Phase transition, interband electronic transitions and enhanced ferroelectric properties in Mn and Sm co-doped bismuth ferrite films

Yalong Liua, Jie Wei*ab, Yaxin Guoa, Tiantian Yanga and Zuo Xuab
aElectronic Materials Research Laboratory, Key Laboratory of Ministry of Education, Xi'an Jiaotong University, Xi'an 710049, P. R. China. E-mail: weij2008@xjtu.edu.cn; jiewei2013wj@gmail.com
bInternational Center for Dielectric Research, Xi'an Jiaotong University, Xi'an 710049, P. R. China

Received 17th August 2016 , Accepted 26th September 2016

First published on 26th September 2016


Abstract

Mn and Sm co-doped BiFeO3 (BFO) polycrystalline films have been fabricated on F-doped SnO2 (FTO) conductive glass substrate by a chemical solution deposition method. X-ray diffraction patterns and Raman spectra both imply a coexistence of rhombohedral and tetragonal phases in Mn–Sm co-doped BFO films. Scanning Electron Microscopy (SEM) images indicate more impact morphology with greater particle size in doped films. X-ray photoelectron spectroscopy (XPS) analysis suggests that high-valence Mn ions induced by Sm-doping lead to efficient decrease in oxygen vacancies, and thus improve leakage and polarization properties. Optical spectra analysis by using a Kubelka–Munk function and corresponding XPS valence band spectra analysis reveal that the optical band gap (Eg, varied from 2.38 eV to 2.26 eV) and Fermi level (EF, varied from 1.18 eV to 1.01 eV) decrease continually with the introduction of Sm dopants, further confirming the decreased defects and vacancies in layer interfaces or grain boundaries. Consequently, larger saturation polarization and lower leakage current densities were obtained in Mn–Sm co-doped BFO films. Furthermore, possible energy band structure schematic proposed on the basis of above analysis shows that the shift in Urbach impurities energy level induced by ions dopants outsiders may be quite crucial in future photovoltaic application.


1. Introduction

BiFeO3 (BFO) has been widely investigated for decades both in fundamental physics and promising application due to its intriguingly large spontaneous polarization, magnetic–electric coupling and anomalous photovoltaic effect, which also make it a potential candidate for future piezoelectric detectors or sensors, nonvolatile lead-free memory storage and novel photovoltaic devices.1–3 For decades, great effort has been made such as structure engineering and ions doping to optimize both crystal structure and multiferroic properties of BFO.4,5 Monocrystalline BFO films with admirable magnetoelectric properties have been deposited successfully by pulse laser deposition,6,7 while such high costs in film preparation impede its industrialization. Therefore, much attention has been transferred to the polycrystalline films by chemical solution deposition, attempting to change its microstructure and improve its multiferroic or optical properties.8–10

Furthermore, various research has been made to investigate both fundamental physics and practical applications of BFO polycrystalline thin films with site-engineering method because of the poor ferroelectric properties observed in pure BFO thin films due to the instability of phase structure and volatility of Bi element, while high performance in electrical properties including both leakage behaviour and ferroelectric properties have been confirmed in B-site Mn ions doped BFO thin films and A-site Sm ions doped BFO thin films respectively, probably resulting from phase structure transition, less defects and compress strains induced by their smaller ions radius than B-site Fe ions and A-site Bi ions respectively.11–14 In our previous work, BFO films with 7.5% Mn ions replacement show the glaring ferroelectric remnant polarization in agreement with the reported 7% by S. K. Singh et al. in ref. 8. Moreover, Xu's research confirms the further non-centrosymmetry distortion by varying iron-oxygen octahedral and long-range ferroelectric order with Sm doping.15 Thus, we facilitate the Bi1−xSmxFe0.925Mn0.075O3 (BSFM, x = 0, 0.025, 0.05, 0.075, and 0.10) films on F-doped SnO2/glass (FTO) substrates by a solution deposition process and expect to obtain further improvement in its electric properties.

However, there still remains some controversies in the investigation of polycrystalline BFO films, including (I) ions doping leading to the microstructure transition: pure BFO crystal structure has been repeatedly considered as nearly-tetragonal phase structure or mixed tetragonal and rhombohedral phase due to its lattice mismatch with the substrate, the existence of morphotropic phase boundary and multilayer interface strain which is sensitive to film thickness.16 Few papers focused on the ions doping structure transition for it's a more puzzling and complicated process as it may introduce new unexpected impurities effect; (II) electronic energy band structure or inter-band electronic transitions: local electron orbits interactive hybridization and its coupling with certain electrons' high-spinning behavior are quite uncertain,17 while ions dopants outsiders probably result in the shifting of impurities energy lever or inter-band electronic transition energies and it's still unclear whether this behavior is favorable to practical applications or not.

All these controversies as the springboard of this paper inspire us to investigate the interplay between the crystalline structure, inter-band electronic transitions and physical properties in Mn–Sm co-doped BFO films.

