France BTU Cottbus, Konrad-Zuse-Str 1, 03046 Cottbus, Germany (Received 22 November 2012; accepted January 2013; published online 23 January 2013) Ternary single crystalline bixbyite PrxY2xO3 films over the full stoichiometry range (x ¼ 0–2) have been epitaxially grown on Si (111) with tailored electronic and crystallographic structure In this work, we present a detailed study of their local atomic environment by extended X-ray absorption fine structure at both Y K and Pr LIII edges, in combination with complementary high resolution x-ray diffraction measurements The local structure exhibits systematic variations as a function of the film composition The cation coordination in the second and third coordination shells changes with composition and is equal to the average concentration, implying that the PrxY2xO3 films are indeed fully mixed and have a local bixbyite structure with random atomic-scale ordering A clear deviation from the virtual crystal approximation for the cation-oxygen bond lengths is detected This demonstrates that the observed Vegard’s law for the lattice variation as a function of composition is based microscopically on a more complex scheme related to local structural distortions which C 2013 American Institute of Physics accommodate the different cation–oxygen bond lengths V [http://dx.doi.org/10.1063/1.4788982] I INTRODUCTION In recent years, the integration of crystalline rare earth (RE) bixbyite oxide RE2O3 thin films on silicon (including the oxides having isomorphic structures like Y2O3 and Sc2O3) has attracted continuous and intense research interest The interest was to identify an alternative high-j dielectric for Si-based complementary metal oxide semiconductor (CMOS) technology.1 Nowadays, RE2O3 thin films also turn out to be a promising candidate for integrating numerous functions in Si technology RE2O3 materials not only exhibit various attractive functionalities of their own (such as radiation tolerance,2 diluted magnetic semiconducting3 and luminescence4 properties, etc.) but also may be used as tailored buffer layers for the growth of other functional oxides (such as ferromagnetic5 and multiferroic6 compounds) as well as semiconductors including Si,7,8 Ge,9,10 and III-V11,12 on Si The heteroepitaxial growth of ternary crystalline RE2O3 alloys on Si with variable stoichiometry provides a new tool and adds extra flexibility to tailor crystallographic and (opto-) electronic properties.13,14 For instance, in our prior studies,15,16 we have shown that the growth of single crystalline PrxY2xO3 films over the full stoichiometry range (x ¼ 0–2) on Si (111) with tailored electronic and crystallographic structures is feasi˚ ) and Y2O3 ble The lattice constants of Pr2O3 (11.152 A ˚ (10.604 A) are 2.7% larger and 2.4% smaller than twice the Si ˚ ), respectively It was thus demonlattice constant (5.431 A strated that fully lattice-matched Pr0.9Y1.1O3 can be grown on Si (111) to serve as a buffer layer for the growth of (almost) fully relaxed Si/insulator/Si heterostructures.7 0021-8979/2013/113(4)/043504/6/$30.00 Despite the great importance of an understanding of the local structure and ordering of such ternary mixed RE2O3 alloy thin films on Si, studies which provide atomistic insights are still limited From a fundamental research point of view, the understanding of chemical disorder effects in solid solutions will permit better comprehension and control of their unique properties which the binary compounds not possess From an application point of view, control over the atomic ordering is for example of importance for Eudoped PrxY2xO3 films which can potentially be used as scintillator materials; in this case, it is important that the Eu3ỵ activator cations reside solely on the noncentrosymmetric cation sites of the mixed PrxY2xO3 crystal in order to optimize its optical activity.17 In this context, we present here a first study to