Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Strain-induced non-linear optical characteristics of pyroelectric PbVO3 epitaxial thin films

Open Access Open Access

Abstract

Highly strained PbVO3 films have potential applications in pyroelectric sensors and optoelectronic devices. To understand how strain influences the nonlinear optical characteristic of PbVO3, highly distorted tetragonal PbVO3 epitaxial thin films were grown on LaAlO3 and SrTiO3 substrates. Using Raman scattering spectroscopy, theoretical and experimental studies on the phonon spectra yield insights on the strained structures of V-O apical bondings. Effects of applied strain on the linear optical properties of PbVO3 thin films were experimentally characterized by spectroscopic ellipsometry. Second-order nonlinear susceptibility tensor components of PbVO3/SrTiO3 thin films were determined to be almost twice as large as those of PbVO3/LaAlO3 thin films.

© 2016 Optical Society of America

1. Introduction

Octahedral rotations influence the B-O bond length and B-O-B angle in ABO3 perovskites, which play a major role in physical properties such as ferroelectricity [1], magnetic interactions, resistive behavior, and bandwidth. There have been some remarkable introductions of octahedral rotation for vanadium-based perovskite oxides, RVO3 (R = Ca, Sr, Y, La, Sm, Eu, Lu) [2,3]. Here, PbVO3 (PVO) is a very interesting perovskite-like pyroelectric because it has a special crystallographic structure that exhibits low-dimensional magnetic properties and giant pyroelectric polarization [4–7].

PVO has a strongly distorted tetragonal structure (c/a = 1.229), which consists of a VO5 square-pyramidal structure with one shortened vanadyl bond of vanadium and apical oxygen [8]. Its unique feature allows two-dimensional antiferromagnetic ordering and strong pyroelectric polarization [4–6]. In this pyramidal configuration of the PVO structure, the 3d electrons of vanadium ions are localized at the dxy orbital, leading to low-dimensional magnetic ordering [5]. Indeed, experimental results showing magnetic susceptibility at a broad maximum at 180 K support this low-dimensional antiferromagnetic ordering [4,5]. However, the orientation of magnetic moment of PVO is rather controversial because the ground energies of C- and G-type ordering for the PVO are similar [6]. In addition, PVO can be considered a potential ferroelectric driven by lone pairs of displaced A-cations, as with Bi3+ in BiFeO3 [6,7]. Highly distorted tetragonal BiFeO3 thin films with a giant axial ratio c/a = 1.23 show clear ferroelectric behavior and coexisting antiferromagnetic order [9,10] The ferroelectric polarization of PVO is enhanced by large tetragonal distortion, which is compatible with well-known ferroelectric, PbTiO3 (c/a = 1.06). Although the estimated value for spontaneous polarization of PVO is as large as 152 μC/cm2 [6], it has not been experimentally demonstrated for bulk or thin-film forms. Some reports indirectly support the possibility of ferroelectric polarization of the PVO. The pressure-induced reversible tetragonal to cubic transition of PVO supports an implicit ferroelectric property [11,12]. The reversible tetragonal to cubic phase transition in PVO is induced at approximately 2 GPa, and is associated with a significant reduction in 4 or 5 orders of magnitude [13]. Linear and nonlinear optical properties of PbVO3 have been theoretically predicted [14,15]. The refraction (n) and extinction (k) coefficients of PVO thin films on (La0.18Sr0.72)(Al0.59Ta0.11)O3 substrate were experimentally obtained [16]. Recently, stable PVO thin films were deposited from lead pyrovanadate, Pb2V2O7, using an off-stoichiometric process under either low oxygen pressure or high argon pressure, unlike other oxide materials with equilibrium targets [17]. These thin films were epitaxially grown along the c-axis and its lattice constant was elongated along the growth direction [18,19].