2. Experimental

Pure BiFeO3 and Bi1−xSmxFe0.925Mn0.075O3 (BSFM, x = 0, 0.025, 0.05, 0.075, and 0.10) films were deposited on F-doped SnO2/glass substrates by a solution deposition process.18 X-ray diffraction (XRD, X'pert PRO) and laser Raman spectrometer (HR800) were used to investigate the crystalline structure and phase transition in BSFM films. The morphology and thickness of these films were observed by scanning electron microscope (SEM, Quanta F250). The ferroelectric properties were measured by using an Aixacct TF Analyzer 2000, and the leakage current densities were characterized by Keithley 4200-SCS. The ionic valences were collected by X-ray photoelectron spectroscopy (XPS, AXIS ULtrabland). The room temperature Ultra-Violet-Visible (UV-Vis) reflectance spectra were measured by using a PerkinElmer spectrometer (Lambda 950).

3. Results and discussion

3.1 XRD patterns

Fig. 1 shows the XRD patterns for pure and Mn–Sm co-doped BFO films, revealing pure phases without any secondary phase in all the samples. All the diffraction peaks in the XRD patterns of pure BFO film can be easily indexed as a rhombohedral and distortion perovskite structure with a space group of R3c (PDF# 86-1518), evidenced by the noticeable doublet in around 32°, extra peak in 37° and splitting of the peak around 67°. However, none of these features maintains in the BSFM films thus suggesting an occurrence of incomplete structure transition possibly from rhombohedral phase (R-phase) to nearly-tetragonal phase (T-phase P4mm, referring to PDF#74-2459).19 Based on our previous work of Mn doping BFO films, a slight phase transition from rhombohedral phase to nearly-tetragonal phase has been confirmed and explained well due to the smaller radius replacement of B-site Fe ions. Furthermore, the smaller A-site Bi radius modification by Sm ions will possibly intensify the asymmetric distortion, so we mainly focused on the further modification of Sm ions to the microstructure of polycrystalline BFO films.
image file: c6ra20740e-f1.tif
Fig. 1 XRD patterns of the pure BFO and BSFMx (x = 0.025, 0.05, 0.075) films deposed on FTO/glass substrate by sol–gel process. Note that PDF#86-1518 card and PDF#74-2459 are used to clarify the phase transition from R3c space group to P4mm space group.

The elaborate fits of 32° peak for pure BFO and BSFM films are shown in Fig. 2. The hints in doublet fusion and broadening can be extracted easily from the peak-fitting–simulating parameters summarized in Table 1. These results further confirm the phase transition from R-phase to nearly T-phase, because T-phase (110) peak strengthens while R-phase (104) peak weakens shown by their increased peak area ratio from 0.67 to 1.53 and decreased interval from 0.274 to 0.248. Furthermore, these peaks broadening and fusion behaviour is probably derived from the introduction of Mn ions,20 as well as the addition of Sm ions.


image file: c6ra20740e-f2.tif
Fig. 2 XRD-peak-fitting–simulating analysis of 32° main peak including (104) and (110) sub-peaks of BFO and BSFMx (x = 0.025, 0.05, 0.075) films. Note that the green line in the left represents for the (104) sub-peak from R-phase while blue line for the (110) sub-peak from T-phase.
Table 1 Unit cell parameters of BFO and BSFMx (x = 0.025, 0.05, 0.075) films extracted from the refinements, peak-fitting–simulating analysis and calculation
Sample category ac (Å) cc (Å) cc/ac Local inhomogeneous distortion (ccac)/cc (110) & (104) peak interval (θ) (110) & (104) peak area ratio
Pure BFO 3.943(3) 3.947(4) 1.0010(1) 0.10(3)% 0.274 0.69
BSFM0.025 3.947(6) 3.955(6) 1.0020(2) 0.20(2)% 0.273 1.35
BSFM0.05 3.947(7) 3.955(8) 1.0020(5) 0.20(4)% 0.248 1.53
BSFM0.075 3.947(8) 3.957(9) 1.0025(6) 0.25(5)% 0.266 1.13


In order to establish the correlation between phase structure transition and polarization property, we carried out the Rietveld refinements for XRD data. The results and analysis reveal that the cubic-like lattice parameter ac and cc were affected by Sm doping concentration. For example, cc/ac ratio and local inhomogeneous distortion vary from 1.0010 to 1.0025 and from 0.10% to 0.25%, respectively. Therefore, the ferroelectric polarization that is directly proportional to the square-root of the distortion is believed to present significantly enhancement, which is confirmed in the later ferroelectric measurements.21 Above all, a complicated mixed phase structure including both R3c group and P4mm group was obtained in all Mn and Sm co-doped films, which is consequently confirmed by the laser Raman spectra in the next section. Simultaneously, this mixed phase structures will probably induce a more complicated electronic band structure and tempting polarization or other physical performance.22