clarify whether the mixed PrxY2xO3 system exhibits random atomic-scale ordering or whether there is a preference for/against cation clustering Moreover, we probe the short scale structure by measuring the cation– oxygen bond lengths as a function of concentration Extended X-ray absorption fine structure (EXAFS) permits obtaining information on the local atomic geometry surrounding the excited atom and therefore it is a powerful tool for determining the local atomic structure of condensed matter and in particular of semiconductor heterostructures and nanostructures.18 It has been successfully used to investigate the local structure (lattice distortion, bond length variation and atomic inter-diffusion at the interface, etc.) of several heterostructure systems such as InGaAs/InP,19–21 GaN/AlN,22 Y2O3/ Si,23 Ag/MgO24 and NiO/Ag,25 and Ge-doped ZrO2.26 X-ray diffraction (XRD) is a suitable technique to determine information such as crystal phase, unit cell dimensions and strain 113, 043504-1 C 2013 American Institute of Physics V 043504-2 Niu et al of thin films In this work, we present a combined EXAFS and XRD study on single crystalline ternary PrxY2xO3 films (x ¼ 0–2) on Si (111) to investigate the local atomic ordering of such mixed bixbyite RE2O3 thin films J Appl Phys 113, 043504 (2013) bulk Y2O3 reference sample (commercial yttria powder), data were recorded at the Y K-edge at room temperature in the transmission mode III RESULTS AND DISCUSSION II EXPERIMENT A XRD analysis All PrxY2xO3 samples were grown on inch p-type Si (111) wafers by co-evaporation using a DCA 600 molecular beam epitaxy (MBE) reactor More details on sample preparation can be found in Ref Four PrxY2xO3/Si (111) samples with different Pr concentrations of x ¼ 0, 0.5, 0.7, 2.0 were fabricated The stoichiometries were determined using quantitative X-ray photoelectron spectroscopy (XPS) analysis by calculating the XPS peak areas below the curves and was cross checked by Rutherford back scattering (RBS) (data not shown) Detailed information was presented in Ref 27 The crystallinity of the samples was characterized using two X-ray diffractometers with different arrangements: a Rigaku DMAX 1500 (Cu-Ka radiation) in medium-resolution setup without crystal monochromator and a Rigaku SmartLab diffractometer ˚ ) in high-resolution setup with (Cu-Ka1 radiation, k ¼ 1.5406 A Ge (400)  crystal collimator X-ray reflectivity was used to determine oxide film thicknesses which are in the range of 1520 nm for all samples discussed here (data not shown) Y K edge and Pr LIII edge EXAFS measurements on the PrxY2xO3/Si (111) thin films as well as a bulk Y2O3 reference sample were carried out at the GILDA beam line (BM08) of the European Synchrotron Radiation Facility (ESRF) in Grenoble, France The epi-layers were measured at 80 K (obtained using a liquid nitrogen cryostat) in the fluorescence mode using a dynamically sagitally focusing Si (311) monochromator,28 a 13-element hyper pure Ge detector and associated digital electronics with a ls peaking time.29 For the Y edge measurements (edge energy 17.038 keV) samples were mounted vertically on a vibrating sample holder to eliminate spectral distortions due to the single crystal substrate and the angle between the impinging photon beam and the sample surface was 45 For the Pr edge measurements (edge energy 5.964 keV) the samples were nearly horizontal in a 2 grazing incidence geometry30 since it was found to be the best solution to eliminate diffraction effects in this case In both cases, higher order harmonics were eliminated with a pair of Pd coated grazing incidence mirrors The energy spacing for the ˚ 1 For the EXAFS spectra was equivalent to less than 0.05 A The crystalline quality of the epitaxial PrxY2xO3 thin films was characterized by XRD measurements Fig 1(a) shows specular h-2h measurements around the Si (222) Bragg diffraction condition (2h ¼ 58.88 ) The