In the present work, non-destructive optical measurements were performed to determine the electronic structure and related physical properties of PVO. Using Raman scattering spectroscopy, the sub-structural distortion of the PVO thin films was studied. This is critically related to the strain and formation of the interface structure, which depends on the lattice mismatch. Linear and nonlinear optical properties for the PVO thin films were characterized through spectroscopic ellipsometry (SE) and second harmonic generation (SHG), respectively. Symmetry breaking along the c-axis in PVO thin films was demonstrated using an SHG signal with nonlinear susceptibility and the Fresnel equations.

2. Experimental methods

2.1 Growth of epitaxial PbVO3 thin films

Epitaxial PVO thin films were deposited on LAO (001) and STO (001) substrates using pulsed laser deposition [19]. This study focused on the tensile train because the c-lattice constant of PVO thin films on LAO substrate is elongated even for tiny tensile strains, with lattice mismatch of 0.26% [19]. The extreme structural distortion in the PVO lattice leads to VO5 square pyramidal structures rather than VO6 octahedral structure with the shortened vanadyl bond V-Oa, which might limit the switchable ferroelectric polarization of PVO. The epitaxial growth of the PVO thin films along the c-axis was determined by X-ray diffraction. Deposition was performed at a constant 570°C, Ar pressure of 200 mTorr, and laser fluence of 1.5 J/cm2 for 15 minutes, with a frequency of 5 Hz and using a KrF excimer laser with a wavelength of 248 nm and a pulse duration of 25 ns. The thickness of the deposited PVO thin films was measured using a surface profilometer (Dektak, Bruker) and was found to be approximately 200 nm. An off-stoichiometric Pb2V2O7 target was ablated to deposit the PVO thin films, owing to the chemical stability of PVO bulk. Before the ablation, all substrates were cleaned with acetone, isopropyl alcohol, ethyl alcohol, and distilled water to prevent any artificial contributions.

2.2 Raman scattering spectroscopy

The micro-Raman spectra of the samples were obtained with a McPherson 207 spectrometer equipped with a nitrogen-cooled charge-coupled device array detector and a 488-nm diode laser at room temperature. The power of the laser illuminating the sample was set to 1 mW and was focused onto a ~1-μm-diameter spot using a microscope objective lens ( × 100).

2.3 Spectroscopic ellipsometry

SE has been applied to determine the complex dielectric function spectra (ε = ε1 + iε2), in the photon energy range of 0.75 to 6.45 eV at room temperature using a spectroscopic rotating compensator-type ellipsometer (M2000-DI model, J. A. Woollam Inc.) [20]. Data were recorded after averaging 1000 cycles of the compensator (1000 revolutions per measurement) to increase the signal-to-noise ratio.

2.4 Second harmonic generation

An optical second harmonic signal is generated when a light source with frequency ω creates a nonlinear polarization by going through a nonlinear medium. The SHG experiments were performed using a Ti-sapphire laser with a wavelength of 800 nm. A typical single pulse has a width of 100 fs, with a repetition rate of 80 MHz. The average power of each fundamental light pulse is about 1.4 mW and is focused on the sample surface with a spot diameter of ~100 μm.

3. Results and discussion

3.1 Crystallinity of epitaxial PVO thin films

Figure 1(a) shows the typical XRD patterns of PVO thin films deposited on two different substrates in the θ-2θ scan, exhibiting only the (00l) reflection of PVO, which indicates highly c-axis-oriented growth of the PVO thin films. Figure 1(b) shows φ scans of the PVO/LAO and PVO/STO thin films for the {110} planes of PVO and LAO. A 4-fold in-plane symmetric structure was demonstrated by the φ scans, presenting a periodicity of 90°. Therefore, the epitaxial growth along the c-axis of the tetragonal PVO thin films was confirmed by the θ-2θ scans and 4-fold symmetric φ scan results. Figure 1(c) and 1(d) show topographic images with root mean square roughness 1.42 nm for PVO/LAO and 8.84 nm for PVO/STO thin films, obtained using atomic force microscopy. Figure 1(e) and 1(f) show HAADF-STEM images of PVO thin films on LAO (001) and STO (001) substrates, respectively. The well-defined crystal structure of the PVO thin films on LAO and STO substrates is obtained at the atomic scale by a fast Fourier transform, which indicates a tetragonal structure as shown inset in Fig. 1(e) and 1(f).

 figure: Fig. 1

Fig. 1 The XRD patterns of the PVO films deposited on LAO and STO substrates. (a) θ-2θ scans and (b) φ scans. The φ scan are for the {110} planes of PVO and LAO {110}. Topographic images of (c) PVO/LAO and (d) PVO/STO. HAADF-STEM images of PVO thin films on (e) LAO and (f) STO with the fast Fourier transform patterns.