3.2 Raman spectra

Further analysis of micro-structure transition was carried out by Raman spectra with 631 nm laser excitation. Standard rhombohedral R3c structure of pure BFO film can be certificated from Fig. 3 due to the appearance of 13 typical Raman active modes (4A1 + 9E) corresponding to R3c. However, only 8 Raman models can be fitted in BSFM films, which looks like much closer to P4mm structure (3A1 + B + 4E).23 In order to obtain the exact peak position, each measured spectrum for BSFM films was properly fitted by decomposing the fitted curves into individual Lorentzian components (Raman spectra of pure BFO film and BSFM (x = 0.05) film are selected to clearly exhibit the fitting procedure, as shown in Fig. 3). Corresponding Raman shift values were summarized in the Table 2. Obviously, the quantitative values of pure BFO film significantly turn in a consistent performance with the R3c group in ref. 24, while BSFM films show dramatic changes in amount, position, and intensity of the corresponding peaks.24,25 It should be mentioned here that the accuracy of the equipment that limits the veracity of the ultra-low wavenumber measurement below 100 cm−1 and this part of fuzziness should not affect the whole variation tendency between 100 cm−1 to 800 cm−1.
image file: c6ra20740e-f3.tif
Fig. 3 Raman spectra and Raman-peak-fitting–simulating analysis of the BFO and BSFMx (x = 0.05) films under excitation 631 nm at room temperature. The red dashed lines show the positions of Raman-active photon modes of BSFMx (x = 0.05). Note that blank lines are the raw intensity and colored lines are the fitting components.
Table 2 Raman active modes of R3c and P4mm structure from references and of as-prepared BFO and BSFMx (x = 0.025, 0.05, 0.075) films
R3c structure Raman active models
E-1 A1-1 A1-2 A1-3 E-2 E-3 E-5 E-6 E-7 E-8 E-9 B (P4mm structure)
R3c[thin space (1/6-em)]:[thin space (1/6-em)]H (ref. 18) 76.0 140 172 214 262 275 345 369 470 521 613  
Pure BFO 75.9 141 172.5 219.5 262.2 273.7 347.9 370.8 470.6 521.0 611.8  
BSFM0.025 73.3 135.1 169.2   247.8     379.6 479.7 543.0   628.9
BSFM0.050 73.5 134.6 169.6   247.5     382.2 479.6 542.3   629.0
BSFM0.075 73.8 134.3 169.4   249.0     385.7 481.5 542.9   628.5

P4mm structure Raman active models
A1-1 A1-3 E-2 E-3 E-4 B
P4mm (ref. 19) 147 227 266 273 369 691


The first-principles calculation26 suggests that Bi atoms only participated in low-frequency modes up to 167 cm−1 and the motion of oxygen atoms dominated the modes above 262 cm−1, whereas Fe atoms were mainly involved in modes between 152 and 262 cm−1 with possible contribution to higher frequency modes. Correspondingly, the two characteristic modes of A1-1 mode at ∼140 cm−1 and E mode at ∼75 cm−1 should be governed by Bi–O covalent bonds, which control the dielectric constant and the ferroelectric phase of BiFeO3.27,28 As seen in Fig. 3, E mode at ∼75 cm−1 presents dramatic increase in relative intensity for all BSFM films compared to that of pure BFO, suggesting a significant change in ferroelectric properties. Especially, a B-mode-like peak appears at the position of about 628 cm−1, as well as some E modes of high frequency above 350 cm−1 (asterisked E peak in Raman spectra, while the asterisked A1-3 mode is probably merged with E models near 250 cm−1) disappeared and the rest Raman modes undergo high frequency shifting, strongly implying the coexistence of R3c and P4mm phases as analysed above in XRD patterns. More details about transformation in electronic band structures will be discussed deeply in subsequent XPS analysis and reflectance spectra sections.

3.3 Morphology

In Fig. 4, BSFM films with an approximately thickness of ∼300 nm were well deposited on FTO glass substrate, having distinct interfaces between different layers as shown in SEM cross-section image. The thickness of BFO ultra-thin BFO films below 100 nm is directly proportional to the ferroelectric properties which results from the interface strain due to the crystal mismatch while the enhanced polarization properties of our polycrystalline specimen are closely related with the grain size, grain distribution and porosity due to their similar thickness around 300 nm with the error of 3%. It's easy to control the thickness of the BFO film above 300 nm by sol–gel coating method and this factor may not be the dominant one in our so-called “thick” films due to their similar section morphology around 300 nm.29
image file: c6ra20740e-f4.tif
Fig. 4 SEM images of BFO film and BSFMx (x = 0.025, 0.05, 0.075) films and cross-section structure of BSFM x = 0.05 film. Note that we obtain a same thickness of ∼300 nm in all films prepared at the same condition.

Since grain boundary and size are of a great significance in ferroelectric properties and leakage conduction, pure BFO film shows smaller grain size ranging from 80–180 nm with numbers cracks or pores which may easily form the traps of charge at crystal interface and boundaries, thus causing the undesirable ferroelectric properties in practical application.30 A small quantity of Mn dopants may expand the particles and improve the leakage property as we reported before in ref. 18. Actually, average grain size doubled to be over 400 nm with 2.5% Sm-doping concentration and extensive small grain appeared to merge with its surroundings, then this grain merging phenomenon persistently happened until almost all grains came to be contiguous over several micrometres like a plane as Sm doping concentration came to be over 7.5% doping concentration. Although BSFM (x = 0.025, 0.05) film shows some pinholes and cracks but it has a better ferroelectric and leakage property mainly for the grain merging resulted in the bigger but less cracks and pinholes than pure BFO film which means the less defects at the boundary. On the other hand, the less but lager grains signify the better uniformity and orientation on account of the less grain boundary in polycrystalline BFO films based on the less oxygen defects by Sm-doping. However, a large number of new small grains with the grain size below 100 nm gradually grew and wrapped at the large particles boundaries in the surface of BSFM x = 0.075 film and somehow lower its polarization property. It should be mentioned that grain size and surface defects probably play an important role in leakage and optical behavior as discussed later.