pure Y2O3 (x ¼ 0) and pure Pr2O3 (x ¼ 2) films exhibit the oxide (444) peak at 2h ¼ 60.46 and 2h ¼ 56.78 , respectively, which correspond to almost entirely relaxed films Both samples were characterized by using a medium-resolution diffractometer (DMAX 1500) thus their Si (222) Bragg peaks show double-peak feature related to Cu-Ka1 and Cu-Ka2 (222) reflections of the Si (111) wafers For the mixed samples with x ¼ 0.5 and x ¼ 0.7, PrxY2xO3 (444) peaks appear between the Y2O3 (444) and the Pr2O3 (444) ones (marked by dashed lines) Moreover, for these two samples, only Cu-Ka1 reflection can be observed for Si (222) peak thanks to the utilization of a high-resolution diffractometer (Rigaku SmartLab) equipped with a crystal monochromator We note that the oxide peak is symmetric and un-split: this indicates that the samples indeed crystallize in the cubic bixbyite phase without phase separation With increasing Pr concentration, the oxide reflection shifts from the Y2O3 side towards the Pr2O3 side, implying an increase of the lattice parameter Furthermore, both mixed PrxY2xO3 films spectra display thickness-related interference fringes at the foot of the main oxide peak, revealing thus a sharp interface structure of the epitaxial films In order to further examine the strain status of the mixed oxide films, a cos2v measurement (Fig 1(b)) was performed In this analysis, we measure the Bragg peaks of different oxide lattice planes in a skew geometry inclined by an angle v to the [111] growth direction A lattice parameter is calculated assuming a cubic lattice structure for each measurement These “cubic” lattice parameters are then plotted as a function of cos2v In this kind of plot, a cubic lattice (relaxed) exhibits a horizontal line because the in-plane (aIP) and out-of-plane (aOP) lattice parameters given at cos2v ¼ and cos2v ¼ 1, respectively, are identical A tetragonal lattice (strained) exhibits an inclined line, the slope of which is either negative (aIP > aOP) for tensile or positive (aIP < aOP) for compressive strain in the layer.31–33 The bulk lattice parameters of Pr2O3, FIG (a) XRD specular h-2h measurements of PrxY2xO3 films with x ¼ 0, 0.5, 0.7 and 2; (b) lattice parameters of PrxY2xO3 as a function of cos2v The bulk lattice parameters of Pr2O3, Y2O3 and the silicon are marked by dotted lines For better comparison with Si, the oxide lattice parameters are divided by The Miller indices of the measured oxide reflections are given in the plots These values have to be divided by for the corresponding indices of Si reflections 043504-3 Niu et al Y2O3 and silicon are marked by dotted lines in Fig 1(b): note that oxide lattice parameters are divided by for better comparison with Si It is observed that both samples (which are Y2O3-rich) have smaller in-plane lattice parameters than Si The experimental points of the x ¼ 0.7 sample (20.5 nm) are positioned on a horizontal line, indicating that this film is fully relaxed The sample with x ¼ 0.5 (14.8 nm), instead, exhibits experimental points lying on an inclined line with a negative slope which corresponds to a small tensile strain, which probably originates from an incomplete relaxation The bulk lattice parameter of this strained sample can be calculated by the relationship between the in-plane and out-of-plane strain during the tetragonal deformation for cubic materials: aOP a0 2 ¼  1  aIPaa , where  is the Poisson’s ratio 0.3, a0 ˚ ˚ ; we find that a0 ¼ 5.375 A ˚, aOP ¼ 5.358 A and aIP ¼ 5.393 A indicating that the residual tensile strain in this sample is thus only 0.33% It is also worth noting that the lattice parameters extracted from direct XRD measurement agree well with those calculated by Vegard’s law, using the stoichiometry determined by XPS and RBS In other words, the variation of lattice parameters of mixed cubic bixbyite PrxY2xO3 films as a function of stoichiometry follows Vegard’s law B EXAFS