Download Full Size | PDF

Applying θ-2θ XRD peaks to Bragg’s law, the c-lattice constants of PVO/LAO and PVO/STO were calculated to be 4.989 and 5.024 Å. The c-lattice parameter of the PVO/STO thin films was contrary to our expectation that it should be reduced owing to the tensile strain of the STO substrate. Therefore, the in-plane strain did not show conventional changes, despite the possibility that the lattice mismatch causes structural transformation. A vanadium-based oxide thin film, LaVO3, showed similar behavior; that is, the expansion of volume rather than the compression of the out-of-plane lattice parameter [21]. The PVO thin films were vertically elongated as much as 6.83% (PVO/LAO) and 7.58% (PVO/STO), compared with the c-lattice constant of bulk PVO, even when the PVO/LAO thin films were deposited on only 0.26% mismatched LAO substrates. The excessive c-elongation of the PVO thin films may have been influenced by the relatively weak interaction between neighboring layers in the PVO structure, compared with the interlayer interaction. Another possibility is that the electron density in the PVO lattice overlaps between Pb and O atoms, unlike other vanadium-based perovskite compounds such as SrVO3 and BaVO3, which do not have covalent bonds between Sr or Ba sites and O atoms [8,22]. In a bulk study containing constant in-plane lattice parameters and carefully controlled out-of-plane lattice parameters under artificial conditions [12], a- and b-lattice parameters showed only slight changes and the c-lattice parameter changed significantly.

3.2 Dynamic phonon structure

The structural distortion in the PVO samples was investigated by characterizing the local structure of the PVO through Raman scattering spectroscopy. Figure 2 illustrates the Raman spectra of the PVO thin films on LAO and STO substrates, with a non-centrosymmetric structure (space group P4mm). In Fig. 2(a), the peaks of the Raman spectra for the PVO/LAO film were observed at 175 cm−1 (Pb-O vibration modes), 220 and 320 cm−1 (V-O-V bending mode), 668 cm−1, and near 950 cm−1, while the Raman signal for PVO/STO film can be observed only at 950 cm−1, because the Raman signal of the STO substrate is large enough to mask the signal of the PVO thin film, as shown in Fig. 2(b). The peak identifications are in agreement with those presented in [18] and [19]. In particular, the peaks at 950 cm−1 split into two peaks, as shown in Fig. 2(c). The peaks were qualitatively identified as V-Oa (apical oxygen) stretching with Ag symmetric vibration mode [6,20], and the dominant intensities can be related to c-axis-oriented PVO films, which differ from bulk PVO [4]. These two peaks are also observed in the cubic KCrFe3, in which there is a melting of the orbital ordering [23].

 figure: Fig. 2

Fig. 2 Raman spectra of the PVO thin films on (a) LAO and (b) STO from 100 to 1,100 cm−1. (c) Raman spectra in specific region for V-O apical vibration signal near 950 cm−1 of PVO/LAO and PVO/STO. (d) Raman shift positions and FWHM depending on the substrate (i.e. lattice mismatch between PVO and the substrates). P and W represent position and FWHM, respectively.