3.4 Leakage properties

Fig. 5 show the leakage current densities at the applied electric field from −333 kV cm−1 to 333 kV cm−1, of which BSFM films lower the leakage current density about one order of magnitude than that of pure BFO. It should be attributed to the limitation to oxygen defects by non-volatile Sm ions and high-valence Mn4+ ions on basis of the possible chemical equation as follow,31
image file: c6ra20740e-t1.tif

image file: c6ra20740e-t2.tif

image file: c6ra20740e-t3.tif

image file: c6ra20740e-f5.tif
Fig. 5 Leakage current density J vs. electric field E of pure BFO and BSFMx (x = 0.025, 0.05, 0.075) films measured at room temperature.

As seen in the above equation, oxygen defects were easily generated due to the Bi volatilization during the preparation of pure BFO. The introduction of non-volatile Sm will combine oxygen ions tightly to form Sm–O bond, and thus restrain the production of oxygen defects. Furthermore, the introduction of Mn2+ ions may probably react with oxidizing oxygen vacancy and consequently lower the density of oxygen defects by generating Mn4+ ions. In other words, more O-bonds and high valence Mn ions (Mn4+) will decrease the leakage current density of BSFM films compared to pure BFO. What's more, all these process can be demonstrated in the XPS analysis of O ions and Mn ions section later.

3.5 Ferroelectric properties

In Fig. 6, ferroelectric measurements were carried out at the frequency of 1 kHz at room temperature to obtain hysteresis loops and polarization switching current of BFO, BFM0.075 (BiFe0.925Mn0.075O3) film and BSFM (x = 0.025, 0.05, 0.075) films. During the ferroelectric measurement, we gradually increased the applied electric field to the critical breakdown value to obtain the maximum withstand capability. So the ferroelectric hysteresis loops were displayed at their respective maximum field corresponding to the varied resistance character. Saturation polarization values of ∼39, 62, 73, 52 μC cm−2 were observed (see Table 3) with increasing Sm dopants at the applied field within 333 kV cm−1, while pure BFO performed quite poor ferroelectric behavior due to its severe leakage behavior32 induced by oxygen defects and abominable morphology. It is clear that the best ferroelectric performance was obtained in the 5% Sm doped BFO film. Moreover, obvious polarization switching behavior shown in Fig. 6(b) was observed in all the BSFM (x = 0.025, 0.05, 0.075) films with the slight increase of the peak current densities. From the leakage analysis in section 3.4, BSFM x = 0.05 film shows the similar leakage properties with the other BSFM films, and it is also demonstrated by the similar diagonal slope of polarization switching behavior in Fig. 6(b), revealing the lager Pr value of BSFM x = 0.05 film mainly results from the different effects of the varied Sm doping concentration, including the change of grain sizes, oxygen defects densities and crystal distortion rather than leakage current compensation. Otherwise, the good performance in ferroelectric properties should be benefited from the lager grain size and the incomplete phase structure transition from R-phase to T-phase by introducing a small quantity of Sm dopants. We can reasonably assume from the analysis above that the crystal distortion with lager cc/ac ratio will probably cause large polarization due to enhanced polarization in our near-tetragonal phase specimen the reported lager cc/ac ratio of ∼1.25 in tetragonal-phase BFO structure.16 Nevertheless, excess Sm doping concentration (over 7.5% in this case) will probably suppress polarization switching process to some content because its particular pinning effect becomes dominant.33 Therefore, enhanced polarization properties have been accessed by Mn and Sm co-doping in this case, forecasting a good future application in low-cost solution-deposited BFO films.
image file: c6ra20740e-f6.tif
Fig. 6 Ferroelectric hysteresis loops (a) and current density hysteresis loops (b) of BFO, BFM0.075 (BiFe0.075Mn0.925O3) and BSFMx (x = 0.025, 0.05, 0.075) films measured at the applied alternating electric field of 1 kHz at room temperature.
Table 3 Key parameters of ferroelectric polarizations for pure BFO, BFM0.075 (BiFe0.075Mn0.925O3) and BSFMx (x = 0.025, 0.05, 0.075) films, their polarization reverse electric fields and corresponding cc/ac value
Sample category PS (μC cm−2) Pr (μC cm−2) EC (kV cm−1) EJ max (kV cm−1) cc/ac (arb. units)
BFO ∼0 ∼20 1.0010(1)
BFM 39 35 106 74
BSFM0.025 62 53 130 83 1.0020(2)
BSFM0.05 73 64 187 105 1.0020(5)
BSFM0.075 52 40 115 67 1.0025(6)