data analysis The initial data processing and background subtraction of the EXAFS data were performed using ATHENA34 and ESTRA packages35 for the Y edge and the Pr edge, respectively The raw Y K edge and Pr LIII edge EXAFS oscillations [kv(k)] for bulk Y2O3 reference and PrxY2xO3 samples with x ¼ 0, 0.5, 0.7 and are, respectively, shown in Figs 2(a) and 2(b) Evidently, PrxY2xO3 samples with different stoichiometries exhibit similar main oscillations at both Y K edge and Pr LIII edges With increasing the Pr concentration, the fine structure oscillations gradually change Moreover, the data of the Y2O3 film (x ¼ 0) demonstrate not only the same oscillations but also a similar level of signal-to-noise ratio as the bulk Y2O3 reference powder sample The magnitude of the Fourier transforms (FT) of the k3 weighted spectra (continuous lines) and the corresponding best fits (dashed lines) are shown in Figs 3(a) and 3(b), ˚ 1 and respectively The ranges for the FT were 4–12 A 1 ˚ 2.5–9 A for the Y and Pr edge spectra, respectively The FTs allow a qualitative interpretation of the spectra Fig 3(a) shows the FT spectra at Y K edge: the overall similarity of FIG Background subtracted raw EXAFS data of PrxY2xO3 samples (a) Y K-edge and (b) Pr LIII-edge J Appl Phys 113, 043504 (2013) FIG Magnitude of the Fourier Transforms of EXAFS presented in Fig (continuous lines) and corresponding fits (dashed lines) (a) Y K-edge and (b) Pr LIII-edge the spectra for films and the bulk reference further confirm the fact that all the mixed PrxY2xO3 films have similar local atomic environment of Y atoms as the bulk Y2O3 one Furthermore, according to the local structure of bixbyite (see below in detail), the first peak can be attributed to cation-O bonds, while the second and third ones are related to cation-cation atomic correlations.23 This interpretation will be confirmed by detailed fitting Spectra gradually evolve with composition, in particular the relative intensity of the second and third peaks relative to the first one change with increasing Pr concentration This could be due, in principle, to the presence of both Y and Pr in the second and third coordination shells and/or to increased structural disorder Similar behavior was also observed for Pr LIII edge spectra (Fig 3(b)), which can be interpreted in a similar way Quantitative analysis was performed by nonlinear fitting of the EXAFS signals using ARTEMIS34 based on theoretical scattering functions simulated by FEFF.36 Simulations were based on the bixbyite sesquioxide structure that belongs to space group Ia3, N 206 Bixbyite is thought of as an oxygen deficient fluorite in which the lattice parameter is doubled by removing in a systematic way a quarter of the oxygen atoms.17 As shown in Fig 4, in the bixbyite structure there are two FIG Schematic of bixbyite RE2O3 structure showing two unique cation sites: 24 d site, where the O vacancy positions are on opposite corner of one cube face leading to a distorted cube and b site, where the O vacancy positions are diagonally opposed thus resulting in an undistorted cube 043504-4 Niu et al J Appl Phys 113, 043504 (2013) TABLE I Average Y K edge distances Sample ˚) RðY  OÞ (A ˚) RðY  Y Þ (A ˚) RðY  Pr Þ (A ˚) RðY  Y Þ (A ˚) RðY  Pr3 Þ (A Pr0.5Y1.5O3 Pr0.7Y1.3O3 Y2O3 Y2O3 bulk Y2O3 XRD 2.305 (0.010) 2.305 (0.015) 2.281 (0.010) 2.266 (0.013) 2.2811 3.519 (0.014) 3.496 (0.024) 3.540 (0.009) 3.523 (0.013) 3.5294 3.576 (0.037) 3.590 (0.047) 2 4.028 (0.032) 3.88 (0.15) 4.024 (0.011) 3.999 (0.019) 3.998 4.120 (0.044) 3.989 (0.065) 2 unique cation sites, named 24 d and b, respectively For the 24 d site, the O vacancy positions are on opposite corners of one cube face leading to a cation coordination in the form of a distorted octahedron with three distinct metal-oxide distances a, b and c Instead, for the b site, the O vacancy positions are diagonally opposed along a space diagonal, thus resulting