Download Full Size | PDF

To obtain a better understanding of the high-frequency modes, we calculated the theoretical phonon modes of PVO with the first principles method using the CRYSTAL09 code. A C-type antiferromagnetic cell was assumed and a hybrid functional B1WC was chosen to reproduce the tetragonal structure and volume [24]. The calculated modes are nearly consistent with the experimental results, within 7.5%. The atom-decomposed eigenmodes are presented in Fig. 3. In particular, doublets at 950 cm−1 and a single line at 760 cm−1 are remarkable. The doublet peaks near 950 cm−1 can be assigned to V-Oa stretching modes, of which the lower- and higher-frequency peaks are anti-symmetric (p1 in Fig. 2(c)) and symmetric (p2 in Fig. 2(c)) motions of two V-sites in the C-AFM cell, respectively. Additionally, two V-Op (planar oxygen) modes are assigned near 760 cm−1 and 550 cm−1. A singlet near 650 cm−1 is assigned as a LO mode corresponding to a V-Op TO mode at near 550 cm−1 by considering some vanadium oxide compounds such as Ca0.33V2O5 [25]. For further analysis, the FWHM of the p1 and p2 peaks was considered, as shown in Fig. 2(d). The two distinguished peaks become closer to each other and overlap as the lattice mismatch increases, indicating that atomic displacement becomes larger. The broadened peaks represent the deterioration of the crystallinity, which is consistent with the mosaic spreads in rocking curves.

 figure: Fig. 3

Fig. 3 Atom-decomposed phonon eigenmodes of PVO. LO modes (in red-lines) were obtained through the non-analytic correction using the calculated anisotropic dielectric constants: εa = 5.5024 and εc = 5.6384.

Download Full Size | PDF

3.3 Linear optical characterization

SE data were analyzed using a multilayer model consisting of ambient pressure, a surface roughness layer, the PVO film, and the substrate. The optical properties of the surface roughness layer are represented by the Bruggeman effective medium approximation (BEMA) [26] using a 50-50 mix of the underlying material and void. The best results with minimum discrepancy between the experimental data and the model were obtained with BEMA layers with thicknesses of 2.6 nm for PVO/LAO and 11.7 nm for PVO/STO systems.

The complex dielectric function ε spectra for PVO layers were constructed using basis (B)-spline formulation [27]. A spline function is a series of polynomial segments constructed in a manner to maintain continuity up to a certain degree of differentiation. This set of polynomial functions can describe the optical structures in the ε2 spectrum, and the corresponding ε1 spectrum is calculated at the same time through the Kramers–Kronig transform. Experimental data (Ψ and Δ) and their best-fit curves, compared in Fig. 4(a) and 4(b), show excellent agreement. The modeled ε1 and ε2 are presented as red and black solid lines, respectively, in Fig. 4(c) and 4(d).

 figure: Fig. 4

Fig. 4 Complex dielectric function spectra (ε = ε1 + iε2) corresponding to epitaxial PVO films prepared on different substrates.

Download Full Size | PDF

The ε2 spectrum for PVO/LAO exhibits three broad spectral features near 2.5, 4.5, and 6.0 eV, which is overall in good agreement with the calculated data [14,15]. In their density-functional theory calculations, Milosevic et al. attributed the first structure at ~2.5 eV to the transition from the top of the valence band consisting mainly of a single V (3d) state to the bottom of the conduction band consisting of the rest of the V (3d) states [14]. The major broad structure spanning from 3.5 eV to 6 eV has been identified as the transition from the O (2p6) states in the valence band to the bottom of the conduction band.

There are two noticeable differences in the ε spectra between PVO/LAO and PVO/STO samples. First, the optical structures in ε2 spectrum for PVO/STO shift to lower energy with respect to the PVO/LAO data. Second, the structure at ~6 eV appears much weaker for PVO/STO. Ju and Cai calculated the optical responses of PVO as a function of the applied strain and found that the peaks in ε exhibit a systematic redshift when the system is under more tensile strain [15]. As PVO/STO experiences larger tensile strain than PVO/LAO, our observation is consistent with the theoretical predictions. Reduction of the optical structure at high energies may be a result of the inferior crystalline quality caused by a large misfit strain.