3.6 XPS analysis

In order to explore electronic structure evolution and co-doping investigated by XPS spectroscopy. Fig. 7 and 8 show further analysis of O 1s and Mn 2p3/2 orbits of BFO film and BSFM (x = 0.025, 0.05, 0.075) films and the corresponding peak-fitting parameters have been concluded in Tables 4 and 5, respectively. Firstly, the fitting result of Fe 2p3/2 orbits display singlet look (not shown here), implying only Fe3+ ions exist in our BFO film and BSFM films, and hence some other factors must be dominant in the leakage mechanism like Bi3+ ions volatilization or oxygen vacancy generation in film preparing process.34 As shown in Fig. 7, the O 1s main peak at ∼529.5 eV at lower binding energy are associated with the O–X bonds (X represents for Bi, Sm, Mn, and Fe), while another one at higher binding energy of ∼531.5 eV are associated with the O defects or vacancies. As seen in Table 4, it is easy to clarify by the increased proportions of O–X bonds from 2.32 to 6.60 in the whole O 1s peaks, implying the formation of more non-volatilize stable O–Sm bonds suppress the generation of the O vacancies serving as the channel for electrons migration, simultaneously leading to the band energy and electronic structure transition.35
image file: c6ra20740e-f7.tif
Fig. 7 O 1s orbits X-ray photoelectron spectra of BFO and BSFMx (x = 0.025, 0.05, 0.075) films and XPS-peak-fitting–simulating analysis of O defect peak (green line) at ∼529.2 eV and O–X bond peak (blue line) at ∼531.4 eV.

image file: c6ra20740e-f8.tif
Fig. 8 Mn 2p3/2 orbit X-ray photoelectron spectra of BSFMx (x = 0.025, 0.05, 0.075) films and XPS-peak-fitting–simulating analysis of Mn2+ peak (green line) at ∼641.2 eV and Mn4+ peak (blue line) at ∼531.4 eV.
Table 4 Primary parameters of XPS-peak-fitting–simulating analysis for O 1s orbits of BFO and BSFMx (x = 0.025, 0.05, 0.075) films
O 1s orbit Chemical bonding Binding energy (eV) Peak area (arb. units) Peak area ratio
Pure BFO O–X bond 529.3 60[thin space (1/6-em)]833 2.32
Defect O 531.5 26[thin space (1/6-em)]259
BSFM0.025 O–X bond 529.2 127[thin space (1/6-em)]045 3.76
Defect O 531.3 33[thin space (1/6-em)]809
BSFM0.05 O–X bond 529.2 105[thin space (1/6-em)]578 4.58
Defect O 531.4 23[thin space (1/6-em)]061
BSFM0.075 O–X bond 529.3 96[thin space (1/6-em)]380 6.60
Defect O 531.4 14[thin space (1/6-em)]616


Table 5 Main parameters of XPS-peak-fitting–simulating analysis for Mn 2p3/2 orbits of BSFMx (x = 0.025, 0.05, 0.075) films
Mn 2p3/2 orbit Chemical valence Binding energy (eV) Peak area (arb. units) Peak area ratio
BSFM0.025 +4 642.7 4409 1.23
+2 641.1 3575  
BSFM0.05 +4 642.7 4460 1.26
+2 641.1 3529  
BSFM0.075 +4 642.3 3202 1.57
+2 641.3 2029  


However, the dividing peaks in Mn 2p3/2 XPS spectra corresponding to ∼642.7 eV and ∼641.3 eV peaks display the variable valences of Mn ions considered as Mn4+ and Mn2+ ions respectively. As seen in Table 5, their area ratio varied from 0.24 to 1.57 shows that more Mn4+ ions appeared and less oxygen vacancies induced with the introduction of Sm dopants, confirming the hypothesis of chemical equations above. Therefore, improved ferroelectric polarization properties may be originated from synergistic effects including the following facts, such as greater lattice distortion, phase transition from rhombohedral to tetragonal phase, less crystal interfaces and smooth morphology, greater grain size, less oxygen vacancies from the limitation of Sm ions, high-valence Mn ions and lower leakage behavior.

3.7 Optical properties

Optical spectroscopy is a widely used to probe electronic band structure in solids, especially it could reveal rich physics in complex oxides. More recently, it has attracted much attention to study the electronic band structure and in-gap states in BiFeO3 compound, since it will reveal the interplay between charge, structure and physical properties in this prototype multiferroic system.