in an undistorted octahedron with six equal metaloxide distances a’ For each b site there are three 24 d sites, so the contribution of the regular b site to the total signal amplitude is 25% and the one of the distorted 24 d site 75% This complex local structure was taken into account by performing two separate FEFF calculations for the two cation sites and subsequently fixing their relative weight to 25% and 75% in the fitting procedure In order to calculate scattering functions for the mixed oxides, we used 2nd and 3rd shells composed of both Y and Pr in equal amounts and we set the lattice parameter of PrxY2xO3 according to Vegard’s law (a ẳ 2x :aPr2 O3 ỵ (1- 2x):aY2 O3 ) in which x is the Pr concentration; cation-O distances were, instead, kept being fixed at the values for the binary compounds We found that the paths that contribute significantly to the signal were the oxygen 1st shell, the 2nd, 3rd cation shells and the 4-leg back and forth multiple scattering from two symmetrical 1st shell oxygen atoms Fitting was per˚ for the ˚ 1 and R ¼ 1.4–4.2 A formed in the ranges k ¼ 4–12 A 1 ˚ ˚ Y edge and k ¼ 2.5–9 A and R ¼ 1.4–4.6 A for the Pr edge, with a k3 weight; extensive attempts indicated that these parameters provided the most consistent fits for all the samples, with reasonable and compatible structural parameters at both edges Fitting variables were a shift of the origin of the k-scale, three distance variations with three Debye-Waller factors for all atomic correlations within the first three coordination shells; the distance variation of the multiple scattering signal was set as twice that of the first shell and the amplitude reduction factors were fixed to the value found for the reference samples The best fits are reported in Figs 3(a) and 3(b) as dashed lines We found that the best fits, with reasonable structural parameters, were obtained at both edges when the relative coordination numbers of the second and third cation shells was equal to the nominal relative composition This is compatible with a random cation arrangement in the alloy samples The numerical results of the fitting procedure are reported in Tables I to IV Numbers in brackets are estimates of the uncertainties and for the sake of simplicity only average distances for each coordination shell are reported We note that for the x ¼ 0.5 sample the small 0.33% strain detected by XRD is expected to give a negligible effect on the first shell distances.37 Cation-oxygen (first shell) distances are plotted as a function of concentration in Fig In Fig 5, the green solid line reports the prediction of the Virtual-crystal approximation (VCA)38 for the cation-oxygen distances, calculated as the linear interpolation between the average Pr-O and Y-O distances The chemical disorder in a solid solution can be theoretically simulated by using first-principles calculation A direct way is to use as large as possible supercells with periodic boundary conditions so that inhomogeneities can be incorporated into perfectly periodic systems, an approach which is, however quite computationally demanding.38 TABLE II Y edge Debye-Waller factors Sample Pr0.5Y1.5O3 Pr0.7Y1.3O3 Y2O3 Y2O3 bulk ˚ 2) r2 ðY  OÞ (103A ˚ 2) r2 ðY  Y Þ (103A ˚ 2) r2 ðY  Pr2 Þ (103A ˚ 2) r2 ðY  Y Þ (103A ˚ 2) r2 ðY  Pr3 Þ (103A ˚ 2) r2 ðMSÞ (103A 4.5 (0.7) 5.3 (0.8) 3.4 (0.9) 3.4 (1.0) 2.1 (1.4) 1.7 (3.0) 2.1 (0.4) 3.8 (0.6) 1.5 (3.8) 0.7 (4.0) 2 3.7 (2.0) (20) 2.8 (0.7) 6.2 (1.5) 1.1 (3.0) (11) 2 (10) (19) 10 (30) (19) TABLE III Average Pr LIII edge distances Sample ˚) RðPr  OÞ (A ˚) RðPr  Pr2 Þ (A ˚) RðPr  Y Þ (A ˚) RðPr  Pr3 Þ (A ˚) RðPr  Y Þ (A Pr2O3 XRD Pr2O3 Pr0.7Y1.3O3 Pr0.5Y1.5O3 2.429 2.448 (0.022) 2.368 (0.013) 2.368 (0.030) 3.7631 3.777 (0.018) 3.67 (0.14) 3.71 (0.28) 2 3.602 (0.058) 3.574 (0.039) 4.248 4.114 (0.097) 4.011 (0.096) 4.116 (0.091) 2 4.011 (0.089) 4.09 (0.10) 043504-5 Niu et al J Appl Phys 113, 043504 (2013) TABLE