3.4 Nonlinear optical characterization

PVO was expected to have large ferroelectric polarization induced by an exotic tetragonal distortion. This polarization can be inferred by checking inversion symmetry breaking. Here, the SHG experiment is a suitable tool because the second harmonic signals can be generated only for non-centrosymmetric samples [28]. The nonlinear optical characteristics of the epitaxial PVO films were examined via SHG with an incident light source of wavelength 800 nm, injected into the sample at incident angle φ = 45° from the surface normal. Incident light is linearly polarized, and its polarization axis is continuously changed by rotating a half-wave plate by angle θ. The SHG experiment is shown in Fig. 5(c). The intensity of the s and p-polarized output SHG signal, Isout and Ipout, at wavelength 400 nm was detected as a function of polarization angle θ of incident light at room temperature. The results show that the PVO/STO thin films have a larger SHG intensity than do the PVO/LAO thin films (Fig. 5(a) and 5(b)). Thus, the PVO/STO thin films undergo harder inversion symmetry breaking than do PVO/LAO thin films. The SHG results are analyzed based on the symmetry analyses considering the 4mm point group where four independent nonzero second-order susceptibility tensor components are given as χxzx = χyzy, χxxz = χyyz, χzxx = χzyy, χzzz [11,16,29] Measured quantities of the SHG intensity Isout and Ipout are expressed as

Isout= |f̃yRPyNL|2
Ipout=|f̃xRPxNL+f̃zRPzNL|2
where

 figure: Fig. 5

Fig. 5 (a) and (b) The s- and p-polarized output SHG signals of the PVO/LAO and PVO/STO thin films, respectively.

Download Full Size | PDF

PxNL=I0sinϕincosϕin2fxfzχxxzcos2θ,
PyNL=I0sinϕin2fyfzχxxzsinθcosθ,
PzNL=I0[χzxx(fx2cos2ϕincos2θ+fy2sin2θ)+χzzzfz2sin2ϕincos2θ].

Here, I0 is the intensity of the incident fundamental light. fi and fiR with i=x,y,z are the Fresnel coefficients for fundamental and second-harmonic light, respectively, at the air–film interface. These coefficients are determined by considering the refractive index of the film and the angle of incidence φi [11,16,29]. The symmetry analysis results, plotted in Fig. 5 as solid lines, reproduce the experimental results quite well. In particular, the tensor components used for the PVO/STO thin films are almost twice as large as those of the PVO/LAO thin films; (χxzx, χzxx, χzzz) = (2.4, 15, −125) for PVO/LAO and (χxzx, χzxx, χzzz) = (4.4, 30, −300) for PVO/STO. Here, all the values are given in arbitrary units. This confirms that the symmetry breaking along the c-axis occurs more strongly for PVO/STO than for PVO/LAO, which could be attributed to the elongation of the c-axis lattice constant for the films grown on the STO substrate.

4. Conclusion

We present the effects of strain on the structural properties of PVO thin films by investigating the dynamic phonon structure, and linear and nonlinear optical properties. The doublet Raman spectra, which indicate V-Oa stretching vibration mode, are caused by octahedral rotation of VO6. An abnormal interfacial layer is induced by the lattice mismatch with the STO (001) substrate. Spectroscopic ellipsometry was used to determine parameters describing the dielectric constant and absorption coefficient of the PVO thin films on LAO and STO substrates. The ellipsometric spectra of the PVO thin films exhibit a redshift tendency with increasing misfit strain. As a result of c-lattice elongation, PVO thin films have stronger inversion symmetry breaking along c-axis, demonstrated by fitting SHG responses with nonlinear susceptibility and the Fresnel equations. This study can be a guideline for further investigation of the magnetic and ferroelectric properties in this material. Highly strained PVO films are potentially applicable to pyroelectric sensors and optoelectronic devices.

Funding

National Research Foundation (NRF) of Korea (NRF-2013H1A8A1004287, 2014R1A2A2A01004070, 2015001948, 2013R1A1A2007951; NRF-2013-Fostering Core Leaders of the Future Basic Science Program; NRF-2010-00453; 2012M3A7B4049888; 2010-0020207; 2015R1A5A1009962, 2015R1A1A1A05001560); Korea Institute of Science and Technology Information Supercomputing Center (KSC-2014-C2-032).