The reflectance spectra collected from UV-Vis spectroscopy was recorded over the range from 200–1000 nm in comparison with the Kubelka–Munk (K–M) models below,35,36

image file: c6ra20740e-t4.tif
where R represents for the reflectivity. Reflectance spectra other than transmittance spectra was used here to derive the band gap due to the some pollutant attached to the back of the glass substrate during the coating process, which may affect the veracity of transmittance data. The direct band gap approach of BFO and BSFM (x = 0.025, 0.05, 0.075) films could be estimated to be 2.38 eV, 2.29 eV, 2.28 eV and 2.26 eV respectively, by extrapolating (F(R) × E)2 to zero of photon energy (in Fig. 9 inset (b)) and the highly noticeable variation of an absorption edge at around 428 nm for all specimens is corresponding to the charge-transition band at ∼2.9 eV (CT band shown in Fig. 9), indicating light absorption within visible range.


image file: c6ra20740e-f9.tif
Fig. 9 Kubelka–Munk functions, F(R) vs. photon energy of BFO and BSFMx (x = 0.025, 0.05, 0.075) films including p–d charge-transfer at ∼2.9 eV and d–d band absorption at ∼2.45 eV. Inset (a) shows two other absorption onsets in low energy band attributed to 6A1g4T1g and 6A1g4T2g crystal-field transitions, respectively. Inset (b) is the energy band gap linearly extrapolated from the low-energy part of (F(R) × )2 to energy axis. Inset (c) displays Urbach energy of BFO and BSFMx (x = 0.025, 0.05, 0.075) by calculating the slope of reciprocal extrapolated from liner part of ln(F(R)) vs. photon energy in low energy side. Note that reflectance (R) is derived from room-temperature UV-Vis reflectance spectra.

In fact, the interatomic transition and coupling effect37 between electron charge (O 2p, Fe 3d, Bi 6s, Bi 5p levels) and absorption bands (d–d bands transition) due to the electron spin of Fe3+ (3d5 high-spin behavior) local lattice field in BiFeO3 system make its local electronic band structure more complicated like the schematic shown in Fig. 10. By considering the C3v local symmetry of Fe3+ ions in BiFeO3 and using the correlation between group and subgroup analysis for the symmetry breaking from Oh to C3v, the typical t32ge2g high-spin configuration of Fe3+ ions in the cubic octahedral environment transforms into an a11e2e2 electronic configuration in the rhombohedral environment.21,38


image file: c6ra20740e-f10.tif
Fig. 10 Schematic energy splitting diagram of Fe3+-d5 orbits and further t2g3 orbits indicating the crystal-field effect with lower symmetry from cubic octahedral Oh to rhombohedral C3V environment.

Despite the intense peak for the charge-transition band, two extra absorption onsets at low energy below 2.0 eV corresponding to 6A1g4T1g and 6A1g4T2g crystal-field d–d transitions have attract much attention because they are the corresponding splitting level. It is believed that these two crystal-field transitions of Fe3+ (3d5) are highly sensitive to slight distortions of octahedron FeO6. Both the transition energy and oscillator strength strongly depend on the Fe–O bond length, site symmetry and Fe–O–Fe exchange interaction.39 The foregoing XRD and Raman results reveal an obvious occurrence of phase transition and distortions in FeO6 octahedron induced by Sm–Mn co-doping. Apparently, these changes should be detected by optical spectroscopy.

As shown in inset (a) of Fig. 9, it's easy to observe two extra absorption onsets at 1.47, 1.39, 1.43, 1.39 eV and 1.77, 1.64, 1.70, 1.65 eV for BFO and BSFM films (x = 0.025, 0.05, 0.075) respectively, corresponding to 6A1g4T1g and 6A1g4T2g crystal-field transitions. More interestingly, both two transitions energies exhibit apparent red-shift behavior in all doped samples. Based on the modified Tanabe–Sugano diagrams,40 this red-shift in the two crystal-field transition energies should be attributed to increase in the e–t2 crystal-field splitting Δ, induced by the smaller size Sm ions substitution of Bi-site in BiFeO3. Furthermore, such a red-shift behavior reflect significant change in the local structure of Fe3+ in BiFeO3 system, which indeed affect its physical properties.

Fermi level (EF) of BFO and BSFM (x = 0.05) films can be decided by XPS valence spectra from Fig. 11(a) and (b). It could be observed that EF is lower than the middle level of the band gap in BSFM (x = 0.05) film, because the dominating Mn4+ ions are kind of accepter dopants. Urbach energy EU was determined through ln(F(R)) = /EU relationship, varying from 0.1397, 0.1192, 0.1190, 0.2060 eV (from Fig. 9 inset (c)) below the conduction band (EC) bottom of BFO and BSFM films (x = 0.025, 0.05, 0.075) respectively. It is known that Urbach energy EU is closely related with the defects and strains.41 The lower Urbach energy value means less oxygen defects and impurities as the unexpected absorption centres existed in BSFM films (except BSFM x = 0.075), which is in good agreement with the foregoing XRD, Raman and SEM analysis. Based on above results, the electronic band structures for BFO and BSFM (x = 0.05) films were schematically proposed in Fig. 11(c) and (d) and their corresponding parameters have been listed in Table 6.