IV Pr edge Debye-Waller factors Sample Pr2O3 Pr0.7Y1.3O3 Pr0.5Y1.5O3 ˚ 2) r2 ðPr  OÞ (103A ˚ 2) r2 ðPr  Pr2 Þ (103A ˚ 2) r2 ðPr  Y Þ (103A ˚ 2) r2 ðPr  Pr3 Þ (103A ˚ 2) r2 ðPr  Y Þ (103A ˚ 2) r2 ðMSÞ (103A 10.4 (3.6) 2.4 (1.2) 4.4 (2.4) 7.4 (2.2) 1.6 (6.4) (36) 0.2 (7.0) 1.5 (4.4) 26 (19) 5.6 (9.6) 10 (10) 9.8 (8.1) 5 (10) (17) (40) (52) Therefore, the VCA, as a simplified approach to address this issue, has been widely used in the last few decades It constructs a composite potential which represents the average of the component atoms and thus implies a linear interpolation between the cation-oxygen distances of the parent compounds In Fig 5, the first shell cation-oxygen distances pffiffiffi (crosses) calculated from the XRD measurement (dM-O ¼ 3aoxide/4)39,40 lie as expected on the VCA line Interestingly, the EXAFS data show that the cation-oxygen distances change with concentration but clearly not follow the VCA line The individual Y-O and Pr-O bond lengths have a smaller variation with concentration respect to that predicted by the VCA In the discussion of the dependence of bond lengths as a function of concentration in solid solutions of the type AxB1xC (AC and BC being the binary compounds), it is customary to identify two limiting cases:41 (i) composition independent according to Bragg’s and Pauling’s notions, namely, RAC(x) and RBC(x) always equal the sum of the covalent (or ionic) radii R0AC and R0BC (marked by straight dotted line in Fig 5); (ii) composition dependent with linear variation, as predicted by VCA In the present case, the bond lengths exhibit a composition dependence which is in between the two limits This behavior is similar to what has been found in III-V semiconductor alloys42–45 and ternary oxide systems like Th1xUxO2 or Th1xPuxO239 and Cd1xCaxO46 This behavior has been understood as due to the fact that a bond angle distortion is energetically less costly than a bond length variation; the linear variation of the lattice parameter FIG Average cation-oxygen bond lengths; the green line reports the prediction of the virtual crystal approximation (VCA) Circles: Pr-O bond length; squares: Y-O bond length; crosses: cation-oxygen average distances calculated from XRD on the samples which follow Vegard’s law (Vegard’s law) is thus accommodated locally by angular distortions and the bond lengths tend to remain relatively close to the sum of the ionic radii exhibiting a bimodal distribution Furthermore, the bimodal distribution of bond lengths is considered to probably result in optical bowing for semiconductor alloys, which is, however, not the case for our mixed bixbyite dielectric oxides because the band gaps of the PrxY2xO3 films seem to exhibit a quite good linear variation as a function of the concentration x (see our prior work Ref 15) The second and third shell interatomic distances, reported in Tables I to IV, are roughly constant as a function of composition and follow a reasonable trend The second and third shell Y-Pr and Pr-Y distances are equal, within the uncertainties On average, distances increase in the order Y-Y, Y-Pr/Pr-Y, Pr-Pr This behavior is consistent with ionic radii, considering that the ionic radii for Y3ỵ, Pr3ỵ and O2 ˚ , 1.13 A ˚ and 1.26 A ˚ , respectively.47 are 1.04 A Turning now to the Debye-Waller factors reported in Tables I to IV and Fig 6, it can be observed that the ones related to the cation–oxygen bonds fall, except for the pure ˚ 20.006 A ˚ 2, Pr2O3 sample (x ¼ 2), in the range of 0.002 A suggesting similar ordered structure in these films A slightly higher Debye-Waller factor of the pure Pr2O3 film suggests an increase of disorder in the film It is likely that the fact that the pure Pr2O3 film exhibits a more disordered structure is related to the complex phase diagram Pr-O which has different non-stoichiometric oxide phases between PrO2 and Pr2O3.48 Some of the second and third shell Debye–Waller