Acknowledgments

This research was supported by a grant from the National Research Foundation (NRF) of Korea, funded by the South Korean government (MSIP) (NRF-2013H1A8A1004287, 2014R1A2A2A01004070, 2015001948, 2013R1A1A2007951). One of the authors (SHO) was supported by an NRF grant funded by the South Korean government (NRF-2013-Fostering Core Leaders of the Future Basic Science Program). This research was also supported by Leading Foreign Research Institute Recruitment Program through an NRF grant funded by the Ministry of Science, ICT and Future Planning of Korea (NRF-2010-00453). The authors SH and JC were supported by NRF grants (Nano Material Technology Development Program 2012M3A7B4049888 and Priority Research Center Program 2010-0020207). The computation (JSA) is supported by the Korea Institute of Science and Technology Information Supercomputing Center through Contract No. KSC-2014-C2-032. This work was supported in part by the Science Research Center and the Basic Science Research Program through the NRF and funded by the Ministry of Science, ICT and Future Planning (Nos. 2015R1A5A1009962 and 2015R1A1A1A05001560).

References and links

1. L. Bellaiche and J. Iniguez, “Universal collaborative couplings between oxygen-octahedral rotations and antiferroelectric distortions in perovskites,” Phys. Rev. B 88(1), 014104 (2013). [CrossRef]  

2. J.-Q. Yan, J.-S. Zhou, and J. B. Goodenough, “Opposing spin-canting mechanism in single-crystal LuVO3 and YVO3,” Phys. Rev. B 72(9), 094412 (2005). [CrossRef]  

3. U. Lüders, Q.-R. Li, R. Feyerherm, and E. Dudzik, “The evolution of octahedral rotations of orthorhombic LaVO3 in superlattices with cubic SrVO3,” J. Phys. Chem. Solids 75(12), 1354–1360 (2014). [CrossRef]  

4. A. Okos, C. Colin, C. Darie, O. Raita, P. Bordet, and A. Pop, “Structure and magnetic properties of the layered perovskite PbVO3,” J. Al. Com. 602, 265–268 (2014). [CrossRef]  

5. K. Oka, I. Yamada, M. Azuma, S. Takeshita, K. H. Satoh, A. Koda, R. Kadono, M. Takano, and Y. Shimakawa, “Magnetic ground-state of perovskite PbVO3 with large tetragonal distortion,” Inorg. Chem. 47(16), 7355–7359 (2008). [CrossRef]   [PubMed]  

6. D. J. Singh, “Electronic structure and bond competition in the polar magnet PbVO3,” Phys. Rev. B 73(9), 094102 (2006). [CrossRef]  

7. J. B. Goodenough and J. Zhou, “Varied roles of Pb in transition-metal PbMO3 perovskites (M = Ti, V, Cr, Mn, Fe, Ni, Ru),” Sci. Technol. Adv. Mater. 16(3), 036003 (2015). [CrossRef]   [PubMed]  

8. R. V. Shpanchenko, V. V. Chernaya, A. A. Tsirlin, P. S. Chizhov, D. E. Sklovsky, E. V. Antipov, E. P. Khlybov, V. Pomjakushin, A. M. Balagurov, J. E. Medvedeva, E. E. Kaul, and C. Geibel, “Synthesis, Structure, and Properties of New Perovskite PbVO3,” Chem. Mater. 16(17), 3267–3273 (2004). [CrossRef]  

9. H. Béa, B. Dupé, S. Fusil, R. Mattana, E. Jacquet, B. Warot-Fonrose, F. Wilhelm, A. Rogalev, S. Petit, V. Cros, A. Anane, F. Petroff, K. Bouzehouane, G. Geneste, B. Dkhil, S. Lisenkov, I. Ponomareva, L. Bellaiche, M. Bibes, and A. Barthélémy, “Evidence for Room-Temperature Multiferroicity in a Compound with a Giant Axial Ratio,” Phys. Rev. Lett. 102(21), 217603 (2009). [CrossRef]   [PubMed]  