image file: c6ra20740e-f11.tif
Fig. 11 X-ray photoelectron spectra of valence band of BFO (a) and BSFMx (x = 0.05) film (b). The intersection plot of two linear fitting parts in their low-energy region are the Fermi-level energy of ∼1.18 eV and ∼1.01 eV respectively. Based on the reflectance spectra and XPS of valence band analysis, splitting energy levels of 6A1g4T1g and 6A1g4T2g transitions and impurities energy level EU are schematically proposed as electronic energy band structure diagram (c) and (d).
Table 6 Corresponding parameters of local absorption level 4T2g, splitting energy level 6A1g4T1g and 6A1g4T2g, absorption onset energy Eg and Urbach energy EU for BFO and BSFMx (x = 0.05) films respectively
Sample 4T2g (eV) 6A1g4T1g (eV) 6A1g4T2g (eV) Eg (eV) EF (eV) EU (eV)
Pure BFO 2.88 1.47 1.77 2.38 1.18 0.139(7)
BSFM0.05 2.88 1.43 1.70 2.29 1.01 0.119(0)


Briefly, Sm and Mn co-doped in BiFeO3 induced remarkable distortions in FeO6 and an incomplete phase transition evidenced by XRD and Raman spectra analysis, and thus led to significant changes in the local structure of Fe3+, which could be clearly reflected by a red-shift behavior in the crystal-field transition energies. Moreover, the lower Urbach energy value means less oxygen defects and impurities as the unexpected absorption centres existed in BSFM films. From a macroscopic point of view, these microstructures changes ultimately resulted in the improved leakage properties and the enhanced ferroelectric properties.

4. Conclusions

Pure BFO and BSFM films have been well fabricated on FTO conductive glass substrate by a solution deposition method. Microstructure analysis in XRD patterns and Raman spectra imply a coexistence of rhombohedral and tetragonal phases. It's clear to observe in XRD patterns that all multimodal peaks show a merging behaviour including R-mode (R, rhombohedral) peaks weakens and T-mode (T, tetragonal) peaks strengthen process. Moreover, amount of Raman active modes reducing from 13 (4A1 + 9E) to 8 modes (3A1 + B + 4E) confirms the incomplete structure phase transition. Scanning electron microscope images indicate more impact morphology by Sm-doping with greater particle size. Higher saturation polarization and lower leakage current density were obtained through the introduction of Sm and Mn. XPS analysis suggests that Sm-doping and high-valence Mn ions transition are beneficial to decrease oxygen vacancy density and improve polarization or leakage properties. Note that the direct optical band gap (Eg, from 2.38 eV to 2.26 eV) and Fermi level (EF, from 1.18 eV to 1.01 eV) decrease with increasing Sm dopants from UV-Vis reflectance spectra by using Kubelka–Munk functions and corresponding XPS valence band spectra, confirming the decreased defects or vacancies in layer interfaces or grain boundaries by Sm modification. Subsequent energy band structure schematic was proposed with Urbach impurities energy level shifting from 0.14 eV to 0.12 eV below the conduction band bottom, indicating less absorption and recombination centres in doped BiFeO3 film, which may be quite crucial in future photovoltaic research.

Acknowledgements

Financial support by the National Natural Science Foundation of China (Grant No. 51272204) is gratefully acknowledged. J. Wei wants to thank the China Scholarship Council (CSC) for funding his stay in France. The authors also thank Ms Dai and Mr Ma for their help in SEM and EDS analysis at International Centre for Dielectric Research (ICDR), Xi'an Jiaotong University, Xi'an, China.