factors have very large errors, sometimes associated to unphysical negative best estimates (in these cases, however, the large error indicates a range of the Debye-Waller factor which includes positive values); this is very likely due to the limited k-range and the correlation among fitting variables and precludes the possibility of extracting meaningful FIG Cation-oxygen Debye-Waller factors (r2) Circles: r2(Pr-O); squares: r2(Y-O) 043504-6 Niu et al physical information from the second and third shell DebyeWaller factors IV CONCLUSION In conclusion, we presented a detailed EXAFS and XRD study on the short and long range order of ternary mixed PrxY2xO3 films on Si (111) The high-quality single crystalline characteristics of the films were confirmed by XRD measurements The variation of lattice parameters of mixed ternary PrxY2xO3 films follows Vegard’s law The EXAFS data at the Y K edge and Pr LIII edge reveals that all the samples have a local structure which bears a close relation to the expected bixbyite one For the alloys no significant increase in static disorder in the cation–oxygen distances is found With the variation of the film composition, the local environment of the excited atom shows an ordered evolution Furthermore, for the two alloy samples, the local cation coordination is random with the composition in the second and third coordination shells being equal to the average concentration This result confirms the formation of fully mixed ternary PrxY2xO3 films on Si (111) A bimodal distribution of the nearest-neighbor bond length exists in the oxide solutions because the individual Y-O and Pr-O bond lengths demonstrate significantly smaller concentrationinduced variation than that predicted by the VCA The second and third shell interatomic distances display a behavior consistent with atomic radii The results presented in this study permit improving the understanding of the local atomic structure of mixed RE2O3 films grown on Si, in order to engineer their properties for future applications ACKNOWLEDGMENTS The authors gratefully acknowledge M Katsikini and C Meneghini for stimulating discussions and Hans-J€urgen Thieme for technical assistance of sample preparation G Niu wishes to thank Alexander von Humboldt foundation for his AvH Post-Doc fellowship The authors furthermore gratefully acknowledge financial funding by DFG (Deutsche Forschungsgemeinschaft) R A McKee, F J Walker, and M F Chisholm, Phys Rev Lett 81, 3014 (1998) M Tang, P Lu, J Valdez, and K Sickafus, J Appl Phys 99, 063514 (2006) J Stankiewicz, F Villuendas, and J Bartolome, Phys Rev B 75, 235308 (2007) S S Kim, J H Moon, B.-T Lee, K.-S Sohn, T S Kang, and J H Je, Appl Surf Sci 221, 231 (2004) F Sanchez, R Bachelet, P d Coux, B Warot-Fonrose, V Skumryev, L Tarnawska, P Zaumseil, T Schroeder, and J Fontcuberta, Appl Phys Lett 99, 211910 (2011) J H Choi, J Y Lee, and Y T Kim, Appl Phys Lett 77, 4028 (2000) A Giussani, P Zaumseil, O Seifarth, P Storck, and T Schroeder, New J Phys 12, 093005 (2010) A Fissel, D K€uhne, E Bugiel, and H J Osten, Appl Phys Lett 88, 153105 (2006) A Giussani, P Zaumseil, P Rodenbach, G Weidner, M A Schubert, D Geiger, H Lichte, P Storck, J Wollschl€ager, and T Schroeder, J Appl Phys 106, 073502 (2009) J Appl Phys 113, 043504 (2013) 10 G Niu, L Largeau, G Saint-Girons, B Vilquin, J Cheng, O Mauguin, and G Hollinger, J Vac Sci Technol A 28, 1187 (2010) 11 L Tarnawska, P Zaumseil, M A Schubert, S Okur, U Ozgur, H Morkoc¸, R Paszkiewicz, P Storck, and T Schroeder, J Appl Phys 111, 073509 (2012) 12 G Saint-Girons, P Regreny, L Largeau, G Patriarche, and G Hollinger, Appl Phys Lett 91, 241912 (2007) 13 V Narayanan, S Guha, N A Bojarczuk, and F M Ross, J Appl Phys 93, 251 (2003) 14 J Zachariae and H Pfn€ ur, Phys Rev B 72, 075410 (2005) 15 O Seifarth, M A Schubert, A Giussani, D O Klenov, D Schmeisser, and T Schroeder, J Appl Phys 108, 103709 (2010) 16 O Seifarth, B