10. D. Sando, B. Xu, L. Bellaiche, and V. Nagarajan, “A multiferroic on the brink: Uncovering the nuances of strain-induced transitions in BiFeO3,” Appl. Phys. Rev. 3(1), 011106 (2016). [CrossRef]  

11. A. Kumar, L. W. Martin, S. Denev, J. B. Kortright, Y. Suzuki, R. Ramesh, and V. Gopalan, “Polar and magnetic properties of PbVO3 thin films,” Phys. Rev. B 75(6), 060101 (2007). [CrossRef]  

12. A. A. Belik, M. Azuma, T. Saito, Y. Shimakawa, and M. Takano, “Crystallographic Features and Tetragonal Phase Stability of PbVO3, a New Member of PbTiO3 Family,” Chem. Mater. 17(2), 269–273 (2005). [CrossRef]  

13. W. Zhou, D. Tan, W. Xiao, M. Song, M. Chen, X. Xiong, and J. Xu, “Structural properties of PbVO3 perovskites under hydrostatic pressure conditions up to 10.6 GPa,” J. Phys. Condens. Matter 24(43), 435403 (2012). [CrossRef]   [PubMed]  

14. A. S. Milosevic, M. V. Lalic, Z. S. Popovic, and F. R. Vukajlovic, “An ab initio study of electronic structure and optical properties of multiferroic perovskites PbVO3 and BiCoO3,” Opt. Mater. 35(10), 1765–1771 (2013). [CrossRef]  

15. S. Ju and T.-Y. Cai, “Effect of misfit strain on the multiferroism and nonlinear optical response in epitaxial PbVO3 thin films,” Appl. Phys. Lett. 93(25), 251904 (2008). [CrossRef]  

16. A. Kumar, N. J. Podraza, S. Denev, J. Li, L. W. Martin, Y.-H. Chu, R. Ramesh, R. W. Collins, and V. Gopalan, “Linear and nonlinear optical properties of multifunctional PbVO3 thin films,” Appl. Phys. Lett. 92(23), 231915 (2008). [CrossRef]  

17. C. Lefevre, R. H. Shin, J. H. Lee, S. H. Oh, F. Roulland, A. Thomasson, E. Autissier, C. Meny, W. Jo, and N. Viart, “Reduced leakage currents and possible charge carriers tuning in Mg doped Ga0.6Fe1.4O3 thin films,” Appl. Phys. Lett. 100(26), 262904 (2012). [CrossRef]  

18. L. W. Martin, Q. Zhan, Y. Suzuki, R. Ramesh, M. Chi, N. Browning, T. Mizoguchi, and J. Kreisel, “Growth and structure of PbVO3 thin films,” Appl. Phys. Lett. 90(6), 062903 (2007). [CrossRef]  

19. S. H. Oh, H.-J. Jin, H.-Y. Shin, R. H. Shin, S. Yoon, Y.-S. Seo, J.-S. Ahn, and W. Jo, “Structural properties and phase formation of epitaxial PbVO3 thin films grown on LaAlO3(001) by pulsed laser deposition,” J. Phys. D Appl. Phys. 47(24), 245302 (2014). [CrossRef]  

20. W. Zhong, R. D. King-Smith, and D. Vanderbilt, “Giant LO-TO Splittings in Perovskite Ferroelectrics,” Phys. Rev. Lett. 72(22), 3618–3621 (1994). [CrossRef]   [PubMed]  

21. A. Masuno, T. Terashima, and M. Takano, “Epitaxial growth of perovskite-type LaVO3 thin films on various substrates by the PLD method,” Solid State Ion. 172(1-4), 275–278 (2004). [CrossRef]  