Notes and references

  1. R. Ramesh and E. Al, et al., Science, 2003, 299, 1719–1722 CrossRef PubMed .
  2. T. J. Park, S. Sambasivan, D. A. Fischer, W. S. Yoon, J. A. Misewich and S. S. Wong, J. Phys. Chem. C, 2008, 112, 10359–10369 CAS .
  3. J. Wu, J. Wang, D. Xiao and J. Zhu, ACS Appl. Mater. Interfaces, 2011, 3, 3261–3263 CAS .
  4. P. Suresh and S. Srinath, J. Appl. Phys., 2013, 113, 17D920 CrossRef .
  5. T. Kawae, Y. Terauchi, H. Tsuda, M. Kumeda and A. Morimoto, Appl. Phys. Lett., 2009, 94, 112904 CrossRef .
  6. K. Y. Yun, D. Ricinschi, T. Kanashima, M. Noda and M. Okuyama, Jpn. J. Appl. Phys., 2004, 43, 647 CrossRef .
  7. S. Y. Yang, F. Zavaliche, L. Mohaddes-Ardabili and V. Vaithyanathan, Appl. Phys. Lett., 2005, 87, 102903 CrossRef .
  8. S. K. Singh, H. Ishiwara, K. Sato and K. Maruyama, J. Appl. Phys., 2007, 102, 094109 CrossRef .
  9. S. K. Singh, H. Ishiwara and K. Maruyama, Appl. Phys. Lett., 2006, 88, 262908 CrossRef .
  10. Y. Guo, B. Guo, W. Dong, H. Li and H. Liu, Nanotechnology, 2013, 24, 86–94 Search PubMed .
  11. S. Singh, C. Tomy, T. Era, M. Itoh and H. Ishiwara, J. Appl. Phys., 2012, 111, 102801 CrossRef .
  12. J. Luo, S. Lin, Y. Zheng and B. Wang, Appl. Phys. Lett., 2012, 101, 062902 CrossRef .
  13. M. Kumar and K. Yadav, Appl. Phys. Lett., 2007, 91, 242901 CrossRef .
  14. C.-J. Cheng, D. Kan, S.-H. Lim, W. McKenzie, P. Munroe, L. Salamanca-Riba, R. Withers, I. Takeuchi and V. Nagarajan, Phys. Rev. B: Condens. Matter Mater. Phys., 2009, 80, 014109 CrossRef .
  15. X. Xu, T. Guoqiang, R. Huijun and X. Ao, Ceram. Int., 2013, 39, 6223–6228 CrossRef CAS .
  16. D. Mazumdar, V. Shelke, M. Iliev, S. Jesse, A. Kumar, S. V. Kalinin, A. P. Baddorf and A. Gupta, Nano Lett., 2010, 10, 2555–2561 CrossRef CAS PubMed .
  17. P. S. V. Mocherla, C. Karthik, R. Ubic, M. S. Ramachandra Rao and C. Sudakar, Appl. Phys. Lett., 2013, 103, 022910–022915 CrossRef .
  18. Y. Liu, J. Wei, Y. Liu, X. Bai, P. Shi, S. Mao, X. Zhang, C. Li and B. Dkhil, J. Mater. Sci.: Mater. Electron., 2016, 27, 3095–3102 CrossRef CAS .
  19. Z. Wen, X. Shen, D. Wu, Q. Xu, J. Wang and A. Li, Solid State Commun., 2010, 150, 2081–2084 CrossRef CAS .
  20. J. Z. Huang, Y. Shen, M. Li and C. W. Nan, J. Appl. Phys., 2011, 110, 094106 CrossRef .
  21. X. Bai, J. Wei, B. Tian, Y. Liu, T. Reiss, N. Guiblin, P. Gemeiner, B. Dkhil and I. C. Infante, J. Phys. Chem. C, 2016, 120, 3595–3601 CAS .
  22. M. K. Singh, S. Ryu and H. M. Jang, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 2101 Search PubMed .
  23. H. M. Tütüncü and G. P. Srivastava, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 78, 1879–1882 CrossRef .
  24. Y. Yang, J. Y. Sun, K. Zhu, Y. L. Liu and L. Wan, J. Appl. Phys., 2008, 103, 093532–093535 CrossRef .
  25. M. N. Iliev, M. V. Abrashev, D. Mazumdar, V. Shelke and A. Gupta, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 175–227 CrossRef .
  26. P. Hermet, M. Goffinet, J. Kreisel and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 220102 CrossRef .
  27. M. K. Singh and R. S. Katiyar, J. Appl. Phys., 2011, 109, 07D916 Search PubMed .
  28. R. Haumont, J. Kreisel, P. Bouvier and F. Hippert, Phys. Rev. B: Condens. Matter Mater. Phys., 2006, 73, 2101 Search PubMed .
  29. Y. Chu, T. Zhao, M. Cruz, Q. Zhan, P. Yang, L. Martin, M. Huijben, C. Yang, F. Zavaliche and H. Zheng, Appl. Phys. Lett., 2007, 90, 252906 CrossRef .
  30. K. Yin, M. Li, Y. Liu and C. He, Appl. Phys. Lett., 2010, 97, 042101–042103 CrossRef .
  31. G. Dong, G. Tan, Y. Luo, W. Liu, A. Xia and H. Ren, Appl. Surf. Sci., 2014, 305, 55–61 CrossRef CAS .
  32. H. Yang, M. Jain, N. A. Suvorova, H. Zhou, H. M. Luo, D. M. Feldmann, P. C. Dowden, R. F. Depaula, S. R. Foltyn and Q. X. Jia, Appl. Phys. Lett., 2007, 91, 072911–072913 CrossRef .
  33. S. M. Wu, S. A. Cybart, D. Yi, J. M. Parker, R. Ramesh and R. C. Dynes, Phys. Rev. Lett., 2013, 110, 640–647 Search PubMed .
  34. G. Catalan and J. F. Scott, Adv. Mater., 2009, 21, 2463–2485 CrossRef CAS .
  35. Q. Ke, X. Lou, Y. Wang and J. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 175–227 Search PubMed .
  36. B. Ramachandran, A. Dixit, R. Naik, G. Lawes and M. S. R. Rao, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 1558–1564 CrossRef .
  37. F. Gao, X. Y. Chen, K. B. Yin, S. Dong, Z. F. Ren, F. Yuan, T. Yu, Z. G. Zou and J. M. Liu, Adv. Mater., 2007, 19, 2889–2892 CrossRef CAS .
  38. J. B. Neaton, C. Ederer, U. V. Waghmare, N. A. Spaldin and K. M. Rabe, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 71, 4113 Search PubMed .
  39. R. G. Burns, Mineralogical applications of crystal field theory, Cambridge University Press, 1993 Search PubMed.
  40. S. Gómezsalces, F. Aguado, F. Rodríguez, R. Valiente, J. González, R. Haumont and J. Kreisel, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 98–102 Search PubMed .
  41. K. Jiang, J. J. Zhu, J. D. Wu, J. Sun, Z. G. Hu and J. H. Chu, ACS Appl. Mater. Interfaces, 2011, 3, 4844–4852 CAS .

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.