Dietrich, P Zaumseil, A Giussani, P Storck, and T Schroeder, J Appl Phys 108, 073526 (2010) 17 C R Stanek, K J McClellan, B P Uberuaga, K E Sickafus, M R Levy, and R W Grimes, Phys Rev B 75, 134101 (2007) 18 F Boscherini, in Characterization of Semiconductor Heterostructures and Nanostructures, edited by C Lamberti (Elsevier, Amsterdam, 2008) 19 M Tormen, D D Salvador, A V Drigo, F Romanato, F Boscherini, and S Mobilio, Phys Rev B 63, 115326 (2001) 20 M Tormen, D De Salvador, F Boscherini, F Romanato, A V Drigo, and S Mobilio, Appl Surf Sci 188, 85 (2002) 21 J C Woicik, J A Gupta, S P Watkins, and E D Crozier, Appl Phys Lett 73, 1269 (1998) 22 F d’Acapito, F Boscherini, S Mobilio, A Rizzi, and R Lantier, Phys Rev B 66, 205411 (2002) 23 M Malvestuto, R Carboni, F Boscherini, F D’Acapito, S S a M Fanciulli, and G V A Dimoulas, and G Mavrou, Phys Rev B 71, 075318 (2005) 24 A M Flank, R Delaunay, P Lagarde, M Pompa, and J Jupille, Phys Rev B 53, R1737 (1996) 25 E Groppo, C Prestipino, C Lamberti, P Luches, C Giovanardi, and F Boscherini, J Phys Chem B 107, 4597 (2003) 26 F Boscherini, F D’Acapito, S F Galata, D Tsoutsou, and A Dimoulas, Appl Phys Lett 99, 121909 (2011) 27 A Wilke, J.-M Yang, J W Kim, O Seifarth, B Dietrich, A Giussani, P Zaumseil, P Storck, and T Schroeder, Surf Interface Anal 43, 827 (2010) 28 S Pascarelli, F Boscherini, F d’Acapito, J Hrdy, C Meneghini, and S Mobilio, J Synchrotron Radiat 3, 147 (1996) 29 G Ciatto, F d’Acapito, F Boscherini, and S Mobilio, J Synchrotron Radiat 11, 278 (2004) 30 C Maurizio, M Rovezzi, F Bardelli, H G Pais, and F D’Acapito, Rev Sci Instrum 80, 063904 (2009) 31 P Zaumseil, J Phys D: Appl Phys 41, 135308 (2008) 32 P Zaumseil and T Schroeder, J Phys D: Appl Phys 44, 055403 (2011) 33 P Zaumseil, L Tarnawska, P Storck, and T Schroeder, J Phys D: Appl Phys 44, 315403 (2011) 34 B Ravel and M Newville, J Synchrotron Radiat 12, 537 (2005) 35 See http://webusers.fis.uniroma3.it/~meneghini/software.html for more instructions 36 A Ankudinov, B Ravel, J J Rehr, and S Conradson, Phys Rev B 58, 7565 (1998) 37 F Romanato, D D Salvador, M Berti, A Drigo, M Natali, M Tormen, G Rossetto, S Pascarelli, F Boscherini, C Lamberti, and S Mobilio, Phys Rev B 57, 14619 (1998) 38 N J Ramer and A M Rappe, Phys Rev B 62, R743 (2000) 39 S Hubert, J Purans, G Heisbourg, P Moisy, and N Dacheux, Inorg Chem 45, 3887 (2006) 40 Crystal structures; Vol 2, edited by R W G Wyckoff (Interscience publishers, New York, 1964) 41 J L Martins and A Zunger, Phys Rev B 30, 6217 (1984) 42 A Zunger and J E Jaffe, Phys Rev Lett 51, 662 (1983) 43 J C Mikkelsen, Jr and J B Boyce, Phys Rev Lett 49, 1412 (1982) 44 J C Mikkelsen, Jr and J B Boyce, Phys Rev B 28, 7130 (1983) 45 A Balzarotti, A Kisiel, N Motta, M Z Starnawska, M T Czycyk, and M Podgorny, Phys Rev B 30, 2295 (1984) 46 V Srihari, V Sridharan, S Chandra, V S Sastry, H K Sahu, and C S Sundar, J Appl Phys 109, 013510 (2011) 47 See http://abulafia.mt.ic.ac.uk/shannon/ptable.php.for Database of ionic radii 48 T Schroeder, T.-L Lee, L Libralesso, I Joumard, J Zegenhagen, P Zaumseil, C Wenger, G Lupina, G Lippert, J Dabrowski, and H.-J M€ ussig, J Appl Phys 97, 074906 (2005) ... 2. 305 (0. 0 10) 2. 305 (0. 015) 2. 281 (0. 0 10) 2. 266 (0. 013) 2. 2811 3.519 (0. 014) 3.496 (0. 024 ) 3.5 40 (0. 009 ) 3. 523 (0. 013) 3. 529 4 3.576 (0. 037) 3.5 90 (0. 047) 2 4 . 02 8 (0. 0 32) 3.88 (0. 15) 4 . 02 4 (0. 011)...JOURNAL OF APPLIED PHYSICS 113, 04 3 504 ( 20 1 3) X- ray diffraction and extended X- ray absorption fine structure study of epitaxial mixed ternary bixbyite PrxY 22 xO3 (x 0? ? ?2) films on Si (111) G Niu,1... Pr0.5Y1.5O3 2. 429 2. 448 (0. 022 ) 2. 368 (0. 013) 2. 368 (0. 0 30) 3.7631 3.777 (0. 018) 3.67 (0. 14) 3.71 (0. 28 ) 2 3. 6 02 (0. 058) 3.574 (0. 039) 4 .24 8 4.114 (0. 097) 4 .01 1 (0. 096) 4.116 (0. 091) 2 4 .01 1 (0. 089)

Ngày đăng: 15/03/2023, 20:11

Tài liệu cùng người dùng

Tài liệu liên quan