22. S. Y. Yan, Y. Xie, T. Liu, and H. T. Yu, “Electronic structures and ferroelectric instabilities of cubic AVO(3) (A = Sr, Ba, and Pb) perovskites by first-principles calculations,” J. Phys. Condens. Matter 22(12), 125501 (2010). [CrossRef]   [PubMed]  

23. S. Margadonna and G. Karotsis, “High temperature orbital order melting in KCrF3 perovskite,” J. Mater. Chem. 17(19), 2013 (2007). [CrossRef]  

24. Y.-S. Seo and J. S. Ahn, “Pressure dependence of the phonon spectrum in BaTiO3 polytypes studied by ab initio calculations,” Phys. Rev. B 88(1), 014114 (2013). [CrossRef]  

25. Z. V. Popovic, M. J. Konstantinovic, V. V. Moshchalkov, M. Isobe, and Y. Ueda, “Raman scattering study of charge ordering in β-Ca0.33V2O5,” J. Phys. Condens. Matter 15(6), L139–L145 (2003). [CrossRef]  

26. G. E. Jellison Jr, L. A. Boatner, D. H. Lowndes, R. A. McKee, and M. Godbole, “Optical functions of transparent thin films of SrTiO3, BaTiO3, and SiOx determined by spectroscopic ellipsometry,” Appl. Opt. 33(25), 6053–6058 (1994). [CrossRef]   [PubMed]  

27. B. Johs and S. H. Jeffrey, “Dielectric function representation by B splines,” Phys. Status Solidi 205(4), 715–719 (2008). [CrossRef]  

28. S. Bancelin, S. Vigne, N. Hossain, M. Chaker, and F. Légaré, “Nonlinear optical properties of calcium barium niobate epitaxial thin films,” Opt. Express 24(15), 17497–17504 (2016). [CrossRef]   [PubMed]  

29. B. Dick, A. Gierulski, G. Marowsky, and G. A. Reider, “Determination of the nonlinear optical susceptibility χ(2) of surface layers by sum and difference frequency generation in reflection and transmission,” Appl. Phys. B 38(2), 107–116 (1985). [CrossRef]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (5)

Fig. 1
Fig. 1 The XRD patterns of the PVO films deposited on LAO and STO substrates. (a) θ-2θ scans and (b) φ scans. The φ scan are for the {110} planes of PVO and LAO {110}. Topographic images of (c) PVO/LAO and (d) PVO/STO. HAADF-STEM images of PVO thin films on (e) LAO and (f) STO with the fast Fourier transform patterns.
Fig. 2
Fig. 2 Raman spectra of the PVO thin films on (a) LAO and (b) STO from 100 to 1,100 cm−1. (c) Raman spectra in specific region for V-O apical vibration signal near 950 cm−1 of PVO/LAO and PVO/STO. (d) Raman shift positions and FWHM depending on the substrate (i.e. lattice mismatch between PVO and the substrates). P and W represent position and FWHM, respectively.
Fig. 3
Fig. 3 Atom-decomposed phonon eigenmodes of PVO. LO modes (in red-lines) were obtained through the non-analytic correction using the calculated anisotropic dielectric constants: εa = 5.5024 and εc = 5.6384.
Fig. 4
Fig. 4 Complex dielectric function spectra (ε = ε1 + iε2) corresponding to epitaxial PVO films prepared on different substrates.
Fig. 5
Fig. 5 (a) and (b) The s- and p-polarized output SHG signals of the PVO/LAO and PVO/STO thin films, respectively.

Equations (5)

Equations on this page are rendered with MathJax. Learn more.

I s out =  | f ̃ y R P y NL | 2
I p out = | f ̃ x R P x NL + f ̃ z R P z NL | 2
P x NL = I 0 sin ϕ in cos ϕ in 2 f x f z χ xxz cos2θ,
P y NL = I 0 sin ϕ in 2 f y f z χ xxz sinθcosθ,
P z NL = I 0 [ χ zxx ( f x 2 cos2 ϕ in cos2θ+ f y 2 sin2θ ) + χ zzz f z 2 sin2 ϕ in cos2θ ].
Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.